investigation of the mechanisms of aromatic amine

245
INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE-INDUCED IDIOSYNCRATIC DRUG REACTIONS by Winnie Wing Yee Ng A thesis submitted in conformity with the requirements for the degree of DOCTOR OF PHILOSOPHY Graduate Department of Pharmaceutical Sciences Faculty of Pharmacy University of Toronto © Copyright by Winnie Wing Yee Ng, 2013

Upload: vokhanh

Post on 29-Jan-2017

217 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

INVESTIGATION OF THE MECHANISMS OF

AROMATIC AMINE-INDUCED IDIOSYNCRATIC

DRUG REACTIONS

by

Winnie Wing Yee Ng

A thesis submitted in conformity with the requirements for the degree of

DOCTOR OF PHILOSOPHY

Graduate Department of Pharmaceutical Sciences

Faculty of Pharmacy

University of Toronto

© Copyright by Winnie Wing Yee Ng, 2013

Page 2: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

ii

Investigation of the Mechanisms of Aromatic Amine-Induced

Idiosyncratic Drug Reactions

Winnie Wing Yee Ng

Doctor of Philosophy

Department of Pharmaceutical Sciences

Faculty of Pharmacy

University of Toronto

2013

ABSTRACT

Drugs with the primary aromatic amine structure are notorious structural alerts in drug

development because of their association with a high incidence of idiosyncratic drug reactions

(IDRs). The toxicity of aromatic amines has been attributed to the formation of reactive

metabolites common to this functional group that can form antigenic substances and redox cycle

to induce oxidative stress. To date, there has not been a systematic study to investigate the

mechanisms of aromatic amine-induced IDRs. Therefore, in the first part of my thesis, the

danger hypothesis was tested through a global screen of hepatic gene expression in rats treated

acutely with the aromatic amine drugs sulfamethoxazole (SMX), dapsone (DDS), and

aminoglutethimide (AMG). Although one gene, Eiih, was up-regulated by all drugs tested and

could be a general biomarker to drugs that induce IDRs, few changes were found that are likely

to represent danger signals. To confirm this, a more comprehensive study was performed on the

liver with AMG after longer treatment intervals. Despite substantial changes in apoptotic and

mitochondrial pathways, there were few immune changes and no histological evidence of

damage. These findings suggest that the liver is not a direct target of aromatic amine drugs or, if

it is, the default response may be immune tolerance.

Page 3: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

iii

The presence of myeloperoxidase in neutrophils and antigen presenting cells provides a

more plausible mechanism of aromatic amine toxicity through direct activation of the immune

system, and this was tested in the second part of my thesis research. A significant increase in

activated neutrophils was observed in AMG-treated rats, which was attributed to an increase in

granulocyte progenitors in the bone marrow. This effect is opposite to the agranulocytosis that is

observed with AMG clinically, but it may have important implications to the early effects

preceding the pathology. In the spleen and lymph nodes, T-cell proliferation was found with

simultaneous indications of immune modulation through differential intracellular cytokine

expression. Moreover, down-regulation of immune activation may be why the AMG-treated rats

do not develop hypersensitivity. Further investigations into tolerogenic mechanisms may have

broader implications into why IDRs only develop in a subset of patients.

Page 4: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

iv

ACKNOWLEDGEMENTS

In addition to satisfying my scientific curiosity, my Ph.D. degree has also been a mental

test of creativity, endurance, and perseverance. Throughout this, I have been fortunate enough to

have amazing support to share both the highs and the lows. To all of you, I am forever grateful

for your patience, kindness, guidance, and unfaltering faith in me.

First and foremost I would like to thank my supervisor, Dr. Jack Uetrecht, for allowing

me the opportunity to experience the breadth of academic research in its entirety. It amazes me

how well-versed he is in so many fields, and his constant barrage of changing ideas has taught

me to think outside of the box both in science and as applied to life. I especially appreciate his

trust in my judgment, but what I admire most is his integrity and commitment to his research

and students; and I hope to carry these influences throughout my career and life. I would also

like to thank my advisory committee members Dr. Denis Grant, Peter O’Brien, and Dana

Philpott for their continued guidance and eagerness to help whenever needed.

My family is very important to me and I could not do without their support. I want to

thank my parents who taught me to always do more and expect more from life. Their constant

encouragement has become something I depend on. As stereotypical Asian parents, they had

always wanted a doctor in the family… I think they have something much better. I would also

like to acknowledge my sister and brother, Judy and Chris, who have always been so

understanding and supportive of my work.

I would like to thank my all of my lab mates both past and present who have experienced

and overcome the same hurdles in our lab. I especially want to thank Ervin Zhu, Xin Chen, Feng

Liu, and Ping Cai for their patience in teaching me, their words of wisdom, and their genuine

Page 5: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

v

friendship. Many scientists, exchange students, and interns have also passed through the lab

during my 5 years, in which I am grateful and privileged to have met and have left me with

significant influences. Lastly, I would like to thank Imir Metushi for his gracious support and

continuous belief in me, and for including me in his crazy ambitions.

Page 6: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

vi

TABLE OF CONTENTS

ABSTRACT ...................................................................................................................................ii

ACKNOWLEDGEMENTS ......................................................................................................... iv

LIST OF THESIS PUBLICATIONS .........................................................................................xii

LIST OF ABBREVIATIONS ................................................................................................... xiii

LIST OF TABLES .....................................................................................................................xvii

LIST OF FIGURES ................................................................................................................. xviii

LIST OF APPENDICES ............................................................................................................ xxi

CHAPTER 1 INTRODUCTION .................................................................................................. 1

1.1 ADVERSE DRUG REACTIONS ..................................................................................... 2

1.2 IDIOSYNCRATIC DRUG REACTIONS ....................................................................... 5

1.2.1 Clinical Characteristics of IDRs ........................................................................... 6

1.2.2 Immune Involvement in IDRs .............................................................................. 7

1.2.3 Mechanisms of IDRs .............................................................................................. 9

1.2.4 Hapten Hypothesis ............................................................................................... 11

1.2.5 Danger Hypothesis ............................................................................................... 13

1.2.6 Pharmacological Interaction with Immune Receptors (P-I) Hypothesis........ 16

1.2.7 Other Hypotheses................................................................................................. 17

1.3 ORGAN-SPECIFIC IDRs............................................................................................... 19

1.3.1 Hepatotoxicity ...................................................................................................... 19

1.3.2 Haematotoxicity ................................................................................................... 21

1.3.3 Cutaneous Toxicity .............................................................................................. 22

1.3.4 Autoimmunity ...................................................................................................... 23

1.4 IMMUNE SYSTEM ........................................................................................................ 24

1.4.1 Tests for Characterizing Immune Changes ...................................................... 26

1.5 AROMATIC AMINE DRUGS ....................................................................................... 28

Page 7: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

vii

1.5.1 Sulfamethoxazole (SMX) .................................................................................... 33

1.5.2 Dapsone (DDS) ..................................................................................................... 33

1.5.3 Aminoglutethimide (AMG) ................................................................................. 34

1.6 ANIMAL MODELS ........................................................................................................ 35

1.6.1 Nevirapine-Induced Skin Rash ........................................................................... 36

1.6.2 D-Penicillamine-Induced Autoimmunity .......................................................... 37

1.6.3 SMX-Induced Immunogenicity .......................................................................... 39

1.7 HAEMATOLOGICAL CONCEPTS RELEVANT TO IDRs ..................................... 41

1.7.1 White Blood Cells (WBCs) .................................................................................. 41

1.7.2 Neutrophils ........................................................................................................... 42

1.7.3 Neutrophil Activation .......................................................................................... 45

1.7.4 Neutrophil Regulation of Adaptive Immunity .................................................. 47

1.7.5 Agranulocytosis .................................................................................................... 48

1.7.6 Clozapine-Induced Agranulocytosis .................................................................. 49

1.7.7 AMG-Induced Agranulocytosis .......................................................................... 49

1.8 OVERVIEW OF THESIS RESEARCH ....................................................................... 51

CHAPTER 2 CHANGES IN GENE EXPRESSION INDUCED BY AROMATIC

AMINE DRUGS: TESTING THE DANGER HYPOTHESIS ........................................... 53

2.1 ABSTRACT ..................................................................................................................... 54

2.2 ABBREVIATIONS .......................................................................................................... 55

2.3 INTRODUCTION ........................................................................................................... 56

2.4 MATERIALS AND METHODS .................................................................................... 61

2.4.1 Reagents ................................................................................................................ 61

2.4.2 Animals ................................................................................................................. 61

2.4.3 Treatments............................................................................................................ 61

2.4.4 RNA Extraction ................................................................................................... 62

Page 8: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

viii

2.4.5 Microarray analysis ............................................................................................. 63

2.4.6 Real-time (RT)-PCR of early insulin-induced hepatic gene (Eiih) ................. 63

2.4.7 Activity assays ...................................................................................................... 64

2.4.8 Statistical Analysis ............................................................................................... 64

2.5 RESULTS ......................................................................................................................... 65

2.5.1 Microarray quality .............................................................................................. 65

2.5.2 SMX-induced hepatic gene expression .............................................................. 65

2.5.3 DDS-induced hepatic gene expression ............................................................... 65

2.5.4 AMG-induced hepatic gene expression ............................................................. 66

2.5.5 Aromatic amine gene expression profile............................................................ 66

2.6 DISCUSSION ................................................................................................................... 74

CHAPTER 3 HEPATIC EFFECTS OF AMINOGLUTETHIMIDE: A MODEL

AROMATIC AMINE ............................................................................................................. 80

3.1 ABSTRACT ..................................................................................................................... 81

3.2 TABLE OF CONTENTS GRAPHIC ............................................................................ 81

3.3 ABBREVIATIONS .......................................................................................................... 82

3.4 INTRODUCTION ........................................................................................................... 83

3.5 MATERIALS AND METHODS .................................................................................... 86

3.5.1 Reagents ................................................................................................................ 86

3.5.2 Animals ................................................................................................................. 86

3.5.3 Treatment ............................................................................................................. 86

3.5.4 Histology ............................................................................................................... 87

3.5.5 Liver Enzymes ...................................................................................................... 87

3.5.6 RNA Extraction ................................................................................................... 89

3.5.7 PCR Arrays .......................................................................................................... 89

3.5.8 Statistical Analysis ............................................................................................... 89

Page 9: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

ix

3.6 RESULTS ......................................................................................................................... 91

3.6.1 Effects of AMG on the Liver............................................................................... 91

3.6.2 AMG-Induced Hepatic Gene Changes .............................................................. 95

3.7 DISCUSSION ................................................................................................................. 102

CHAPTER 4 EFFECT OF AMINOGLUTETHIMIDE ON NEUTROPHILS IN

RATS: IMPLICATIONS FOR IDIOSYNCRATIC DRUG-INDUCED BLOOD

DYSCRASIAS ....................................................................................................................... 107

4.1 ABSTRACT ................................................................................................................... 108

4.2 TABLE OF CONTENTS GRAPHIC .......................................................................... 109

4.3 ABBREVIATIONS ........................................................................................................ 110

4.4 INTRODUCTION ......................................................................................................... 111

4.5 MATERIALS AND METHODS .................................................................................. 114

4.5.1 Chemicals and Reagents ................................................................................... 114

4.5.2 Antibodies ........................................................................................................... 114

4.5.3 Animals ............................................................................................................... 114

4.5.4 Treatments.......................................................................................................... 115

4.5.5 Measurement of AMG Blood Levels ................................................................ 115

4.5.6 Leukocyte Counts .............................................................................................. 115

4.5.7 Measurement of Cytokines ............................................................................... 116

4.5.8 Leukocyte Isolation from Peripheral Blood .................................................... 116

4.5.9 Measurement of New Neutrophil Release Using 5-bromo-2′-deoxyuridine

(BrdU) ................................................................................................................. 117

4.5.10 Bone Marrow Histology .................................................................................... 117

4.5.11 Characterization of Bone Marrow Cells ......................................................... 117

4.5.12 Investigation of Bone Marrow Progenitor Cells ............................................. 118

4.5.13 Determination of Neutrophil Apoptosis........................................................... 118

4.5.14 Determination of Neutrophil Activation .......................................................... 119

Page 10: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

x

4.5.15 Flow Cytometry Analysis .................................................................................. 119

4.5.16 Data Analysis ...................................................................................................... 119

4.6 RESULTS ....................................................................................................................... 120

4.6.1 Effect of AMG on Peripheral Blood Leukocyte Counts ................................ 120

4.6.2 Effect of AMG on Bone Marrow ...................................................................... 125

4.6.3 Effect of AMG on Neutrophils.......................................................................... 131

4.7 DISCUSSION ................................................................................................................. 135

CHAPTER 5 INVESTIGATION OF THE IMMUNE CHANGES INDUCED BY

AMINOGLUETHIMIDE ..................................................................................................... 140

5.1 ABSTRACT ................................................................................................................... 141

5.2 ABBREVIATIONS ........................................................................................................ 142

5.3 INTRODUCTION ......................................................................................................... 143

5.4 MATERIALS AND METHODS .................................................................................. 147

5.4.1 Chemicals and Reagents ................................................................................... 147

5.4.2 Antibodies ........................................................................................................... 147

5.4.3 Animals ............................................................................................................... 148

5.4.4 Treatments.......................................................................................................... 148

5.4.5 Measurement of Serum Cytokines ................................................................... 148

5.4.6 Detection of Cell Proliferation in Lymphoid Organs ..................................... 149

5.4.7 Phenotyping of Immune Cells........................................................................... 149

5.4.8 Characterization of an Immune Response through the Lymphocyte

Transformation Test (LTT). ............................................................................. 150

5.4.9 Flow Cytometry Analysis .................................................................................. 151

5.4.10 Data Analysis ...................................................................................................... 151

5.5 RESULTS ....................................................................................................................... 152

5.5.1 Systemic Immune Response .............................................................................. 152

5.5.2 Effect of AMG on Lymphoid Organs .............................................................. 155

Page 11: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

xi

5.5.3 AMG-Induced Immune Cell Changes ............................................................. 159

5.6 DISCUSSION ................................................................................................................. 162

CHAPTER 6 CONCLUSIONS AND FUTURE DIRECTION ............................................. 167

6.1 DISCUSSION AND CONCLUSIONS ......................................................................... 168

6.2 FUTURE DIRECTIONS .............................................................................................. 176

REFERENCES .......................................................................................................................... 178

APPENDIX 1 .............................................................................................................................. 211

APPENDIX 2 .............................................................................................................................. 215

APPENDIX 3 .............................................................................................................................. 221

APPENDIX 4 .............................................................................................................................. 223

APPENDICES REFERENCES ................................................................................................ 224

Page 12: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

xii

LIST OF THESIS PUBLICATIONS

1. Ng, W., Lobach, A. R. M., Zhu, X., Chen, X., Liu, F., Metushi, I. G., Sharma, A., Li, J., Cai,

P., Ip, J., Novalen, M., Popovic, M., Zhang, X., Tanino, T., Nakagawa, T., Li, Y., and

Uetrecht, J. Animal models of idiosyncratic drug reactions. Adv. Pharmacol., 2012. 63: p.

81-135.

2. Ng, W., and Uetrecht, J. Changes in gene expression induced by aromatic amine drugs:

Testing the Danger Hypothesis. J. Immunotox., 2013. 10: p. 178-191.

3. Ng, W., Metushi, I.G., and Uetrecht, J. Hepatic effects of aminoglutethimide: a model

aromatic amine. (Submitted to J. Immunotox., 2013).

4. Ng, W., and Uetrecht, J. Effect of aminoglutethimide on neutrophils in rats: implications for

idiosyncratic drug-induced blood dyscrasias. Chem. Res. Toxicol., 2013. 26: p. 1272-1281.

Informa Healthcare and the American Chemical Society granted permission to include my

article in this thesis.

Page 13: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

xiii

LIST OF ABBREVIATIONS

ADRs, adverse drug reactions

ALT, alanine aminotransferases

AMG, aminoglutethimide

APCs, antigen presenting cells

APC, allophycocyanin

β2M, β2-microglobulin

BrdU, 5-bromo-2’-deoxyuridine

CFUs, colony forming units

Cxcl1, chemokine (C-X-C) 1

P450, cytochrome P450

DAMPs, damage-associated molecular patterns

DAPI, 4', 6-diamidino-2-phenylindole dihydrochloride

DDS, dapsone

DMSO, dimethyl sulfoxide

DRESS, drug reaction with eosinophilia and systemic symptoms

Dusp1, dual specificity phosphatase 1

Eiih, early insulin-induced hepatic gene

FACS, florescence-activated cell sorting

FBS, fetal bovine serum

Page 14: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

xiv

FDR, false discovery rate

FITC, fluorescein isothiocyanate

G-CSF, granulocyte-colony stimulating factor

GLDH, glutamate dehydrogenase

GSH, glutathione

GSH-S-Tr, glutathione-S-Transferase

H&E, hematoxylin and eosin

HIV, human immunodeficiency virus

HLA, human leukocyte antigen

IDILI, idiosyncratic drug-induced liver injury

IDRs, idiosyncratic drug reactions

IL, interleukin

IMDM, Isocove’s Modified Dulbecco’s Medium

LPS, lipopolysaccharide

LTT, lymphocyte transformation test

MHC, major histocompatibility complex

MPO, myeloperoxidase

NAT, N-acetylation

NK, natural killer

NOD-like receptors, nucleotide-binding oligomerization domain receptors

Page 15: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

xv

PABA, para-aminobenzoic acid

PAMPs, pathogen-associated molecular patterns

PBMCs, peripheral blood mononuclear cells

PBS, phosphate buffered saline

PE, phycoerythrin

PerCP, peridinin chlorophyll

P-I hypothesis, pharmacological interaction with immune receptors hypothesis

PMA, phorbol myristate acetate

PRRs, pattern recognition receptors

RAGE, receptors for advanced glycosylation end products

RBCs, red blood cells

SDH, sorbitol dehydrogenase

SEM, standard error of the mean

Sgk, serum/glucocorticoid-regulated kinase 1

SMX, sulfamethoxazole

SMX-OH, hydroxyl metabolite of sulfamethoxazole

SMX-NO, nitroso metabolite of sulfamethoxazole

SOD2, superoxide dismutase 2

TLRs, Toll-like receptors

TNF-α, tumor necrosis factor alpha

Page 16: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

xvi

Txnrd, thioredoxin reductase

WBCs, white blood cells

Page 17: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

xvii

LIST OF TABLES

Table 1.1. Drugs associated with ADRs under restricted use and/or patient monitoring programs

in the US from 1990-2002 . ........................................................................................................... 3

Table 1.2. Drugs associated with IDRs . ....................................................................................... 7

Table 1.3. Potential danger signals .............................................................................................. 16

Table 1.4. Comparison of the characteristics between the innate and adaptive immunity. ......... 24

Table 1.5. Aromatic amine drugs and associated IDRs. .............................................................. 30

Table 1.6. Pharmacokinetic parameters of the aromatic amine drugs SMX, DDS, and AMG . . 32

Table 1.7. WBC composition in normal rats and humans . ......................................................... 42

Table 1.8. Components of neutrophil granules ........................................................................... 46

Table 2.1. Similar genes differentially regulated at least 1.4-fold change in the same direction in

all aromatic amine drugs tested 12 hr after treatment.................................................................. 67

Table 2.2. Keap1-Nrf2-ARE-regulated genes that were changed ≥ 1.4 fold by at least one of the

aromatic amine drugs (SMX, DDS, or AMG) tested 12 hr after treatment................................. 71

Table 3.1. Hepatic changes in gene expression induced by AMG in the apoptosis pathway.. ... 95

Table 3.2. Mitochondrial gene expression changes in the liver of AMG-treated rats. ................ 99

Table 3.3. AMG-induced hepatic gene changes in chemokines and receptors ......................... 100

Table 3.4. Hepatic gene changes induced by AMG in various immune pathways ................... 101

Table A1.1. SMX-induced hepatic gene changes at 12 hr……………………………………. 211

Table A1.2 DDS-induced hepatic gene changes at 12 hr…………………………………….. 212

Table A1.3. AMG-induced hepatic gene changes at 12 hr ……………………………………213

Page 18: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

xviii

LIST OF FIGURES

Figure 1.1. Mechanistic scheme for the hapten and danger hypotheses. ..................................... 11

Figure 2.2. Pharmacological interaction of drugs with immune receptors (PI) hypothesis. ....... 17

Figure 1.3. Scheme for the general metabolism of primary aromatic amine drugs..................... 30

Figure 1.4. Chemical structures of the aromatic amine drugs SMX, DDS, and AMG. .............. 32

Figure 1.5. Overview of haematopoiesis in the bone marrow focusing on granulocyte

progenitors. .................................................................................................................................. 44

Figure 1.6. Mechanistic scheme for the production of oxidants by neutrophils......................... 45

Figure 2.1. The Danger Hypothesis as applied to IDRs... ........................................................... 57

Figure 2.2. Metabolic scheme for the formation of reactive metabolites of aromatic amine drugs

..................................................................................................................................................... 58

Figure 2.3. Structures of the aromatic amine drugs used in this study and some associated IDRs

..................................................................................................................................................... 60

Figure 2.4. Eiih expression induced by aromatic amine drugs 12 hr after treatment.. ................ 68

Figure 2.5. Time-course of hepatic Eiih expression induced by AMG. ...................................... 69

Figure 2.6. Hepatic gene expression of Eiih induced by several other drugs associated with

IDRs. .......................................................................................................................................... 70

Figure 2.7. Hepatic Txnrd activity after treatment with aromatic amine drugs. ......................... 72

Figure 2.8. Time-course of hepatic GSH-S-Tr activity in AMG-treated rats .............................. 73

Figure 3.1. Histological changes in the liver of AMG-treated rats. ............................................ 92

Figure 3.2. Comparison of hepatocyte proliferation in control and AMG-treated animals……..93

Figure 3.3. AMG-induced changes in serum levels of liver enzymes……………..……………94

Page 19: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

xix

Figure 3.4. Pathway analysis of significantly changed hepatic genes in the apoptotic pathway in

AMG-treated rats……………………………………………………………………………..…97

Figure 4.1. Structure of AMG and the formation of its reactive metabolites. ........................... 113

Figure 4.2. Blood levels of AMG in rats. .................................................................................. 121

Figure 4.3. Acute AMG-induced changes on peripheral blood neutrophils in rats given an oral

dose of 80 mg/kg/day.......................................................................................................................

Figure 4.4. Acute AMG-induced changes on peripheral blood neutrophils in rats given an oral

dose of 125 mg/kg/day............................................................................................................... 122

Figure 4.5. Changes in peripheral leukocyte cell counts induced by treatment of rats with AMG

at a dose of 125 mg/kg/day for 14 days. .................................................................................... 123

Figure 4.6. AMG-induced changes in neutrophil-associated cytokines in the blood of rats. .... 124

Figure 4.7. Effect of AMG on new neutrophil release from the bone marrow. ........................ 126

Figure 4.8. Bone marrow changes induced by AMG. ............................................................... 127

Figure 4.9. AMG-induced changes in bone marrow cells. ........................................................ 128

Figure 4.10. Effect of AMG on bone marrow granulocyte and macrophage CFUs .................. 130

Figure 4.11. AMG-induced changes in neutrophil apoptosis .................................................... 132

Figure 4.12. Changes in neutrophil CD62L expression induced by AMG. .............................. 133

Figure 5.1. Early changes in serum levels of Grow/KC, MCP-1, and IP-10 induced by AMG

treatment.. .................................................................................................................................. 153

Figure 5.2. Later changes in serum cytokines induced by AMG treatment. ............................. 154

Figure 5.3. Cell proliferation detected by Ki-67 staining induced in the white pulp of the spleen

after 14 days of AMG treatment. ............................................................................................... 156

Figure 5.4. Analysis of proliferating (Ki-67+) splenocytes after 14 days of AMG treatment. .. 157

Page 20: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

xx

Figure 5.5. LTT using lymphocytes from the lymph nodes of AMG-treated rats. .................... 158

Figure 5.6. Early changes in the fraction of CD4+ T cells from the spleen and lymph nodes that

express IL-17 or IL-10 induced by AMG treatment.. ................................................................ 160

Figure 5.7. AMG-induced changes in macrophage subsets in lymph nodes after 14 days of

treatment .................................................................................................................................... 161

Figure A2.1. Metabolism of AMG by rat and mouse neutrophils.…………………………….218

Figure A2.2. Mass fragmentation of (A) AMG and (B) product of m/z 249.0.……………….219

Figure A2.3. Potential structures for m/z 249 hydroxylation of AMG…………………..……220

Figure A3.1. Peripheral blood changes induced by SMX and DDS. ………………………....221

Figure A4.1. Changes in peripheral blood WBCs induced by AMG in mice.………………...223

Page 21: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

xxi

LIST OF APPENDICES

Appendix 1: Supplementary material from Chapter 2: Changes in Gene Expression Induced by

Aromatic Amine Drugs… …………………………………………………………………...211

Appendix 2: In Vitro Metabolism of AMG by Neutrophils to a Hydroxylated Metabolite…215

Appendix 3: Effect of SMX and DDS on Peripheral Blood Neutrophils……………………221

Appendix 4: Haematological Effects of AMG in Mice……………………………………...223

Page 22: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

1

CHAPTER 1

INTRODUCTION

Page 23: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

2

1.1 ADVERSE DRUG REACTIONS

Adverse drug reactions (ADRs) are defined as an appreciably harmful or unpleasant

reaction, resulting from an intervention related to the use of a medicinal product, which predicts

hazard from future administration and warrants prevention or specific treatment, or alteration of

the dosage regimen, or withdrawl of the product [1]. ADRs are a significant issue of concern

because approximately 6% of hospital admissions are related to ADRs [2], and most ADRs go

unreported. The most commonly implicated drugs are non-steroidal anti-inflammatory drugs,

warfarin, diuretics, and ACE inhibitors; however; over half of the drugs on the market are

associated with serious ADRs, which may not be detected prior to market release [3]. ADRs

were ranked the fourth leading cause of death; trailing behind only heart disease, cancer, and

stroke; with an estimated 106 000 deaths from ADRs in 1994 in the United States alone [4].

Therefore, it is essential to better understand these reactions so that it is easier to prevent or treat

them.

In addition to the risk to patients, ADRs can lead to the attrition of prime drug candidates

resulting in significant financial losses and they play an important role in the risk of drug

development. In 2000, the price of developing a new drug was estimated to be over $400 million

US [5], although there is much debate and higher numbers have been projected. Thus

determining the safety of a drug as early as possible during the development pipeline is crucial to

decreasing the financial risks to a pharmaceutical company, and the development of preclinical

tests to detect ADRs prior to market release would have profound implications for drug

development. Approximately 3% of newly approved drugs were withdrawn from the market

from 1975-2000 due to safety issues [6], although some have been given restricted use rather

than withdrawn (Table 1.1).

Page 24: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

3

Table 1.1. Drugs associated with ADRs under restricted use and/or patient monitoring

programs in the US from 1990-2002 [7].

Pharmacovigilance programs have been helpful in monitoring ADRs in the general population;

however, there are still significant limitations on the current ability to predict and detect ADRs

and to determine the offending drug culprit. This may, in part, be due to the varying clinical

characteristics of ADRs and the fact that many ADRs are not related to the pharmacology of the

drug and can be dependent on other factors such as underlying host pathologies or drug-drug

interactions. Thus, ADRs have been commonly classified based on their characteristics as

follows [adapted from 8]:

Type A (Augmented): Type A reactions represent 80% of ADRs and are usually dose-

related and due to an augmentation of the pharmacological effect

of the drug. These reactions are typically predictable and recovery

usually occurs upon drug withdrawl. Examples include

Page 25: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

4

hemorrhaging with warfarin toxicity; hypotension with anti-

hypertensives.

Type B (Bizarre): Type B reactions are also known as idiosyncratic drug reactions.

These reactions are uncommon and unpredictable. They do not

have a simple dose-response nor are they related to the

pharmacology of the drug. Examples are sulfonamide

hypersensitivity; halothane-induced hepatitis.

Type C (Chemical): Type C reactions are usually due to, and can be predicted by, the

chemical structure of the drug. Examples include penicillin

hypersensitivity and acetaminophen hepatotoxicity.

Type D (Delayed): Type D reactions are delayed upon initiating drug therapy and can

occur as late as years after initiating the drug. Examples are

teratogenesis with thalidomide; tumorigenesis with carcinogens.

Type E (End-of-Treatment): Type E reactions are usually due to (sudden) withdrawl of

treatment. Examples include adrenal insufficiency with cortisol

withdrawl; seizures upon anticonvulsant withdrawl.

No drug is free from the risk of ADRs. Even if the drug is considered very safe, there is

usually some level of risk of developing ADRs. Since the majority of ADRs are Type A and can

be predicted, careful monitoring by both patient and medical practitioners can usually mitigate or

manage ADRs such that effective drugs with a high risk of ADRs can still have a positive

Page 26: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

5

risk/benefit ratio. However, the more significant concerns are the ADRs that are unpredictable;

with these reactions often there is little prior warning before their development. Thus, Type B or

idiosyncratic drug reactions are the main focus of my thesis research.

1.2 IDIOSYNCRATIC DRUG REACTIONS

Idiosyncratic drug reactions (IDRs) are often referred to as hypersensitivity reactions;

however, the definition of IDRs that we use is an adverse reaction that occur uncommonly in the

population at doses that are therapeutically achievable, and do not involve the therapeutic

mechanisms of the drug [9]. The incidence of IDRs is quite low, only representing 6-10% of all

ADRs; yet, they are important to study because many drugs are able to induce IDRs, and a large

number of people take these drugs. These reactions can also be quite severe and even fatal with

the annual economic burden for hospital admissions due to IDRs in the United States estimated

to be between $275-600 million [10], and this does not include outpatient costs. Contrary to the

common perception, these reactions are dose dependent because IDRs are more likely to develop

with drugs that are given at high doses. In general, drugs given below 10 mg per day rarely lead

to the development of IDRs [11], and patients that are slowly titrated to a full dose can often be

desensitized to these reactions [12]. Moreover, there is always a dose at which no IDR will

occur, and the dose-response curve for IDRs is typically different from that of the curve for the

therapeutic effect [13], suggesting that the mechanism of IDRs differs from the therapeutic

effects of the drug.

Unlike Type A reactions, the major concern for IDRs is the inability to effectively predict

or detect which drugs will induce IDRs. Preclinical safety studies have not been effective at

Page 27: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

6

detecting IDRs [14], and these reactions typically only surface in the general population after

market release. Thus the low incidence of IDRs hinders our ability to systematically study these

reactions and poses significant issues for the development of new drug candidates and the ability

to effectively treat patients. Despite the potential to develop serious reactions, many drugs

associated with IDRs are still very pharmacologically effective for the majority of the

population. Thus if factors could be identified that predispose certain individuals to IDRs, this

would have profound implications for the prediction and prevention of these reactions in both

new drug candidates and patient populations.

1.2.1 Clinical Characteristics of IDRs

The clinical manifestations of IDRs vary quite substantially, even between different

patients that take the same drug. In fact, one drug can induce a range of different IDRs, even

simultaneously, while one type of reaction can be induced by many different drugs (Table 1.2).

The varying clinical characteristics suggest that there are differences in the mechanisms that lead

to the initiation of IDRs. The most common organ systems affected by IDRs include the liver,

bone marrow, and skin. In general, there is a delay between initiating the drug therapy and the

onset of IDRs. Typically skin rashes occur the earliest, about 1-2 weeks after starting the drug

and can range from mild to severe cutaneous reactions such as Stevens-Johnson syndrome and

toxic epidermal necrolysis. Liver toxicity and blood dyscrasias usually take more time to develop

and typically occur approximately after 1-2 months after initiating drug treatment, and

autoimmune reactions can occur more than 1 year after starting drug treatment. Although the

delay in onset is common with most IDRs, there are rare exceptions in which the onset is quite

short such as in cases of telithromycin-induced hepatitis that can develop within a few days of

treatment [15]. Interestingly, when patients are re-challenged with drug, IDRs usually, but not

Page 28: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

7

always, occur with a quicker onset. Drug reaction with eosinophilia and systemic symptoms

(DRESS) is one type of IDR. Underlying host factors such as sex, age, weight, and disease state

are also important risks factors that predispose individuals to IDRs, but are unlikely to

substantially increase the risk to IDRs [9].

Table 1.2. Drugs associated with IDRs [16].

1.2.2 Immune Involvement in IDRs

The majority of the clinical characteristics displayed by patients with IDRs suggest the

involvement of the immune system. Antigen recognition by T-cells in a primary response, and

subsequent clonal expansion requires time; this could account for the delayed onset of IDRs.

Moreover, a memory T-cell response could explain why the onset of reaction upon re-challenge

is quicker. Other common features of IDRs include fever, rash, and eosinophilia, and these

symptoms are typically taken as evidence of activation of an immune response. Perhaps the most

profound evidence of an immune response is the presence of anti-drug antibodies in the serum of

Page 29: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

8

IDR patients. Antibody production requires immune activation through T- and B-cell dependent

processes to produce plasma cells to secrete drug-specific immunoglobulins. Drugs can also

induce the formation of auto-antibodies, as in the case of tienilic acid-induced hepatitis where

antibodies were found against the cytochrome P450 (P450), CYP2C9 [17]. However, despite

activation of the immune system, the presence of antibodies may or may not necessarily be

pathogenic, as was found with halothane hepatitis [18]. Furthermore, it is likely that these

biochemical changes may occur in the majority of patients, and are precursors to IDRs in which

there are some other additional factors that make only certain individuals more susceptible.

Genetic associations with IDRs have also been investigated and the most convincing of

these findings suggest the involvement of immune mechanisms. An increased odds of

developing diclofenac hepatotoxicity was observed in individuals with variant alleles for the

cytokines interleukin (IL)- 10 and IL-4 [19]. Genetic differences in human leukocyte antigen

(HLA) have also been associated with the risk that an individual will develop an IDR to certain

drugs. All patients with immunologically confirmed cases of abacavir hypersensitivity reactions

were positive for the HLA-B*5701 allele [20]. Moreover, the HLA-B*5801 allele was found in

all patients who developed severe cutaneous adverse reactions to allopurinol; however, this

seemed to be restricted to the Han Chinese, and HLA-B*5801 was observed in 15% of

allopurinol tolerant patients as well [21]. This suggests that although HLA associations can be

one determinant of whether an individual is at risk, there are likely other factors that are required

to initiate an IDR. Interestingly, polymorphisms in the T-cell repertoire have also been

associated with IDRs because the VB-11-ISGSY clonotype was found in carbamazepine-specific

CD8+ T-cells from 84% of patients with carbamazepine-induced Stevens-Johnson syndrome and

toxic epidermal necrolysis but not found in tolerant patients [22], and this, taken together with

Page 30: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

9

HLA associations, may be strong determinants to whether individuals are predisposed to IDRs.

Thus, there is extensive clinical evidence to support the involvement of the immune system, and

most IDRs are proposed to be due to the initiation of an adaptive immune response [23].

1.2.3 Mechanisms of IDRs

In addition to immune-mediated mechanisms, metabolic idiosyncrasy has also been

proposed as a mechanism for the initiation of IDRs. Polymorphisms in drug metabolizing

enzymes could influence the disposition of a drug, specifically how much and to what chemical

species the body is exposed. A slow acetylator phenotype is a risk factor for the development of

severe reactions to sulfonamides [24], suggesting a shift from drug clearance to oxidation to

form a reactive metabolite. Similarly polymorphisms in P450 have also been implicated because

individuals with the genotype CYP2E1 c1/c1 were at a higher risk of isoniazid-induced

hepatotoxicity [25]. Although differences in drug metabolizing enzymes could contribute to the

development of IDRs, the increase in risk in individuals with a metabolic polymorphism is too

small to be responsible for the idiosyncratic nature of IDRs [26]. This is in contrast to the

increased risks observed with some HLA associations such as the odds ratio of 960 associated

with the HLA-B*5701 allele in abacavir hypersensitivity [27]. Thus, there is currently no IDR

that can be fully explained by differences in drug metabolizing enzymes [28].

In contrast, the differences in the clinical manifestations of IDRs can be rationalized by

the diversity of the immune response and variations in the antigens that the T-cell repertoire

recognizes. Immune mechanisms of IDRs also provide an explaination as to why most

individuals do not get IDRs and why some IDRs do not occur rapidly upon re-challenge. Most

patients taking a drug associated with idiosyncratic drug-induced liver injury will often

experience a mild and transient increase in transaminases; however, these levels will typically

Page 31: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

10

return to normal and transaminases levels will only continue to increase in a few susceptible

individuals [29]. This suggests that in the majority of the population the default response is

adaptation or tolerance, and IDRs may result only when this tolerance is overcome. This can be

further illustrated with cases of aminoglutethimide-induced agranulocytosis that have resolved

on their own without discontinuing the drug [30]. Furthermore, some IDRs have a long time to

onset upon re-challenge, and this has been taken to be a lack of immune memory response.

However, this does not necessarily mean that an IDR is not immune-mediated. A lag in response

upon re-challenge is commonly observed in idiosyncratic autoimmune diseases, which in its self-

reactive nature may inhibit the anamnestic response by deleting immune memory cells, and this

may be typical of the drugs D-penicillamine and propylthiouracil that are associated with

idiosyncratic autoimmunity [28]. In light of the large body of clinical evidence to support the

involvement of the immune system, several immune mechanisms have been proposed for IDRs

including the hapten, danger, and pharmacological interaction of drugs with immune receptors

hypotheses.

Page 32: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

11

Figure 1.1. Mechanistic scheme for the hapten and danger hypotheses.

1.2.4 Hapten Hypothesis

The theory behind the hapten hypothesis was first introduced by Landsteiner and Jacobs

in 1935 in studies of animal sensitization to simple chemicals [31]. The hapten hypothesis

proposes that small molecules less than 1000 Da are not immunogenic on their own but must

form stable conjugates with endogenous macromolecules in order to be recognized by the

immune system as “foreign” and elicit an immunogenic response. This concept has been

commonly illustrated with studies on the penicillin family of drugs. Penicillin is chemically

reactive in the absence of metabolism due to the ring strain in its beta-lactam ring, and it reacts

with lysine groups on proteins to form antigenic substances. Penicillin-protein conjugates were

found to induce skin rashes in guinea pigs and rabbits and also led to the production of drug-

Page 33: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

12

specific antibodies [32]. In terms of IDRs, many drugs are not chemically reactive as the parent

molecule, but rather they are prohaptens that require metabolic bioactivation in vivo to gain

reactivity and covalently bind to proteins to become antigenic and induce an immune response.

This concept has been widely adapted in the pharmaceutical industry, and reactive metabolite

screens are generally routine tests in new drug development. Moreover, covalent binding of

drugs and/or reactive metabolites have been used to predict the potential of a drug to develop

IDRs, and some pharmaceutical companies have utilized this as a standard preclinical test for

new drug candidates. Typically the dose is correlated to the amount of covalent binding with an

in vivo threshold limit of 50 pmol drug equivalent/mg total liver protein [33]. However, there are

some exceptions in which drugs that form reactive metabolites are not associated with a

significant risk of IDRs. An example is olanzapine, which forms a reactive nitrenium ion but

does not lead to a high incidence of IDRs [34], presumably because the daily dose is low. Other

drugs such as ximelagatran and allopurinol are associated with IDRs but do not appear to form

reactive metabolites [9]. The magnitude of the binding may also not necessarily correlate to the

severity of the reaction [35]. In the case of metabolism to a very reactive metabolite that is a

suicide inhibitor; covalent binding less likely to cause toxicity because binding is limited to the

enzyme that formed the reactive metabolite. Covalent binding of the reactive metabolite of

tienilic acid was believed to be exclusively to P450 [36], and thus was thought to be relatively

harmless. However, tienilic acid was able to induce more changes in gene expression that were

representative of cell stress than sulfamethoxazole, which is a primary aromatic amine that can

be very metabolically reactive [37]. Moreover, tienilic acid was later found to bind to many

hepatic proteins [38], and the functionality of the protein to which the drug modifies may be an

important determinant of whether an IDR occurs. The localization of the covalently-modified

Page 34: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

13

protein and whether it is organ specific may also factor into which system the IDR is targeted

[13]. In addition, the type of immune response induced by a drug may also depend on what

portion of the drug-modified protein the T-cell receptor recognizes. For example, an adaptive

immune response may arise from recognition of the drug portion of the adduct, whereas

autoimmunity may result from recognition of the protein portion [28]. Thus, reactive metabolites

and covalent binding are very important factors to consider in the mechanism of IDRs.

1.2.5 Danger Hypothesis

The Danger Hypothesis was first proposed by Matzinger in 1994 in response to

discrepancies in the field of immunology as to what and how an immune response is initiated

[39]. At that time, the self-nonself theory was generally accepted in which the immune system

responded only to substances that were “foreign”. These antigenic substances would be detected

by antigen presenting cells (APCs) and undergo uptake and processing before being presented to

T-cells; this represents signal 1. In addition to this, costimulatory molecules such as CD80 and

CD86 needed to be expressed on the surface of APCs to interact with receptors on the surface of

the T-cell in order to initiate an immune response, and this is known as signal 2. Janeway had

proposed that the expression of costimulatory molecules was regulated by pattern recognition

receptors (PRRs) on APCs that were activated by pathogen-associated molecular patterns

(PAMPs) in what is known as the infectious-nonself model that dictates immunity to pathogens

[40]. Examples of PRRs include the Toll-like receptors (TLRs), receptors for advanced

glycosylation end products (RAGE), and nucleotide-binding oligomerization domain receptors

(NOD-like receptors). However, it did not fully explain some discrepancies such as why some

autologous transplants are rejected, nor did it explain the effectiveness of non-bacterial adjuvants

such as alum [41]. Moreover, immunity is rarely directed against new proteins produced during

Page 35: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

14

natural processes such as puberty, pregnancy, and aging, to which the body had never been

exposed. Thus, Matzinger proposed that the immune system does not react to foreign substances

but rather reacts to substances that are deemed dangerous. Similar to the infectious-nonself

model, PRRs on APCs would recognize so-called danger signals released from cells undergoing

damage or stress (known as damage-associated molecular patterns or DAMPs) and this would

lead to the upregulation of costimulatory molecules on APCs to stimulate an immune response.

In the absence of signal 2, T-helper cells would not be activated and this would lead to tolerance.

The postulates of the danger hypothesis can also be applied to the mechanisms of IDRs

[42]. Oxidative activation of drugs to reactive metabolites may induce cell stress or damage, and

this could lead to the release of danger signals and the upregulation of costimulatory molecules

on APCs to produce signal 2. In addition, covalent drug-proteins adducts could be antigenic and

act as signal 1. Moreover, the danger hypothesis is attractive because if danger signals were to be

identified for drugs associated with IDRs, they may provide a better mechanistic understanding

as to how these reactions are initiated, and could potentially be biomarkers for drugs predisposed

to causing these reactions.

However, danger signals are very vaguely defined with the only set characteristic being

their release under situations of cell stress or damage and their ability to activate APCs through

the upregulation of costimulatory molecules to initiate an immune response [43]. They can be

constitutive and released upon damage to the cellular membrane, or inducible and secreted upon

cell stress. They can also be a result of underlying host factors such as disease or invasive

procedures such as surgery. For example, human immunodeficiency virus (HIV)-infected

individuals have 10 times greater incidence of hypersensitivity reactions to co-trimoxazole than

patients who are not infected [44], and this is presumably due to a greater number of danger

Page 36: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

15

signals associated with HIV infections. Various pathological factors have also been found to

increase sulfamethoxazole-protein adducts in human APCs [45], implying that danger signals

play a role in cell activation that may lead to an increase in drug metabolism and covalent

binding. However, in general inflammation inhibits the synthesis of metabolic enzymes and

decreases drug metabolism [46]. Post-translational modifications of proteins can also be danger

signals, such has been reported with the hyperacetylation or phosphorylation of high mobility

group box proteins [47]. Although many potential danger signals have been identified (Table

1.3); only a handful of studies have confirmed their role in the induction of an immune response.

Uric acid (monosodium urate) was found to upregulate CD86 on bone marrow-derived dendritic

cells from rodents and stimulated a CD8 T-cell response in vivo [48]. Interestingly, cell death

signals have also been implicated in immune activation because APCs only expressed

costimulatory molecules in the presence of necrotic splenocytes in a rodent model of

sulfamethoxazole immunogenicity [49].

The danger hypothesis also has the potential to answer questions as to what type of

immune response is initiated. In this respect, the immune response is thought to be initiated at the

site at which the damage or stress occurs; therefore, the injured tissue is responsible for dictating

what type of response is elicited [50]. This has important implications for IDRs, because for

drugs that require metabolic activation, the site of drug metabolism may be the primary site of

immune activation and a major target of IDRs. Although not initially acknowledged, the danger

hypothesis is now widely accepted and provides a mechanistic basis from which to model our

studies on how drugs may initiate immune-mediated IDRs.

Page 37: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

16

Table 1.3. Potential danger signals. [Adapted from 42, 43, 51].

1.2.6 Pharmacological Interaction with Immune Receptors (P-I) Hypothesis

The P-I hypothesis was first introduced by Pichler based on observations in which drug-

specific T-cell clones from sulfamethoxazole hypersensitive individuals were activated by the

parent drug (in the absence of drug metabolism) and without uptake or processing by APCs [52].

Thus, the P-I hypothesis proposes that some drugs can interact directly with the major

histocompatibility complex (MHC) of APCs and the T-cell receptor in a labile, non-covalent,

MHC-restricted interaction to activate T-cells [53]. In addition, the underlying assumption of the

P-I hypothesis is that what the T-cell recognizes is what induced the initial primary response.

However, this is not always true because in a model of nevirapine-induced skin rash in which the

rash was induced by the 12-hydroxy-nevirapine metabolite; T-cells proliferated with a greater

response to the parent drug, nevirapine, even though they had never been previously exposed.

Furthermore, we know that it is actually a reactive sulfate metabolite that induces the rash, not

nevirapine [54]. Although the P-I hypothesis may be valid for some drugs such as lamotrigine, it

Page 38: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

17

is more likely due to a cross-reactivity of the T-cell receptor on memory cells rather than direct

activation by the parent drug as a super antigen, and T cells specific for both sulfamethoxazole

and its nitroso metabolite have been found in naive individuals [55].

Figure 2.2.. Pharmacological interaction of drugs with immune receptors (PI) hypothesis.

1.2.7 Other Hypotheses

In terms of an immune response, the hapten and P-I hypotheses focus on signal 1;

whereas the danger hypothesis concentrates on the importance of signal 2. Thus, the above

hypotheses are not mutually exclusive, and it is likely that more than one or all of them are

simultaneously involved with the initiation of IDRs. The hapten and danger hypotheses are the

most logical for how a drug may elicit an immune response, and thus they have been the

backbone to my thesis research. In addition, other hypotheses have been proposed such as the

inflammagen hypothesis and mitochondrial toxicity; however, although they may play a role,

they do not involve the adaptive immune system, and the likelihood for these mechanisms to be

significant determinants of an IDR is quite low.

Page 39: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

18

The inflammagen hypothesis was proposed by Roth and states that concomitant

inflammatory stress during drug therapy may be required for the development of IDRs [56]. In

studies focusing on idiosyncratic drug-induced liver injury (IDILI), co-treatment of rats with

lipopolysaccharide (LPS) as an inflammagen and with drugs such as ranitidine and trovafloxacin

have induced liver injury that is not present when drugs are administered without the

inflammatory stress [57, 58]. However, LPS is a strong immuno-stimulant, and the clinical

characteristics of this model are not representative of the IDILI that occurs in humans. In the

inflammagen model, liver injury occurs very quickly, within hours, and the injury is not

sustained. In contrast, the response to LPS is rapidly down-regulated repeated exposure and

therefore a sustained response is unlikely [59]. In addition, the inflammagen model is associated

with an infiltration of neutrophils into the liver [56], whereas the more common histology

observed in IDILI is infiltration of lymphocytes and eosinophils [60]. Some of the drugs tested,

such as ranitidine, are also quite safe; therefore, there is a high potential for false positives in the

inflammagen model. Thus, although it is likely that inflammatory stress may be a risk factor for

IDRs, the relative significance is unknown, and the inflammagen model, as it stands, is not a

valid representation of IDRs [61].

Mitochondrial injury has also been implicated in the mechanism of IDRs, specifically for

drug-induced liver toxicity [62]. Mitochondria are functionally very important to energy

production, lipid metabolism, and programmed cell death such that even subtle insults by a drug

may tip the balance towards toxicity. Mitochondrial toxicity has been studied in the context of

IDRs with heterozygous superoxide dismutase 2 deficient (SOD2+/-) mice. Superoxide

dismutase 2 (SOD2) is a mitochondrial associated enzyme involved with modulating cellular

oxidative stress, which can be quite high in the mitochondria because of the electron transport

Page 40: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

19

chain. Heterozygous mice have impaired mitochondrial function and increased mitochondrial

oxidative stress [63]. Troglitazone treatment induced liver injury in SOD2+/- mice but not in

wildtype control [64]; however, despite similar clinical characteristics to IDILI in humans, such

as midzonal hepatic necrosis [65], the changes were very mild and were not reproducible in

independent studies [66]. Furthermore, if mitochondrial injury were to play a significant role in

the initiation of IDRs an increase in dose would likely lead to an IDR. Thus, although

mitochondrial toxicity could contribute to the initiation of IDRs, it does not seem to be a

dominant mediator.

1.3 ORGAN-SPECIFIC IDRs

1.3.1 Hepatotoxicity

Liver injury is a common reaction observed with drugs that are associated with IDRs, and

IDILI is a major cause for drug withdrawl [29]. The liver is a common target presumably

because of its high content of drug metabolizing enzymes such as the P450s, which have the

potential to form reactive metabolites that may lead to the initiation of IDRs. As such, the liver is

conventionally used as the primary organ for metabolism tests during new drug development.

The main forms of liver injury are either hepatocellular or cholestatic; however, the clinical

characteristics of IDILI tend to be more hepatocellular injury rather than cholestatic injury [13].

Hepatocellular injury typically involves hepatocyte cell death, and histologically there is an

infiltration of mononuclear cells and eosinophils [60]. Cholestatic injury is typically less severe

than hepatocellular injury, and it is characterized by an accumulation of bile acids that can be

toxic to the hepatocytes and potentially lead to the release of danger signals and an immune

response [67].

Page 41: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

20

Although the time to onset of IDILI is typically 1-3 months after initiation of treatment,

there are some drugs that such as fluoroquinolones that develop liver injury after only a few days

of treatment [68]. Measurement of enzymes released from hepatocytes such as alanine

aminotransferases (ALT), sorbitol dehydrogenase (SDH), and glutamate dehydrogenase (GLDH)

have been used as biomarkers of hepatotoxicity, and elevated ALT levels are the gold standard

for hepatotoxicity [69]. Clinically, mild elevations in liver injury enzymes in patients that take

drugs that cause IDILI are common; however, the injury usually resolves on its own without

discontinuing the drug, suggesting an adaptive response, and only in a small percentage of

patients does serious IDILI occur [13].

Mechanistically, there is evidence to suggest that IDILI is caused by an adaptive immune

response [70]. Anti-drug and antinuclear antibodies are common with IDILI, and some

antibodies such as in cases of halothane hepatitis have been able to recognize liver-specific

antigens [71]. The liver is an ideal organ for immune activation because a proportion of the non-

parenchymal cells are lymphocytes [72]. In addition, innate cells are also present in the liver, e.g.

Kupffer cells, which are hepatic macrophages, as well as a large number of natural killer (NK) T-

cells [73]. There is evidence that innate immune cells can be protective and decrease liver injury

caused by directly toxic drugs such as acetaminophen [74]; specifically, depletion of Kupffer

cells increased the severity of acetaminophen-induced liver injury presumably due to a decrease

in release of regulatory cytokines and mediators [75] and the phagocytosis of dying cells which

decreases inflammation [76]. Thus, the liver is an important site for immune activation, and this

may be responsible for the hepatotoxicity associated with drugs that can induce IDRs.

Page 42: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

21

1.3.2 Haematotoxicity

Blood cells are another common target of drugs that induce IDRs, and the most common

drug-induced blood dyscrasias are agranulocytosis, aplastic anemia, and thrombocytopenia [77].

Agranulocytosis is a severe drop in the granulocyte population primarily characterized by the

depletion of neutrophils, and drugs that can induce idiosyncratic agranulocytosis include

aminopyrine, clozapine, vesnarinone, amodiaquine, and aminoglutethimide. Aplastic anemia is

the depletion of hematopoietic cells in the bone marrow, and it is usually idiopathic rather than

caused by drugs [78]. Drugs that induce aplastic anemia include felbamate and remoxipride.

Thrombocytopenia is a decrease in platelets, and drugs associated with idiosyncratic

thrombocytopenia include heparin and quinidine.

Acute and predictable haematotoxicity usually occurs with anticancer drugs, which is an

extension of their therapeutic effect of killing cancer cells, although it may take some time to

develop because the target is dividing neutrophil precursor cells and agranulocytosis is not

present until mature cells are depleted. As with other IDRs, idiosyncratic drug-induced blood

dyscrasias are typically delayed in onset. Drugs can target the bone marrow in addition to the

blood cells themselves; therefore, studies on the mechanism of blood dyscrasias have often

focused on the progenitors and cellular reserve of the bone marrow. There is also evidence to

suggest that many idiosyncratic blood dyscrasias are immune-mediated, primarily through the

formation of drug-dependent antibodies. Aminopyrine-induced agranulocytosis was found to be

mediated by antibodies [79]. Heparin and quinidine have both been shown to produce antibodies

against platelets, and this has been implicated in their ability to induce thrombocytopenia [80].

Drug metabolism has also been implicated in idiosyncratic blood dyscrasias; however, in place

of P450s in hepatocytes, formation of reactive metabolites can occur through oxidizing enzymes

Page 43: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

22

found within immune cells such as NADPH oxidase or myeloperoxidase (MPO) found within

neutrophils [81]. In activated neutrophils, clozapine was metabolized to a reactive nitrenium ion

that formed covalent adducts to cells [82], and vesnarinone was metabolized to the reactive

quinine iminium ion by activated neutrophils, but this was not observed in metabolism studies in

the liver [83]. Thus, the localized formation of reactive metabolites within immune cells may

play an important role in the development of idiosyncratic blood dyscrasias; however, other than

covalent binding, it is still not clear exactly how these reactive metabolites may elicit a response

leading to a blood dyscrasia. Because a large portion of my thesis research focuses on

idiosyncratic drug-induced agranulocytosis, a more detailed discussion on agranulocytosis will

be presented in later sections (see Section 1.7.5).

1.3.3 Cutaneous Toxicity

Many IDRs are manifested as skin rashes, presumably because the skin is a very

immunologically sensitive organ, and its immune function has been extensively reviewed [84].

Although their expression is low, many metabolizing enzymes such as P450s, flavin

monooxygenases, sulfotransferases, etc., have been found to be expressed in skin [85], and these

could potentially lead to localized bioactivation of drugs, which may play a role in the induction

of cutaneous reactions. Drug-induced skin rashes include maculopapular rashes, urticaria, fixed

drug eruptions, and acute generalized exanthematous pustulosis. Typically skin rashes manifest

fairly early, approximately 1-2 weeks after initiation of treatment. Drug reactions with

eosinophilia and systemic symptoms (DRESS) is included in this category, and it is characterized

by fever, skin rash, and some additional organ involvement such as hepatotoxicity,

haematotoxicity, lymphadenopathy, etc.; this is commonly considered to be a generalized

hypersensitivity reaction. Anticonvulsants such as carbamazepine are the most common drugs

Page 44: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

23

implicated in causing DRESS [86]. Furthermore, some of these reactions can be quite severe;

among these are Stevens-Johnson syndrome and toxic epidermal necrolysis. Toxic epidermal

necrolysis is characterized by sloughing of the epidermis and mucositis, which may lead to fatal

infections. The fatality rate of toxic epidermal necrolysis is fairly high: approximately 30% [87].

However, the mortality rate of toxic epidermal necrolysis can be decreased to 4% with early

treatment in a burn centre [88]. Toxic epidermal necrolysis appears to be involve keratinocyte

apoptosis caused by activated CD8+ T cells [87]. The etiology of Stevens-Johnson is similar to

toxic epidermal necrolysis except the extent of skin involvement is less than 10%.

1.3.4 Autoimmunity

Autoimmune-type IDRs can resemble idiopathic systemic lupus erythematosus. Typically

the onset of autoimmunity occurs quite late, usually more than a year after starting the drug, and

patients tend to develop antinuclear antibodies. Drugs can also induce organ-specific

autoimmunity such as vasculitis in the skin or autoimmune hepatitis. Drugs associated with

idiosyncratic autoimmunity include carbamazepine, procainamide, minocycline, D-

penicillamine, propylthiouracil, and isoniazid. Anti-neutrophil cytoplasmic antibodies have been

found in drug-induced lupus [89], and this suggests a potential role for metabolic activation of

drugs within leukocytes in the induction of idiosyncratic drug-induced autoimmunity.

Page 45: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

24

1.4 IMMUNE SYSTEM

The immune system is a complicated network of cells and lymphoid organs that work

together to defend the body against insults and maintain homeostasis. The immune system is

subdivided into two broad categories: the innate and adaptive immunity, and their main

characteristics are highlighted in Table 1.4.

Table 1.4. Comparison of the characteristics between the innate and adaptive immunity.

Innate immunity usually occurs quite early in response to a pathogen or injury and is

relatively non-specific. In contrast, adaptive immunity occurs later, and is antigen-specific based

on what the T-cell receptor recognizes. Although canonically the innate immune system is

activated first, APCs bridge the gap between the innate and adaptive immunity by taking up,

Page 46: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

25

processing, and presenting antigens to the adaptive immune system [90]. The clinical

characteristics of IDRs suggest that most are due to an adaptive immune response [91]. In

addition, a fine balance and appropriate regulatory mechanisms must be maintained to ensure

that the appropriate immune response is elicited when needed, and that the response is resolved

when there is no longer a threat. T cells can be activated to elicit an immunogenic response by

antigenic substances that include: endogenous macromolecules sequestered within tissues

released through processes such as cell death, covalently-modified proteins, molecular mimicry

of endogenous macromolecules by immunogenic macromolecules, and epitope spreading. To

counteract T-cell activation, tolerogenic mechanisms have been developed to ensure that the

immune response remains in check. They include: clonal deletion of B-cells that produce auto-

reactive antibodies, clonal anergy of T-cells through inhibitory costimulatory molecules such as

CTLA-4 and PD1, and also T-regulatory cells that dampen the immune response. Inappropriate

responses to antigens may lead to susceptibility to pathogenic agents; however, inadequate

induction of immune tolerance may lead to self-reactivity and autoimmunity. In terms of IDRs, if

they are immune-mediated, the default response in the majority of the population seems to be

immune tolerance, and only in some individuals is tolerance overcome, which could explain the

rare incidence of IDRs [13]. The exact mechanisms of how immune tolerance is overcome are

still unknown and are currently being investigated.

The primary role of the immune system is to defend against insults; therefore, modulation

of the response is important, and it is only when these responses are uncontrolled that a

pathological response occurs. Thus, it is quite possible that the immune system may be activated

without observed pathology if it is able to efficiently deal with the offending agent. The

interactions of the immune system are quite complex, and a precise interplay between different

Page 47: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

26

immune factors is required to maintain tight regulation of the response. Yet, if we could monitor

these changes and detect when the immune threshold is exceeded, this would be particularly

useful for predicting drug hypersensitivity.

1.4.1 Tests for Characterizing Immune Changes

The status of the immune system can be determined using various diagnostic tests. The

systemic immune response can be determined through the measurement of cytokine and

chemokine mediators in the serum; however, changes in expression levels that differ from basal

levels tend to be difficult to detect if the magnitude of the response to drug is low. In addition,

for optimal control of the immune response, the changes in these mediators may be localized to

the tissues involved. Therefore, identifying cellular changes at the level of the target organ or

lymphoid organs, such as the lymph node and spleen, may be better for detecting an immune

response. Characterizing the changes in immune cell populations can be done through

immunohistochemistry to determine the relative localization within the tissue. To quantify the

exact number of immune cell changes, florescence-activated cell sorting (FACS) analysis can be

used to detect the expression of activation markers on the cell surface or production of cytokines

and chemokines intracellularly. These methods have been employed throughout my research to

characterize the immune changes induced by aromatic amine drugs.

Determining the immune changes related to drug exposure provides an idea as to whether

there is an immune response; however, it does not determine whether the response is due to the

drug itself or if the immune is elicited in response to tissue damage. Detection of anti-drug

antibodies implies that the drug is recognized by the immune system as antigenic and is able to

initiate an immunogenic response to activate plasma cells to produce drug-specific antibodies.

These antibodies usually take a minimum of 2 weeks of drug treatment to be detected; however,

Page 48: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

27

drug-modified protein is required in order to develop an assay to detect anti-drug antibodies, and

in most cases such antibodies have not been tested for. However, some ex vivo and in vitro

assays have been developed specifically to determine whether drugs can activate an adaptive

immune response.

The lymphocyte transformation test (LTT) is a diagnostic tool used to detect drug

hypersensitivity by determining whether lymphocytes can be activated by the drug in a

secondary response that is dependent on the formation of drug-specific memory immune cells

from a prior primary immune sensitization event. This is analogous to an ex vivo re-challenge to

test the drug culprit responsible for the reaction. Peripheral blood mononuclear cells (PBMCs)

from drug hypersensitive patients are taken and re-stimulated with the drug in question, and

lymphocyte proliferation is measured by the incorporation of 3H-thymidine or 5’-bromo-2’-

deoxyuridine to test the anamnestic response as a stimulation index that compares the response

of cells to the drug and the proliferation in the absence of drug. However, the LTT is limited in

sensitivity and has been found to be more useful in identifying certain types of drug

hypersensitivity reactions such as generalized exanthema and DRESS, than other reactions such

as blood dyscrasias [92]. Thus, a negative LTT may not necessarily indicate that a specific IDR

is not caused by a specific drug. The inconsistencies can also be a result of the immune response

directed not to the parent drug, but to the reactive metabolite bound to proteins, in which case

stimulation with a drug-modified protein would be required. Another disadvantage of the LTT is

that in order to detect a response, the individual must have had previous exposure and a robust

immunogenic episode, such that, although it can detect the drug responsible, it does not have

predictive value as to whether some individuals will develop a specific drug reaction without

Page 49: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

28

prior exposure. Nevertheless, despite false negatives, if a patient has a positive LTT to a drug it

is very likely that if the patient is given the drug again they will have an IDR.

Recently, another assay has been reported to test immune system activation without the

necessity of prior drug exposure. Similar to the LTT, this method involves a cell culture drug

stimulation system; in short, naive T-cells are obtained from healthy individuals and co-

incubated with dendritic cells and the drug in question to simulate an initial priming event after

which T cell proliferation is measured. Using this system it was found that the nitroso

sulfamethoxazole metabolite (SMX-NO) was able to expand CD45RO+ memory T cells, and

these cells expressed the chemokine receptors CCR2, CCR3, and CXCR3 as well as the

cytokines IFN-γ, IL-5, and IL-13 [93]. Thus, this method may be more predictive of whether a

drug is able to elicit primary sensitization. Furthermore, this assay can be useful in predictive

testing of patients prior to treatment to determine their susceptibility instead of a confirmatory

test after the development of hypersensitivity. However, this assay is quite labour-intensive in

terms of its practicality, and more studies will have to be performed to determine its applicability

to other drugs that cause IDRs.

1.5 AROMATIC AMINE DRUGS

Drugs containing the primary aromatic amine moiety are notorious structural alerts in

drug development because they are almost always associated with a high incidence of IDRs

(Table 1.5). It has been assumed that a hypersensitivity reaction to one sulfonamide means that a

patient will have a hypersensitivity reaction to all sulfonamide drugs; however, this is a common

misconception because it appears that only sulfonamide drugs with a primary aromatic amine

Page 50: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

29

moiety are associated with IDRs, and challenge of patients with a history of a hypersensitivity

reaction to an aromatic amine sulfonamide does not predict an IDR to sulfonamides that do not

contain the arylamine moiety [94]. Examples of drugs that are sulfonamides but do not contain

the aromatic amine moiety are the non-steroidal inflammatory, celecoxib, and the diuretic,

acetazolamide.

The toxicity of aromatic amine drugs is presumably attributed to their ability to form

reactive metabolites common to the group [94]. Nitrogen contains a lone pair of electrons that

makes it easily oxidized. Oxidation produces an electron deficient or electrophilic species that

can react with endogenous macromolecules such as proteins and DNA that are nucleophilic. In

addition, nitrogen-containing molecules can undergo redox cycling, and this may lead to

oxidative stress. These chemical characteristics may predispose aromatic amines to cause IDRs.

Moreover, the majority of aromatic amine drugs contain an electron withdrawing group in the

para position to the arylamine. This decreases the electron density of the aromatic ring and

makes it more difficult to oxidize the nitrogen. In addition, the electron withdrawing group also

prevents the formation of a reactive nitrenium ion, which is responsible for the toxicity of some

drugs such as clozapine [95], and also appears to play a role in the carcinogenic potential of more

electron rich aromatic amines [96].

Page 51: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

30

Table 1.5. Aromatic amine drugs and associated IDRs.

*withdrawn from the market

Figure 1.3. Scheme for the general metabolism of primary aromatic amine drugs.

Page 52: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

31

In the general metabolic scheme (Figure 1.3), primary aromatic amines are detoxified by

N-acetylation, and low expression of N-acetyltransferases could be a risk factor for toxicity to

sulfonamide drugs [24]. Aromatic amines can also undergo oxidation by enzymes such as P450

and myeloperoxidase to form a hydroxylamine metabolite. Loss of the hydroxide can lead to the

formation of the reactive nitrenium ion; however, the presence of the electron-withdrawing

group prevents this from occurring, and the hydroxylamine is not very reactive per se. However,

the hydroxylamine can also be further oxidized to a nitrosoamine, which is electrophilic and can

undergo nucleophilic attack by endogenous protein thiols to form sulfinamide adducts [97].

Under reducing conditions such as in the presence of ascorbic acid, the hydroxylamine is quite

stable, and it can distribute throughout the body as a precursor to the reactive nitrosoamine to

elicit distal tissue toxicity. In addition, the nitrosamine metabolite can be reduced back to the

parent drug by antioxidants such as ascorbic acid and glutathione, and this could potentially lead

to redox cycling. The consequence of redox cycling is usually an increase in oxidative stress in

the cell. Drugs containing an aromatic nitro group can also be reduced back to the same

oxidation products as the products of aromatic amine oxidation [94]. Thus, the metabolic

disposition of aromatic amine drugs provides mechanisms by which IDRs may be elicited, but it

is still not exactly known how. It is hypothesized that aromatic amines induce IDRs through

immune-mediated mechanisms through the formation of antigenic substances and the production

of danger signals by the induction of oxidative cell stress. Therefore, my thesis research has

focused on investigating the mechanisms of aromatic amine-induced IDRs in the context of

understanding the underlying immune mechanisms by specifically focusing on the aromatic

amine drugs sulfamethoxazole (SMX), dapsone (DDS), and aminoglutethimide (AMG).

Page 53: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

32

Figure 1.4. Chemical structures of the aromatic amine drugs SMX, DDS, and AMG.

Table 1.6. Pharmacokinetic parameters of the aromatic amine drugs SMX, DDS, and AMG

[98].

Page 54: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

33

1.5.1 Sulfamethoxazole (SMX)

SMX is a broad spectrum bacteriostatic sulfonamide antimicrobial used to treat infections

and as a prophylaxis for Pneumocystis pneumonia in HIV-infected individuals. It is usually given

in conjunction with trimethoprim which is marketed as co-trimoxazole under the trade names

Bactrim (Roche) and Septra (Glaxosmithkline). SMX is an analog of para-aminobenzoic acid

(PABA), and therefore its pharmacological mechanism of action is to inhibit folic acid synthesis

that is required by bacteria. It does so my mimicking PABA and inhibiting dihydropteroate

synthase. Typical IDRs associated with SMX include generalized hypersensitivity, a lupus-like

syndrome, and agranulocytosis. SMX is also one of the drugs associated with the highest

incidences of Stevens-Johnson syndrome and toxic epidermal necrolysis [99]. The incidence of

SMX hypersensitivity is approximately 3% in normal patients, whereas in HIV-infected

individuals the incidence increases ~10-fold [100]. Although their immune system is

compromised, the increase in hypersensitivity could potentially be due to an increase in danger

signals within these individuals. SMX-modified proteins have been found in humans treated with

SMX [101], and anti-SMX antibodies have also been detected [102].

1.5.2 Dapsone (DDS)

DDS was first discovered in the late 1930s as a sulfone derivative, and it is still currently

in use. DDS is structurally and functionally very similar to SMX. It is a sulfonamide analog of

PABA that blocks folate synthesis. In contrast to SMX, DDS also has some anti-inflammatory

properties [103, 104], and can also inhibit the oxidative burst of neutrophils [105]. It is also

indicated for the treatment of leprosy and dermatologic diseases [106]. In terms of IDRs, DDS

has a similar profile to SMX with the exception of a greater number of blood dyscrasias such as

hemolytic anemia, methemoglobinemia, and agranulocytosis [107].

Page 55: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

34

1.5.3 Aminoglutethimide (AMG)

AMG is an analog of the sedative hypnotic, glutethimide, and it was first approved by the

Food and Drug Administration in the 1960’s as an anticonvulsant. However, AMG had poor

efficacy as an anticonvulsant, and it was found to induce adrenal suppression [108]. Further

studies found that it inhibited two main enzymes: desmolase (CYP11) and aromatase (CYP19).

CYP11 is the limiting enzyme for the production of steroid hormones, and aromatase converts

androgens to estrogens through successive hydroxylations. However, AMG has a higher affinity

for the aromatase enzyme such that it has been used for the treatment of hormone responsive

breast and prostate cancer. AMG is usually given in conjunction with hydrocortisone to

circumvent the effects on adrenal suppression, yet this effect has also been useful for treating

Cushing’s syndrome because of its ability to inhibit cortisol production.

AMG was marketed as Cytadren (Novartis) and has since been discontinued. Despite its

effectiveness in cancer treatment, AMG is not commonly used because of its propensity to

induce ADRs, and it has been replaced by the safer 3rd

generation aromatase inhibitors,

anastrozole and letrozole. Approximately 50% of patients taking AMG experience mild ADRs

such as lethargy and ataxia [109]; however, this is thought to be due to the high daily dosages of

drug (up to 1 gram per day), and a decreased toxicity profile has been observed with a decrease

in dose with similar efficacy [110]. AMG has been reported to induce smooth endoplasmic

reticulum proliferation and induce P450s in rats [111], and an upregulation of drug metabolism

could account for increased drug clearance. Interestingly, this can affect AMG itself, and the

half-life of AMG tends to decrease during the course of AMG treatment [112]. This could affect

the amount of parent drug vs. metabolite that the body is exposed to, and it may contribute to

drug tolerance. In terms of IDRs, AMG is associated with skin rashes, cholestasis, systemic

Page 56: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

35

lupus erythematosus, and blood dyscrasias [109]. Of these reactions, agranulocytosis is the most

commonly reported, occurring in approximately 2% of treated patients [113]. Unfortunately,

other than scattered case-reports, very few studies have been performed on the mechanism of

AMG-induced IDRs. Although AMG contains an aromatic amine moiety, it is quite different

from SMX and DDS because it does not contain an electron withdrawing group; therefore, the

electron density within the aromatic ring is higher than that of SMX and dapsone, and AMG is

more easily oxidizable than the other two drugs. Thus, AMG was studied for its uniqueness to

compare drugs containing the same aromatic amine moiety, despite varying chemical structures,

to determine whether these drugs induce IDRs through similar mechanisms.

1.6 ANIMAL MODELS

IDRs are very difficult to study in humans because the incidence of a specific IDR is

typically very low, and there are many ethical issues associated with obtaining patient samples

and performing controlled experiments. Thus animal models represent a crucial tool to study the

mechanisms of IDRs in a system containing a complete immune system. However, the issue

remains that IDRs are just as idiosyncratic in animals as they are in humans, and treating animals

with a drug that may cause IDRs in humans rarely leads to an IDR in animals. The animal model

must also have similar characteristics as to what is observed in patients with an IDR. Animal

models of IDRs that are clinically different to what occurs in humans are not useful in

elucidating the mechanisms of these adverse reactions in humans. Thus development of valid

animal models is difficult. Our lab has attempted to develop various animal models, and the

majority of these attempts have been unsuccessful, even with various interventions to increase

susceptibility such as the depletion of antioxidants or attempts to stimulate the immune system

Page 57: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

36

and overcome tolerance [61]. However, there are a few good animal models that we have been

able to use to rigorously test hypotheses related to the mechanisms of IDRs, and this could

potentially lead to the discovery of biomarkers or pre-clinical diagnostics that could be used to

predict whether a drug candidate is predisposed to causing IDRs. The following is a discussion

of specific animal models: nevirapine-induced skin rash and D-penicillamine-induced

autoimmunity, which have been extensively studied in our lab and are each important to

highlighting clinical differences in the characteristics of IDRs.

1.6.1 Nevirapine-Induced Skin Rash

Nevirapine is a non-nucleoside reverse transcriptase inhibitor used to treat HIV-1

infections and as a prophylaxis in pregnant women to prevent transmission of postnatal HIV

infections. In terms of IDRs, nevirapine is associated with hepatotoxicity, but the more serious

concern is skin rashes including Stevens-Johnson syndrome and toxic epidermal necrolysis,

which can be fatal [114]. An animal model of nevirapine-induced skin rash has been developed

in our lab in the female Brown Norway rat, and it has characteristics that are similar to what is

observed in patients [115]. Clinically, the time to onset is delayed, typically 1-3 weeks [114], and

the incidence tends to be higher in females than males [116].

Female Brown Norway rats treated with nevirapine (150 mg/kg/day) developed red ears

at about 7 days, and a skin rash by day 21. The rash was characterized by an inflammatory

infiltrate of CD4 and CD8 T cells as well as macrophages [115]. This rash was not observed in

males and was strain dependent because the incidence in female Sprague Dawley rats was only

20%. Re-challenged rats developed red ears within 24 hours, and splenocytes from re-challenged

animals were able to transfer susceptibility to naive animals, which provides strong evidence that

the reaction is immune-mediated. In addition, a decrease in CD4 T-cells was able to prevent the

Page 58: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

37

skin rash in female Brown Norway rats [117], and this is similar to what is reported in humans

whereby a low CD4 T-cell count appears to be associated with a lower risk of skin rash [118].

In terms of mechanistic studies, hydroxylation of nevirapine to 12-OH-nevirapine is

required for the rash to occur in rats, and deuterium substitution at the 12 position of nevirapine

inhibited the skin rash [119]. However, 12-OH-nevirapine is not chemically reactive, and it is

sulfated to a reactive benzylic sulfate. This reactive sulfate has been shown to be responsible for

the skin rash because a topical sulfotransferase inhibitor prevented covalent binding and the rash

where it was applied [120]. Immunologically, lymphocytes from re-challenged nevirapine-

treated rats responded to nevirapine by expressing interferon gamma and IL-10 [54], and the

immunosuppressants, cyclosporine and tacrolimus, were also able to prevent the skin rash [117].

The IL-1β inflammasome pathway has been implicated in contact hypersensitivity [121], and this

pathway is implicated in the nevirapine-induced rash because the chemically reactive sulfate

conjugate induces IL-1β release by keratinocytes in vitro, and an anti-IL-1β antibody markedly

suppresses the rash. These promising results could lead to a biomarker to predict the risk that a

drug candidate would cause severe dermatologic reactions. Thus, the nevirapine-induced skin

rash in Brown Norway rats represents a valid animal model that can be used to test further

hypotheses and gain a better understanding of the metabolic pathway that leads to the rash and to

understand the biological mediators of these reactions.

1.6.2 D-Penicillamine-Induced Autoimmunity

D-penicillamine has been used for the treatment of rheumatoid arthritis and as a chelating

agent to treat Wilson’s Disease; however, it is associated with a high incidence of adverse

reactions [122], and it causes autoimmune disease in 2% of treated patients [123]. Clinical

characteristics include rash, gastrointestinal upset, stomatitis, proteinuria, leucopenia, and

Page 59: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

38

thrombocytopenia. In addition, antinuclear antibodies against DNA have been reported in

patients treated with D-penicillamine [124].

An animal model of D-penicillamine-induced autoimmunity has been developed; it is

characterized by dermatitis, granulomatous and necrotic lesions, antinuclear antibodies, immune

complexes, and immunoglobulin G deposits on the glomerular basement membrane [125]. D-

penicillamine was also able to induce liver injury in the rats. Interestingly, these responses occur

after 3 weeks of treatment and were dose-dependent. Dosages less than 20 mg/kg produced little

response, in fact it induces tolerance to higher doses, and the higher dose of 50 mg/kg did not

increase the incidence of autoimmunity. The response to D-penicillamine is also strain

dependent, with an incidence of autoimmunity in approximately 70% of Brown Norway rats, but

no effect in Lewis or Sprague-Dawley rats. Using this model, our lab has been able to investigate

many of the immune mechanisms that may play a role in the initiation of IDRs.

D-penicillamine is reactive on its own without the need for prior drug metabolism. D-

penicillamine treatment in rats induced the infiltration of macrophages into the spleen prior to the

onset of reaction, and macrophage depletion decreased the incidence of autoimmunity [126].

Furthermore, D-penicillamine was found to covalently bind to aldehyde groups on macrophages

[127], and this interaction may mediate macrophage activation because the RAW264.7

macrophage cell line upregulated the cytokines TNF-α, IL-6, and IL-23 when treated with D-

penicillamine [128]. In addition, another inflammatory cell type, the T-helper 17 cell, was found

to be significantly upregulated by D-penicillamine treatment of rats, and Th17 cells appear to be

involved in the pathogenesis of D-penicillamine-induced autoimmunity [129].

Page 60: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

39

Interestingly, the time to onset of reaction is not quicker upon re-challenge with D-

penicillamine in rats, and this characteristic has been taken to indicate a lack of an immune

mechanism. However, D-penicillamine-induced autoimmunity is obviously immune-mediated. A

lack of an anamnestic response has been observed with other drugs that can induce autoimmune

diseases, which may be due to a deletion of memory T-cells [28]. Additional evidence for an

immune response includes the transfer of tolerance from the splenocytes of a tolerant animal to a

naive animal [130], and the TLR activator poly (I:C) was able to increase the severity of the

autoimmune reaction [131]. Thus, D-penicillamine-induced autoimmunity represents a good

animal model with clinical characteristics similar to what occurs in humans, which we are able to

utilize to study the immune mechanisms of one IDR and to determine factors that modulate this

type of immune response.

1.6.3 SMX-Induced Immunogenicity

Dogs have been reported to develop sulfonamide hypersensitivity that is characterized by

fever, thrombocytopenia, hepatotoxicity, and neutropenia [132]. This has been commonly

attributed to the lack of N-acetyltransferase expression in dogs [133]. These changes are similar

to what occurs in humans who have sulfonamide hypersensitivity reactions, and therefore, this is

a good naturally occurring animal model. However, the use of dogs for experimental studies

raises significant ethical and practical issues, and the low incidence makes this model of limited

practical use. To date there are no practical animal models of aromatic amine-induced IDRs. In

light of the high incidence of SMX hypersensitivity found in patients, many studies have been

performed on the immunogenicity of SMX in rodents, and some of these findings have even

been translatable to patients.

Page 61: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

40

In terms of the metabolic disposition, SMX incubated with rat liver microsomes only lead

to covalent adduct formation in the presence of an NADPH oxidizing system, while incubation

with the hydroxyl metabolite of SMX (SMX-OH) readily induced covalent binding [134]. The

nitroso metabolite of SMX (SMX-NO) was found to covalently bind preferentially to the APCs

rather than the T-cells in the spleen, and haptenation over a certain threshold led to cell death

[135]. These results suggest that the nitroso metabolite can form an antigenic substance and a

danger signal. Treatment of rats with the SMX-NO led to splenocyte proliferation in response to

the nitroso metabolite; however, treatment with SMX alone did not induce a response, and

splenocyte proliferation was only observed in the presence of a Freund’s adjuvant and only upon

re-stimulation with SMX-NO [136], which suggests that the nitroso metabolite is responsible for

the immunogenicity rather than the parent drug. Anti-SMX antibodies were found in rats after 4

weeks of treatment with SMX-NO, and to a lesser extent with SMX-OH, but not in rats treated

with SMX alone [137]. These studies provide evidence for the immunogenicity of the SMX

metabolites rather than the parent drug.

Mechanistically, treatment of normal human epidermal keratinocytes with the SMX-OH

led to up-regulation of danger signals such as heat shock protein 70, TNF-α, and IL-1β [138].

Stimulation of dendritic cells with SMX and SMX-NO induced the expression of the

costimulatory molecule CD40, and when mice were given an anti-CD40 antibody it prevented T-

cell activation [139]. This suggests the requirement of Signal 2 for the induction of an immune

response. In addition, adoptive transfer of dendritic cells treated with SMX and SMX-NO into

naive mice induced T-cell proliferation in response to SMX- and SMX-NO-derived metabolites

[49]. Despite evidence that SMX induces an immunogenic response, rodents treated with SMX

or SMX-NO do not develop hypersensitivity reactions; however, this model of SMX-induced

Page 62: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

41

immunogenicity has been useful for understanding how IDRs may be initiated by reactive

metabolites. Perhaps what is more useful is that similar changes occur in rodents as have been

observed in humans, such as an increase in expression of CD40 on the dendritic cells of HIV-

infected individuals that are hypersensitive to SMX [139].

1.7 HAEMATOLOGICAL CONCEPTS RELEVANT TO IDRs

The following is a discussion of the haematological concepts relevant to understanding

my thesis research on the mechanisms of idiosyncratic drug-induced agranulocytosis.

1.7.1 White Blood Cells

Monitoring the changes in white blood cells (WBCs), especially in patients, can provide

an early indicator of whether a drug is more prone to haematological toxicity, and for some drugs

such as clozapine, monitoring the WBC differential has led to a decrease in drug-associated

fatalities [140]. Changes in the WBC differential can also provide insight into the response of the

bone marrow because it is the main site of haematopoiesis. The majority of the WBCs in humans

are neutrophils, whereas the lymphocytes only constitute 20-45% of the WBCs [141]. In

contrast, the relative numbers of WBCs in rodents differs from what is found in humans.

Specifically, in mice and rats, lymphocytes constitute approximately 69-90% of WBCs, whereas

neutrophils comprise only 5-30% [142]. Nevertheless, rodents contain the same types of WBCs

as found in humans, and therefore they can be used to study haematological toxicities. The WBC

differentials in humans and rats are presented in Table 1.7.

Page 63: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

42

Table 1.7. WBC composition in normal rats and humans [142, 143].

1.7.2 Neutrophils

Neutrophils or polymorphonuclear cells are typically the first immune cells to be

recruited and activated in response to an inflammatory or immune insult. As such, they play an

important role in the development and resolution of an immune response, and they have

implications as to how drugs may lead to an IDR. Morphologically, neutrophils have a

distinctive hyper-segmented/multi-lobular nucleus, and the status of neutrophils can be

determined through the number of lobes of the nucleus. For example, a shift towards an increase

in immature neutrophils with fewer lobes is often described during bouts of neutrophilia [142].

The rate of neutrophil production is fairly high at approximately 1010

cells per day [144] to

compensate for their short circulating half-life of approximately 8-10 hours in humans [145] and

11 hours in mice [146]. Neutrophils originate from the bone marrow from pluripotent stem cells

that differentiate into the colony forming unit (CFU) - granulocyte-macrophage and subsequently

Page 64: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

43

to CFU-granulocyte (Figure 1.5). Within the bone marrow the neutrophils are categorized into

the mitotic pool, which can maintain cell division and includes the myeloblasts, promyelocytes,

and myelocytes, and the post-mitotic or maturation pool consisting of the metamyelocytes,

immature band cells, and mature neutrophil cells that are released into the circulation. From the

time neutrophils enter the post-mitotic pool to their release from the bone marrow into the

peripheral blood usually takes about 7 days in humans [145]; however in mice it is faster (about

4 days) [146]. Granulocyte-colony stimulating factor (G-CSF) plays a major role in the

proliferation, maturation, release, survival, and activation of neutrophils [146]. Moreover, the

chemokine CXCL12 and its corresponding receptor CXCR4 are involved with regulating the

release of neutrophils from the bone marrow [147].

Neutrophils are a terminally differentiated cell type and are programmed to undergo

apoptosis readily. They are then scavenged through uptake by macrophages in the liver, spleen,

and bone marrow [144]. Phagocytosis of apoptotic neutrophils by macrophages leads to the

expression of anti-inflammatory mediators such as TGF-β and IL-10 [148]. In addition,

neutrophils can also release lipid mediators such as lipoxins, resolvins, and protectins that can

activate the reticuloendothelial system and down-regulate its own chemotactic and oxidative

functions to mediate the resolution of the inflammatory effects [149]. The requirement for the

meticulous control of neutrophils is presumably due to their ability to mount an intense

inflammatory response, which if left unchecked, could lead to excessive tissue damage.

Page 65: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

44

Figure 1.5. Overview of haematopoiesis in the bone marrow focusing on granulocyte

progenitors.

Page 66: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

45

1.7.3 Neutrophil Activation

When stimulated by inflammatory signals, neutrophils from the blood extravasate into

tissues through chemokine-receptor gradients for homing to the site of inflammation. Neutrophil

recruitment involves an initial capture and rolling of the neutrophil on the endothelium of the

blood vessel, and this is mediated by selectins such as CD62L and CD62P, L- and P-selectin,

respectively. However, in order to extravasate into tissues, the selectins must be shed, and

intracellular adhesion molecules such as β2-integrins must be up-regulated to facilitate firm

adhesion and transmigration from the endothelium into the tissue. The hallmark of neutrophil

activation is the respiratory burst (Figure 1.6), which can be activated by PRRs such as the TLRs

on neutrophils [150] and by cytokines such as TNF-α [151]. The respiratory burst begins with an

increase in oxygen uptake by neutrophils and the assembly of NADPH oxidase machinery on its

endosomal membrane. NADPH oxidase converts molecular oxygen to superoxide, which then is

converted to hydrogen peroxidase by superoxide dismutase. The enzyme myeloperoxidase

(MPO) then uses the hydrogen peroxide to oxidize chloride ions to hypochlorous acid, which is a

very reactive oxidant that can have microbicidal activities and can damage endogenous

macromolecules.

Figure 1.6. Mechanistic scheme for the production of oxidants by neutrophils.

Page 67: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

46

Neutrophils also have other granules that contain enzymes that are involved with immunological

defense (Table 1.8); moreover, some of these enzymes such as MPO can be involved with the

oxidation of drugs, and this may play a role in the initiation of IDRs.

Table 1.8. Components of neutrophil granules [adapted from 149].

The major enzyme within neutrophils that has been extensively studied in terms of its

ability to metabolize drugs is MPO [152]. MPO is 150 kDa heme-containing enzyme that

contains a catalytic histidine site that can form the antimicrobial oxidant hypochlorous acid.

Similar to other peroxidases, it is also prone to oxidizing compounds with a phenol or aniline

moiety [153]. Myeloperoxidase makes up 5% of the neutrophil [154] and is found within the

azurophilic granules. Drugs associated with IDRs such as SMX, DDS, carbamazepine, and

ticlopidine have been found to be oxidized by MPO of activated neutrophils [155-157]. This

Page 68: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

47

suggests that localized reactive metabolite formation within the immune cells may play a role in

the activation of an immune response and the induction of an IDR.

1.7.4 Neutrophil Regulation of Adaptive Immunity

Neutrophil function is not restricted to its role in innate defense. Recently, studies have

shown that neutrophils can significantly contribute to adaptive immunity through the expression

of an array of different cytokines including CXC- and CC-chemokines, pro- and anti-

inflammatory cytokines, immunoregulatory cytokines, colony-stimulating factors, angiogenic

factors, and TNF superfamily members [reviewed in 158]. Neutrophils can crosstalk to adaptive

cells and regulate their function. In the spleen, neutrophils were found to enhance the

immunoglobulin production of marginal zone B-cells by releasing the factors BAFF, APRIL, and

IL-21, and these phenotypically distinct neutrophils were defined as B-helper neutrophils [159].

In humans treated with endotoxin, a subset of neutrophils expressing

CD11cbright

/CD62Ldim

/CD11bbright

/CD16bright

were found to inhibit T-cell proliferation that was

dependent on Mac-1 (alpha M subunit of the integrin alpha-M beta-2) and the localized

production of hydrogen peroxide [160]. This has important implications for the induction of

immune tolerance in response to drugs that can cause IDRs. A similar group of cells, the

myeloid-derived suppressor cells have also been reported to inhibit T-cells and are potential

cancer targets because such cells appear to prevent the immune system from eliminating tumors.

These cells inhibit the immune response through mechanisms such as the production of arginase,

reactive oxygen species, inducible nitric oxide synthase, IL-10, and through the depletion of

cysteine [161]. Alternatively, neutrophil plasticity has also been described in tumour

immunology where N1 neutrophils were found to be more activated and inhibited tumour growth

through the release of proinflammatory mediators such as TNF-α and Fas. In contrast, N2

Page 69: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

48

neutrophils were more immunosuppressive and pro-tumour growth; this plasticity was dependent

on TGF-β expression [162]. Thus, the neutrophil is a dynamic cell type; which could play a very

important role in the development of immunogenic responses and should be investigated in the

context of IDRs.

1.7.5 Agranulocytosis

Agranulocytosis is a blood dyscrasia characterized by a severe depression of

granulocytes, in particular neutrophils, below 0.5x109 cells/mL, which can be very serious and

even fatal. Chemotherapeutic drugs can induce agranulocytosis; however this is thought to be

mediated by its pharmacological effect of cytotoxicity. The most common non-chemotherapeutic

drugs that have been associated with agranulocytosis include clozapine, DDS, dipyrone,

methimazole, procainamide, and propylthiouracil [163]. The incidence of fatality due to

agranulocytosis is quite high, approximately 4-16% in clozapine-induced agranulocytosis [140],

and many patients develop complications due to infection and septicemia.

The onset of drug-induced agranulocytosis usually occurs more than 1 month after

initiating the drug [163]. Mechanistically, agranulocytosis is thought to be either due to direct

drug toxicity to the granulocyte progenitors in the bone marrow or an immune-mediated

mechanism [164]. In patients treated with aminopyrine, anti-neutrophil antibodies have been

found [165]. Peripheral destruction of neutrophils could lead to a compensatory increase in

production by the bone marrow that is eventually overwhelmed leading to agranulocytosis. Since

the clinical characteristics are relevant to understanding the changes observed in my thesis

research, the following is a discussion on clozapine- and AMG-induced agranulocytosis.

Page 70: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

49

1.7.6 Clozapine-Induced Agranulocytosis

Clozapine is an atypical antipsychotic associated with agranulocytosis in 0.8% of

patients with a higher incidence in female patients and patients of increased age [166]. The onset

of agranulocytosis typically occurs after about 2 months of treatment [163], and patients’ WBC

differential must be monitored in order to be prescribed the drug [140]. An initial neutrophilia

has been reported in patients treated with clozapine, which only leads to agranulocytosis in some

patients [167]. Many patients have also been observed to have elevated expression of the

cytokines TNF-α, sIL-2R, IL-6, and G-CSF as early as after 1 week of treatment [168, 169].

Many studies have been performed to understand the metabolic disposition of clozapine

and its relationship to agranulocytosis. In vitro, neutrophil apoptosis was induced by the

nitrenium ion metabolite of clozapine [170]. Activated neutrophils metabolize clozapine to a

reactive nitrenium ion, and this metabolite covalently binds to neutrophils in patients [171].

Moreover, covalent binding was observed in the bone marrow upon treatment of rats with

clozapine [172]. Thus, although the exact mechanism is unknown, the reactive nitrenium

metabolite of clozapine seems to be involved in the induction of agranulocytosis, and localized

formation of the reactive metabolite may play an important role in the pathogenesis.

1.7.7 AMG-Induced Agranulocytosis

AMG-induced agranulocytosis has been reported in 1.9% of treated patients [113].

Unfortunately, much less is known about what occurs prior to the onset of agranulocytosis

because, in contrast to clozapine, patients are not typically monitored prior to the development of

agranulocytosis, and AMG is not commonly used in the clinic anymore because of the high risk

of ADRs. Previous case reports describe agranulocytosis as occurring within 4 weeks of starting

the AMG treatment, and the bone marrow aspirate was hypocellular; however, the cellularity of

Page 71: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

50

the bone marrow usually returns to normal upon cessation of the drug [30]. Moreover, adaptation

has been described wherein AMG-induced agranulocytosis recovered on its own without

discontinuing drug. If the reaction is immune-mediated, this could potentially be immune

tolerance. Nevertheless, our knowledge of the clinical characteristics of AMG treatment remains

limited.

Few studies have focused on the mechanisms of AMG-induced agranulocytosis. An

animal model of AMG-induced leucopenia has been described in mice treated with AMG (50

mg/kg/day) in both B6C3F1 hybrids [173] and ICI-derived mice [174]. However, this was not

successfully replicated in CD1 mice [175], and it is characterized by a decrease in the total WBC

count, not just a change in the granulocyte population. Leucopenia occurred as early as 2 weeks

and was quicker in onset upon re-challenge [174]. In contrast, leucopenia was not observed in

Wistar rats treated with AMG at a dose of 50 mg/kg/day, suggesting that there are species

differences in the effect of AMG on WBCs [173]. This effect was also dependent on the

aromatic amine structure because glutethimide (an anticonvulsant analog of AMG without the

aromatic amine) did not produce similar changes in WBCs in mice [174]. In terms of the

metabolic disposition, AMG can be metabolized to a hydroxylamine metabolite [176], and the

subsequent formation of the reactive nitroso metabolite is presumably responsible for the

toxicity. Although there are few studies that focus on the ability of neutrophils to metabolize

AMG, MPO has been found to oxidize AMG in vitro into a MPO-protein free radical [177], and

this could be involved in the mechanism of agranulocytosis.

Page 72: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

51

1.8 OVERVIEW OF THESIS RESEARCH

The objective of my thesis research has been to investigate the mechanisms of IDRs

caused by drugs containing the aromatic amine moiety, which is notorious for its ability to

induce IDRs. In doing so, the main hypothesis has been that aromatic amine drugs may induce

IDRs through immune mechanisms that involve their ability to form reactive species and

covalently bind to endogenous proteins. In addition, they are able to redox cycle which has the

potential to induce oxidative cell stress. However, because there is no valid animal model of

aromatic amine-induced IDRs, we were unable to fully test this hypothesis. Our laboratory has

made many attempts to develop models by introducing various immune interventions to see how

this may affect the immune response such as the severity of IDRs, but the majority of these

attempts have not been successful. Thus, although there is substantial indirect clinical evidence

for immune involvement, the main goal was to characterize the immune changes induced by

aromatic amine drugs, and perhaps this may lead to the discovery of biomarkers and to the

development of an animal model.

In the initial studies, the objective was to determine whether aromatic amines drugs

induced changes in the liver that could be danger signals through a screen of hepatic gene

expression. Given that most aromatic amine drugs are metabolized to common reactive

metabolites, then they should induce a similar pattern of injury and similar changes in gene

expression that could potentially be biomarkers to drugs predisposed to IDRs. To our knowledge,

this is the first systematic study of the mechanisms of IDRs induced by aromatic amine drugs in

terms of detecting similar gene expression profile changes. We also performed more detailed

studies with AMG to investigate the downstream effects of gene expression changes, in

comparison to physiological changes in the liver, to determine whether the liver is a target organ

Page 73: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

52

of aromatic amine drugs. The second half of my thesis focuses on investigating the immune

effects induced by one representative aromatic amine drug, AMG. Since most aromatic amine

drugs are also associated with agranulocytosis, the effect of AMG on neutrophils was

investigated to gain a better understanding of aromatic amine-induced agranulocytosis. Studies

on the immune changes induced by AMG were also performed to gain mechanistic insight into

how immune responses may be induced by aromatic amine drugs. Brown Norway rats were

chosen because we have experience with this species, and we have developed several good

animal models with this species. Our findings suggest that the default response to aromatic

amine drugs may be tolerance and this may explain why only a small fraction of the population

is susceptible to IDRs.

Page 74: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

53

CHAPTER 2

CHANGES IN GENE EXPRESSION INDUCED BY

AROMATIC AMINE DRUGS: TESTING THE DANGER

HYPOTHESIS

Winnie Ng and Jack Uetrecht

(Reproduced with permission from: Ng, W., and Uetrecht, J. Journal of Immunotoxicology,

2013. 10: p. 178-191. Copyright 2012 Informa Healthcare USA Inc. Informa Healthcare granted

permission to include my own article in this thesis).

Page 75: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

54

2.1 ABSTRACT

Virtually all drugs that contain a primary aromatic amine are associated with a high

incidence of idiosyncratic drug reactions (IDRs), suggesting that this functional group has

biological effects that may be used as biomarkers to predict IDR risk. Most IDRs exhibit

evidence of immune involvement and the ability of aromatic amines to form reactive metabolites

and redox cycle may be responsible for initiation of an immune response through induction of

cell stress as postulated by the Danger Hypothesis. If true, danger signals could be biomarkers of

IDR risk. Our previous attempt to test the Danger Hypothesis found that sulfamethoxazole

(SMX), the only aromatic amine tested, was also the only drug not associated with an increase of

cell stress genes in mice. To ensure that these observations were not species specific and to

determine biomarkers of IDR risk common to aromatic amines, rats were treated with SMX and

two other aromatic amine drugs, dapsone (DDS) and aminoglutethimide (AMG), and hepatic

gene expression was determined using microarrays. As in mice, SMX induced minimal gene

changes in the rat and none indicated cell stress, whereas DDS and AMG induced several

changes including up-regulation of enzymes such as aldo-keto reductase, glutathione-S-

transferase, and aldehyde dehydrogenase, which may represent danger signals. Early insulin-

induced hepatic gene (Eiih) was up-regulated by all three drugs. Some mRNA changes were

observed in the Keap-1-Nrf2-ARE pathway; however, the pattern was significantly different for

each drug. Overall, the most salient finding was that the changes in the liver were minimal even

though aromatic amines cause a high incidence of IDRs. The liver generates a large number of

reactive species; however, the ability of aromatic amines to be bioactivated by cells of the

immune system may be why they cause a high incidence of IDRs.

Page 76: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

55

2.2 ABBREVIATIONS

AMG, aminoglutethimide

2M, 2-microglobulin

P450, cytochrome P450

DDS, dapsone

Dusp1, dual specificity phosphatase 1

Eiih, early insulin-induced hepatic gene

FDR, false discovery rate

GSH-S-Tr, glutathione-S-Transferase

IDR, idiosyncratic drug reaction

SMX, sulfamethoxazole

SMX-OH, hydroxyl metabolite of sulfamethoxazole

SMX-NO, nitroso metabolite of sulfamethoxazole

Sgk, serum/glucocorticoid-regulated kinase 1

Txnrd, thioredoxin reductase

Page 77: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

56

2.3 INTRODUCTION

Drugs that contain the primary aromatic amine functional group are almost always

associated with a high incidence of idiosyncratic drug reactions (IDRs) and this is considered a

structural alert for drug development [94]. The inability to predict which drug candidates will

cause such reactions, and the fact that they are usually not detected until very late in

development or after the drug has been marketed, markedly increases the risks associated with

the development of new/novel drugs.

The exact mechanisms of IDRs remain unclear; however, many lines of evidence suggest

that most IDRs are mediated by the adaptive immune system [23]. The Danger Hypothesis has

been proposed as a mechanism for the initiation of immune-mediated IDRs [42]. This hypothesis

posits that the tissues are responsible for initiating an immune response through the release of

danger signals in response to cell stress or damage [39]. This hypothesis compliments other

hypotheses such as the hapten hypothesis because it may result in activation of antigen-

presenting cells leading to costimulatory signals, referred to as signal 2, that are required for

activation of T-helper cells (Figure 2.1). In the context of IDRs, a drug, or more likely its

reactive metabolite, could induce cell stress or damage and this might contribute to the ability of

a drug to cause IDRs.

Page 78: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

57

Figure 2.1. The Danger Hypothesis as applied to IDRs. The drug or its reactive metabolite

can form antigenic adducts with endogenous molecules that are presented to T-cells on major

histocompatibility complexes of antigen-presenting cells (Signal 1). Additionally, the drug or

reactive metabolite may induce cell stress or damage through processes such as redox cycling

and oxidative damage. This may lead to the release of danger signals to up-regulate co-

stimulatory molecules such as B7 and CD40 on antigen-presenting cells to activate T-cells

(Signal 2). Signals 1 and 2 are required concurrently to initiate an adaptive immune response and

certain drugs have the ability to induce both signals.

To date, many potential danger signals have been identified such as heat shock proteins,

high-mobility group protein B1, ATP, cytokines, and nuclear DNA [178]. However, it is difficult

to rigorously test whether the induction of danger signals is an important biomarker of IDR risk.

Danger signals vary widely in their characteristics; the only consistent feature being that they are

always endogenous molecules, either actively secreted or released due to destruction of cell

membrane integrity in the event of cell stress or damage. If the release of danger signals could be

used as a biomarker to predict the risk that a drug candidate would cause IDRs, it would have a

profound effect on the process of drug development.

Page 79: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

58

The IDRs associated with aromatic amine drugs are thought to be due to the ease with

which they form reactive metabolites (Figure 2.2). Aromatic amines can be oxidized by

cytochrome P450s (P450s) and also myeloperoxidase to hydroxylamine metabolites, which are

not very reactive, but can auto-oxidize to electrophilic nitroso metabolites that can covalently

bind endogenous proteins. The nitroso metabolite can be easily reduced back to the parent amine

with antioxidants and this redox cycling could lead to oxidative stress. Thus, aromatic amine

drugs have the potential to form antigenic substances and to cause cell stress and damage, which

could produce danger signals and initiate immune-mediated IDRs.

Figure 2.2. Metabolic scheme for the formation of reactive metabolites of aromatic amine drugs.

Oxidation of aromatic amines to hydroxylamines can occur through enzymes such as P450s or

myeloperoxidase and this is prevented by N-acetylation. Hydroxylamines are not very reactive but can

auto-oxidize to electrophilic nitroso metabolites. The nitrosoamine can react with thiol-containing

nucleophiles such as glutathione or endogenous proteins to form sulfinamide adducts. Additionally, the

nitrosoamine can be reduced back to the parent drug leading to redox cycling and oxidative stress.

Page 80: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

59

Sulfamethoxazole (SMX) is an aromatic amine antimicrobial that is widely used in

combination with trimethoprim for many indications including prophylaxis against and to treat

Pnuemocystis carinii pneumonia in HIV-infected individuals. SMX can cause a generalized

hypersensitivity reaction and is also associated with among the highest incidences of Stevens-

Johnson syndrome and toxic epidermal necrolysis, which can be very serious and even fatal

[179, 180].

Previously we tried to test the Danger Hypothesis with a study of changes in gene

expression induced by tienilic acid, which causes idiosyncratic liver injury. Its reactive

metabolite binds principally to P450 that is unlikely to cause much cell stress. However, it binds

to several other proteins and it induced several changes in gene expression in the liver that likely

represented danger signals [37]. In contrast, treatment of mice with SMX, which was meant as a

positive control because it can covalently bind to a wider variety of proteins and also induce

oxidative stress, did not lead to up-regulation of genes that were likely to represent danger

signals, and the few changes that were induced in the liver were mostly down-regulation of gene

expression [37]. It is possible that down-regulation of mRNAs represents a danger signal.

However, metabolism of SMX in mice is limited [181], and the metabolism of SMX in rats more

closely resembles its metabolism in humans [182]. Thus a global screen of hepatic gene

expression was performed to determine the effects of SMX in the rat to investigate whether it can

induce cell stress leading to the release of danger signals. In addition, two other aromatic amine

drugs associated with a high incidence of IDRs, dapsone (DDS) and aminoglutethimide (AMG;

Figure 2.3), were tested to determine whether similar expression profiles exist between aromatic

amine drugs that could potentially indicate a similar mechanism of action and be useful as

biomarkers for drugs with the aromatic amine functional group. As the major site of drug

Page 81: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

60

metabolism and target for IDRs, the liver (or hepatocytes) are extensively used for reactive

metabolite and toxicity studies during drug development. Therefore, we chose the liver in which

to study changes in gene expression induced by aromatic amines; possible downstream effects of

these gene changes were also tested.

Figure 2.3. Structures of the aromatic amine drugs used in this study and some associated IDRs.

Page 82: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

61

2.4 MATERIALS AND METHODS

2.4.1 Reagents

Sulfamethoxazole (SMX) and dapsone (DDS) were purchased from Sigma (Oakville,

Ontario). AMG was purchased from Toronto Research Chemical (North York, Ontario).

Nevirapine was obtained from Boehringer Ingelheim (Ridgefield, Connecticut). Clozapine was

provided by Novartis (Dorval, Quebec). RNeasy Mini kits were obtained from Qiagen (Missis-

sauga, Ontario). Gene chips were purchased from Affymetrix (Santa Clara, CA). Light Cycler

FastStart DNA Master SYBR Green I was purchased from Roche Applied Science (Laval,

Quebec). Primers were obtained from Integrated DNA Technologies (Coralville, IA).

2.4.2 Animals

Male Brown Norway rats (≈ 8-wk-old, 200-250 g) were purchased from Charles River

(Montreal, Quebec). Rats were housed under standard conditions (doubly housed in plastic

cages, standard rat chow, automatic watering, 12:12 hr light:dark cycle, and 22°C). Food and

water were provided ad libitum. Rats were acclimatized for one week before the start of

experiments. In other studies performed with penicillamine and nevirapine in our laboratory, the

Brown Norway rat was more susceptible to IDRs induced by these drugs (i.e., autoimmunity and

skin rash, respectively) than other strains; for this reason, this strain was selected for use in the

current study. The study protocol was approved by and performed in accordance with the

University of Toronto Faculties of Medicine and Pharmacy Animal Care Committee.

2.4.3 Treatments

The treatment protocol was similar to Pacitto et al. with some modifications [37]. Each

treatment had 4 rats per group unless otherwise stated. The drugs were administered by oral

gavage at doses of 150 mg/kg (SMX), 20 mg/kg (DDS), and 80 mg/kg (AMG). These dosages

Page 83: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

62

were chosen from previous rat studies [182-185] to mimic peak drug concentrations in the blood

of patients taking these drugs for therapy (≈ 37 µg/mL for 1 000 mg SMX [98], 1.7 µg/mL for

100 mg DDS [98], and 3 µg/mL for 250 mg AMG [186]).

Drugs were suspended in 0.5% methylcellulose due to the poor solubility of DDS and

AMG. For the microarray study, rats were euthanized by CO2 asphyxiation after 12 hr drug

treatment. Treatments for follow-up studies were performed as described previously using

several different time points, with the exception that for time points >24 hr, rats were given an

additional dose of drug every 24 hr. At necropsy, the liver was extracted using aseptic

techniques, immersed in RNAlater RNA Stabilization Reagent (Qiagen), and stored according to

manufacturer instructions at -80°C until RNA extraction. Liver tissue was also immersed in

phosphate buffered saline (PBS, pH 7.4) with protease inhibitor (Sigma) for activity assays. PBS

was used instead of lysis buffer to minimize possible interferences by the solvent with assay

matrices.

2.4.4 RNA Extraction

Liver tissue (≈20-30 mg) stored in RNAlater RNA Stabilization Reagent was homo-

genized with a rotor stator homogenizer (IKA Ultra-Turrax T25 S1, Janke & Kunel, Staufen,

Germany). Total RNA was extracted using an RNeasy Mini Kit (Qiagen) as per manufacturer

protocol. Extracted RNA was eluted in 40 µl of RNase-free water and stored at -80°C. The

quality and purity of the extracted RNA was initially measured using UV spectrophotometry

based on the A260/A280 ratio before further quality check through capillary electrophoresis using

the Agilent Bioanalyzer (Affymetrix).

Page 84: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

63

2.4.5 Microarray analysis

Affymetrix RatGene 1.0 ST chips were used for detection of changes in gene expression.

Microarray processing of the liver samples was performed at the microarray facility in The

Centre for Applied Genomics (Hospital for Sick Children, Toronto, Ontario, Canada) based on

Affymetrix protocols. For microarray data, Expression Console software was used for quality

control analysis before fold-changes were determined using Partek software. The Robust Multi-

array Analysis method was used for background correction, which included Quantile Normaliza-

tion and Median Polish summarization. Data was log-base 2 transformed and fold-change values

were obtained for treatments as compared to the control group. One-way ANOVA analysis was

performed to determine the statistical significance based on differences between the treatment

and control mean intensities for each gene. Attempts were made to obtain more accurate p-values

by performing multiple test corrections using the false discovery rate (FDR).

2.4.6 Real-time (RT)-PCR of early insulin-induced hepatic gene (Eiih)

Extracted RNA samples were converted to cDNA using the Omniscript RT Kit (Qiagen),

with oligo(dT15) primers (Roche) and RNase inhibitor (Roche), as per the manufacturer’s

protocol. Quality was checked using the spectrophotometry method as previously described. For

each sample, 2 µg of RNA was converted to cDNA in a total reaction volume of 20 µl. Using

Light Cycler FastStart DNA Master SYBR Green 1 (Roche), PCR was performed using a Light

Cycler instrument (Roche) as per the following conditions: pre-incubation at 95°C for 10 min

and amplification for 45 cycles (denaturation at 95°C for 15 sec, annealing at 60°C for 5 sec, and

elongation at 72°C for 10 sec). Primers (Integrated DNA Technologies) for Eiih were as follows:

forward primer 5’-AGCTCTCCAGCTCTGGATTCTT-3’, reverse primer 5’-CACACCCAG-

AACAGAGTCTTAC-3’. 2-Microglobulin (2M) was used as the housekeeping gene and the

Page 85: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

64

primers (Integrated DNA Technologies) were as follows: forward primer 5’-TCAGTTCCACCC-

ACCTCAGATAGA-3’, reverse primer 5’-TGTGAGCCAGGATGTAGAAAGAC-3’. Data was

analyzed using the RelQuant software and normalized to a calibrator control.

An additional two drugs, nevirapine and clozapine, were both tested in the rat for Eiih

expression using RT-PCR. The treatment protocol, tissue extraction, and PCR analysis were

similar to previously stated. Nevirapine and clozapine were administered at 150 mg/kg and 50

mg/kg, respectively, through oral gavage and samples were taken 6 and 12 hr after treatment.

2.4.7 Activity assays

Thioredoxin reductase (Txnrd) and glutathione-S-transferase (GSH-S-Tr) activity in the

liver was measured using a Txnrd assay kit from Cayman Chemical (Ann Arbour, MI) and a

GSH-S-Tr activity colorimetric assay kit from Abcam (Cambridge, MA), respectively, as per

manufacturer instructions. Assayed samples contained a total protein concentration of 4 mg/mL

in PBS with proteinase inhibitor. The levels of sensitivity of the Txnrd and GSH-S-Tr kits/assays

were 0.015 and 0.025 µmol/min/mL, respectively.

2.4.8 Statistical Analysis

Data obtained from PCR and activity assays were analyzed by ANOVA using GraphPad

Prism 5. Bonferroni post-tests were used to determine statistical significance between treatment

and control groups.

Page 86: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

65

2.5 RESULTS

2.5.1 Microarray quality

The extracted rat liver RNA demonstrated good quality as demonstrated via UV spectro-

photometry and the Agilent Bioanalyzer methods. The A260/A280 ratios of the samples were in the

range of 1.96-2.29 indicating good RNA quality. Only one sample in the SMX treatment had a

RNA Integrity Number < 7, but was still sufficient for microarray processing. Expression

Console software revealed no outlier arrays and the Spearman Rank Correlation (r2) was > 0.982,

indicating a good correlation between signal values of different gene chips. Thus, all arrays were

used for microarray data analysis.

2.5.2 SMX-induced hepatic gene expression

SMX induced very few gene changes > 2-fold expression in the rat 12 hr after treatment

(see Appendix 1, Table A1.1). When FDR was applied for multiple test corrections, none of the

changes were significant (p < 0.05); of the gene changes that occurred, none seemed to explicitly

indicate cell stress or initiation of an immune response. Interestingly, genes involved with the

acute cell response were down-regulated including dual specificity phosphatase 1 (Dusp1) and

serum/glucocorticoid-regulated kinase (Sgk).

2.5.3 DDS-induced hepatic gene expression

In contrast to SMX, a greater number of genes were up-regulated > 2-fold with DDS

treatment in the rat (see Appendix 1, Table A1.2). However, only changes in six genes were

statistically significant (p < 0.05) with the FDR applied: aldo-keto reductase 7A3 (Akr7a3, 5.38-

fold), glutathione-S-transferase Yc2 subunit (Yc2, 3.18-fold), aldehyde oxidase 1 (Aox1, 1.63-

fold), ferritin light polypeptide (Ftl1, 1.44-fold), UDP glucosyltransferase 1A3 (Ugt1a3, 1.39-

fold), and glutathione reductase (Gsr, 1.38-fold). Up-regulation of genes such as Akr7a3, Yc2,

Page 87: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

66

Ugt1a3, and Gsr could indicate cell stress due to their function in drug metabolism and

cytoprotection against oxidative stress. Again, Dusp1 and Sgk were among the most down-

regulated genes similar to what was observed with SMX.

2.5.4 AMG-induced hepatic gene expression

AMG induced the most gene changes of the aromatic amine drugs tested (see Appendix

1, Table A1.3). Among these changes, 552 genes were significantly different using the FDR. Of

these genes, quite a few were involved in antioxidant and drug metabolising pathways, including

P450 oxidoreductase (Por), aldehyde dehydrogenase 1A1 (Aldh1a1), UDP

glucuronosyltransferase 2B1 (Udpgtr2), epoxide hydrolase 1(Ephx1), aldo-keto reductase 7A3

(Akr7a3), glutathione-S-transferase Yc2 subunit (Yc2), thioredoxin reductase 1 (Txnrd1),

glutathione reductase (Gsr), and UDP glycosyltransferase 1A3 (Ugt1a3), which could indicate

that cell stress was induced. Additionally, several heat shock proteins were up-regulated such as

Hspca, Hsph1, Hspa1b, and Hspb8 that could be potential danger signals to initiate an immune

response. The pattern of decreased gene expression was also similar to the other aromatic amine

drugs, as Dusp1 and Sgk were among the most down-regulated.

2.5.5 Aromatic amine gene expression profile

Although the gene changes were different for each aromatic amine drug, in general,

aromatic amines induced greater down-regulated than up-regulated gene changes (Table 2.1).

Page 88: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

67

Table 2.1. Similar genes differentially regulated at least 1.4-fold change in the same

direction in all aromatic amine drugs tested 12 hr after treatment.

Gene Function

Fold-change

SMX DDS AMG

Eiih (hepatic protein EIIH)

Unknown function; induced early in rats upon insulin treatment; may play role in carbohydrate metabolism through association with glucose

transporter 2 [187].

2.86 4.23 12.94*

Socs2

(suppressor of cytokine signaling 2)

Inhibits growth hormone activity; role in mediating the anti-inflammatory

effects of lipoxins [188].

1.48 2.08 1.41

Dusp1

(dual specificity

phosphatase 1)

Involved with dephosphorylation of tyrosine and serine/threonine residues;

role in signalling pathways due to interaction with p38, JNK, ERK;

inducible early response to LPS, hypoxia, heat shock, oxidative stress

[189].

-3.58 -3.01 -3.61

Sgk

(serum/glucocorticoid regulated kinase)

Serine/threonine protein kinase involved with cellular signalling pathways

of metabolism, proliferation, and differentiation in response to glucocorticoids and serum; induced transiently during early response to

stress [190].

-2.69 -3.72 -5.37*

G0s2

(G0/G1 switch gene 2)

Involved with mitochondrial induction of apoptosis and growth arrest in

human fibroblasts; target of retinoic acid and peroxisome proliferator-activated receptor agonists; may play role in adipogenesis [191, 192].

-2.11 -2.03 -1.59

Jun (Jun proto-oncogene)

Activated by JNK phosphorylation; interacts with c-Fos to form AP-1 transcription factor that regulates cell proliferation and apoptosis

pathways in response to stress [193].

-1.97 -2.49 -1.51

Cyr61

(cysteine rich protein 61)

Heparin binding angiogenesis inducer, involved with endothelial adhesion

and migration and tumour growth [194].

-1.87 -2.5 -1.94

Slc25a25

(solute carrier family 25,

member 25)

Regulates ATP homeostasis as a shuttle across inner mitochondrial

membrane; involved with maintaining metabolic efficiency [195].

-1.86 -2.67 -4.04*

Acat2 (acetyl-Coenzyme A

acetyltransferase 2)

Cytosolic acetyl-Coenzyme A acetyltransferase activity; involved with synthesis of cholesteryl esters and cholesterol absorption; mainly

localized to liver and intestines; may play a role in atherosclerosis [196].

-1.63 -1.63 -2.28

Slc34a2 (solute carrier family 34,

member 2)

Sodium-phosphate transporter found in intestines, kidney, and lung; may be involved with regulating phosphate levels in intestinal epithelium [197].

-1.59 -1.82 -3.02*

Sqle

(squalene epoxidase)

Involved in rate-limiting step of cholesterol biosynthesis; oxidizes squalene

to squalene epoxide; negatively regulated by cholesterol [198].

-1.51 -1.42 -1.72

Inhbe

(inhibin beta E) Member of TGF-β family; may induce apoptosis in hepatoma cells and

down-regulated in hepatocarcinogenesis [199].

-1.44 -1.51 -2.01*

Data expressed as fold-change between treated and controls. *p < 0.05 with FDR applied (n = 4).

Page 89: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

68

Among these, hepatic protein EIIH (Eiih; aka early insulin-induced hepatic gene) was the

most highly up-regulated gene in all three drugs tested. Increased Eiih gene expression was

confirmed through RT-PCR for both DDS and AMG treatment, but not SMX (Figure 2.4).

Further time-course investigation using RT-PCR found that Eiih was expressed only acutely,

with the highest expression at 8 and 12 hr after AMG treatment (Figure 2.5). Other drugs known

to cause IDRs such as nevirapine and clozapine also increased expression of Eiih at early time

points in treated rats (Figure 2.6).

Figure 2.4. Eiih expression induced by aromatic amine drugs 12 hr after treatment. Among the

treatment groups (solid bars), AMG induced a significant increase in Eiih expression as compared to

control (open bars). DDS also induced an increase in Eiih; however, no change was observed for SMX.

Rats were given 150 mg SMX/kg, 20 mg DDS/kg, or 80 mg AMG/kg, and liver samples were taken for

RT-PCR analysis. Eiih expression was normalized to 2-microglobulin (B2M) as a housekeeping gene and

a calibrator control. RT-PCR was run in triplicate (n = 4; ***p < 0.001).

Page 90: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

69

Figure 2.5. Time-course of hepatic Eiih expression induced by AMG. Eiih was expressed the greatest

8 and 12 hr after AMG treatment (solid bars) as compared to control (open bars). Rats were treated with

80 mg AMG/kg, and liver samples were taken for RT-PCR analysis. Eiih expression was normalized to

2-microglobulin (B2M) as a housekeeping gene and a calibrator control. RT-PCR was run in triplicate (n

= 2).

Page 91: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

70

Figure 2.6. Hepatic gene expression of Eiih induced by several other drugs associated with

IDRs. Clozapine (CLZ), an atypical antipsychotic drug, induced an increase in Eiih 12 hr after

treatment of rats, and nevirapine (NVP), an anti-retroviral drug, increased Eiih expression at both 6

and 12 hr after treatment of rats. Data expressed as fold-change between treated and controls (n =

4; * p < 0.05, ** p < 0.01).

Page 92: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

71

A greater number of similarities were also observed between DDS and AMG in terms of

genes regulated by the Keap1-Nrf2-ARE pathway (Table 2.2).

Table 2.2. Keap1-Nrf2-ARE-regulated genes that were changed ≥ 1.4 fold by at least one of

the aromatic amine drugs (SMX, DDS, or AMG) tested 12 hr after treatment.

Keap-1-Nrf2-ARE-Regulated Genes Fold-change

SMX DDS AMG

Xenobiotic Metabolism

Glucuronidation:

Udpgtr2 (UDP glucuronosyltransferase 2 family, polypeptide B1) 1.12 1.28 3.59*

Ugt1a3 (UDP glycosyltransferase 1 family, polypeptide A3) 1.09 1.39* 2.06*

Ugt2b4 (UDP glycosyltransferase 2 family, polypeptide B4) -1.09 1.10 1.45

Glutathione Conjugation:

Yc2 (glutathione S-transferase Yc2 subunit) 1.01 3.19* 2.23*

Gstm5 (glutathione S-transferase, mu 5) 1.08 1.53 1.14

Gsta2 (glutathione-S-transferase, alpha type2) -1.02 1.31 1.49*

Other:

Aldh1a1 (aldehyde dehydrogenase family 1, member A1) -1.23 2.33 4.15*

Ephx1 (epoxide hydrolase 1, microsomal) 1.02 1.71 2.33*

Nqo1 (NAD(P)H dehydrogenase, quinone 1) 1.13 1.57 1.53

Regulation of Cellular Redox/Antioxidant Status:

Txnrd1 (thioredoxin reductase 1) 1.16 1.40 2.23*

Gsr (glutathione reductase) 1.08 1.38* 2.06*

Energy Production:

Me1 (malic enzyme 1) 1.09 1.42 2.76*

Heme Catabolism:

Hmox1 (heme oxygenase (decycling) 1) 1.09 1.06 1.72

Data expressed as fold-change between treated and controls. *p < 0.05 with FDR applied (n = 4).

Page 93: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

72

Further testing of downstream effects of these gene changes found that enzyme activity did not

always correlate with changes in mRNA expression. Greater Txnrd activity was found after

SMX treatment when Txnrd1 gene expression was low compared to that induced by DDS or

AMG, whereas gene expression of Txnrd1 was higher with the other two drugs, when Txnrd

activity was low (Figure 2.7a). Nevertheless, a time-course study with AMG found that Txnrd

activity was only increased at 48 hr after treatment (Figure 2.7b). GSH-S-Tr activity was also

tested, but no significant changes were observed except for a possible increase in activity 48 hr

after AMG treatment (Figure 2.8).

Figure 2.7. Hepatic Txnrd activity after treatment with aromatic amine drugs.

Thioredoxin reductase activity was increased at both 12 and 24 hr after SMX treatment (A)

but not with the other two aromatic amine drugs (n = 4; ***p < 0.001). However, a time-

course study (B) found that AMG appeared to increase thioredoxin reductase activity 48 hr

after treatment (n = 2). Rats were treated with 150 mg SMX/kg, 20 mg DDS/kg, or 80 mg

AMG/kg before liver samples were taken for activity assay. For time points > 24 hr, rats

were given an additional dose of drug every 24 hr.

Page 94: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

73

Figure 2.8. Time-course of hepatic GSH-S-Tr activity in AMG-treated rats. No significant

change was observed in GST activity after treatment with 80 mg AMG/kg (squares) in the liver

as compared to controls (open circles). For time points > 24 hr, rats were given an additional

dose of drug every 24 hr (n = 2).

Page 95: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

74

2.6 DISCUSSION

If biomarkers that predict the risk that a drug candidate will cause an unacceptable risk of

IDRs could be found it would have a profound effect on the process of drug development. The

liver is a logical place to look because it is the site of most reactive metabolite formation, and it

is the target of IDRs that are most likely to lead to drug withdrawl. Therefore, when drug

candidates are screened for reactive metabolite formation, this is done almost exclusively in the

liver or with hepatocytes. Accordingly, much of the screening for toxicities of drug candidates is

focused on the liver.

Drugs that are primary aromatic amines represent a good starting point because this

functional group is notorious for being associated with IDRs. Microarray technology represents

the easiest way to examine a wide variety of possible biomarkers. The initial goal of these

experiments was to verify whether the changes in hepatic gene expression induced by SMX were

similar to our previous study performed in mice and whether there were changes that could be

danger signals, biomarkers, or both. We chose here to focus on early timepoints because subtle

signs of cell stress, through changes in mRNA expression, would likely happen rapidly upon

drug treatment and thus be good predictors. In comparison, effects on host immune responses

would not be expected to occur - or to be detected - as early. Our findings indicate that, as in

mice, SMX induced minimal hepatic gene changes in the rat. Further, of these changes, none met

the strict criteria set here for denoting statistical significance.

It was surprising that there was a significant difference between the gene expression

profile of SMX and DDS because they are structurally very similar with similar pharmacology,

although DDS causes hemolytic anemia and methemoglobinemia and SMX does not [200];

which presumably reflects differences in the amount of redox cycling and distribution to red

Page 96: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

75

cells [201]. AMG is a first generation aromatase inhibitor used to treat estrogen-responsive

breast and prostate cancer. AMG is structurally different from the other two drugs; in particular,

there is no electron-withdrawing group on the aromatic ring to decrease the electron density of

the aromatic amine and this should change its redox potential. AMG was included to determine

the extent of the involvement of the aromatic amine group in the induction of IDRs to determine

whether the mere presence of the moiety is enough to induce similar changes. Interestingly,

AMG induced the most significant number and magnitude of gene changes. This is not

surprising because AMG is associated with a variety of adverse reactions in ≈ 50% of patients

that take this drug, although only 1% of patients experience idiosyncratic hematologic toxicities

including thrombocytopenia, leucopenia, pancytopenia, and agranulocytosis [201]. To some

degree the similarities between DDS and AMG correlate with the fact that they are more prone

to induce blood dyscrasias and bone marrow toxicity than SMX [107, 202]. Covalent adducts of

SMX hydroxylamine have been found to localize on the surface of normal human epidermal

keratinocytes in vitro, whereas covalent adducts of DDS hydroxylamine were found to localize

intracellularly [203]. These distinct covalent binding patterns may play a role in the observed

differences and types of IDRs induced because intracellular adducts may have a greater ability to

induce a strong immune response.

DDS and AMG induced some genes that could be considered danger signals. Our

previous microarray studies on drugs including tienilic acid, carbamazepine, phenytoin, and D-

penicillamine have all shown induction of cell stress through the up-regulation of genes involved

with the Keap1-Nrf2-ARE pathway [37, 204, 205], and this may potentially be a sign of danger

to the immune system. The Keap1-Nrf2-ARE pathway is responsible for regulating the

transcription of a variety of genes involved with detoxification and cytoprotection. The up-

Page 97: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

76

regulation of these genes suggests that cell stress occurred and there is an increased need to

modulate potential damage that could eventually lead to an immune response.

In the present study, the same general pattern of increased expression of Keap1-Nrf2-

ARE-regulated genes was observed, although mainly for DDS and AMG. Nevertheless, although

the changes induced by SMX were not statistically significant, the same Keap1-Nrf2-ARE-

regulated genes as DDS and AMG appeared to be slightly elevated. Txnrd is an enzyme

analogous to glutathione reductase that functions to provide reducing equivalents to thioredoxin

to act as an antioxidant, and it is crucial in maintaining cell redox status [206]. Surprisingly, it

was SMX that induced the greatest increase in Txnrd activity. Furthermore, there was a lag

period between gene induction and Txnrd activity during AMG treatment. This implies that gene

expression does not necessarily correlate with functional changes and that gene expression was

likely induced before protein expression, which may explain the inconsistent relationship

between gene and protein activity. This also raises an important issue as to whether gene changes

are meaningful on their own. The activities of certain proteins are governed by post-translational

modifications or by cytoplasmic “regulators”. In the canonical pathway, Nrf2 is kept inactivated

by Keap1 in the cytoplasm, and only when Keap1 undergoes proteolytic cleavage is Nrf2

released to migrate into the nucleus where it then binds to the antioxidant response element and

initiates transcription [207]. Thus, changes in gene expression of Nrf2 may not be useful because

activation is regulated post-translationally.

The acute cell response genes Sgk and Dusp1 were consistently the most down-regulated

with all aromatic amine drugs tested. Sgk, a serine/threonine protein kinase involved in

regulating cellular metabolism, proliferation, and differentiation, can be induced transiently by

Page 98: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

77

glucocorticoids and serum [190]. Dusp1, a nuclear phosphatase involved in regulating cellular

signaling pathways, has been reported as inducible in early response to LPS, heat shock, and

oxidative stress [189]. Down-regulation of these genes opposes what would be expected upon

drug treatment; however, decreased Sgk expression upon tienilic acid treatment was attributed to

the short half-life of Sgk mRNA [37, 190]. It is also plausible that a lack of signal could alert the

cell to stress and initiate an immune response, although the exact mechanism is unclear.

Possibly the most interesting change was the increase in Eiih mRNA that was observed

with all three drugs, although we failed to confirm increases in Eiih expression with RT-PCR for

SMX. Additional testing of two other drugs associated with IDRs: clozapine and nevirapine,

found a rapid increase in Eiih expression. These other drugs are not primary aromatic amines,

and therefore it is possible that this is a more general biomarker of IDR risk. However, without

testing a variety of drugs including negative controls that are not associated with a significant

risk of IDRs it would be premature to conclude that this represents a useful biomarker of IDR

risk. Unfortunately, little is known about the function of this gene. To date, there has only been

one published study on Eiih that found it was induced early and acutely upon insulin treatment in

rats with expression mainly localized to the liver, intestine, and islets of Langerhan, and there is

some indication for its involvement in carbohydrate metabolism [187]. Its predicted protein

arrangement suggests that it has the potential to be secreted or membrane bound, which implies

that Eiih could possibly be a danger signal if secreted or act as a receptor-signalling complex on

cellular membranes.

Although there were changes in gene expression induced by these primary aromatic

amines, especially AMG, which could represent a danger signal, the most salient finding of this

Page 99: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

78

study is that primary aromatic amines do not appear to cause many changes in gene expression in

the liver. Given that the primary aromatic amine is such a notorious structural alert and is readily

oxidized to reactive metabolites in the liver and can also redox cycle, this is surprising. SMX and

AMG frequently cause liver injury: the incidence of elevated γ-glutamyltransferase is greater

than 60% in AMG-treated patients [208]. However, SMX-induced liver injury is usually part of a

more generalized hypersensitivity IDR, and the more severe AMG-induced liver injury is more

commonly cholestatic rather than hepatocellular. The most likely explanation for these results is

that, although this functional group is readily oxidized to reactive metabolites in the liver, the

liver is well equipped to deal with reactive metabolites. For example, in the case of an animal

model of nevirapine-induced skin rash, there was more covalent binding in the liver than in the

skin, but there were many more significant changes in gene expression in the skin (>400,

unpublished observation, manuscript in preparation). It is likely that what makes the aromatic

amine functional group a structural alert is that it can also be oxidized by myeloperoxidase,

which is present in antigen presenting cells. The formation of reactive metabolites by antigen

presenting cells has been shown to lead to their activation [49], and this is likely to be a strong

stimulus for the induction of an immune response. Specifically, dendritic cells treated with SMX

and SMX-NO were found to increase expression of the co-stimulatory molecule CD40 [209].

Similarly, keratinocytes were found to up-regulate their expression of heat shock protein 70,

which could be a potential danger signal, upon treatment with the hydroxylamine metabolite of

SMX [210].

Even in the case of immune-mediated liver injury, it appears that activation of the

immune system outside of the liver is necessary in order to cause an immune response in the

liver that results in injury [211]. It is notable that all three of these drugs are associated with a

Page 100: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

79

relatively high incidence of idiosyncratic agranulocytosis; this presumably involves oxidation of

the drug to reactive metabolites by myeloperoxidase, the major oxidizing enzyme in neutrophils.

This clinical picture is most consistent with activation of the immune system outside of the liver

with the immune response sometimes extending to the liver. Therefore, the present focus on the

liver by the pharmaceutical industry may miss signals that predict IDR risk, especially for drugs

that also cause IDRs outside of the liver. In particular, bioactivation of easily oxidized drugs,

such as those containing aromatic amines, by myeloperoxidase in neutrophils and antigen-

presenting cells may play an important role in the induction of many types of IDRs.

Page 101: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

80

CHAPTER 3

HEPATIC EFFECTS OF AMINOGLUTETHIMIDE: A MODEL

AROMATIC AMINE

Winnie Ng, Imir G. Metushi, and Jack Uetrecht

(Submitted to the Journal of Immunotoxicology for publication, 2013).

Page 102: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

81

3.1 ABSTRACT

Primary aromatic amine drugs are structural alerts in drug development because of their

association with a high incidence of idiosyncratic drug reactions (IDRs). Although the

mechanisms of IDRs remain unclear, there is a large amount of evidence to suggest that most are

immune-mediated. If biomarkers could be found that predict IDR risk, it would have a major

impact on drug development. Previous attempts to do this through screening of hepatic gene

expression profiles in rodents treated with aromatic amine drugs found limited changes. Of the

drugs studied, aminoglutethimide (AMG) induced the most changes, and this led to a more

comprehensive study of its effects on the liver. Brown Norway rats treated with AMG for up to

14 days showed hepatocyte hypertrophy and hyperplasia, and only a transient elevation of

glutamate dehydrogenase was observed. The transient nature of the liver injury is similar to the

adaptation observed in many patients treated with a drug that can cause idiosyncratic liver injury.

Pathway-specific PCR arrays found few AMG-induced gene changes associated with an immune

response, and of these changes, the majority were involved with innate immunity such as Tlr2,

Ticam2, CD14, and C3. AMG treatment did lead to significant changes in the apoptosis and

mitochondrial panel of genes. Although these changes may be precursors to IDRs, it is difficult

to discern their mechanistic importance because AMG-treated rats did not exhibit any obvious

liver pathology. Moreover, tolerogenic mechanisms could prevent liver injury, and that may be

why most people do not develop IDRs.

Page 103: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

82

3.2 ABBREVIATIONS

ALT, alanine aminotransferases;

AMG, aminoglutethimide;

DAB, 3’3 diaminobenzidine

GLDH, glutamate dehydrogenase;

H&E, hematoxylin and eosin

HLA, human leukocyte antigen

HRP, horseradish peroxidase

IDRs, idiosyncratic drug reactions;

PBS, phosphate buffered saline

TUNEL, terminal deoxynucleotidyl transferase dUTP nick end labeling

Page 104: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

83

3.3 INTRODUCTION

The primary aromatic amine functional group is a notorious structural alert in drug

development because drugs with this group are associated with a high incidence of idiosyncratic

drug reactions [IDRs; 94]. Despite the potential for toxicity, many aromatic amine drugs remain

pharmacologically very effective; therefore, understanding the mechanism of IDRs or

determining biomarkers associated with these reactions could provide a tremendous benefit to

drug development. Although the types of IDRs vary from drug to drug, and one drug can cause

more than one type of IDR, the toxicity of aromatic amine drugs is presumably due to their

ability to form common reactive metabolites that could potentially induce cellular damage and

lead to an adverse drug response [212].

Immune involvement has been implicated in the mechanisms of IDRs [13]; however, our

current understanding of IDRs is still far from complete. IDRs pose significant issues to drug

development because of their lack of predictability, and the inherent risks posed to patients once

the drug is out in the market. Ideally, the ability of drug candidates to induce IDRs would be

detected early in preclinical studies to decrease the enormous financial risk of drug attrition and

to ensure drug safety to patients. However, in reality, IDRs are much more difficult to predict,

largely due to our lack of understanding of the underlying mechanisms, and as a result, although

there have been strong human leukocyte antigen (HLA) gene associations; there are currently no

reliable tests to predict IDR risk. In addition, development of valid animal models of IDRs with

clinical characteristics that closely mimic what occurs in humans has proven to be a daunting

task despite many attempts at various immune interventions [61]. Analyzing changes in gene

expression of drugs that induce IDRs has been proposed as a method of determining biomarkers

Page 105: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

84

to predict drugs predisposed to these reactions, and this has had some success in distinguishing

model hepatotoxins in vitro [213]. However, it is difficult to determine the relevancy of these

gene changes and how well these signatures would apply to predicting IDR risk for new drug

candidates.

In an attempt to understand the mechanisms underlying aromatic amine-induced IDRs,

microarray studies were previously performed in both mice and rats to look for common hepatic

gene changes induced acutely by aromatic amine drugs that could provide mechanistic clues or

act as potential biomarkers to identify drugs predisposed to causing IDRs. The rationale for

choosing early time points was that it is usually early cell injury, which may occur in most

patients and animals, that presumably stimulates an immune response and later presents as an

IDR in rare patients. It was also anticipated that the predictive value of these results would have

significant implications for the assessment of preclinical drug safety. However, in mice treated

with sulfamethoxazole, few gene changes were observed in the liver except for down-regulation

of the expression of a few genes [37]. Yet, it was uncertain whether these results were species- or

drug-specific. Additional testing in rats with three different aromatic amine drugs - including

sulfamethoxazole, dapsone, and aminoglutethimide - found few common gene changes, and

again the majority were down-regulated [212]. Among these drugs, aminoglutethimide induced

the most significant gene changes, and several were associated with the Keap-1-Nrf2-ARE

pathway, which suggests that some cellular stress occurred. This is consistent with gene changes

observed with other drugs associated with IDRs such as carbamazepine, phenytoin, tienilic acid,

and D-penicillamine [37, 204, 205]. Given that the liver is the conventional target of drug

metabolism studies because of its predominant function in drug metabolism, and aromatic

amines are readily oxidized to reactive metabolites, the paucity of changes in gene expression

Page 106: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

85

was unexpected. Furthermore, given the postulates of the danger hypothesis, the liver would be

expected to experience more adverse effects because it is assumed to be the site where most

reactive metabolites would form and induce damage.

To ensure that any aromatic amine-induced changes in the liver were not overlooked

because of experimental limitations, a more comprehensive study combining physiological tests

with gene arrays at successive time-points was performed in the liver of rats to confirm whether

aromatic amine drugs affect the liver as a target organ. The aromatase inhibitor

aminoglutethimide (AMG) was used in these studies as a prototypical aromatic amine drug

because it induced the most changes in the previous microarray study and was more likely to

result in observable effects. AMG is primarily indicated for breast and prostate cancer treatment

and is associated with a variety of IDRs including cholestasis, agranulocytosis, lupus, and skin

rashes [214]. AMG is associated with a higher incidence of IDRs than sulfamethoxazole and

dapsone, the other two aromatic amines that were studied previously, and because of this it is not

currently used clinically. This higher incidence of IDRs with AMG may be a result of a higher

electron density which would facilitate the oxidation of the aromatic amine. Thus, the effect of

AMG on the liver was tested including pathway-specific gene expression to gain a more detailed

understanding of the mechanisms of aromatic amine-induced IDRs.

Page 107: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

86

3.4 MATERIALS AND METHODS

3.4.1 Reagents

AMG was purchased from Toronto Research Chemical (North York, ON). Epitope

Unmasking Solution and Antibody Dilution Buffer were acquired from ProHisto (Columbia,

SC). 3’3 diaminobenzidine (DAB) was obtained from Vector Labs (Burlingame, CA). Mouse

anti-rat Ki-67 (clone: MIB-5), rabbit anti-mouse biotinylated IgG, and streptavidin-horseradish

peroxidase (HRP) was purchased from Dako (Burlington, ON). Mouse anti-rat CD8 (clone: OX-

8) and mouse anti-rat CD68 (clone: ED1) were obtained from Abcam (Cambridge, MA). Alanine

Aminotransferase (ALT) Liquid Stable Reagent was purchased from Thermo Scientific

(Middletown, VA). Glutamate dehydrogenase (GLDH) Assay Kit was from Randox Laboratories

Ltd. (Crumlin, UK). RNAlater RNA Stabilization Solution, Trizol Reagent, and RNase free

water were from Life Technologies (Burlington, ON). RNeasy Mini Kit was purchased from

Qiagen (Mississauga, ON). RT2 Profiler PCR Arrays and RT

2 First Strand Kit were obtained

from SABiosciences (Valencia, CA) and were supplied by Roche.

3.4.2 Animals

Male Brown Norway rats (≈ 9-wk-old, weighing 200-250 g) were purchased from

Charles River (Montreal, QC). Rats were housed at the Department of Comparative Medicine

(University of Toronto) under standard care conditions (i.e., doubly-housed, 12:12 hr light:dark

cycle, and 22°C room temperature). Standard rat chow and water were provided ad libitum. Rats

were given 1 wk to acclimatize prior to the start of experiments. The protocol was approved by

the University of Toronto Faculties of Medicine and Pharmacy Animal Care Committee.

Page 108: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

87

3.4.3 Treatment

AMG were suspended in 0.5% methylcellulose and administered by oral gavage at a dose

of 125 mg/kg/day. This dosage is an increase from that which was used in our previous study

[212] to ensure that maximal changes were induced, while maintaining the blood levels of AMG

within the range found in humans taking AMG therapeutically [≈ 4.7-32.4 µg/ml; 215]. Control

rats were given 0.5% methylcellulose vehicle only. Blood levels of AMG in rats were measured

in our previous studies and are within the range of AMG blood levels found in humans [216].

Control animals were given methylcellulose as a vehicle control. Four rats were tested for each

treatment group unless otherwise stated. Body weight and general health of rats were monitored

daily. At various time points (Day 0, 3, 7, and 14), blood samples isolated for use in the

protocols below. In addition, at Day 1, 7, and 14, subsets of rats were euthanized by CO2

asphyxiation and liver (as well as a final blood) samples isolated for use in the protocols below.

3.4.4 Histology

At necropsy, livers were excised and immediately immersed in 10% formalin solution.

H&E-stained and unstained paraffin-embedded slides were prepared at the Division of Pathology

at the Hospital for Sick Children (Toronto, ON). Immunohistochemistry was carried out using

standard procedures. Briefly, slides were de-paraffinized with xylene and rehydrated with serial

dilutions of ethanol. Heat-induced antigen retrieval was performed for 20 min (Ki-67 staining) or

incubated at 71°C overnight (for CD8 and CD68 staining) in Epitope Unmasking Solution.

Endogenous peroxidases were blocked with 3% hydrogen peroxide in phosphate-buffered saline

(PBS, pH 7.4) for 10 min. Washes were then performed using PBS-0.05% Tween-20 solution

and antibodies were diluted in Antibody Dilution Buffer. Slides were incubated with primary

antibodies against Ki-67, CD68, and CD8 at room temperature. For mouse anti-rat Ki-67

Page 109: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

88

antibody, the dilution used was 1:25 and incubation time was 30 min. For mouse anti-rat CD68

and anti-rat CD8 antibodies, the dilution used was 1:100 and incubation time was 90 min. Rabbit

anti-mouse biotinylated IgG was used as a secondary antibody (1:250 dilution) and incubated for

30-40 min at room temperature. Streptavidin-HRP (1:500) was then added and the slide

incubated 15 min after which the signal was developed with DAB solution; haematoxylin was

used as counterstain. Slides were dehydrated with ethanol and xylene before coverslips were

attached with Cytoseal. Histological images were taken at the Microscopy Imaging Lab (Faculty

of Medicine, University of Toronto) on a Zeiss fluorescence microscope with deconvolution.

Immunohistochemical grading was performed by counting the number of cells/field of view; at

least two slices of tissue (3-6 mm2) were mounted on glass slides. Five areas from each slice

were manually counted under the microscope. All slides were read in a blinded manner.

3.4.5 Liver Enzymes

Serial samples of blood were taken from the tail vein of rats to determine liver enzyme

activities in the serum. Blood was processed within 1 hr of collection, and ALT and GLDH

activities were determined using ALT Liquid Stable Reagent and a GLDH Assay Kit,

respectively, according to manufacturer protocols. For GLDH, minor modifications were made

to the protocol as previously described [217], wherein 200 µL of Reagent 1 was premixed with 8

µL of Reagent 2, and 200 µL of this mixture was loaded into each well of a 96-well plate before

serum samples were added. All reagents were reconstituted per manufacturer specifications and

absorbance values at 340 and 405 nm monitored on a SpectraMax Plus384 (Molecular Devices,

Sunnyvale, CA) plate reader for at least 5 min at 25°C.

Page 110: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

89

3.4.6 RNA Extraction

At necropsy, the livers were removed, rinsed in PBS, and immediately immersed in

RNAlater RNA Stabilization Solution to preserve RNA for extraction. Liver samples were then

homogenized with a IKA Ultra-Turrax T25 S1 rotor stator homogenizer (Janke and Kunel,

Staufen, Germany), and RNA was extracted using Trizol Reagent as per manufacturer protocols.

An additional RNA clean-up step was performed using an RNeasy Mini Kit. Ultimately,

extracted RNA was eluted in 40 µl RNase-free water and stored at -80°C. The quality and purity

of the extracted RNA was measured using UV spectrophotometry based on the A260/A280 ratio.

3.4.7 PCR Arrays

Extracted RNA was converted to cDNA using the RT2 First Strand Kit with 400 ng RNA

per sample, following manufacturer protocols. Six different panels of genes were analyzed using

the RT2 Profiler PCR Arrays, including: apoptosis (PARN-012), chemokines and receptors

(PARN-022), common cytokines (PARN-021), mitochondria (PARN-087), toll-like receptor

signaling (PARN-018), and Th17 and autoimmunity (PARN-073). However, limitations on the

number of plates prevented analysis of every gene panel at all time points (1, 7, 14 days);

specifically, only Day 14 data were collected for common cytokine, toll-like receptor signaling,

and Th17 response pathways. Array studies were performed according to manufacturer protocols

on 384-well plates that held four samples simultaneously. For each sample, cDNA converted

from 400 ng total RNA was loaded on the plate, and each plate had samples staggered between

AMG-treated and control animals. Arrays were run on an ABI7900HT real-time PCR instrument

(Applied Biosystems, Grand Island, NY) according to the SABiosciences protocols. A threshold

cycle value of 0.2 was assigned to each array plate for data normalization. Data was analyzed

using RT² Profiler PCR Array Data Analysis Template Version 3.2 (SABiosciences); lactate

Page 111: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

90

dehydrogenase A and ribosomal protein P1 were used as housekeeping genes. Pathway analyses

were performed using Gene Network Central Pro software (SABiosciences).

3.4.8 Statistical Analysis

Data were analyzed with GraphPad Prism 5 (GraphPad Software Inc., La Jolla, CA)

using a Mann-Whitney U test or two-way analysis of variance (ANOVA), depending on the

constraints of the data being analyzed. PCR array data were analyzed with RT² Profiler PCR

Array Data Analysis Template v3.2 (SABiosciences) using Student’s t-test.

Page 112: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

91

3.5 RESULTS

3.5.1 Effects of AMG on the Liver

Rats tolerated the 125 mg/kg/day AMG dose well for a period of up to 14 days without

significant changes in body weight or evidence of distress. Histological analysis of the liver of

AMG-treated rats found enlarged hepatocytes in Zone 3 as compared to control rats (Figure

3.1A, B). In the representative H&E staining of the liver, the noticeable presence of red blood

cells in sinusoids of the control is likely an artifact from lack of perfusion prior to fixation rather

than a treatment effect. An increase in the liver weight of AMG-treated rats was also found, and

this was most significant after 7 days of treatment (Figure 3.1C). Although no data on liver

weight was collected at the 14-day endpoint, the histological findings suggest that the increase in

liver weight with AMG was sustained. In addition, there appeared to be a slight increase in cell

proliferation, as determined by the proliferation marker Ki-67, in the liver of AMG rats after 7

days compared to the untreated group (Figure 3.2); however, this did not reach statistical

significance.

Examination of H&E-stained slides did not reveal any evidence of liver injury (neither

hepatocellular nor cholestatic) in AMG-treated rats after 14 days, nor were immune cell

infiltrates detected in the liver. Staining for specific immune cells such as CD8 T-lymphocytes

and CD68 macrophages in the liver found no differences using immunohistological grading

between AMG-treated and control animals (data not shown).

In serum, few changes were observed in enzymes associated with liver injury in AMG-

treated rats. A subtle but transient increase in ALT was observed with AMG treatment at Day 3

(Figure 3.3A). A significant increase in GLDH was also observed with AMG at Day 3; however,

Page 113: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

92

these changes were transient and returned back to baseline/control levels at Day 7 (Figure 3.3B).

No changes were observed in ALT or GLDH at Day 14 of AMG treatment (data not shown).

Figure 3.1. Histological changes in the liver of AMG-treated rats. Representative H&E

staining of (A) control and (B) AMG-treated (125 mg/kg/day, 14 days) rat. Evidence of Zone 3

hepatocyte hypertrophy was found in rats treated with AMG as compared to controls. 20X

magnification. (C) Rats treated with AMG had an increase in liver weight that was most

significant on Day 7 of treatment. Values shown are mean (± SEM) for each group (n = 4). *p <

0.05 compared to control (Mann Whitney U-test).

Page 114: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

93

Figure 3.2. Comparison of hepatocyte proliferation in control and AMG-treated animals.

Representative Ki-67 staining in tissue from (A) control and (B) AMG-treated (125 mg/kg/day,

14 days) rats. 20X magnification. (C) A possible increase in Ki-67 was observed at Day 7 of

AMG treatment. Though some evidence of hepatocyte proliferation was found in the AMG-

treated rats, the main effect was hypertrophy rather than hyperplasia. Open bars = control rats;

solid bars = AMG-treated rats. Values shown are mean (± SEM) number of Ki-67-positively-

stained nuclei per 10X field of view (n=4). No results were statistically significant (two-way

ANOVA).

Page 115: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

94

Figure 3.3. AMG-induced changes in serum levels of liver enzymes. (A) ALT and (B) GLDH

in rats treated with AMG (125 mg/kg/day) compared to control rat values. Open bars = control

group; solid bars = AMG-treated group. Values shown are mean (± SEM) (n=4). **p < 0.01 as

compared to control (two-way ANOVA).

Page 116: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

95

3.5.2 AMG-Induced Hepatic Gene Changes

Hepatic gene expression was determined at various time points of AMG treatment using

PCR arrays that were pathway-specific. The greatest number of AMG-induced differential

changes in gene expression was primarily found within the apoptotic panel of genes, but both

pro- and anti-apoptotic regulatory pathways were induced (Table 3.1). Although the majority of

changed genes were involved with the induction of apoptosis, significant up-regulation of anti-

apoptotic genes was also quite notable including Aven, Bag1, Bcl2l1, Birc3, and Nol3. At earlier

time points, a general down-regulation of apoptotic genes was observed, whereas up-regulation

of genes was mainly observed at the latter 14-day endpoint. A pathway analysis of the up-

regulated genes induced by AMG did not reveal any particular dominant signaling pathway,

although most genes were associated with Nfkb1 (Figure 3.4). ). In terms of downstream effects

of the gene changes, TUNEL staining for apoptotic cells in the liver did not show any differences

between AMG-treated and control rats (data not shown). This may be due to the activation of

genes in opposing apoptosis pathways, which may act to modulate apoptotic events.

Page 117: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

96

Table 3.1. Hepatic changes in gene expression induced by AMG in the apoptosis pathway.

Gene

Symbol

Gene Name Day 1 Day 7 Day 14

Aven Apoptosis, caspase activation inhibitor -1.16 -1.05 2.39*

Bad BCL2-associated agonist of cell death -2.01* -1.29 2.53

Bag1 BCL2-associated athanogene -1.10 1.13 2.43*

Bax Bcl2-associated X protein -1.25 1.03 2.28*

Bcl2l1 Bcl2-like 1 -2.95 -2.07 2.39*

Bcl2l2 Bcl2-like 2 -2.24* -1.09 1.14

Bid BH3 interacting domain death agonist -1.08 1.21 2.65*

Hrk Harakiri, BCL2 interacting protein (contains only BH3 domain) -3.26* -1.71 -1.27

Birc3 Baculoviral IAP repeat-containing 3 -2.03 1.28 3.42*

Birc5 Baculoviral IAP repeat-containing 5 1.12 2.35* -1.52

Casp1 Caspase 1 -1.13 -1.22 -3.71*

Casp7 Caspase 7 1.31 1.88* 6.51*

Casp9 Caspase 9, apoptosis-related cysteine peptidase -1.94 -1.26 2.67*

Cidea Cell death-inducing DFFA-like effector a 1.99 1.90* 5.85*

Dapk1 Death associated protein kinase 1 -3.46* -3.99 2.68

Dffa DNA fragmentation factor, alpha subunit -2.13* 1.16 1.37

Fadd Fas (TNFRSF6)-associated via death domain 1.03 1.09 2.62*

Ltbr Lymphotoxin beta receptor (TNFR superfamily, member 3) -1.06 1.12 5.55*

Nfkb1 Nuclear factor of kappa light polypeptide gene enhancer in B-cells 1 -2.17 -1.44 2.21*

Nol3 Nucleolar protein 3 (apoptosis repressor with CARD domain) 7.72* 11.61* 7.18*

Prlr Prolactin receptor -3.81* 1.91 3.15

Prok2 Prokineticin 2 1.11 3.36* -1.33

Tnfrsf1a Tumor necrosis factor receptor superfamily, member 1a 1.40 1.49 2.89*

Cd40 CD40 molecule, TNF receptor superfamily member 5 -1.13 1.08 -2.02*

Fas Fas (TNF receptor superfamily, member 6) -1.73 2.32* -1.07

Tnfsf10 Tumor necrosis factor (ligand) superfamily, member 10 -3.09* -1.12 1.10

Tnfsf12 Tumor necrosis factor ligand superfamily member 12 -1.58 -1.37 4.50*

Tradd TNFRSF1A-associated via death domain -1.29 -1.28 2.61*

Values are expressed as fold-change in AMG-treated compared to control rats (n = 4). *p < 0.05

compared to control.

Page 118: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

97

Figure 3.4. Pathway analysis of significantly changed hepatic genes in the apoptotic

pathway in AMG-treated rats. Only genes with changes >2-fold (up-regulation) after 14 days

of AMG treatment were included in analysis for tends in the apoptotic panel of genes. Arrows

indicate known associations and do not represent how genes are regulated by AMG. Green = up-

regulation; red = down-regulation; yellow = physical interaction. Analysis was performed using

Gene Network Central Pro software (SABiosciences).

Page 119: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

98

Many mitochondria-associated genes were also significantly changed upon AMG

treatment (Table 3.2). Again, down-regulation of genes was primarily observed at the earlier

time points whereas up-regulation was observed after 14 days of AMG treatment. Of these

AMG-induced genes, many were associated with mitochondrial transportation such as the solute

carrier family 25 (Slc25) and translocase (Tomm34, Tomm40, Tspo, Timm17b) genes.

Mitochondria-associated apoptotic genes such as Aifm2, Bbc3, and Bid were also significantly

up-regulated with AMG treatment; the function of these genes suggested induction of apoptosis

via the intrinsic mitochondrial pathway.

In terms of changes in immune-associated genes, AMG induced fewer significant

changes than that observed for the other two pathways, and most of these changes were more

suggestive of changes in innate immunity. In the chemokine and receptor pathways there were

significant increases in C3, Ccl19, and Nfkb1 (Table 3.3); in contrast, a sustained down-

regulation was observed for Cxcl13 upon AMG treatment. Increases in the toll-like receptor Tlr2

and the adaptor molecule Ticam2 genes were also observed with an increase in Cd14 and

changes in the IL-1 pathway (Il1f5, IL1rn) upon AMG treatment (Table 3.4). Up-regulation of

Cd4 was the predominant change associated with adaptive immunity.

Page 120: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

99

Table 3.2. Mitochondrial gene expression changes in the liver of AMG-treated rats.

Gene

Symbol Gene Name 7 Days 14 Days

Aifm2 Apoptosis-inducing factor, mitochondrion-associated 2 1.52 2.54*

Bbc3 Bcl-2 binding component 3 2.67* 2.80*

Bid BH3 interacting domain death agonist 2.37* 2.00

LOC691853

Similar to COX10 homolog, cytochrome c oxidase assembly protein,

heme A: farnesyltransferase 1.53 -2.26*

Mfn1 Mitofusin 1 -1.63 2.21*

Mipep Mitochondrial intermediate peptidase 1.62* 2.05*

Msto1 Misato homolog 1 (Drosophila) 1.35 3.30*

Ppargc1a Peroxisome proliferator-activated receptor gamma, coactivator 1 alpha -3.97* 2.51

Rnf135 Ring finger protein 135 1.17 2.38*

Sfn Stratifin 2.98* -1.44

Slc25a12 Solute carrier family 25 (mitochondrial carrier, Aralar), member 12 -1.26 2.64*

Slc25a13 Solute carrier family 25, member 13 (citrin) -1.23 2.20*

Slc25a19

Solute carrier family 25 (mitochondrial thiamine pyrophosphate carrier),

member 19 -1.13 3.61*

Slc25a23

Solute carrier family 25 (mitochondrial carrier; phosphate carrier),

member 23 1.13 2.91*

Slc25a25

Solute carrier family 25 (mitochondrial carrier, phosphate carrier),

member 25 -2.23* 2.08*

Slc25a30 Solute carrier family 25, member 30 -3.79* -1.29

Slc25a5

Solute carrier family 25 (mitochondrial carrier: adenine nucleotide

translocator) member 5 -1.13 2.00*

Tomm34 Translocase of outer mitochondrial membrane 34 1.33 2.06*

Tomm40 Translocase of outer mitochondrial membrane 40 homolog (yeast) -1.02 2.69*

Tspo Translocator protein 27.30 4.95*

Timm17b Translocase of inner mitochondrial membrane 17 homolog B (yeast) -1.31 2.43*

Values are expressed as fold-change in AMG-treated compared to control rats (n = 4). *p < 0.05

compared to control.

Page 121: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

100

Table 3.3. AMG-induced hepatic gene changes in chemokines and receptors.

Gene

Symbol

Gene Name Day 1 Day 7 Day 14

C3 Complement component 3 -1.08 1.90* 2.16*

Ccbp2 Chemokine binding protein 2 -3.54* -2.03 1.62

Ccl19 Chemokine (C-C motif) ligand 19 1.64* 2.24* 1.30

Ccl3 Chemokine (C-C motif) ligand 3 -4.53* 1.09 1.25

Ccr5 Chemokine (C-C motif) receptor 5 -5.08* -1.13 -2.03

Cxcr7 Chemokine (C-X-C motif) receptor 7 -1.75 -4.66* 3.46*

Cmtm3 CKLF-like MARVEL transmembrane domain containing 3 -2.47* -1.40 1.03

Cx3cl1 Chemokine (C-X3-C motif) ligand 1 1.19 2.06* -1.42

Ppbp Pro-platelet basic protein (chemokine (C-X-C motif) ligand 7) -2.11* 1.49 2.26*

Tymp Thymidine phosphorylase -3.87* -2.80* 1.64

Cxcl13 Chemokine (C-X-C motif) ligand 13 -5.05* -2.21 -5.61*

Mmp2 Matrix metallopeptidase 2 -2.11* -1.67 -1.99

Mmp7 Matrix metallopeptidase 7 1.41 2.32* -1.80

Nfkb1 Nuclear factor of kappa light polypeptide gene enhancer in B-cells 1 -2.21* 1.11 2.79*

Tlr2 Toll-like receptor 2 -3.76* -1.43 1.30

Tnf Tumor necrosis factor (TNF superfamily, member 2) -2.41* -1.05 1.08

Tnfrsf1a Tumor necrosis factor receptor superfamily, member 1a -1.16 1.11 2.57*

Trem1 Triggering receptor expressed on myeloid cells 1 1.32 2.42* -1.26

Values are expressed as fold-change in AMG-treated compared to control rats (n = 4). *p < 0.05

compared to control.

Page 122: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

101

Table 3.4. Hepatic gene changes induced by AMG in various immune pathways after 14

days of treatment.

Gene

Symbol

Gene Name p-value Fold-

Change

Common Cytokines Pathway

Bmp1 Bone morphogenetic protein 1 0.019 2.48

Il16 Interleukin 16 0.035 -2.89

Il1f5 Interleukin 1 family, member 5 (delta) 0.001 -2.38

Il1rn Interleukin 1 receptor antagonist 0.004 2.49

Ltb Lymphotoxin beta (TNF superfamily, member 3) 0.001 3.28

Tnfsf12 Tumor necrosis factor ligand superfamily member 12 0.019 2.83

Toll-Like Receptor Signalling Pathway

Cd14 CD14 molecule 0.005 6.48

Hmgb1 High mobility group box 1 0.032 -3.27

Irak1 Interleukin-1 receptor-associated kinase 1 0.008 3.23

Map2k3 Mitogen activated protein kinase kinase 3 0.047 2.86

Map4k4 Mitogen-activated protein kinase kinase kinase kinase 4 0.028 4.26

Nfkbil1 Nuclear factor of kappa light polypeptide gene enhancer in B-cells inhibitor-like 1 0.017 -1.75

Ppara Peroxisome proliferator activated receptor alpha 0.006 3.55

Sarm1 Sterile alpha and TIR motif containing 1 0.032 5.43

Ticam2 Toll-like receptor adaptor molecule 2 0.022 4.14

Tlr2 Toll-like receptor 2 0.035 2.28

Tnfrsf1a Tumor necrosis factor receptor superfamily, member 1a 0.006 3.27

Th17 Response Pathway

Cd4 Cd4 molecule 0.015 4.39

Icam1 Intercellular adhesion molecule 1 0.005 3.04

Il17rb Interleukin 17 receptor B 0.009 5.57

Il18 Interleukin 18 0.008 -2.91

Nfatc2 Nuclear factor of activated T-cells, cytoplasmic, calcineurin-dependent 2 0.004 3.41

Stat3 Signal transducer and activator of transcription 3 0.004 2.73

Stat5a Signal transducer and activator of transcription 5A 0.004 2.40

Stat6 Signal transducer and activator of transcription 6 0.016 3.17

Values are expressed as fold-change in AMG-treated compared to control rats (n = 4). *p < 0.05

compared to control.

Page 123: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

102

3.6 DISCUSSION

The current findings are consistent with our previous experiments in that aromatic amine

drugs do not appear to cause significant liver injury even though AMG treatment resulted in

more changes in hepatic mRNA expression that are consistent with liver injury than the other

aromatic amines that were tested. In retrospect, although AMG is associated with a very high

incidence of IDRs including cholestatic liver injury, the non-steroidal anti-inflammatory,

bromfenac, is the only primary aromatic amine drug to be withdrawn from the market because of

idiosyncratic hepatocellular liver injury [218]. Sulfamethoxazole often causes idiosyncratic liver

injury, but it is usually part of a more generalized hypersensitivity reaction than liver-specific

toxicity. Furthermore, because the liver is so extensively exposed to xenobiotics that can be

oxidized to reactive metabolites, the primary response of the liver to drug-modified proteins may

be tolerance [211]; thus, liver toxicity caused by aromatic amines may be more a consequence of

a generalized hypersensitivity reaction rather than direct liver injury.

It has been proposed that primary activation of T-lymphocytes in the liver leads to

immune suppression through CD8+ T-lymphocyte apoptosis - the liver has been referred to as a

T-lymphocyte graveyard [219] - whereas primary activation outside of the liver may lead to a

robust immune response in the liver [211]. This may be an appropriate response for an organ

with high xenobiotic exposure and reactive metabolite formation. Although inflammatory

infiltrates were not observed in the liver histologically, the transient increase in GLDH, which is

a marker of liver injury, may be a sign of mild injury followed by adaptation events. Thus, if

these AMG-induced changes in the liver are immune-mediated, then the response to AMG in the

liver in the majority of patients may be immune tolerance.

Page 124: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

103

Approximately 60% of AMG-treated patients experience elevations in γ-

glutamyltransferse [208], which may be an indicator of cholestatic liver injury. The more serious

form of idiosyncratic liver injury is hepatocellular rather than cholestatic [13]. Although

enzymes of cholestatic liver injury were not measured, there was no histological evidence of

cholestatic injury or bile accumulation in the AMG-treated rats. The hypertrophy and hyperplasia

of hepatocytes observed in the AMG-treated rats is consistent with the ability of AMG to induce

cytochrome P450 [215], and this is presumably the reason that the half-life of AMG decreases

with sustained AMG therapy [112].

The largest number of AMG-induced gene changes were within the apoptotic panel of

genes, and this suggests that apoptosis may be involved in AMG-induced IDRs. Apoptosis is a

dynamic process; as cells undergo apoptotic cell death they are cleared by phagocytic cells, and

if this process is not overwhelmed then apoptotic changes may not be observed. However, in the

context of IDRs, excessive apoptotic cell death that cannot be efficiently cleared by phagocytes

may lead to secondary necrosis and the release of danger signals that activate the immune system

[51].

Although at later time points, AMG up-regulated more pro-apoptotic than anti-apoptotic

genes, it was difficult to discern which of the pathways was more dominant. Interestingly, a

pathways analysis found many associations with Nfkb1, which is a part of the nuclear factor-

kappaB (NF-κB) family of transcription factors. NF-κB is involved in immune responses and

inflammation, which ranges from involvement in haematopoiesis and organogenesis to

transcription of cytokines and chemokines [220]. Furthermore, NF-κB can also regulate

programmed cell death, and it has a predominant pro-survival effect because of its ability to

Page 125: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

104

transcribe anti-apoptotic genes such as inhibitor of apoptosis proteins and anti-apoptotic Bcl-2

proteins [221]. This suggests that these changes may modulate the potential effects of the pro-

apoptotic gene expression. Consistent with this, TUNEL staining found no difference between

AMG-treated rats and control rats in the number of apoptotic cells in the liver. Since NF-κB is

regulated in the cytoplasm by the inhibitory protein, IκB, there are limitations in the

interpretation of changes in gene expression because the effect of NF-κB is dependent on

cytoplasmic activation [221]. There is also evidence to suggest that NF-B may play a role in

immune tolerance since lack of NF-B1 expression in dendritic cells activates CD8+ T-

lymphocytes [222], which could be an additional factor contributing to the tolerogenic response

in the liver. However, based on our results, it seems that AMG exposure in the liver induces

modest gene expression in apoptotic pathways that is balanced by anti-apoptotic gene changes

and prevents the induction of apoptosis.

Mitochondrial injury has been proposed to contribute to the mechanisms of IDRs as a

danger signal [13]. Given the importance of mitochondria in maintaining cellular energy and

regulating programmed cell death, it is not surprising that AMG also induced a significant

number of mitochondrial gene changes. Although the majority of mitochondrial gene changes

are involved with solute carriers and transporters, all of the apoptotic genes that were up-

regulated by AMG positively regulate the intrinsic apoptotic pathway, and these changes may

compliment the observed changes in the apoptosis pathways.

AMG induced fewer changes in immune-related genes at these early time points, and the

majority of these changes involved genes associated with the innate immune response. Although

they provide some evidence for the up-regulation of immune signaling pathways, the only

Page 126: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

105

conclusive evidence for activation of an adaptive response was an increase in Cd4 gene

expression. However, a histological increase in lymphocytes was not observed in the liver with

AMG. Thus, activation of the innate immune response upon AMG treatment may be sufficient to

deal with the drug insult without activating the adaptive immune response in the majority of

patients. Moreover, the innate immune system has been proposed to be protective, such as in

cases of hepatic inflammation induced by acetaminophen [74].

Similar to previous gene array studies of aromatic amine drugs, successive time point

testing found that, at least for aromatic amine drugs, the majority of genes are down-regulated at

early time points. The true implications for these findings are still unknown; however, there may

be some compensatory mechanisms in which a threshold needs to be crossed before gene

expression is elicited. As seen with these studies, the mere presence of changes in gene

expression does not necessarily correlate with pathology. Furthermore, we did not investigate

downstream protein expression or signaling pathways because our main goal was only to detect

early biomarkers to predict drug risk because, according to the danger hypothesis, it is early

injury of the target cells, in this case the hepatocytes, that would determine the immune response.

Overall, the implications of this study are that the effects of AMG on the liver are small,

and that the very high incidence of IDRs associated with this drug may originate from outside the

liver. The finding that immune cells can also bioactivate aromatic amines may be the key to their

ability to cause IDRs [49]. If this is true it also suggests that in addition to screening for reactive

metabolite formation in the liver, bioactivation by immune cells, especially macrophages, may

provide important biomarkers for detecting the potential of whether a drug candidate can cause

IDRs. Conversely, perhaps the primary response in the liver is adaptation and/or tolerance. Even

Page 127: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

106

though there was no pathological response, the changes in gene expression induced by AMG

could be mechanistic clues, and these changes may occur in most individuals, and it is only when

some additional factor is present that an IDR occurs. The most significant findings were that the

changes in the apoptosis and mitochondrial panel of genes may play a role in the mechanisms of

AMG-induced IDRs, potentially as danger signals. Further-more, because the predominant

immune changes were associated with innate immunity, it is quite possible that innate immune

cells are able to effectively deal with the injury caused by the drug and effectively down-regulate

any potential adaptive immune response. We have performed high through-put arrays with

multiple IDR-associated drugs and found differential changes suggestive of cellular stress and an

immune response at early-time points with most. The major exceptions are the primary aromatic

amine drugs, which emphasizes the uniqueness of this chemical moiety and the need for its

study.

Page 128: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

107

CHAPTER 4

EFFECT OF AMINOGLUTETHIMIDE ON

NEUTROPHILS IN RATS: IMPLICATIONS FOR

IDIOSYNCRATIC DRUG-INDUCED BLOOD

DYSCRASIAS

Winnie Ng and Jack Uetrecht

(Reproduced with permission from: Ng, W., and Uetrecht, J. Chemical Research in Toxicology,

2013. 26: p.1272-1281. Copyright 2013 American Chemical Society. The American Chemical

Society granted permission to include my own article in this thesis).

Page 129: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

108

4.1 ABSTRACT

Aminoglutethimide (AMG) is an aromatic amine aromatase inhibitor associated with a

high incidence of idiosyncratic blood dyscrasias, especially agranulocytosis. Animal models of

idiosyncratic drug reactions (IDRs) represent essential tools to study these reactions; however,

there is currently no valid model of idiosyncratic drug-induced agranulocytosis. Although AMG

does not cause agranulocytosis in most animals or humans, drugs associated with serious IDRs

generally cause a higher incidence of mild reactions that resolve despite continued treatment.

Therefore, the effects of AMG on neutrophils and bone marrow in rats were studied to

understand the mechanisms of more serious IDRs. An increase in peripheral blood neutrophils

occurred as early as 24 h after AMG treatment with minimal changes to the total leukocyte

count. Further investigation using 5-bromo-2’-deoxyuridine (BrdU) found an increased release

of neutrophils from the bone marrow. Histologically, this corresponded to an increase in myeloid

cells in the bone marrow, which was confirmed by differential staining with CD45 and CD71.

AMG treatment stimulated an increase in colony forming unit granulocyte-macrophage and

colony forming unit granulocyte ex vivo. There was also a marked increase in the number of

activated neutrophils in the circulation expressing the extravasation marker CD62L. These

findings indicate that AMG affects neutrophil production, release, and function. Similar effects

on neutrophil kinetics in clozapine-treated rats have previously been found, and transient

neutrophilia has been observed in patients taking other drugs associated with idiosyncratic

agranulocytosis; therefore, the changes observed with AMG may be biomarkers to predict the

risk that a drug will cause agranulocytosis.

Page 130: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

109

4.2 TABLE OF CONTENTS GRAPHIC

Page 131: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

110

4.3 ABBREVIATIONS

AMG, aminoglutethimide

APC, allophycocyanin

BrdU, 5-bromo-2’-deoxyuridine

CFUs, colony forming units

Cxcl1, chemokine (C-X-C) 1

DAPI, 4',6-diamidino-2-phenylindole dihydrochloride

FACS, fluorescence-activated cell sorting

FBS, fetal bovine serum

FITC, fluorescein isothiocyanate

G-CSF, granulocyte colony stimulating factor

GSH, glutathione

H&E, hematoxylin and eosin

IDRs, idiosyncratic drug reactions

IMDM, Isocove’s Modified Dulbecco’s Medium

MPO, myeloperoxidase

NAT, N-acetylation

PE, phycoerythrin

PI, propidium iodide

RBCs, red blood cells

TNF-α, tumor necrosis factor alpha

Page 132: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

111

4.4 INTRODUCTION

Aminoglutethimide (AMG) is a first-generation aromatase inhibitor used primarily for

the treatment of hormone-responsive breast and prostate cancers. Structurally, AMG contains a

primary aromatic amine. This chemical moiety is almost always associated with a high incidence

of idiosyncratic drug reactions (IDRs); therefore, this is a notorious structural alert for drug

development [223]. The toxicity of aromatic amines is presumably due to the ease with which

they can be oxidized to reactive electrophilic metabolites, which can covalently bind to cellular

macromolecules and potentially induce damage [212]. However, the exact mechanisms of IDRs

are still poorly understood. More importantly, it is unknown why only a subset of patients

develops these reactions whereas most do not. Animal models, with similar pathogenesis as

occurs in humans, are essential for studying the underlying mechanisms of IDRs; yet, valid

animal models of IDRs are rare and quite difficult to develop [61].

AMG is associated with a wide range of IDRs including serious skin rashes, cholestasis,

and haematological toxicities [224]. Of these reactions, the most commonly reported are blood

dyscrasias such as agranulocytosis, which is characterized by a decrease in neutrophils to below

500/µL of blood; and if sustained for a prolonged period of time, usually results in a fatal

infection. Agranulocytosis occurs in approximately 1.9% of AMG-treated patients and typically

manifests after more than 4 weeks of treatment [113]. In these patients, bone marrow aspirates

are usually hypocellular with decreased myeloid cells, which return to normal after

discontinuation of the drug. However, the WBC counts sometimes return to normal without

cessation of AMG therapy [225], which suggests the induction of tolerogenic mechanisms that

may be why most patients do not develop agranulocytosis. In fact, it is a common characteristic

that agents that can induce serious IDRs always cause a much higher incidence of mild reactions

Page 133: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

112

that often resolve despite continued treatment. This suggests that such drugs have biological

effects that are quite common and asymptomatic even if the severe reactions are rare.

Nevertheless, data is lacking on the effects in most patients, in terms of the early events leading

up to agranulocytosis, because blood monitoring of AMG-treated patients does not usually occur

before symptoms arise.

In general, aromatic amines can be oxidized to a N-hydroxylamine metabolite by

cytochrome P450 enzymes or myeloperoxidase, which is relatively stable and un-reactive.

However, the hydroxylamine can auto-oxidize to a nitrosoamine metabolite that is electrophilic

and reactive and can covalently bind to proteins. Compared to other aromatic amine-containing

drugs, such as the more common sulfonamides, AMG has a higher electron density, which

allows it to be even more easily oxidizable (Figure 4.1). N-hydroxyl-AMG has been detected in

both humans and mice [226, 227]. Neutrophils contain high concentrations of myeloperoxidase

[228]; therefore, it is quite likely that AMG can be oxidized by this cell type and lead to cell

damage. Alternatively, myeloperoxidase protein free radicals have been detected in vitro upon

incubation with AMG [177], which further suggests a role for oxidation by neutrophils in the

pathogenesis of agranulocytosis.

To date very few studies, if any, have focused on the effect of AMG on neutrophils, and a

true animal model of AMG-induced agranulocytosis has not yet been established. Leucopenia

was observed in both ICI-derived and B6C3F1 hybrid mice as early as after 2 weeks of AMG

treatment [174, 229]; however, these mouse strains are uncommon, and this model was not

reproducible in CD-1 mice [175]. Given our negative results in mice and our previous experience

Page 134: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

113

with Brown Norway rats, we investigated the effects of AMG on neutrophils and bone marrow in

this rat strain.

Figure 4.1. Structure of AMG and the formation of its reactive metabolites. AMG can be

detoxified through N-acetylation (NAT). Alternatively, AMG can be oxidized by cytochrome

P450 (P450) or myeloperoxidase (MPO) to a hydroxylamine, which can spontaneously form a

highly electrophilic and reactive nitrosoamine that can covalently bind to protein thiols. The

nitrosoamine can also be reduced back to the parent drug by ascorbate or glutathione (GSH) and

redox cycle to cause oxidative stress.

Page 135: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

114

4.5 MATERIALS AND METHODS

4.5.1 Chemicals and Reagents

AMG was purchased from Toronto Research Chemicals (North York, ON). Phenacetin,

dextran-500, BrdU, and 10% neutral buffered formalin were obtained from Sigma (Oakville,

ON). Heat-inactivated fetal bovine serum (FBS), Iscove’s Modified Dulbecco’s Medium

(IMDM), and 4',6-diamidino-2-phenylindole dihydrochloride (DAPI) were purchased from Life

Technologies (Burlington, ON).

4.5.2 Antibodies

Anti-rat granulocyte phycoerythrin (PE; clone: RP-1), 7-amino-actinomycin D (7-AAD),

anti-BrdU fluorescein isothiocyanate (FITC), annexin V FITC, propidium iodide (PI), anti-rat

CD18 FITC (clone: WT.3), and anti-rat CD11b allophycocyanin (APC; clone: WT.5) were

acquired from BD Biosciences (San Jose, CA). Anti-rat CD71 FITC (clone: OX-26), and anti-rat

CD62L FITC (clone: OX-85) were purchased from Cedarlane (Burlington, ON). Anti-rat CD45

PE (clone: OX-1) was obtained from BioLegend (San Diego, CA).

4.5.3 Animals

Male Brown Norway rats (176-200 g) were purchased from Charles River (Montreal,

QC). Rodents were housed doubly under standard conditions with automatic watering and a

12:12 h light/dark cycle at 22°C. Food was provided ad libitum, and all animals were given

standard rodent chow. Animals were acclimatized for one week before the start of experiments.

The experiment protocol was approved by and performed in accordance with the University of

Toronto Faculties of Medicine and Pharmacy Animal Care Committee.

Page 136: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

115

4.5.4 Treatments

Rats were treated with 80, 125, or 160 mg/kg/day of AMG in 0.5% methylcellulose

through oral gavage. Control animals were given methylcellulose vehicle only. Each group had 4

rats each unless otherwise stated. The 80 mg/kg AMG dose was chosen as a starting dose based

on previous rat studies [230] to mimic the plasma concentrations found in patients taking

therapeutic doses of AMG [231].

4.5.5 Measurement of AMG Blood Levels

Blood levels of AMG were determined 1 h and 24 h after a single 80 or 160 mg/kg dose

of AMG, and analysis was performed similar to Adam et al. [232]. Briefly, blood was collected

from the tail vein of rats into heparinised tubes, and plasma was obtained by centrifugation. To

20 µl of plasma, 10 µl of phenacetin (10 µg/ml stock in methanol) was added as an internal

standard and 80 µl of methanol. The solution was vortexed vigorously for 30 s and then

incubated at -20 °C for 30 min. The sample was centrifuged at 16000g for 10 min, and the

supernatant was transferred to a clean tube and evaporated to dryness under nitrogen. Samples

were then reconstituted in 50 µl of mobile phase (42% methanol: 58% water) and 10 µl was

injected into the HPLC (Hewlett-Packard Series 1050) for analysis with an Ultracarb column (5

µ ODS (30) 100 x 2.0 mm; Phenomenex) at a flow rate of 0.2 ml/min. AMG was detected by UV

absorption at 245 nm, and a standard curve was produced to quantify AMG. The retention times

were 3.8 and 7.3 min for AMG and phenacetin, respectively, and a standard curve was produced

with a correlation coefficient greater than 0.99. The limit of detection was 0.96 µg/ml.

4.5.6 Leukocyte Counts

Blood was collected from the tail vein into EDTA-coated tubes. Whole blood was diluted

20-fold in Turks solution (Ricca Chemical Company, Arlington, TX) to lyse red blood cells

Page 137: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

116

(RBCs), and the total WBC count was obtained manually using a haemocytometer. A 5 µl

aliquot of whole blood was also smeared on a slide, fixed, and Wright-Giemsa stained using

CAMCO Stain Pak (Cambridge Diagnostic Products Inc., Fort Lauderdale, FL) as per the

manufacturer’s instructions, and the WBC differential was determined manually under a light

microscope by characterizing 100 leukocytes per slide.

4.5.7 Measurement of Cytokines

Peripheral blood cytokines, chemokine (C-X-C) 1 (Cxcl1), and tumor necrosis factor

alpha (TNF-α) were measured using the Rat Quantikine ELISA kits (R&D Systems,

Minneapolis, MN) according to manufacturer’s protocol. Granulocyte colony stimulating factor

(G-CSF) was also measured in the peripheral blood with a Mouse Quantikine ELISA kit (R&D

Systems). Although the G-CSF ELISA was not specifically developed for use in rats, a BLAST

search shows greater than 80% gene sequence identity to rat G-CSF.

4.5.8 Leukocyte Isolation from Peripheral Blood

Collected whole blood was mixed with equal volumes of 3% dextran (prepared in saline)

and incubated for 18 min at room temperature. The straw-coloured upper layer was collected and

centrifuged at 350g for 5 min and the supernatant discarded. RBCs were lysed with red cell lysis

buffer by incubation for 10 min. Cells were washed twice in florescence-activated cell sorting

(FACS) buffer (5% FBS in PBS) before final resuspension in FACS buffer, and cell counting

was performed using trypan blue and a Countess Automated Cell Counter (Invitrogen, Life

Technologies).

Page 138: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

117

4.5.9 Measurement of New Neutrophil Release Using 5-bromo-2′-deoxyuridine

(BrdU).

Rats were treated with either methylcellulose vehicle control or AMG for 10 days. BrdU

is known to be carcinogenic and was handled using proper procedures for carcinogenic

substances as outlined by the University of Toronto Faculties of Medicine and Pharmacy Animal

Care Committee. On day 4 of treatment, BrdU (100 mg/kg dissolved in warm saline) was

injected intraperitoneally, and blood samples were taken from the tail vein to monitor neutrophil

incorporation of BrdU after leukocyte isolation using the FITC BrdU Flow Kit (BD Biosciences)

as per the manufacturer’s protocol. Anti-rat granulocyte PE antibody was included as a surface

marker for granulocytes. Cells were analyzed by BD FACSCalibur (BD Biosciences) at a flow

rate of no more than 400 events/s.

4.5.10 Bone Marrow Histology

The femurs of the rats was isolated using standard procedures and fixed in neutral

buffered formalin (Sigma) for 2-5 days. Tissues were submitted to the Toronto Centre for

Phenogenomics Histology Laboratory (Toronto, ON) for paraffin embedding and H&E stained

slide preparation. Microscopic pictures were taken at the Microscopy Imaging Lab (Faculty of

Medicine, University of Toronto) on a Zeiss fluorescence microscope with deconvolution.

4.5.11 Characterization of Bone Marrow Cells

The extracted rat femur was flushed with 5 ml of IMDM containing 10% FBS with a 22

gauge needle. Aggregates were broken up through repeated aspiration, and cells were passed

through a 40 µm filter to make a single cell suspension. Cells were centrifuged at 300g for 5 min

and resuspended in 5.0 ml of FACS buffer before cell counting as per above. Bone marrow cells

were characterized similar to the protocol of Saad et al [233]. For surface staining, 5 µl of anti-

Page 139: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

118

rat CD71 FITC and 1.25 µl of anti-rat CD45 PE antibodies were added to 1x106 cells in a total

volume of 100 µl and incubated in the dark for 30 min at 4°C. Cells were washed twice with

FACS buffer before cells were fixed with IC Fixation Buffer (eBioscience) for 20 min, washed,

and left overnight in FACS buffer. Cells were then re-fixed and permeablized with

Fixation/Permeablization Buffer (eBioscience) for 20 min, washed once in FACS buffer, and

then resuspended in 200 µl of staining buffer (100 mM Tris pH 7.4, 150 mM NaCl, 1 mM CaCl2,

0.1% Tween20, 0.5 mM MgCl2 ). DAPI was added to the sample in a final concentration of 3

µM in 100 µl volume and incubated in the dark for 15 min at room temperature before samples

were analyzed immediately using flow cytometry.

4.5.12 Investigation of Bone Marrow Progenitor Cells

Bone marrow cells were extracted as previously described and resuspended in IMDM

containing 2% FBS. Cells were counted manually with a haemocytometer using Turk’s solution.

Progenitor cells in the bone marrow were studied in a colony-forming cell (CFC) assay for rat

cells using the MethoCult GF R3774 medium (StemCell Technologies, Vancouver, BC)

according to the manufacturer’s protocols. Briefly, cells were plated at 1.5x104 per 1.1 ml culture

in 35 mm dishes and incubated at 37 °C, 5% CO2, with ≥95% relative humidity for 10 days

before colony forming unit-granulocyte-macrophage (CFU-GM), colony forming unit-

granulocyte (CFU-G), and colony forming unit-macrophage (CFU-M) were manually assessed

under a light microscope. All colonies on the plate were counted, and each sample was

duplicated in a separate dish.

4.5.13 Determination of Neutrophil Apoptosis

Apoptosis was measured using the Annexin V Apoptosis Kit (BD Biosciences) as per the

manufacturer’s protocol. Specifically, isolated blood leukocytes were washed twice with PBS

Page 140: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

119

and then reconstituted at 2x105 cells per 100 µl binding buffer. Cells were then incubated with 5

µl each of annexin V FITC and PI for 15 min in the dark at room temperature. Binding buffer

(400 µl) was added to each sample to stop the reaction and cells were analyzed by flow

cytometry within 2 h.

4.5.14 Determination of Neutrophil Activation

Leukocytes were isolated as described previously and cell surface markers were stained

using standard flow cytometry procedures. Anti-rat CD62L FITC, anti-rat CD11b APC, anti-rat

CD18 FITC, and 7-AAD were incubated in the dark for 30 min at 4°C, washed twice in FACS

buffer, and then analyzed by flow cytometry.

4.5.15 Flow Cytometry Analysis

Unless otherwise stated, flow cytometry experiments were run at the Faculty of Medicine

Flow Cytometry Facility (University of Toronto) on the BD FACSCanto II (BD Biosciences)

using BD FACSDiva Software (BD Biosciences) with a flow rate of no more than 400 events/s.

Compensation was corrected using BD Compensation Beads (BD Biosciences). FlowJo Software

(Tree Star, Inc., Ashland, OR) was used for detailed analysis of the flow cytometry results.

4.5.16 Data Analysis

Statistical analysis was performed using GraphPad Prism 5 (GraphPad Software Inc., La

Jolla, CA). The Mann-Whitney U test or two-way ANOVA with Bonferroni post-tests were used

depending on the constraints of the data to determine statistical significance between AMG and

control groups. Error is represented as standard error of the mean (SEM).

Page 141: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

120

4.6 RESULTS

4.6.1 Effect of AMG on Peripheral Blood Leukocyte Counts

Plasma levels of AMG 1 h after a single gavage dose of 80 and 160 mg/ml AMG were

15.0 (± 2.0) µg/ml and 18.4 (± 2.1) µg/ml, respectively (Figure 4.2), which is within the

therapeutic range (4.7-32.4 µg/ml) found in humans during chronic AMG therapy [234]. Upon

AMG treatment at a dose of 80 mg/kg/day, an increase in peripheral blood neutrophils was

observed as early as 24 h and remained elevated at 48 h with minimal changes to the total WBC

count (Figure 4.3). Further investigation of the dose-response relationship found no clear

changes in the WBC differential of rats treated with 160 mg/kg/day of AMG (data not shown).

This was presumably due to direct toxicity because the rats did not tolerate the high dose of

AMG well, and treatment had to be withdrawn. However, at a lower dose of 125 mg/kg/day of

AMG, a similar elevation in peripheral blood neutrophils was found compared to the dose of 80

mg/kg/day; this was also significantly increased after 24 h (Figure 4.4) and sustained through 14

days of treatment (Figure 4.5). Again, changes in total WBC counts were minimal in the AMG-

treated animals because the neutrophil increase seemed to be compensated for by a slight

decrease in the lymphocyte count, most noticeable after 14 days AMG treatment (Figure 4.5C).

Because the most significant neutrophil response was observed at 125 mg/kg/day of AMG, this

dose was used for subsequent experiments. To provide further evidence for the effect of AMG on

neutrophils, blood cytokines associated with neutrophil production (G-CSF), chemotaxis

(Cxcl1), and activation (TNF-α) were measured; they were all found to be noticeably elevated

within a day of AMG treatment (Figure 4.6).

Page 142: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

121

Figure 4.2. Blood levels of AMG in rats. Blood was obtained from the tail vein of rats at 1 and

24 h after a single oral dose of 80 or 160 mg/kg of AMG. Values expressed are mean ± SEM for

each group (n = 4).

Figure 4.3. Acute AMG-induced changes on

peripheral blood neutrophils in rats given an

oral dose of 80 mg/kg/day. Blood was taken

from the tail vein of rats and the WBC

differential was determined by Wright-Giemsa

staining. (A) No major changes were observed

in the total WBC count; however, (B) an

increase in neutrophils was observed as early as

24 h after AMG treatment. Values expressed are

mean ± SEM for each group (n=4). None of the

results were statistically significant by Mann

Whitney U test.

Page 143: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

122

Figure 4.4. Acute AMG-induced changes on peripheral blood neutrophils in rats given an

oral dose of 125 mg/kg/day. Blood was taken from the tail vein of rats, and the WBC

differential was determined by Wright-Giemsa stain. (A) No differences were observed in the

total WBC count between AMG-treated and control rats; whereas, (B) significant increases in

peripheral blood neutrophils were observed after both 1 and 7 days of AMG treatment, which

seemed to be compensated for by (C) a slight decrease in peripheral blood lymphocytes. Open

bars represent the control group; solid bars represent the AMG-treated group. Values expressed

are mean ± SEM (n=4). *p<0.05 and ***p<0.001 compared to controls by two-way ANOVA.

Page 144: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

123

Figure 4.5. Changes in peripheral leukocyte cell counts induced by treatment of rats with

AMG at a dose of 125 mg/kg/day for 14 days. (A) No changes were observed in total WBC

between control and AMG-treated rats; whereas, (B) a noticeable increase in the neutrophil

population was found after both 7 and 14 days of AMG treatment and (C) a corresponding slight

decrease in lymphocyte counts. Open circles and dotted lines represent the control group; solid

squares and lines represent the AMG-treated group. Values expressed are mean ± SEM (n=4).

None of the results were significant by two-way ANOVA.

Page 145: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

124

Figure 4.6. AMG-induced changes in neutrophil-associated cytokines in the blood of rats.

(A) G-CSF, involved with stimulating neutrophil production, was significantly elevated after 7

days of AMG treatment. (B) The neutrophil chemoattractant Cxcl1 was noticeably increased

after both 1 and 7 days of AMG treatment, and (C) TNF-α, involved with neutrophil activation,

was up-regulated after both 1 and 7 days of AMG treatment. Values are expressed as mean ±

SEM (n=4). *p<0.05 and ***p<0.001 compared to baseline measurements by Mann Whitney U

test.

Page 146: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

125

4.6.2 Effect of AMG on Bone Marrow

Given that all neutrophils originate from bone marrow and BrdU is a thymidine analogue

that gets incorporated into newly formed DNA, BrdU was used to track new neutrophils released

from the bone marrow into the circulation by flow cytometry and double staining with an anti-rat

granulocyte antibody. Consistent with the increase in peripheral blood neutrophils, an increase in

newly formed neutrophils labeled with BrdU was observed in the AMG-treated rats with the

most significant changes occurring at 120 h post-BrdU injection (Figure 4.7). Furthermore, the

increase in newly formed neutrophils was sustained throughout AMG treatment, which suggests

that AMG prolongs the half-life of circulating neutrophils. H&E staining of bone marrow from

the femurs of rats found hypercellularity induced by AMG that was attributed to an increase in

the myeloid cells and was not observed in the controls (Figure 4.8). To confirm this, nucleated

cells of the bone marrow, detected with DAPI, were phenotyped by flow cytometry through

differential staining of CD45 and CD71 to identify changes in the myeloid, lymphoid, and

erythroid cell populations. A significant increase in the myeloid cell population was found with

AMG treatment, and this corresponded to a simultaneous decrease in the lymphoid cell

population (Figure 4.9). The myeloid to erythroid ratio of the bone marrow in AMG-treated rats

was also significantly elevated compared to controls (2.36 ± 0.31 vs. 1.48 ± 0.13, respectively,

p=0.03). To determine which myeloid cell populations were affected by AMG, a MethoCult

assay was performed ex vivo to evaluate the proliferative capacity of granulocyte and

macrophage colony forming units (CFUs) in response to AMG treatment. The effect of AMG

seemed to be more specific to granulocytes because there was an increase in the number of CFU-

granulocyte-macrophage (CFU-GM) and CFU-granulocyte (CFU-G) colonies, but there was not

much of a change in CFU-macrophage (CFU-M) colonies (Figure 4.10).

Page 147: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

126

Figure 4.7. Effect of AMG on new neutrophil release from the bone marrow. Rats were

treated with 125 mg/kg/day of AMG for 10 days. On the fourth day, 100 mg/kg BrdU was

injected i.p., and blood from the tail vein was monitored for changes in neutrophils by flow

cytometry. Neutrophils were gated by forward and side scatter, and newly formed neutrophils

were characterized as cells double positive for anti-granulocyte antibody and BrdU. Both the (A)

percent BrdU-stained neutrophils, and the (B) absolute number of newly formed neutrophils

were increased by AMG treatment. Open circles and the dotted line are from control animals;

solid squares and lines represent AMG-treated animals. Values expressed are mean ± SEM (n=5

control, n=6 AMG). *p<0.05 compared to control group by two-way ANOVA.

Page 148: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

127

Figure 4.8. Bone marrow changes induced by AMG. Representative H&E staining of bone

marrow from (A) control rat femur and (B) rat femur after 14 days of AMG at a dose of 125

mg/kg/day. 40x magnification.

Page 149: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

128

Figure 4.9. AMG-induced changes in bone marrow cells. Bone marrow cells were extracted

from rats treated for 14 days with 125 mg/kg/d of AMG and analyzed through FACS analysis.

Representative flow plots are shown for (A, B) control and (C, D) AMG-treated rats. (A, C)

From the nucleated cells, stained as DAPI positive, nucleated erythroid cells were defined as

CD71 positive, CD45 medium expression, whereas the myeloid and lymphoid cells were defined

as CD71 low, CD45 medium-high. (B, D) From this, myeloid and lymphoid cells were further

separated based on side scatter where myeloid cells were more granular than lymphoid cells. (E)

Page 150: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

129

No change was observed in the erythroid population between treated and control animals;

however, in the AMG-treated rats there was a significant increase in myeloid cells with a

simultaneous significant decrease in the lymphoid population. Open bars represent the control

group; solid bars represent the AMG-treated group. Values expressed are mean ± SEM (n=4).

**p<0.01 compared to the control group by two-way ANOVA.

Page 151: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

130

Figure 4.10. Effect of AMG on bone marrow granulocyte and macrophage CFUs. Bone

marrow cells were extracted from rats treated with 125 mg/kg/day AMG for 14 days. Cells were

then seeded into culture dishes at 1x105 cells per 1.1 ml MethoCult medium and incubated at

37°C, 5% CO2, and ≥95% humidity. After 10 days colonies were counted under a light

microscope and classified as either CFU-GM, CFU-G, or CFU-M. AMG significantly increased

the total number of CFUs and CFU-GM. An increase in the CFU-G was also observed with

AMG treatment; however, no change was observed in CFU-M. Open bars represent the control

group; solid bars represent the AMG-treated group. Samples were performed in duplicate.

Values expressed are mean ± SEM (n=4). *p<0.05 and **p<0.001 compared to control by two-

way ANOVA.

Page 152: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

131

4.6.3 Effect of AMG on Neutrophils

Since AMG induced a sustained increase in newly formed neutrophils in rats, further

investigation into the effect of AMG on neutrophil survival was warranted. Neutrophil apoptosis

was investigated through annexin V and PI staining using flow cytometry whereby neutrophils

were gated based on forward and side scatter, and cells that were annexin V positive/PI negative

were considered early apoptotic cells. It is interesting to note the increase in density of the

neutrophil population in the AMG-treated rats (Figure 4.11A, B), which corresponds to the

increase in blood neutrophils that was previously observed. Using this approach, AMG treatment

induced a significant reduction in the number of early apoptotic neutrophils compared to controls

(62.5% at 7 days treatment, p=0.01; and 55.6%, after 12 days of treatment, p=0.03) as shown in

representative flow plots (Figure 4.11C, D). In terms of activation, AMG significantly down-

regulated the expression of CD62L in neutrophils over time (Figure 4.12), while a simultaneous

increasing trend was observed in CD11b expression in neutrophils, most noticeable at day 7 and

12 (data not shown). However, no changes were found in CD18 expression (data not shown).

Both CD62L and CD11b are involved with the infiltration of leukocytes into tissues, and these

changes suggest neutrophil activation.

Page 153: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

132

Figure 4.11. AMG-induced changes in neutrophil apoptosis. Blood taken from the tail vein of

rats treated with AMG at a dose of 125 mg/kg/day was incubated with annexin V and PI to

determine the number of apoptotic cells. As compared to (A) control, representative flow plots

show (B) an increase in the percentage of neutrophils upon AMG treatment. From the neutrophil

population, (C) the percent annexin V positive/PI negative early apoptotic cells are high in the

control rats, whereas (D) there is a decrease in the number of apoptotic neutrophils in the AMG-

treated rats.

Page 154: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

133

Figure 4.12. Changes in neutrophil CD62L expression induced by AMG. Blood taken from

the tail vein of rats treated with 125 mg/kg/day of AMG were analysed by flow cytometry for

CD62L expression on neutrophils. Neutrophils were identified by their distinct forward and side

scatter characteristics as shown previously in Figure 12. Representative flow plots from (A)

control and (B) AMG-treated rats show that CD62 is expressed in neutrophils under normal

physiological conditions; however, upon AMG treatment there is an increase in the number of

neutrophils that have decreased CD62L expression, as shown by the increase in cells in the

CD62L low gate. (C) The mean florescence intensity (MFI) for the CD62L channel is

significantly decreased at both days 7 and 12 in the AMG-treated as compared to controls. (D) A

significant increase was also observed in the percentage of cells in the CD62L low gate in the

Page 155: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

134

AMG-treated at days 7 and 12. Open bars represent control; solid bars represent AMG-treated.

Values expressed are mean ± SEM (n=4). *p<0.05, **p<0.01 compared to control by two-way

ANOVA.

Page 156: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

135

4.7 DISCUSSION

To date, there have been no detailed studies on the effect of AMG on neutrophils despite

the many reports of AMG-induced agranulocytosis. The IDRs induced by AMG are not

surprising given the fact that it is an aromatic amine drug, and aromatic amines are readily

oxidized by neutrophil myeloperoxidase [235, 236]. In the current studies, the major finding was

that AMG increased the number of peripheral neutrophils. The effect on neutrophils was

observed as early as 24 h, which suggests that AMG mobilized the marginal zone neutrophils;

however, this increase was sustained by an increase in granulocyte production for at least 14

days. This change was compensated for by a slight decrease in peripheral blood lymphocytes

such that no significant changes were found in the total WBC count. In rodents, the neutrophils

make up only approximately 10-30% of the total WBCs in the blood [142], whereas in humans

the majority of peripheral WBCs are neutrophils, thus the compensatory effect of lymphocytes is

consistent to a subtle change.

The dosages of AMG given to rats in this study produced blood levels of AMG similar to

what was found in humans. However, the dosage of AMG used in human therapy is fairly high

(can be up to 1000 mg per day) and it would not be surprising if AMG can induce neutrophilia

similar to what occurs commonly during stress or inflammation. The stress response, specifically

corticosterone is known to increase neutrophils; however this effect is unlikely to be a significant

issue because AMG has been found to decrease corticosterone levels in the blood [237].

Moreover, inflammation may be a central feature to induce immune activation, and if these

reactions are immune-mediated, this may lead to an immune response to drugs associated with

agranulocytosis.

Page 157: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

136

It is interesting that the effect of AMG on neutrophils in rats is the opposite of the

agranulocytosis that is observed in some patients [225] and the leucopenia that occurs in mice

[174]. However, in general, AMG patients are not monitored until blood dyscrasias arise;

therefore, it is unknown whether patients experience earlier changes in their WBC differential. In

the previous mouse models, leucopenia was observed as early as after 2 weeks of treatment

[174], and it may be possible that if we had continued treatment for a much longer period of time

it would result in agranulocytosis. A previous study with AMG-treated rats found no changes in

WBC differential even after 3 weeks of treatment [229]; however, the dose was much lower (50

mg/kg/day) than the dose currently used. In addition, rats had more sustained levels of AMG in

the blood 24 h after a single dose than mice treated with similar doses (data not shown). For

these reasons, the rat may be a better model of AMG-induced blood dyscrasias and be more

useful as a tool for mechanistic studies.

Interestingly, AMG-induced changes in neutrophils share common characteristics to what

is observed in humans upon clozapine therapy. Clozapine is an atypical antipsychotic that is

associated with agranulocytosis in approximately 1% of patients, and the mechanism is thought

to be immune-mediated [77]. Moreover, clozapine patients experience an initial neutrophilia

before the onset of agranulocytosis [167] and increases in CD34+ hematopoietic stem and

progenitor cells has also been reported within 2 weeks of initiating therapy [238], suggesting the

mobilization of neutrophil precursors. A similar increase in neutrophils has been observed in

rabbits [239] and rats (unpublished observations) upon treatment with clozapine, and this is

consistent with the increase in neutrophils that we have observed in AMG-treated rats. Clozapine

also appears to decrease neutrophil half-life in rabbits [239], and it may be the lack of ability of

the bone marrow reserve to keep up with demand that leads to agranulocytosis. In contrast, the

Page 158: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

137

increased release of newly formed neutrophils from the bone marrow and the kinetic profile of

the BrdU-labelled neutrophils suggest that AMG may increase neutrophil survival and half-life.

Although further studies using annexin V and PI found a decrease in apoptotic neutrophils in the

AMG-treated rats, which is consistent with an increase in neutrophil survival, this change may be

caused by a greater number of younger neutrophils being released from the bone marrow that

express less phosphatidylserine. The basis for this observation may be the high levels of

apoptosis in the control rats (although partially due to the handling procedures) that decreases

upon AMG treatment concomitant with a greater number of neutrophils. Furthermore, in vitro

studies on neutrophil survival in the presence of AMG resulted in no difference in neutrophil

apoptosis from vehicle-treated (data not shown). Therefore, more detailed studies are required to

determine how AMG alters neutrophil survival, although the rapid turnover rate and clearance of

apoptotic neutrophils may make such studies difficult. In the bone marrow of the AMG-treated

rats there was also an increase in CFU-GM and CFU-G colonies, which are precursors to

neutrophils, and this is consistent with an increase in the production of progenitor cells from the

bone marrow.

In terms of cytokine expression, clozapine-induced fever is typically mediated by IL-6,

and patients characteristically express increased levels of TNF-α and sIL-2r that are exacerbated

by fever [169]. Clozapine has also been found to transiently increase the levels of G-CSF in

patients that was independent of the development of fever [168]. Thus, these changes in serum

cytokines upon clozapine treatment are common in the majority of patients and could be

biomarkers to identify drugs predisposed to causing agranulocytosis. In our study, an increase in

neutrophil-associated cytokines, G-CSF and TNF-α, was observed in the blood upon AMG

treatment in rats. The release of G-CSF stimulates the production of granulocytes, and it has

Page 159: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

138

been given to patients to treat neutropenia. Moreover, TNF-α can stimulate the neutrophil

respiratory burst by activating NADPH oxidase [151]. Thus there seems to be similarities in the

expression of cytokines by both clozapine and AMG and these changes could be biomarkers to

identify drugs predisposed to causing agranulocytosis.

Although the rats, under our current experimental protocols, do not develop

agranulocytosis or show evidence of any other pathologies, the changes in the blood upon AMG

treatment may occur in the majority of treated patients and may also provide clues to the

mechanism of more serious IDRs. One common explanation for the lack of development of a

pathological response is the induction of immune tolerance. Evidence for this is provided by the

resolution of clozapine-induced agranulocytosis without discontinuing treatment [166].

Activated neutrophils may lead to an immune response that leads to agranulocytosis; however,

the usual response may be immune tolerance. From the current studies, AMG appears to

stimulate the bone marrow to release new neutrophils that are activated in terms of the

extravasation marker CD62L. In humans challenged with endotoxin, CD62Ldim

neutrophils were

found to be more hypersegmented than the CD62Lbright

neutrophils, and this population was also

able to suppress T-cell activation [160], which may be a mechanism of immune tolerance.

However, it is unknown how this is affected in AMG-induced IDRs, and more studies will be

required to determine the link between neutrophil activation and agranulocytosis. Another

limitation that we were unable to sufficiently address was the extent of the involvement of the

arylamine group of AMG to the effects observed on the neutrophils, as we were unable to obtain

glutethimide because it is classified as a controlled substance. Nevertheless, numerous studies

have shown additional effects of neutrophils on the adaptive immune system in addition to their

traditional role in innate immunity; specifically, their ability to interact and regulate a variety of

Page 160: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

139

different immune cells [240]. Thus, it is imperative to consider the role of neutrophils when

investigating the initiation of IDRs.

Page 161: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

140

CHAPTER 5

INVESTIGATION OF THE IMMUNE CHANGES

INDUCED BY AMINOGLUTETHIMIDE

Winnie Ng, Imir G. Metushi, Libia Vega-Loyo, and Jack Uetrecht

Page 162: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

141

5.1 ABSTRACT

Idiosyncratic drug reactions (IDRs) are a significant issue of concern primarily because

of their unpredictable nature. The mechanisms of IDRs are not fully understood, although most

IDRs exhibit characteristics that suggest they are immune-mediated. The primary aromatic

amines are associated with a high incidence of IDRs. This is presumably because they can be

easily oxidized to reactive metabolites that covalently bind to proteins, and they also redox cycle,

which could induce oxidative stress. Protein modification and cell stress may activate the

immune system, which in some patients may lead to an IDR. Previous studies of rodents treated

with aromatic amine drugs found few immune changes in the liver; however, aromatic amines

can also be oxidized in cells of the immune system, and this may be more important for the

induction of an immune response. The objective of this study was to determine the immune

response to one representative aromatic amine, aminoglutethimide (AMG) in rats. AMG

treatment increased serum levels of the chemotactic proteins Gro/KC, MCP-1, and IP-10.

Increased cell proliferation was observed in the white pulp of the spleen after 14 days of AMG

treatment, and this was attributed to an increased proliferation of CD4 T-cells. Interestingly, a

lymphocyte transformation test (LTT) using splenocytes was negative, but it was positive when

using lymphocytes from the auricular lymph nodes. There were also transient increases in

intracellular IL-17 and IL-10 in lymphocytes from the spleen and lymph nodes following AMG.

An increase in M2 macrophages was observed and this may be part of the immune response that

prevents a pathogenic immune response. In short, although AMG does not cause overt toxicity, it

induces an immune response that is consistent with the development of immune tolerance.

Page 163: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

142

5.2 ABBREVIATIONS

AMG, aminoglutethimide

PC, allophycocyanin

APCs, antigen presenting cells

DMSO, dimethyl sulfoxide

FACS, florescence-activated cell sorting

FITC, fluorescein isothiocyanate

H&E, hematoxylin and eosin

IDRs, idiosyncratic drug reactions

LTT, lymphocyte transformation test

MPO, myeloperoxidase

P450, cytochrome P450

PBS, phosphate buffered saline

PE, phycoerythrin

PerCP, peridinin chlorophyll

PMA, phorbol myristate acetate

Page 164: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

143

5.3 INTRODUCTION

Idiosyncratic drug reactions (IDRs) are of great concern to the pharmaceutical industry

primarily due to their unpredictable nature. Promising new drug candidates often have to be

withdrawn from the market or achieve a “black box” warning because of a serious IDR that was

not detected during clinical trials. If the risk that a drug candidate would cause serious IDRs

could be predicted, or if the patients at high risk could be identified it would have a profound

effect on drug development and therapy.

The mechanisms of IDRs have yet to be fully elucidated; however, there is increasing

evidence that most IDRs are caused by reactive metabolites and are immune-mediated [13]. The

delay in onset and rapid onset upon re-challenge, seen commonly with IDRs, provides indirect

evidence of a T-cell response that requires time for T-cell clonal expansion and a subsequent

memory response upon re-stimulation with the drug. Moreover, the formation of anti-drug and

anti-nuclear antibodies in some patients further implicates the involvement of the immune

system. Thus, the majority of IDRs are thought to be due to an adaptive immune response to

drug-modified proteins that can become antigenic or induce danger signals through the induction

of cellular stress [241].

Traditionally the liver has been a main focus of drug metabolism and toxicity studies

because it has high levels of metabolic enzymes such as the cytochrome P450s (P450s) that can

oxidize drugs to chemically reactive metabolites. Although idiosyncratic drug-induced liver

injury is a major cause for drug withdrawl [242], the relevancy of using liver tests as a universal

indicator of drug toxicity is questionable, especially with drugs for which the primary target of

toxicity is not the liver.

Page 165: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

144

Drug metabolizing enzymes such as P450s have been found in peripheral blood

mononuclear cells in humans [243]. Although this could potentially contribute to reactive

metabolite formation, their expression is presumably much less than in the liver. In contrast,

neutrophil activation was observed upon stimulation with prohaptenic chemical sensitizers, and

this effect was attributed to the presence of myeloperoxidase (MPO) in these cells that was not

present in the other cell types tested [244]. MPO is a heme-containing enzyme primarily

involved in the production of strong microbicidal oxidants, and it is found within the azurophilic

granules of a variety of immune cells including neutrophils, monocytes, and macrophages [245].

Oxidation of drugs to reactive metabolites by MPO has been implicated in the mechanisms of

IDRs [246]. Alternatively, there are many enzymes other than MPO that are found within

immune cells which may have the capability to oxidize drugs [149]. Thus, the localized

production of reactive metabolites and oxidative stress within immune cells may provide a

stronger stimulus for the initiation of an immune response.

The primary aromatic amine moiety is a notorious structural alert for drug development

because the majority of aromatic amine drugs are associated with a high incidence of IDRs [94].

Although aromatic amine drugs can induce liver injury, it is one of the less common reactions

induced by these drugs, and liver injury is more likely a result of a more generalized

hypersensitivity reaction. Studies in rodents to test the changes in hepatic gene expression

induced by several aromatic amine drugs: sulfamethoxazole (SMX), dapsone (DDS), and

aminoglutethimide (AMG), found few changes that were suggestive of liver injury or the

initiation of a cellular immune response in the liver [37, 212]. Additionally, a comprehensive

study of the liver to determine the effects of AMG, an aromatic amine aromatase inhibitor, found

few patho-histological changes, and only minor changes in genes involved with the innate

Page 166: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

145

immune system (Ng et al.; submitted). Despite the capacity for the liver to mount an immune

response, in comparison to other organ systems, the dominant immune response in the liver is

immune tolerance [247].

Aromatic amines are easily oxidized and can be oxidized by enzymes other than P450s

such as MPO [248], and some aromatic amine drugs have been shown to be oxidized in vitro by

activated neutrophils to hydroxylated metabolites [235, 249]. Interestingly, the aromatic amine

anti-microbial SMX has been extensively studied in models of immunogenicity [250]. In

particular, antigen presenting cells (APCs) were found to metabolize SMX to reactive species,

and these drug-modified cells were able to induce an immunogenic response in lymphocytes

from SMX hypersensitive individuals [49]. Thus, testing the direct effect of aromatic amines on

immune cells may be a better approach to understanding the mechanisms of IDRs caused by

these drugs.

In general, the reactivity of aromatic amines is thought to be due to their ability to form

common reactive metabolites such as the highly electrophilic nitrosoamine. Since aromatic

amines can be metabolized to reactive species that could covalently bind to endogenous

molecules to produce antigenic substances and can also redox cycle to cause oxidative damage, it

is quite possible that aromatic amine drugs could induce IDRs through danger mechanisms

[251]. Given that aromatic amine drugs can be oxidized within immune cells, the objective of

this study was to investigate the role of the immune system in aromatic amine-induced IDRs by

characterizing the AMG-induced immune changes and determining how they may be involved in

initiating a cellular immune response. In these studies, AMG was used as a prototypical aromatic

Page 167: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

146

amine drug because our previous experiments found that AMG induced the most biological

changes compared to other aromatic amine drugs tested.

Page 168: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

147

5.4 MATERIALS AND METHODS

5.4.1 Chemicals and Reagents

AMG was purchased from Toronto Research Chemical (North York, ON).

Methylcellulose, dextran, 5-bromo-2′-deoxyuridine (BrdU), ethylenediaminetetraacetic acid

(EDTA), Tris-base, sodium chloride, calcium chloride, magnesium chloride, phorbol myristate

acetate (PMA), ionomycin, sodium bicarbonate, 10% neutral buffered formalin, xylene,

hematoxylin, and Cytoseal were obtained from Sigma-Aldrich (Oakville, ON). Phosphate

buffered saline (PBS), heat inactivated fetal bovine serum (FBS), RPMI-1640, 2-

mercaptoethanol, and trypan blue were purchased from Life Technologies (Burlington, ON).

Potassium bicarbonate, dimethylsulfoxide (DMSO), hydrogen peroxide, and Tween20 were

obtained from BioShop Canada Inc. (Burlington, ON). Ammonium chloride was purchased from

ACP Chemical (Montreal, QC). Methanol was acquired from Caledon Laboratory Chemicals

(Georgetown, ON). Ethanol was purchased from Commercial Alcohols (Brampton, ON). Saline

was obtained from Baxter (Mississauga, ON). Epitope Unmasking Solution and Antibody

Dilution buffer were obtained from ProHisto (Columbia, SC). Lymphoprep was obtained from

Axis-Shield (Oslo, Norway).

5.4.2 Antibodies

Anti-rat IL-10 phycoerythrin (PE; clone: JES5-16E3), anti-rat CD32 (FcγII; clone: D34-

485), 7-amino-actinomycin D (7-AAD), anti-BrdU fluroescein isothiocyanate (FITC), and

GolgiStop were acquired from BD Biosciences (San Jose, CA). Anti-rat CD4 FITC (clone:

W3/25), anti-rat NKRP1 PE (clone: 10-78), and anti-rat CD45RA/B FITC (clone:MRC OX-33)

were purchased from Cedarlane (Burlington, ON). Anti-rat CD3 Pacific Blue (clone: IF4), anti-

rat CD163 Alexa647 (clone: ED2), anti-rat CD68 (clone: ED1), and anti-rat CD11c Alexa647

Page 169: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

148

(clone: 8A2) were obtained from AbD Serotec (Raleigh, NC). Anti-mouse/rat Ki-67 peridinin

chlorophyll (PerCP; clone: SolA15), anti-rat CD8a PECy7 (clone: OX8), anti-mouse/rat IL-17

allophycocyanin (APC; clone: ebio17B7), anti-rat MHCII APC (clone: HIS19), and anti-

mouse/rat TNF-α PECy7 (clone: TN3-19) were purchased from eBiosciences (San Diego, CA).

Anti-rat Ki-67 (clone MIB-5), rabbit anti-mouse biotinylated IgG, and streptavidin-HRP were

acquired from Dako (Burlington, ON).

5.4.3 Animals

Male Brown Norway rats (176-200 g) were purchased from Charles River (Montreal,

QC). Animals were housed under standard conditions with automatic watering, 12:12 hr

light/dark cycle and a temperature of 22 °C. Food and water were provided ad libitum and all

animals were given standard rodent chow. Animals were acclimatized for one week before the

start of experiments. The experiment protocol was approved by and performed in accordance

with the University of Toronto Faculties of Medicine and Pharmacy Animal Care Committee.

5.4.4 Treatments

Rats were treated with AMG (125 mg/kg/day in 0.5% methylcellulose) through oral

gavage. Control animals were given methylcellulose vehicle only. Each group had 4 rats each

unless otherwise stated.

5.4.5 Measurement of Serum Cytokines

Blood was collected from the tail vein of rats. Samples were analyzed at the University

Health Network Microarray Centre (Toronto, ON) using a Rat Cytokine/Chemokine Kit

(MILLIPLEX MAP, Millipore) that contained a 23-Plex Premix bead set according to

Page 170: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

149

manufacturer’s protocol. Data was read with a Luminex 100 instrument and analyzed using Bio-

Plex Manager 6.0 software (Bio-Rad Laboratories, Mississauga, ON).

5.4.6 Detection of Cell Proliferation in Lymphoid Organs

Spleen and auricular lymph node sections were excised from the rat and immediately

immersed in 10% formalin solution. H&E and unstained paraffin-embedded slides were prepared

by the Division of Pathology at the Hospital for Sick Kids (Toronto, Ontario).

Immunohistochemistry was carried out using standard procedures. Briefly, slides were

deparaffinised with xylene and serial dilutions of ethanol. Heat-induced antigen retrieval was

performed using Epitope Unmasking Solution (ProHisto, Columbia, SC) and boiling the slides

for 20 min. Endogenous peroxidases were blocked with 3% hydrogen peroxide in PBS for 10

min. Washes were performed using PBS-0.05% Tween20 solution and antibodies were diluted in

Antibody Dilution Buffer (ProHisto). The primary antibody for Ki-67 (1:25 dilution) was

incubated for 30 min. Rabbit anti-mouse biotinylated IgG (1:250 dilution) was used as a

secondary antibody for 30-40 min. Streptavidin-HRP (1:500 dilution) was incubated for 15 min

after which 3’3-diaminobenzidine (DAB, Burlingame, CA) was used for visualization and

haematoxylin was used as a counter-stain. Slides were dehydrated with ethanol and xylene

before slip covers were mounted with Cytoseal (Sigma). A Zeiss fluorescence microscope with

deconvolution was used to obtain histological pictures at the Microscopy Imaging Lab (Faculty

of Medicine, University of Toronto).

5.4.7 Phenotyping of Immune Cells

Spleen and lymph nodes were removed, placed in florescence-activated cell sorting

(FACS) buffer (5% FBS in PBS solution), and a single cell suspension was made using the blunt

end of a syringe through a 40 µm nylon mesh filter. Red blood cells were lysed with red blood

Page 171: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

150

cell lysis buffer (155 mM NH4Cl, 10 mM KHCO3, 0.1 mM EDTA) after which FACS buffer was

added and cells were washed twice before cell counting using trypan blue and a Countess

Automated Cell Counter. Each sample consisted of 1.0 x 106 cells in a total volume of 100 µL

per well. Non-specific binding was blocked by the addition of anti-rat CD32 (FcγII Receptor)

antibody for 15 min. After several washes using FACS buffer, primary antibodies were

incubated for 30 min before washing and resuspension in FACs buffer for analysis.

For intracellular cytokines, 1.0x106 cells were stimulated for 4 h with in culture medium

containing RPMI 1640, 10% FBS, 50 ng/mL PMA, 1 µg/mL inomysin, and GolgiStop prior to

the staining of extracellular antibodies. Cells were then fixed and permeablized with Fix/Perm

solution (BD) for 15 min before the addition of intracellular antibodies in Permeabilization

Buffer for 1 h. After several washes cells were re-suspended in FACs buffer before analysis.

5.4.8 Characterization of an Immune Response through the Lymphocyte

Transformation Test (LTT).

After 14 days of AMG treatment, a single cell suspension of spleen and auricular lymph

node cells was prepared as previously described. Cells extracted from the lymph nodes were

used neat, whereas, in the spleen, lymphocytes were extracted using Lymphoprep (Axis-Shield)

according to manufacturer’s instructions. For each reaction, 1x106 cells were incubated with 10

µM BrdU and various concentrations of AMG (in 30 µL DMSO) in a final reaction volume of 1

mL culture medium (RPMI 1640, 10% FBS, antibiotics, 50 µM β-mercaptoethanol, 1 µg/mL

indomethacin). Cells were incubated for 72 h at 37°C with 5% CO2 after which the reaction was

stopped. Cells were washed with FACS buffer and BrdU was detected using the FITC BrdU

Flow Kit (BD Biosciences) as per the manufacturer’s protocol.

Page 172: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

151

5.4.9 Flow Cytometry Analysis

Flow cytometry experiments were run at the Faculty of Medicine Flow Cytometry

Facility (University of Toronto) on the BD FACSCanto II (BD Biosciences) using BD

FACSDiva Software (BD Biosciences). The flow rate was no more than 400 events/s, and

FlowJo Software (Tree Star, Inc., Ashland, OR) was used for detailed analysis of the flow

cytometry output.

5.4.10 Data Analysis

Statistical analysis was performed using GraphPad Prism 5 (GraphPad Software Inc., La

Jolla, CA). Unless otherwise stated, ANOVA and Bonferroni post-test were used to determine

the statistical significance between AMG and control groups.

Page 173: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

152

5.5 RESULTS

5.5.1 Systemic Immune Response

Rats did not develop any overt signs of an immune or toxic response to AMG treatment

up to 14 days. At early time-points an increasing trend in innate immune factors was observed in

the AMG-treated as compared to control (Figure 5.1), which suggests early immune changes

because Gro/KC, MCP-1, and IP-10 are associated with neutrophil, monocyte, and T-cell

chemotaxis, respectively. Although these results were not statistically significant, this could be a

type 2 error because there does look to be a difference between AMG-treated and control and

perhaps an increase in sample size would make this change more apparent. Treatment with AMG

for up to 14 days also induced innate cytokine changes (Figure 5.2). Again Gro/KC and MCP-1

were up-regulated; however, a significant increase in IL-5 was also observed with a concurrent

significant down-regulation of GM-CSF. IL-5 is a cytokine associated with immunoglobulin

production in B-cells, and its transient increase by AMG may suggest induction of a tolerogenic

mechanism. We have previously reported an increase in blood neutrophils in rats treated with

AMG [216]. Thus, it is possible that the down-regulation of GM-CSF in the current study may

represent a negative feedback mechanism to inhibit granulocyte and monocyte production and

maturation. Therefore the AMG-induced systemic changes are predominantly of an innate

nature.

Page 174: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

153

A B

C

24 480

200

400

600

800

Control

AMG

Time (h)

Gro

/KC

(pg

/ml)

24 480

2000

4000

6000

Control

AMG

Time (h)

MC

P-1

(p

g/m

l)

24 480

50

100

150

200

250

Control

AMG

Time (h)

IP-1

0 (

pg

/ml)

Figure 5.1. Early changes in serum levels of Grow/KC, MCP-1, and IP-10 induced by AMG

treatment. Open bars represent control; solid bars represent AMG-treated. Values are expressed

as mean ± SEM (n=3 control, n=4 AMG). None of the results were statistically significant.

Page 175: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

154

A B

0 7 140

1000

2000

3000

**

Control

AMG

Day

Gro

/KC

(p

g/m

l)

0 7 140

1000

2000

3000Control

AMG

Day

MC

P-1

(p

g/m

l)

C D

0 7 140

20

40

60

80

Day

*

Control

AMG

IL-5

(p

g/m

l)

0 7 140

50

100

150

200

*

Control

AMG

Day

GM

-CS

F (

pg

/ml)

Figure 5.2. Later changes in serum cytokines induced by AMG treatment. (A) Gro/KC, (B)

MCP-1, and (C) IL-5. Alternatively, AMG induced a significant decrease in GM-CSF. Open bars

represent control; solid bars represent AMG-treated rats. Values are the mean ± SEM (n=4

control, n=6 AMG). *p<0.05, **p<0.01 compared to control.

Page 176: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

155

5.5.2 Effect of AMG on Lymphoid Organs

Cellular proliferation was measured in the auricular lymph nodes and the spleen through

the expression of the nuclear marker Ki-67. Although no changes in cellular proliferation were

observed in the auricular lymph nodes, a significant increase in Ki-67-stained cells was observed

in the spleen (Figure 5.3). The Ki-67 positive cells were clustered in the white pulp of the spleen

in AMG-treated rats, whereas in the control the proliferating cells were sparsely distributed

throughout the spleen. This suggests that AMG induces immune cell proliferation because the

white pulp is important for immune activation, and that is where the B- and T-cells are primarily

found within this organ. FACS analysis subsequently found that the proliferating cells were

CD4+ CD3

+ T-lymphocytes (Figure 5.4). It should be noted that, although the magnitude of the

changes between the methodologies vary quite noticeably, this is presumably due to the focal

nature of the proliferation, which was not as easily detected by analyzing the whole organ with

FACS. More importantly, the functional impact of these changes was investigated to determine

whether the proliferation of T-lymphocytes induced by AMG translated into an adaptive immune

response as measured by the lymphocyte transformation test (LTT). Despite the T-lymphocyte

proliferation found in the spleen, there was no change upon re-stimulation with AMG; however,

lymphocytes from the auricular lymph nodes did stimulate a slight increase in proliferation in

response to re-stimulation with AMG concentrations of 50 and 100 µg/mL (Figure 5.5),

suggesting a small secondary adaptive response to the drug; however, the magnitude of these

changes were too small to be certain.

Page 177: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

156

Figure 5.3. Cell proliferation detected by Ki-67 staining induced in the white pulp of the

spleen after 14 days of AMG treatment. (A) control, (B) AMG-treated rat. 20x magnification.

A

B

Page 178: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

157

A

B CD3+ CD45RAB

0

5

10

15Control

AMG

Percen

t K

i-67+

Cells in

Lym

ph

ocyte

Su

bp

op

ula

tio

n

CD4+ CD8+ NKRP1+0

5

1015

20

2560

70

80Control

AMG

Perc

en

t K

i-67

hig

h/C

D3+

C

CD11c NKRP1 CD68 CD1630

20

40

60

80

**Control

AMG

Perc

en

t K

i-67

hig

h C

ells

Figure 5.4. Analysis of proliferating (Ki-67+) splenocytes after 14 days of AMG treatment.

(A) Lymphocyte subpopulation (B) T-cell subpopulation and (C) innate cell subpopulations.

Open bars represent control, solid bars represent AMG-treated rats. Values are the mean ± SEM

(n=4). **p<0.01 as compared to control.

Page 179: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

158

Control 1

6.25 12.5 25 50 1000.0

0.2

0.4

0.6

0.8

DMSO

AMG (ug/ml)

Brd

U (

Perc

en

t S

tain

ed

in

Lym

ph

ocyte

Gate

)

Control 2

6.25 12.5 25 50 1000.0

0.2

0.4

0.6

0.8

DMSO

AMG (ug/ml)

Brd

U (

Perc

en

t S

tain

ed

in

Lym

ph

ocyte

Gate

)

AMG 1

6.25 12.5 25 50 1000.0

0.2

0.4

0.6

0.8

DMSO

AMG (ug/ml)

Brd

U (

Perc

en

t S

tain

ed

in

Lym

ph

ocyte

Gate

)

AMG 2

6.25 12.5 25 50 1000.0

0.2

0.4

0.6

0.8

DMSO

AMG (ug/ml)

Brd

U (

Perc

en

t S

tain

ed

in

Lym

ph

ocyte

Gate

)

Figure 5.5. LTT using lymphocytes from the lymph nodes of AMG-treated rats. Rats were

treated with AMG (125 mg/kg/day) for 14 days before lymphocytes were extracted from the

lymph nodes. Ex vivo cell proliferation was measured using BrdU after 72 h re-stimulation with

various concentrations of AMG. Each graph represents one animal (n=2).

Page 180: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

159

5.5.3 AMG-Induced Immune Cell Changes

AMG induced very subtle changes in the phenotype of immune cells despite various

attempts to characterize changes in immune cell subsets. No significant changes were found in

CD4-, CD8-, CD45RAB-, CD11c-, or NKRP1-positive cells at both early (24 h) and late (14

days) time-points in the spleen or lymph nodes upon AMG treatment (data not shown). However,

subtle changes were observed in intracellular cytokines. A slight elevation in both IL-17- and IL-

10-positive cells at early time-points was observed upon AMG treatment in the spleen and

auricular lymph nodes (Figure 5.6). Although not statistically significant, up-regulation of both

would suggest that there may be a modulating effect occurring fairly early upon treatment.

Furthermore, an increase was observed in the M2 alternative macrophage subset upon AMG

treatment, which may be modulating the immune response as well (Figure 5.7).

Page 181: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

160

A B

Control 24 h AMG 48 h AMG0.0

0.2

0.4

0.6

0.8

IL-1

7+

(perc

en

t C

D3

+C

D4

+ c

ell

s)

*

Control 24 h AMG 48 hAMG0.0

0.1

0.2

0.3

0.4

0.5

*

IL-1

7+

(pe

rce

nt

CD

3+C

D4

+ c

ells)

C D

Control 24 h AMG 48 h AMG0.0

0.1

0.2

0.3

0.4

0.5

IL-1

0+

(Pe

rce

nt

CD

3+C

D4

+cells)

Control 24 h AMG 48 h AMG0.0

0.1

0.2

0.3

0.4

0.5

IL-1

0+

(Pe

rce

nt

CD

3+C

D4

+cells)

Figure 5.6. Early changes in the fraction of CD4+ T cells from the spleen and lymph nodes

that express IL-17 or IL-10 induced by AMG treatment. (A and C) cells from the spleen and

(B and D) cells from auricular lymph nodes. Values are mean ± SEM (n=4). *p<0.05 compared

to control.

Page 182: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

161

M1 M2ab M2c0

5

10

15

20

25* Control

AMG

Perc

en

t L

N M

acro

ph

ag

es

Figure 5.7. AMG-induced changes in macrophage subsets in lymph nodes after 14 days of

treatment. For phenotyping, M1 was defined as MHCII+/CD68

-, M2ab was defined as

MHCII+/CD68

+, and M2c was defined as CD68

+/MHCII

-. Open bars represent control; solid bars

represent AMG-treated rats. Values are the mean ± SEM (n=4 control, n=6 AMG). *p<0.05

compared to control.

Page 183: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

162

5.6 DISCUSSION

The fact that primary aromatic amines are associated with a high incidence of IDRs

independent of their therapeutic class presumably reflects the fact that they are readily oxidized

to reactive metabolites. The liver is the major site of drug metabolism and reactive metabolite

formation; however, with the exception of bromfenac, liver injury associated with aromatic

amines is usually part of a more general hypersensitivity reaction rather than being confined to

the liver. The liver contains high concentrations of P450s that can oxidize a wide variety of

xenobiotics to reactive metabolites including components of the diet; therefore, the liver also

possesses multiple protective mechanisms such as high levels of glutathione and detoxifying

enzymes. In addition, the major immune response to covalent binding in the liver appears to be

immune tolerance. Even when a xenobiotic is able to cause an immune response leading to liver

injury, it appears that activation of the immune system outside of the liver is required to

overcome immune tolerance in the liver [211].

The ease of aromatic amine oxidation also means that it can be oxidized by enzymes

other than P450s such as MPO [94]. The fact that neutrophils and their precursors contain high

levels of MPO [252] is presumably why most aromatic amines can cause agranulocytosis. APCs

also contain MPO and bioactivation of SMX has been demonstrated in APCs [49, 155].

AMG is a first generation aromatase inhibitor and is primarily indicated for the treatment

of breast and prostate cancer. Severe IDRs associated with AMG include: skin rashes,

cholestasis, blood dyscrasias (most notably agranulocytosis), and systemic lupus erythematosus

[253]. The more commonly reported AMG-induced IDRs are related to its effects on specific

immune cells such as the granulocyte progenitors and is delayed in onset [225]. Compared to

other aromatic amine drugs, AMG does not have an electron withdrawing group para to the

Page 184: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

163

aromatic amine moiety, such that the electron density in the aromatic ring is higher and AMG

may be more susceptible to oxidation than other drugs with an aromatic amine moiety. In

preliminary studies, neutrophils have been found to be capable of oxidizing AMG to a

hydroxylated species in vitro (unpublished observations). Thus, AMG has the potential to elicit

immune effects to initiate IDRs.

Previously we had found that treatment of rats with AMG did not lead to liver injury;

however, in this study we found an immune response to AMG including proliferation of

leukocytes in the white pulp of the spleen and an increase in Gro/KC (CXCL1) and IL-5 after 7

days of treatment that returned to normal at 14 days. IL-5 is involved with immunoglobulin

production in B-cells; however it takes time for antibodies to be produced and it is unlikely that

an antibody response would be detected as early as 7 days. In contrast, elevated IL-5 has been

found in patients with eosinophil-associated drug hypersensitivity, and the increase in IL-5

preceded the elevation in eosinophils, which is known to be driven by IL-5, suggesting that it

could be a biomarker to predict the onset of drug hypersensitivity syndrome [254]. Interestingly,

GM-CSF is also associated with the activation of eosinophils and the successive decrease in GM-

CSF observed with AMG is consistent to a decrease in or modulation of a response. GM-CSF is

a cytokine involved primarily with granulopoiesis but is also associated with the activation of

granulocytes such as neutrophils and implicated in a variety of inflammatory diseases [255]. The

decrease of GM-CSF with AMG may also be a negative feedback mechanism to the increase in

granulocyte progenitors induced by AMG in the bone marrow that was previously observed

[216].

Page 185: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

164

Immune cell infiltration was not observed within the lymphoid organs; however, a

significant increase in cellular proliferation was found in the white pulp of the spleen upon AMG

treatment, and this was attributed to the proliferation of CD4 T cells. Further investigation found

that these changes may potentially be more of a localized effect rather than an immune response

to drug treatment, as re-stimulation of splenocytes through the LTT found no proliferation. The

LTT is a diagnostic tool to determine drug hypersensitivity and is dependent on the formation of

memory lymphocytes after a subsequent primary immune sensitization. It is likely that the

immune response is to the reactive metabolite of AMG, rather than to the parent drug and the

immune response was insufficient to lead to a large number of memory T cells that cross react

with the parent drug. Unfortunately, a drug-modified protein is currently unavailable for testing,

and a negative LTT does not necessarily indicate lack of immune response. In contrast, no

cellular proliferation was observed in the auricular lymph nodes, yet these cells were able to

induce a subtle proliferative response to AMG upon secondary re-stimulation. The doses of

AMG that increased lymphocyte proliferation were approximately two times higher than the

blood levels of AMG found in patients [256], and only one animal had only a marginal

stimulation index at this concentration so the results are of questionable significance. An in vitro

method was recently reported to detect drug hypersensitivity in naive T cells co-cultured with

dendritic cells [93], and this method may be a better method to determine whether the drug is

able to elicit primary sensitization. Nevertheless, these immune changes may be below the

threshold to elicit a systemic immune response, which may be why rats do not experience

pathology.

In terms of a cell-specific response, a subtle up-regulation of IL-17- and probably IL-10-

producing CD4+ T cells was observed in the spleen at 24 h, as well as an increase in M2

Page 186: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

165

macrophages in lymph nodes after 14 days. Early increases in both IL-17 and IL-10 upon AMG

treatment suggest modulation of an immune response. IL-17 is a proinflammatory cytokine that

can be released from Th17 cells to mediate inflammatory downstream effects [257], and has

been found to be increased as early as 2 h in mice with acetaminophen-induced liver injury,

which suggests that this cell type has innate functions in addition to its adaptive functions [258].

In contrast, IL-10 is an anti-inflammatory cytokine released from T-regulatory cells to dampen

immune responses such that early up-regulation by AMG may suppress immune activation.

Macrophages have also been implicated in the pathogenesis of drug-induced reactions in

both an inflammatory and protective capacity due to the heterogeneity of different macrophage

subsets. AMG induced an increase in the M2 macrophage subset, which may prevent a

pathological immune response. M2 alternatively activated macrophages were found to be

induced in the liver of acetaminophen-treated mice and depletion of these macrophages

prolonged acetaminophen-induced liver injury [259]. This suggests that macrophages may play a

protective role to prevent the initiation of a response, and this may be what is occurring with

AMG. In our studies, M2 macrophages were identified only by differential expression of CD68

and MHCII. Further investigation and more detailed phenotyping of these macrophages with M2

markers such as CCR2, arginase 1, Ym1, etc. will be required to determine more accurately what

role these macrophages play in AMG-induced IDRs.

From our findings, there was data to suggest that the main deterrent to AMG-induced

IDRs may be early immune modulating events that prevent immune activation. These changes

are consistent with an immune response that ends in immune tolerance. Given that AMG-

induced IDRs are idiosyncratic in humans it is not surprising that a clinically evident adverse

Page 187: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

166

reaction did not occur in these rats but evidence of an immune response may be a biomarker to

predict which drug candidates are likely to cause IDRs in susceptible individuals.

Page 188: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

167

CHAPTER 6

CONCLUSIONS AND FUTURE DIRECTION

Page 189: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

168

6.1 DISCUSSION AND CONCLUSIONS

As our understanding of the events that lead up to an IDR are limited, any mechanistic

clues as to how these reactions may lead to an immune response and subsequent pathology

would have profound implications on the prediction, prevention, detection, and treatment of

IDRs. In our studies, the approach has not been to focus on a specific type of organ injury, but

rather to concentrate on a group of drugs that are frequently associated with IDRs. The general

hypothesis of my thesis research is that aromatic amine drugs may induce IDRs through immune

mechanisms. However, currently there is no valid animal model of aromatic amine-induced IDRs

in which to fully test this hypothesis; animal models of IDRs are more difficult to develop than

merely administering the drug alone [61]. Nevertheless, monitoring the immune changes induced

by aromatic amine drugs may reveal how these drugs activate or suppress the immune system,

which may perhaps lead to the development of an animal model. In the case of the absence of a

reaction, this information may be very important to understanding why the majority of the

population are tolerant to IDRs, and identifying the factors that overcome this suppression will

be essential to understanding individual susceptibility to IDRs.

Our initial studies were an extension of previous findings in our lab that tested the danger

hypothesis in SMX-treated mice in which few hepatic gene changes were observed [37]. These

findings were surprising, because aromatic amines can be fairly easily oxidized to reactive

metabolites, and we were curious to see whether other aromatic amine drugs induced similar

changes and whether these changes were species-specific. However, even in Brown Norway rats

treated with SMX, the hepatic gene changes were minimal and varied quite significantly from

the other aromatic amine drugs tested: DDS and AMG. Since SMX and DDS are structurally

quite similar, it is disconcerting to see such differences. One explanation may have been the

Page 190: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

169

differences in the dosages, although DDS was given at a much lower dose than SMX, and

dosages were designed to produce similar blood levels to those found in patients for each drug.

Another possible explanation may be the pharmacology of the drugs wherein DDS has a higher

volume of distribution and protein binding than SMX [98]. Variations in drug localization could

also account for these differences because SMX-OH was found to bind to the cell surface

whereas the N-hydroxyl DDS metabolite was found to form adducts intracellularly [203].

The similarities observed between DDS and AMG in the upregulation of genes involved

with the Keap1-Nrf2-ARE pathway are commonly found with drugs that induce IDRs [37, 204,

205]. Although, changes in genes regulated by the Keap1-Nrf2-ARE pathway are not direct

indicators of cellular stress, expression of these genes implies the presence of a cellular insult

that requires these genes to respond. We also focused on the downstream effects of protein

expression, and more importantly, protein functionality. However, our studies on thioredoxin

reductase and glutathione S-transferase activity found that gene changes do not necessarily

translate to functional changes. Thus, although gene changes could be biomarkers to predict

drugs associated with IDRs, focusing on protein pathways may be a better approach for

elucidating mechanisms.

Interestingly, despite the lack of many changes in gene expression, Eiih was up-regulated

at acute time points by all aromatic amine drugs tested and also by other drugs associated with

IDRs such as clozapine and nevirapine. Although the function of this gene is yet to be

determined, our findings show that it has the potential to act as a biomarker of drugs associated

with IDRs. However, due to the lack of commercially available reagents, we are unable to

investigate the mechanistic implications of Eiih expression further than testing several drugs

Page 191: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

170

associated with IDRs and also testing for the lack of Eiih up-regulation in drugs not associated

with IDRs. Eiih was found to be up-regulated by insulin in rats [187], and it has been reported

that insulin plays a role in maintaining innate immunity [260]. Thus, Eiih could potentially be

involved in the immune response. It should also be noted that the change in expression of Eiih

induced by SMX using RT-PCR was not significant; therefore, a lack of Eiih response does not

demonstrate that a drug cannot cause IDRs. This illustrates the difficulty of finding good

biomarkers that are universally applicable. Nevertheless, this is the first systematic study of

drugs containing the primary aromatic amine moiety in the context of their common association

with IDRs through a global screen of hepatic gene expression. Not surprisingly, even for drugs

with the same chemical moiety and the potential to form common reactive metabolites, it seems

that most of the effects are drug-specific. Given the lack of commonality between drugs that are

chemically quite similar, finding universal tests to predict IDRs may not be possible, and it is

more likely that each drug elicits characteristic effects based upon where it is accumulated within

the body.

Our concern for the lack of significant changes in the liver induced by aromatic amine

drugs led to a comprehensive study in the liver to test the changes induced by AMG. AMG is

structurally and functionally significantly different from SMX and DDS, and it was chosen for

its uniqueness despite containing the aromatic amine moiety. Greater changes had been expected

with AMG because it induced the most significant changes in gene expression, and so it was

used as a prototypical aromatic amine drug. Even so, few physiological changes were found in

the liver with AMG despite monitoring up to 14 days. AMG has been reported to induce P450

and decrease its own half-life as well as that of other drugs given concurrently [176, 215];

therefore, it may be possible that increased AMG clearance led to a decrease in response. On the

Page 192: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

171

other hand, increased metabolism could also lead to the generation of more reactive metabolites

leading to greater hepatocyte damage. However, no matter what mechanism, the liver seems to

be able to handle the insult because AMG did not induce liver injury at the time points tested.

In terms of gene expression, AMG induced differential gene changes associated with the

apoptotic and mitochondrial pathways, although there was no evidence for apoptosis as

determined by the TUNEL assay or mitochondrial injury in the form of microvesicular steatosis.

Despite this, these changes could be very important as danger signals to initiate an immune

response because APCs were only activated in the presence of cell death in a rodent model of

SMX-induced immunogenicity [49]. Apoptosis in the liver also has broader implications because

apoptotic hepatic leukocytes may mediate a tolerogenic effect whereas apoptosis of hepatocytes

may stimulate the immune system through the release of DAMPs. Unfortunately we were unable

to differentiate which cells undergo apoptosis because the gene expression of the liver was tested

as a whole rather than separating the hepatocytes from immune cells. Hepatic gene changes were

also observed in innate immune genes, which also suggests that innate immunity is able to

mitigate the toxic effects of the drug.

Interestingly, aromatic amine drugs rarely induce direct liver injury in humans, but rather

hepatotoxicity is usually due to indirect effects of generalized hypersensitivity. Although AMG

does induce symptoms of cholestasis [208], this type of liver injury is not usually life-

threatening. The basis for our focus on the liver as a target organ of aromatic amine drugs was

due to the abundance of metabolizing enzymes in the liver, and the ability of aromatic amines to

be oxidized to reactive metabolites; however, this seems to be disconnected from what is

Page 193: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

172

clinically observed. In spite of this, studying the effects of aromatic amines in the liver has given

us another perspective and shifted our attention to metabolism by immune cells.

Although drug metabolism by immune cells does not make a significant contribution to

the clearance of most drugs, they do possess drug metabolizing enzymes. Neutrophils incubated

with AMG in the presence of ascorbic acid were able to oxidize AMG to a hydroxylated

metabolite (see Appendix 2). It is likely the hydroxylamine but we were unable to identify the

exact location of the oxidation because of a lack of an authentic N-hydroxyl-AMG standard;

however, this data suggest that enzymes within neutrophils are able to oxidize AMG.

The most profound and interesting findings for AMG were the elevation of peripheral

blood neutrophils in rats that was not observed with the other aromatic amine drugs, SMX and

DDS (see Appendix 3). Although it did seem that DDS induced an increase in neutrophils, this

was not statistically significant; however, this is consistent with DDS being associated with a

higher incidence of agranulocytosis than SMX. These changes are opposite to the idiosyncratic

agranulocytosis that is observed in patients treated with AMG [261] and to a mouse model of

AMG-induced leucopenia [262]. We have treated mice with AMG and found a similar decrease

in WBCs as previously reported (see Appendix 4); however, we believe that the rat may be a

more clinically relevant model of idiosyncratic drug-induced agranulocytosis. This is because

rats were found to have more sustained blood levels of AMG than mice (data not shown), and

often patients taking other drugs that are associated with idiosyncratic agranulocytosis such as

clozapine experience an initial neutrophilia that occurs before the onset of agranulocytosis [263].

Thus, we believe that the rat is a better model of the early changes in AMG-induced

agranulocytosis. The AMG studies ended at 14 days because mice were shown to develop

Page 194: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

173

leucopenia as early as 14 days, and so it was assumed that this would be enough time to develop

some cellular response in the peripheral blood; yet it may be possible that if AMG treatment in

the rats were prolonged, the rats may eventually develop a decrease in neutrophils.

Although the kinetics vary slightly, the elevated neutrophils induced by AMG in rats are

similar to what is observed in clozapine-treated rats [264] and rabbits [265]. AMG induced an

increased release of neutrophils from the bone marrow; however, the BrdU-labeled neutrophils

appeared to have a longer half-life in the peripheral blood; this is in contrast to the decrease in

neutrophil half-life observed in clozapine-treated rabbits [265]. We were unable to perform a

more detailed study on neutrophil kinetics because of the small blood volume of rats; however,

we did find a decrease in apoptotic neutrophils upon AMG treatment using annexin V and PI.

This is likely confounded by a decreased expression of phosphatidylserine in newly formed

neutrophils and may not indicate an intrinsically prolonged survival of neutrophils. It is also

unknown whether AMG directly affects the neutrophils and leads to a compensatory production

of granulocyte progenitors in the bone marrow or whether AMG directly effects the bone

marrow. From these findings, it is more likely that AMG affects the bone marrow directly

because there did not seem to be a change in neutrophil clearance. In addition, no change was

observed in the total WBC count because the elevation in neutrophils was accompanied with a

decrease in lymphocytes; the exact implications of this are not known. Neutrophils in AMG-

treated rats also expressed lower levels of CD62L suggesting activation and extravasation into

tissue targets, and future effort should be put forth to monitor neutrophil trafficking in relation to

bone marrow neutrophil production to determine whether this is an important effect of AMG.

The effect of AMG on neutrophils is intriguing and more research should be invested into

Page 195: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

174

determining how this may be related to drug-induced blood dyscrasias, specifically

agranulocytosis.

The majority of the immune changes induced by AMG in the rat were subtle, and even

the most profound immune change, cell proliferation in the spleen, did not lead to a positive

LTT. In isolation, this suggests that AMG may be mitogenic rather than immunostimulatory; yet,

the LTT is not entirely specific and depends essentially on the development of an immune

memory response, which may not be very strong because the rats were only treated for 14 days.

Thus, testing the ability of AMG to induce primary T cell sensitization may be a better approach.

Conversely, in the auricular lymph nodes, no cell proliferation was observed, but slight increases

in T cell proliferation were found upon re-stimulation at higher AMG concentrations. These

results are difficult to interpret because they suggest small localized immune reactions, which

may not be sufficient to stimulate an immune response that becomes pathological. However,

these small immune changes are mechanistically important because they are associated with an

early elevation of IL-17 and IL-10, which suggests a modulating effect that may prevent immune

activation. Throughout these studies there have been changes to suggest immune tolerance that

prevents the initiation of an IDR in the rats. However, it must be noted that although we loosely

call this immune tolerance, immunologists might disagree and call it immune suppression or

modulation of basal immune responses because we have not determined whether these changes

induced by AMG are in response to antigen through activation of the T cell receptor. Perhaps a

more useful approach would be in vitro studies to determine the effect of AMG on specific cell

types to determine whether immune cells are directly activated. In vivo studies are useful for

detecting an overall physiological response; however, it may mask some very important and

Page 196: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

175

essential mechanisms, especially for drugs such as AMG in which the immune response is quite

mild.

Moving forward, it is anticipated that it will be difficult to continue studying AMG in the

context of IDRs because it is no longer commonly used, and obtaining clinical samples to

confirm whether these findings in rats are applicable to humans will be difficult. Although our

initial attempts to discover the factors common to aromatic amine drugs that are responsible for

inducing IDRs were not entirely successful, as the projects evolved, we have found various clues

as to how the body may be responding to aromatic amine drugs and how these changes may be

implicated in IDRs. This is especially true with AMG as our focus gradually shifted to it as a

prototypical aromatic amine drug and greater evidence was found to suggest immune modulation

as a mechanism to prevent the initiation of an immune-mediated IDR.

Page 197: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

176

6.2 FUTURE DIRECTIONS

The current findings for AMG have provided a strong foundation for further investigation

of the mechanisms of aromatic amine-induced IDRs; however, the biggest deterrent for the

success of future studies is the lack of an antibody against AMG that could be used to detect

covalent binding. In fact, this has been a major limitation throughout my thesis research. The

generation of AMG-modified protein was attempted using several methods including through the

formation of a diazonium salt and through the targeting of the heterocyclic ring of AMG; both of

which were not successful. The major chemical limitation seems to be maintaining the stability

of the heterocyclic ring while introducing a reactive site required for protein conjugation.

Nevertheless, if a drug-modified protein were available, an anti-AMG antibody could be

generated. The antibody could then be used for covalent binding studies to determine exactly

where AMG covalently binds and which target organs to focus on. It would also allow for the

determination of which cells and where within the cell AMG binding is localized. In addition,

the AMG-modified protein would allow us to perform more detailed studies on the metabolism

of AMG and the involvement of the reactive metabolite in immune activation in assays such as

the LTT.

Another area of focus would be to determine whether AMG directly activates immune

cells and the functional consequences of such activation. These studies would primarily be

carried out in vitro before confirming the findings in vivo. The effect of AMG on neutrophils is

worth pursuing, and it is essential to determine whether other drugs associated with

agranulocytosis also induce similar changes, for example in CD62L expression. In doing so,

perhaps an expression profile could be developed for drug-induced agranulocytosis that could be

used to test new drug candidates. It is also important to determine the effect of the neutrophils in

Page 198: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

177

the scheme of immune activation, and how it may be involved in an adaptive cellular response.

Previously it was reported that endotoxin-induced CD11cbright

/CD62Ldim

/CD11bbright

/CD16bright

neutrophils were able to inhibit T-cell proliferation in a mechanism dependent on the release of

hydrogen peroxide [160]. These results are quite interesting because the implications are that the

innate immune system may suppress a T cell response, and this protective effect may prevent the

initiation of IDRs and should be investigated as a potential mechanism for AMG.

It is also essential to determine whether a reactive metabolite of the drug is involved with

the immune response and whether this can lead to primary immune sensitization. Park et al. have

had been successful doing this with SMX-NO, in a method that involves co-culture of naive T-

cells with APCs pre-treated with drug [93], but they have not as yet determined whether other

drugs that induce IDRs may act through a similar mechanism. Thus it would be interesting to see

whether this is true for AMG and other aromatic amine drugs, and if it works, this would be

widely applicable to preclinical drug safety testing.

Lastly, development of an animal model of AMG-induced IDRs would provide an

invaluable tool to rigorously test new hypotheses. Given that IDRs do not occur in most patients

that take a drug like AMG it is not surprising that simply giving the drug to animals does not

produce an IDR. A better understanding of the immune effects of drugs, especially their effects

on immune tolerance, may make it easier to develop animal models that could be utilized to

study the detailed steps involved in the initiation of an IDR.

Page 199: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

178

REFERENCES

1. Edwards, I.R. and J.K. Aronson, Adverse drug rections: definitions, diagnosis, and

management. Lancet, 2000. 356: p. 1255-1259.

2. Pirmohamed, M., et al., Adverse drug reactions as a cause of admission to hospital:

prospective analysis of 18 820 patients. BJM, 2004. 329(15-19).

3. Pezalla, E., Preventing adverse drug reactions in the general population. Manag Care

Interface, 2005. 18(10): p. 49-52.

4. Lazarou, J., B.H. Pomeranz, and P.N. Corey, Incidence of adverse drug reactions in

hospitalized patients. JAMA, 1998. 279(15): p. 1200-1205.

5. Light, D.W. and R. Warburton, Demythologizing the high costs of pharmaceutical

research. BioSocieties, 2011. 6: p. 34-50.

6. Lasser, K.E., et al., Timing of new blakc box warnings and withdrawals for prescription

medications. JAMA, 2002. 287: p. 2215-2220.

7. Wysowski, D.K. and L. Swartz, Adverse drug event surveillance and drug withdrawals in

the United States, 1969-2002. Arch. Intern. Med., 2005. 165: p. 1363-1369.

8. Naisbitt, D.J., et al., Immunological principles of adverse drug reactions. Drug Safety,

2000. 23(6): p. 483-507.

9. Uetrecht, J., Idiosyncratic drug reactions: current understanding. Ann. Rev. Pharmacol.

Toxicol., 2007. 47: p. 513-539.

Page 200: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

179

10. Adkinson, N.F., et al., Task force report: future research needs for the prevention and

management of immune-mediated drug hypersensitivity reactions. J. Allergy Clin.

Immunol., 2002. 109: p. S461-478.

11. Uetrecht, J.P., New concepts in immunology relevant to idiosyncratic drug reactions: the

"danger hypothesis" and innate immune system. Chem. Res. Toxciol., 1999. 12: p. 387-

395.

12. Uetrecht, J.P., New concepts in immunology relevant to idiosyncratic drug reactions: the

"Danger Hypothesis" and innate immune system. Chem. Res. Toxicol., 1999. 12: p. 387-

395.

13. Uetrecht, J. and D.J. Naisbitt, Idiosyncratic adverse drug reactions: current concepts.

Pharmacol. Rev., 2013. 65: p. 779-808.

14. Olson, H., et al., Concordance of the toxicity of pharmceuticals in humans and in

animals. Regul. Toxicol. Pharmacol., 2000. 31: p. 56-67.

15. Clay, K.D., et al., Brief communication: severe hepatotoxicity of telithromycin: three

case reports and literature review. Ann. Intern. Med., 2006. 144: p. 415-420.

16. Park, B.K., M. Pirmohamed, and N.R. Kitteringham, Role of drug disposition in drug

hypersensitivity: a chemical molecular, and clinical perspective. Chem. Res. Toxicol.,

1998. 11(9): p. 969-988.

17. Beaune, P.H. and M. Bourdi, Autoantibodies against cytochrome P-450 in drug induced

autoimmune hepatitis. Ann. N.Y. Acad. Sci., 1993. 685: p. 641-645.

Page 201: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

180

18. Neuberger, J.M., J.G. Kenna, and R. Williams, Halothane hepatitis: attempt to develop

an animal model. Int. J. Immunopharmacol., 1987. 9: p. 123-131.

19. Aithal, G.P., et al., Hepatic adducts, circulating antibodies, and cytokine polymorphisms

in patients with diclonfenac hepatotoxicity. Hepatology, 2004. 39: p. 1430-1440.

20. Mallal, S., et al., HLA-B*5701 screening for hypersensitivity to abacavir. N. Engl. J.

Med. , 2008. 358: p. 568-579.

21. Hung, S.L., et al., HLA-B*5801 allele as a genetic marker for severe cutaneous adverse

reactions caused by allopurinol. PNAS, 2005. 102(4134-4139).

22. Ko, T.M., et al., Shared and restricted T-cell receptor use is crucial for carbamazepine-

induced Stevens-Johnson syndrome. J. Allergy Clin. Immunol., 2011. 128(6): p. 1266-

1276.

23. Uetrecht, J., Immune-mediated adverse drug reactions. Chem. Res. Toxciol., 2009. 22: p.

24-34.

24. Shear, N.H., et al., Differences in metabolism of sulfonamides predisposing to

idiosyncratic toxicity. Ann. Intern. Med., 1986. 105: p. 179-184.

25. Huang, Y.S., et al., Cytochrome P450 2E1 genotype and the susceptibility to

antituberculosis drug-induced hepatitis. Hepatology, 2003. 37: p. 924-930.

26. Huang, Y.S., et al., Polymorphism of the N-acetyltransferase 2 gene as a susceptibility

risk factor for antituberculosis drug-induced hepatitis. Hepatology, 2002. 35: p. 883-889.

Page 202: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

181

27. Martin, A.M., et al., Predisposition to abacavir hypersensitivity conferred by HLA-

B*5701 and a haplotype Hsp70-Hom variant. PNAS, 2004. 101: p. 4180-4185.

28. Uetrecht, J., Immunoallergic drug-induced liver injury in humans. Semin. Liver Dis.,

2009. 29: p. 383-392.

29. Watkins, P.B., Idiosyncratic liver injury: challenges and approaches. Toxicologic

Pathol., 2005. 33: p. 1-5.

30. Harris, A.L., et al., Agranulocytosis associated with aminoglutethimide: pharmacological

and marrow studies. Br. J. Cancer, 1986. 54: p. 119-122.

31. Landsteiner, K. and J. Jacobs, Studies on the sensitization of animals with simple

chemical compounds. J. Exp. Med., 1935. 61: p. 643-656.

32. de Weck, A.L. and H.N. Eisen, Some immunochemical properties of penicillenic acid. J.

Exp. Med., 1960. 112: p. 1227-1247.

33. Evans, D.C., et al., Drug-protein adducts: an industry perspective on minimizing the

potential for drug bioactivation in drug discovery and development. Chem. Res. Toxciol.,

2004. 17: p. 3-16.

34. Gardner, I., et al., A comparison of the oxidation of clozapine and olanzapine to reactive

metabolites and the toxicity of these metabolites to human leukocytes. Mol. Pharmacol.,

1998. 53(991-998).

35. Uetrecht, J., Evaluation of which reactive metabolite, if any, is responsible for a specific

idiosyncratic reaction. Drug Met. Rev., 2006. 38: p. 745-753.

Page 203: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

182

36. Lecoeur, S., et al., Specificity of in vitro covalent binding of tienilic acid metabolites to

human liver microsomes in relationship to the type of hepatotoxicity: comparison with

two directly hepatotoxic drugs. Chem. Res. Toxciol., 1994. 7: p. 434-442.

37. Pacitto, S.R., et al., Changes in gene expression induced by tienilic acid and

sulfamethoxazole: testing the Danger Hypothesis. J. Immunotox., 2007. 4(4): p. 253-266.

38. Lopez-Garcia, M.P., P.M. Dansette, and J. Coloma, Kinetics of tienilic acid bioactivation

and functional generation of drug-protein adducts in intact rat hepatocytes. Biochem.

Pharmacol., 2005. 70: p. 1870-1882.

39. Matzinger, P., Tolerance, danger, and the extended family. Ann. Rev. Immunol., 1994.

12: p. 991-1045.

40. Janeway, C.A., Approaching the asymptote? Evolution and revolution in immunology.

Cold Spring Harbor Symp. Quant. Biol., 1989. 54: p. 1-13.

41. Matzinger, P., The danger model: a renewed sense of self. Science, 2002. 296: p. 301-

305.

42. Li, J. and J.P. Uetrecht, The danger hypothesis applied to idiosyncratic drug reactions.

Handb Exp Pharmacol, 2010(196): p. 493-509.

43. Gallucci, S. and P. Matzinger, Danger signals: SOS to the immune system. Curr. Opin.

Immunol., 2001. 13: p. 114-119.

44. Van der Ven, A.J., et al., Adverse reactions to co-trimoxazole in HIV infection. Lancet,

1991. ii: p. 431-433.

Page 204: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

183

45. Lavergne, S.N., et al., "Danger" conditions increase sulfamethoxazole-protein adduct

formation in human antigen-presenting cells. J. Pharmacol. Exp. Ther., 2009. 331(372-

381).

46. Aitken, A.E., T.A. Richardson, and E.T. Morgan, Regulation of drug-metabolizing

enzymes and transporters in inflammation. Annu. Rev. Pharmacol. Toxicol., 2006. 46: p.

123-149.

47. Lotze, M.T. and K.J. Tracey, High-mobility group box 1 protein (HMGB1): nuclear

weapon in the immune arsenal. Nat. Rev. Immunol., 2005. 5: p. 331-342.

48. Shi, Y., J.E. Evans, and K.L. Rock, Molecular identification of a danger signal that

alerts the immune system to dying cells. Nature, 2003. 425: p. 516-521.

49. Elsheikh, A., et al., Drug antigenicity, immunogenicity, and costimulatory signaling:

evidence for formation of a functional antigen through immune cell metabolism. J.

Immunol., 2010. 185(11): p. 6448-6460.

50. Matzinger, P. and T. Kamala, Tissue-based class control: the other side of tolerance. Nat.

Rev. Immunol., 2011. 11: p. 221-230.

51. Kono, H. and K.L. Rock, How dying cells alert the immune system to danger. Nat. Rev.

Immunol., 2008. 8: p. 279-289.

52. Schnyder, B., et al., Direct, MHC-dependent presentation fo the drug sulfamethoxazole to

human T-cell clones. J. Clin. Invest. , 1997. 100(136-141).

Page 205: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

184

53. Pichler, W.J., Pharmacological interaction of drugs with antigen-specific immune

receptors: the p-i concept. Curr. Opin. Allergy Clin. Immunol., 2002. 2: p. 301-305.

54. Chen, X., et al., A study of the specificity of lymphocytes in nevirapine-induced skin rash.

J. Pharmacol. Exp. Ther., 2009. 331(3): p. 836-41.

55. Engler, O.B., et al., A chemically inert drug can stimulate T cells in vivo by their T cell

receptor in non-sensitised individuals. Toxicology, 2004. 197(47-56).

56. Deng, X., et al., Inflammatory Stress and Idiosyncratic Hepatotoxicity: Hints from

Animal Models. Pharmacological Reviews, 2009. 61(3): p. 262-282.

57. Luyendyk, J.P., et al., Ranitidine treatment during a modest inflammatory response

precipitates idiosyncrasy-like liver injury in rats. J. Pharmacol. Exp. Ther., 2003. 307: p.

9-16.

58. Waring, J.F., et al., Microarray analysis of lipopolysaccharide potentiation of

trovafloxacin-induced liver injury in rats suggests a role for proinflammatory chemokines

and neutrophils. J. Pharmacol. Exp. Ther., 2005. 316: p. 1080-1087.

59. Fan, H. and J.A. Cook, Molecular mechanisms of endotoxin tolerance. J. Endotoxin Res.,

2004. 10: p. 71-84.

60. Zimmerman, H., Hepatotoxicity: the adverse effects of drugs and other chemicals on the

liver second ed. 1999, Philadelphia: Lippincott Williams and Wilkins.

61. Ng, W., et al., Animal models of idiosyncratic drug reactions. Advances Pharmacol.,

2012. 63: p. 81-135.

Page 206: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

185

62. Kass, G.E., Mitochondrial involvement in drug-induced hepatic injury. Chemico-Biol.

Interact., 2006. 163: p. 145-159.

63. Kokoszka, J.E., et al., Increased mitochondrial oxidative stress in the Sod2 (+/-) mouse

results in age-related decline of mitochondrial function culminating in increased

apoptosis Proc. Acad. Nat. Sci., 2001. 98: p. 2278-2283.

64. Ong, M.M.K., C. Latchoumycandane, and U.A. Boelsterli, Troglitazone-induced hepatic

necrosis in an animal model of silent genetic mitochondrial abnormalities. Toxicol. Sci.,

2007. 97: p. 205-213.

65. Smith, M.T., Mechanisms of troglitazone hepatotoxicity. Chem. Res. Toxciol., 2003. 16:

p. 679-687.

66. Fujimoto, K., et al., Sensitivity of liver injury in heterozygous Sod2 knockout mice treated

with troglitazone or acetaminophen. Toxicologic Pathol., 2009. 37: p. 193-200.

67. Copple, B.L., H. Jaeschke, and C.D. Klaassen, Oxidative stress and the pathogenesis of

cholestasis. Semin. Liver Dis. , 2010. 30: p. 195-204.

68. Orman, E.S., et al., Clinical and histopathologic features of fluoroquinolone-induced

liver injury. Clin. Gastroenterol. Hepatol. , 2011. 9: p. 517-523.

69. Ozer, J., et al., The current state of serum bioarkers of hepatotoxicity. Toxicology, 2008.

245: p. 194-205.

70. Ju, C. and T. Reilly, Role of immune reactions in drug-induced liver injury (DILI). Drug

Met. Rev., 2012. 44: p. 107-115.

Page 207: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

186

71. Kenna, J.G., et al., Metabolic basis for a drug hypersensitivity: antibodies in sera from

patients with halothane hepatitis recognize liver neoantigens that contain the

trifuoroacetyl group derived from halothane. J. Pharmacol. Exp. Ther., 1988. 245: p.

1103-1109.

72. Noris, S., et al., Resident human hepatic lymphocytes are phenotypically different from

circulating lymphocytes. J. Hepatol., 1998. 28: p. 84-90.

73. Knolle, P.A. and G. Gerken, Local control of the immune response in the liver. Immunol.

Rev., 2000. 174: p. 21-34.

74. Jaeschke, H., et al., Acetaminophen hepatotoxicity and repair: the role of sterile

inflammation and innate immunity. Liver Int., 2012. 32: p. 8-20.

75. Ju, C., et al., Protective role of Kupffer cells in acetaminophen-induced hepatic injury in

mice. Chem. Res. Toxciol., 2002. 15: p. 1504-1513.

76. Serhan, C.N. and J. Savill, Resolution of inflammation: the beginning programs the end.

Nat. Immunol. 6: p. 1191-1197.

77. Ip, J. and J.P. Uetrecht, In vitro and animal models of drug-induced blood dyscrasias.

Environ. Toxicol. Pharmacol., 2006. 21: p. 135-140.

78. Young, N.S., P. Scheinberg, and R.T. Calado, Aplastic anemia. Curr. Opin. Hematol. ,

2008. 15: p. 162-168.

79. Moeschlin, S. and K. Wagner, Agranulocytosis due to the occurrence of leukocyte-

agglutinins. Acta Haematol., 1952. 8: p. 29-41.

Page 208: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

187

80. Aster, R.H., Drug-induced immune thrombocytopenia: an overview of pathogenesis.

Semin. Haematol., 1999. 36: p. 2-6.

81. Uetrecht, J.P., The role of leukocyte-generated metabolites in the pathogenesis of

idiosyncratic drug reactions. Drug Met. Rev., 1992. 24: p. 299-366.

82. Liu, Z.C. and J.P. Uetrecht, Clozapine is oxidized by activated human neutrophils to a

reactive nitrenium ion that irreversibly binds to cells. J. Pharmacol. Exp. Ther., 1995.

275: p. 1476-1483.

83. Uetrecht, J., N. Zahid, and D. Whitfield, Metabolism of vesarinone by activated

neutrophils; implictions for vesnarinone-induced agranulocytosis. J. Pharmacol. Exp.

Ther., 1994. 270: p. 865-872.

84. Di Meglio, P., G.K. Perera, and F.O. Nestle, The multitasking organ: recent insights into

skin immune function. Immunity, 2011. 35: p. 857-869.

85. Svensson, C.K., Biotransformation of drugs in human skin. Drug Met. Dispos., 2009. 37:

p. 247-253.

86. Shear, N.H. and S.P. Spielberg, Anticonvulsant hypersensitivity syndrome. In vitro

assessment of risk. J. Clin. Invest., 1988. 82: p. 1826-1832.

87. Downey, A., et al., Toxic epidermal necrolysis: rview of pathogenesis and management.

J. Am. Acad. Dermatol., 2012. 66: p. 995-1003.

Page 209: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

188

88. McGee, T. and A. Munster, Toxic epidermal necrolysis syndrome: mortality rate reduced

with early referral to regional burn center. Plast. Reconstr. Surg., 1998. 102: p. 1018-

1022.

89. Elkayam, O., M. Yaron, and D. Caspi, Minocycline-induced autoimmune syndromes: an

overview. Semin. Arthritis Rheum., 1990. 28: p. 392-397.

90. Rossi, M. and J.W. Young, Human dendritic cells: potent antigen-presenting cells at the

crossroads of innate and adaptive immunity. J. Immunotox., 2005. 175: p. 1373-1381.

91. Adams, D.H., et al., Mechanisms of immune-mediated liver injury. Toxicol. Sci., 2010.

115(2): p. 307-21.

92. Pichler, W.J. and J. Tilch, The lymphocyte transformation test in the diagnosis of drug

hypersensitivity. Allergy, 2004. 59: p. 809-820.

93. Faulkner, L., et al., The development of in vito culture methods to characterize primary

T-cell responses to drugs. Toxicol. Sci., 2012. 127: p. 150-158.

94. Uetrecht, J., N-oxidation of drugs associated with idiosyncratic drug reactions. Drug

Metab. Rev., 2002. 34(3): p. 651-65.

95. Liu, Z.C. and J.P. Uetrecht, Clozapine is oxidized by activated human neutrophils to a

reactive nitrenium ion that irreversibly binds to the cells. J. Pharmacol. Exp. Ther., 1995.

275: p. 1476-1483.

96. Skipper, P.L., et al., Monocyclic aromatic amines as potential human carcinogens: old is

new again. Carcinogenesis, 2010. 31: p. 50-58.

Page 210: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

189

97. Uetrecht, J.P., Reactivity and possible significance of hydroxylamine and nitroso

metabolites of procainamide. J. Pharmacol. Exp. Ther., 1985. 232: p. 420-425.

98. Goodman, L.S., et al., Goodman and Gilman's the Pharmacological Basis of

Therapeutics. 11 ed. 2006, New York: McGraw-Hill.

99. Roujeau, J.C., et al., Medication use and the risk of Stevens-Johnson syndrome or toxic

epidermal necrolysis. NEJM, 1995. 333: p. 1600-1607.

100. van der Ven, A.J., et al., Adverse reaction s to co-trimoxazole in HIV infection. Lancet,

1991. ii: p. 431-433.

101. Meekins, C.V., T.J. Sullivan, and R.S. Gruchalla, Immunochemical analysis of

sulfonamide drug allergy: identification of sulfamethoxazole-substituted human serum

proteins. J. Allergy Clin. Immunol., 1994. 94: p. 1017-1024.

102. Daftarian, M.P., et al., Immune response to sulfamethoxazole in patients with AIDS. Clin.

Diagn. Lab Immunol., 1995. 2: p. 199-204.

103. Wozel, G., et al., Dapsone hydroxylamine inhibits the LTB4-induced chemotaxis of

polymorphonuclear leukocytes into human skin: results of pilot study. Inflam. Res., 1997.

46: p. 420-422.

104. Wozel, G. and B. Lehmann, Dapsone inhibits the generation of 5-lipoxygenase products

in human polymorphonuclear leukocytes. Skin Pharmacol., 1995. 8: p. 196-202.

105. Suda, T., et al., Dapsone supresses human neutrophil superoxide production and elastase

release in a calcium-depende t manner. Br. J. Dermatol., 2005. 152: p. 887-895.

Page 211: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

190

106. Zhu, Y.I. and M.J. Stiller, Dapsone and sulfones in dermatology: overview and update. J.

Am. Acad. Dermatol., 2001. 45: p. 420-434.

107. Coleman, M.D., Dapsone-mediated agranulocytosis: risks, possible mechanisms and

prevention. Toxicology, 2001. 162: p. 53-60.

108. Santen, R.J., et al., adrenal suppression with aminoglutethimide. I. differential effects of

aminogleutehimide on glucocorticoid metabolism as a rationale for use of

hydrocortisone. J. Clin. Endocrinol. Metab., 1977. 45: p. 469-479.

109. Young, J.A., L.N. Newcomer, and A.M. Keller, Aminoglutethimide-induced bone

marrow injury: report of a case and review of the literature. Cancer Cell, 1984. 54: p.

1731-1733.

110. Stuart-Harris, R.C. and I.E. Smith, Aminoglutethimide in the treatment of advanced

breast cancer. Cancer Cell, 1984. 11: p. 189-204.

111. Murray, M., E. Cantrill, and G.C. Farrell, Inductionof cytochrome P4502B1 in rat liver

by the aromatase inhibitor aminoglutethimide. J. Pharmacol. Exp. Ther., 1993. 265: p.

477-481.

112. Murray, F.T., et al., Serum aminoglutethimide levels: studies of serum half-life,

clearance, and patient compliance. J. Clin. Pharmacol., 1979. 19: p. 704-711.

113. Vincent, M.D., et al., Aminoglutethimide (with hydrocortisone) induced agranulocytosis

in primary breast cancer. Br. Med. J., 1985. 291: p. 101-106.

Page 212: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

191

114. Pollard, r.B., P. Robinson, and K. Dransfield, Safety profile of nevirapine, a

nonnucleoside reverse transcriptase inhibitor for the treatment of human

immunodeficiency virus infection. Clin. Therap., 1998. 20: p. 1071-1092.

115. Shenton, J.M., et al., Characterization of a potential animal model of an idiosyncratic

drug reaction: nevirapine-induced skin rash in the rat. Chem. Res. Toxciol., 2003. 16: p.

1078-1089.

116. Bersoff-Matcha, S.J., et al., Sex differences in nevirapine rash. Clin. Infect. Dis., 2001.

32: p. 124-129.

117. Shenton, J.M., et al., Evidence of an immune-mediated mechanism for an idiosyncratic

nevirapine-induced reaction in the female Brown Norway rat. Chem. Res. Toxciol.,

2005. 18: p. 1799-1813.

118. Patel, S., et al., Serious adverse cutaneous and hepatic toxicities associated with

nevirapine use by non-HIV-infected individuals. J. Acquir. Immune Defic. Syndr, 2004.

35: p. 120-125.

119. Chen, J., et al., Demonstration fo the metabolic pathway responsible for nevirapine-

induced skin rash. Chem. Res. Toxciol., 2008. 21: p. 1862-1870.

120. Sharma, A.M., et al., 12-OH-Nevirapine sulfate, formed in the skin, is responsible for

nevirapine-induced skin rash. Chem. Res. Toxciol., 2013. 26: p. 817-827.

121. Watanabe, H., et al., Activation of the IL-1b-processing inflammasome is involved in

contact hypersensitivity. J. Invest. Dermatol., 2007. 127: p. 1956-1963.

Page 213: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

192

122. Stein, H.B., et al., Adverse effects of D-penicillamine in rheumatoid arthritis. Ann. Intern.

Med., 1980. 92: p. 24-29.

123. Bigazzi, P.E., Autoimmunity induced by chemicals. Clin. Toxicol., 1988. 26: p. 125-156.

124. Enzenauer, R.J., S.G. West, and R.L. Rubin, D-penicillamine-induced lupus

erythematosus. Arthritis Rheumatism, 1990. 33: p. 1582-1585.

125. Donker, A.J., et al., Effects of prolonged administration of D-penicillamine or catopril in

various strains of rats. Brown Norway rats treated with D-penicillamine develop

autoantibodies, circulating immune complexes, and disseminated, intravascular

coagulation. Clin. Immunol. Immunopathol., 1984. 30: p. 142-155.

126. Masson, M.J., et al., Investigation fo the involvement of macrophages and T-cells in D-

penicillamine-induced autoimmunity int eh Brown Norway rat. J. Immunotoxicol., 2004.

1: p. 79-93.

127. Li, J., B. Mannargudi, and J.P. Uetrecht, Covalent binding of penicillamine to

macrophages: implications for penicillamine-induced autoimmunity. Chem. Res.

Toxciol., 2009. 22: p. 1277-1287.

128. Li, J. and J. Uetrecht, D-penicillamine-induced autoimmunity:relationship to macrophage

activation. Chem. Res. Toxciol., 2009. 22: p. 1526-1533.

129. Zhu, X., et al., Involvement of the T helper 17 cell in D-penicillamine-induced

autoimmune disease in Brown Norway rats. Toxicol. Sci., 2011. 120: p. 331-338.

Page 214: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

193

130. Masson, M.J. and J. Uetrecht, Tolerance induced by low dose D-penicillamine in the

Brown Norway rat model of drug-induced autoimmunity is immune mediated. Chem. Res.

Toxciol., 2004. 17: p. 82-94.

131. Sayeh, E. and J.P. Uetrecht, Factors that modify penicillamine-induced autoimmunity in

Brown Norway ras: failure. Toxicology, 2001. 163: p. 195-211.

132. Trepanier, L.A., Delayed hypersensitivity reactions to sulphonamides: syndromes,

pathogenesis and managment. Vet. Dermatol., 1999. 10: p. 241-248.

133. Trepanier, L.A., et al., Cytosolic arylamine N-acetyltransferase (NAT) deficiency in the

dog and other canids due to absence of NAT genes. Biochem. Pharmacol., 1997. 54: p.

73-80.

134. Cribb, A.E., et al., Covalent binding of sulfamethoxazole reactive metabolites to human

and rat liver subcellular fractions assessed by immunochemical detection. Chem. Res.

Toxciol., 1996. 9: p. 500-507.

135. Naisbitt, D.J., et al., Covalent binding of the nitroso metabolite of sulfamethoxazole leads

to toxicity and major histocompatibility complex-restricted antigen presentation. Mol.

Pharmacol., 2002. 62: p. 628-637.

136. Naisbitt, D.J., et al., Antigenicity and immunogenicity of sulfamethoxazole:

demonstration of metabolism-dependent haptenation and T-cell proliferation in vivo. Br.

J. Pharmacol., 2001. 133: p. 295-305.

Page 215: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

194

137. Gill, H.J., et al., The relationship between the disposition and immunogenicity of

sulfamethoxazole in the rat. J. Pharmacol. Exp. Ther., 2002. 282: p. 795-801.

138. Khan, F.D., et al., Effect of arylhydroxylamine metabolites of sulfamethoxazole and

dapsone on stress signal expression in human keratinocytes. J. Pharmacol. Exp. Ther.,

2007. 323: p. 771-777.

139. Sanderson, J.P., et al., Sulfamethoxazole and its metabolite nitroso sulfamethoxazole

stimulate dendritic cell costimulatory signaling. J. Immunol., 2007. 178: p. 5533-5542.

140. Schulte, P.F.J., Risk of clozapine-associated agranulocytosis and mandatory white blood

cell monitoring. Ann. Pharacother., 2006. 40: p. 683-688.

141. Blumenreich, M.S., The white blood cell and differential count, in Clinical Methods: The

History, Physical, and Laboratory Examinations, H.K. Walker, W.D. Hall, and J.W.

Hurst, Editors. 1990, Butterworths: Boston.

142. Evans, G.O., Animal Hematotoxicology: a practical guide for toxicologists and

biomedical researchers. 2009, Boca Raton: CRC Press Taylor & Francis Group.

143. Hannigang, B.M., C.B.T. Moore, and D.G. Quinn, eds. Immunology. 2009, Scion

Publishing Ltd.: Bloxham, Oxfordshire.

144. Summers, c., et al., Neutrophil kinetics in health and disease. Trends Immunol., 2010. 31:

p. 318-324.

Page 216: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

195

145. Price, T.H., G.S. Chatta, and D.C. Dale, Effect of recombinant granulocyte colony-

stimulating factor on neutrophil kinetics in normal young and elderly humans. Blood

Rev., 1996. 88: p. 335-340.

146. Basu, S., et al., Evaluation of role of G-CSF in the production survival and release of

neutrophils from bone marrow into circulation. Blood, 2002. 100: p. 854-861.

147. von Vietinghoff, S. and K. Ley, Homeostatic regulation of blood neutrophil counts. J.

Immunol., 2008. 181: p. 5183-5188.

148. Kennedy, A.D. and F.R. DeLeo, Neutrophil apoptosis and the resolution of infection.

Immunol. Res., 2009. 43: p. 25-61.

149. Amulic, B., et al., Neutrophil function: from mechanisms to disease. Annu. Rev.

Immunol., 2012. 30: p. 459-489.

150. Parker, L.C., et al., The expression and roles of Toll-like receptors in the biology of the

human neutrophil. J. Leukoc. Biol. , 2005. 77: p. 886-892.

151. Dusi, S., et al., Mechanisms of stimulation of the respiratory burst by TNF in

nonadherent neutrophils. J. immunol., 1996. 157: p. 4615-4623.

152. Hofstra, A.H. and J.P. Uetrecht, Myeloperoxidase-mediated activation of xenobiotics by

human leukocytes. Toxicology, 1993. 82: p. 221-242.

153. O'Brien, P.J., Peroxidases. Chem.-Biol. Interact. , 2000. 129: p. 113-139.

Page 217: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

196

154. Brown, K.E., E.M. Brunt, and J.W. Heinecke, Immunohistochemical detection of

myeloperoxidase and its oxidation products in Kupffer cells of human liver. Am. J.

Pathol., 2001. 159: p. 2081-2088.

155. Uetrecht, J., N.H. Shear, and N. Zahid, N-chlorination of sulfamethoxazole and dapsone

by the myeloperoxidase system. Drug Met. Dispos., 1993. 21: p. 830-834.

156. Furst, S.M. and J.P. Uetrecht, Carbamazepine metabolism to a reactive intermediate by

the myeloperoxidase system of activated neutrophils. Biochem. Pharmacol., 1993. 45: p.

1267-1275.

157. Liu, Z.C. and J.P. Uetrecht, Metabolism of ticlopidine by activated neutrophils:

implications for ticlopidine-induced agranulocytosis. Drug Met. Dispos., 2000. 08: p.

726-730.

158. Mantovani, A., et al., Neutrophils in the activation and regulation of innate and adaptive

immunity. Nat. Rev. Immunol., 2011. 11(519-531).

159. Puga, I., et al., B-helper neutrophils stimulate the diversification and production of

immunoglobulin in the marginal zone of the spleen. Nat. Immunol., 2011. 13: p. 170-180.

160. Pillay, J., et al., A subset of neutrophils in human systemic inflammation inhibits T cell

responses through Mac-1. J. Clin. Invest., 2012. 122: p. 327-336.

161. Srivastava, M.K., et al., Myeloid-derived suppressor cells inhibit T-cell activation by

depleting cystine and cysteine. Cancer Res., 2010. 70: p. 68-77.

Page 218: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

197

162. Fridlender, Z.G., et al., Polarization of tumor-associated neutrophil phenotype by TGF-

"N1" versus "N2" TANCancer Cell, 2009. 16: p. 183-194.

163. Andersohn, F., C. Konzen, and E. Garbe, Systematic review: agranulocytosis induced by

nonchemotherapy drugs. Ann. Intern. Med., 2007. 146: p. 657-665.

164. Pisciotta, A.V., Drug induced agranulocytosis peripheral destruction of

polymorphonuclear leukocytes and their marrow precursors. Blood Rev. , 1990. 4: p.

226-237.

165. Moeschlin, S. and K. Wagner, Agranulocytosis due to the occurrence of leukocyte-

agglutinins. Acta Haematol., 1952. 8(29-41).

166. Alvir, J.M.J., et al., Clozapine-induced agranulocytosis - incidence and risk factors in the

United States. NEJM, 1993. 329: p. 162-167.

167. Alvir, J.M., J.A. Lieberman, and A.Z. Safferman, Do what-cell count spikes predict

agranulocytosis in clozapine recipients? Psychopharmacol. Bull., 1995. 31: p. 311-314.

168. Pollmacher, T., et al., The influence of clozapine treatment on plasma granulocyte

colony-stimulating (G-CSF) levels. Pharmacopsychiatry, 1997. 30: p. 118-121.

169. Pollmacher, T., D. Hinze-Selch, and J. Mullington, Effects of clozapine on plasma

cytokine and soluble cytokine receptor levels. J. Clin. Psychopharmacol., 1996. 16: p.

403-409.

170. Williams, D.P., et al., Induction of metabolism-dependent and -independent neutrophil

apoptosis by clozapine. Mol. Pharmacol., 2000. 58: p. 207-216.

Page 219: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

198

171. Liu, Z.C. and J.P. Uetrecht, Clozapine is oxidized by activated human neutrophils to a

reactive nitrinium ion that irreversibly binds to the cells. J. Pharmacol. Exp. Ther., 1995.

275: p. 1476-1483.

172. Guest, I., et al., Examination of possible toxic and immune mechanisms of clozapine-

induced agranulocytosis. Toxicology, 1998. 131: p. 53-65.

173. Ali, H., et al., Comparison of in vitro and in vivo haemotoxic effects of aminoglutethimide

and glutethimide. Toxicology In Vitro, 1990. 4: p. 381-83.

174. Coleman, M.D., L.F. Khalaf, and P.J. Nicholls, Aminoglutethimide-induced leucopenia in

a mouse model: effects of metabolic and structural determinates. Environ. Toxicol.

Pharmacol., 2003. 15: p. 27-32.

175. Shenton, J.M., J. Chen, and J.P. Uetrecht, Animal models of idiosyncratic drug reactions.

Chem. Biol. Interact., 2004. 150(1): p. 53-70.

176. Goss, P.E., M. Jarman, and L.J. Griggs, Metabolism of aminoglutethimide in humans:

quantification and clinical relevance of induced metabolism. British Journal of Cancer,

1985. 51(2): p. 259-62.

177. Siraki, A.G., et al., Aminoglutethimide-induced protein free radical formation on

myeloperoxidase: a potential mechanism of agranulocytosis. Chem. Res. Toxicol., 2007.

20(7): p. 1038-45.

178. Gallucci, S. and P. Matzinger, Danger signals SOS to the immune system. Curr. Opin.

Immunol., 2001. 13: p. 113-119.

Page 220: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

199

179. Chan, H.L., et al., The incidence of erythema multiforme, Stevens-Jonson syndrome, and

toxic epidermal necrolysis. Arch. Dermatol., 1990. 126: p. 43-47.

180. Mockenhaupt, M., The current understanding of Stevens-Johnson syndrome and toxic

epidermal necrolysis. Expert Rev. Clin. Immunol., 2011. 7: p. 803-815.

181. Farrell, J., et al., Characterization of sulfamethoxazole and sulfamethoxazole metabolite-

specific T-cell responses in animals and humans. J. Pharmacol. Exp. Ther. , 2003. 306: p.

229-237.

182. Gill, H.J., et al., The relationship between the disposition and immunogenicity of

sulfamethoazole in the rat. J. Pharmacol. Exp. Ther., 1997. 282: p. 795-801.

183. Helton, D.R., et al., Pharmacokinetic profiles in rats after intravenous, oral, or dermal

administration of dapsone. Drug Metab. Dispos., 2000. 28: p. 925-929.

184. Nicholls, P.J., et al., Aminoglutethimide: absorption, physiological disposition and

pharmacokinetics, in Aminoglutethimide as an aromatase inhibitor in the treatment of

cancer: an international symposium held during the 13th International Congress of

Chemotherapy, G.A. Nagel and R.J. Santen, Editors. 1984, Hans Huber, Berne,

Switzerland: Vienna, Austria. p. 58-67.

185. Schwartz, D.E. and J. Rieder, Pharmacokinetics of sulfamethoxazole and trimethoprim in

man and their distribution in the rat. Chemother., 1970. 15: p. 337-355.

186. Lonning, P.E., et al., Single-dose and steady -state pharmacokinetics of

aminoglutethimide. Clin. Pharmacokinet., 1985. 10: p. 353-364.

Page 221: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

200

187. Coffy, S., et al., Identificaiton of a novel rat hepatic gene induced early by insulin,

independently of glucose. Biochem. J., 2005. 385: p. 165-171.

188. Machado, F.S., et al., Anti-inflammatory actions of lipoxin A4 and aspirin-triggered

lipoxin are SOCS-2 dependent. Nature Med., 2006. 12(3): p. 330-334.

189. Patterson, K.I., et al., Dual-specificity phosphatases:critical regulators with diverse

cellular targets. Biochem. J., 2009. 418: p. 475-489.

190. Webster, M.K., L. Goya, and G.L. Firestone, Immediate-early transcriptional regulation

and rapid mRNA turnover of a putative serine/threonine protein kinase. J. Biol. Chem.,

1993. 268: p. 11482-11485.

191. Mandard, S., et al., The G0/G1 switch gene 2 is a novel PPAR target gene. Biochem. J.,

2005. 392(2): p. 313.

192. Welch, C., et al., Identification of a protein, G0S2, that lacks Bcl-2 homology domains

and interacts with and antagonizes Bcl-2. Cancer Res., 2009. 69(17): p. 6782-6789.

193. Leppa, S. and D. Bohmann, Diverse function of JNK singaling and c-Jun in stress

response and apoptosis. Oncogene, 1999. 18: p. 6158-6162.

194. Babic, A.M., et al., CYR61 a product of a growth factor-inducible immediate early gene,

promotes angiogenesis and tumor growth. Proc. Natl. Acad. Sci. USA, 1998. 95: p. 6355-

6360.

Page 222: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

201

195. Anunciado-Koza, R.P., et al., Inactivation of the mitochondrial carrier SLC25A25 (ATP-

Mg2+/Pi Transporter) reduces physical endurance and metabolic efficiency in mice. J.

Biol. Chem., 2011. 286(13): p. 11659-11671.

196. Lada, A.T., et al., Identification of ACAT1- and ACAT2-specific inhibitors using a novel,

cell-based fluorescence assay: individual ACAT uniqueness. J. Lipid Res., 2003. 45(2): p.

378-386.

197. Xu, H., et al., Molecular cloning, functional characterization, tissue distribution, and

chromosomal localization of a human, small intestinal sodium-phosphate (Na+-Pi)

transporter (SLC34A2). Genomics, 1999. 62: p. 281-284.

198. Chugh, A., A. Ray, and J.B. Gupta, Squalene epoxidase as hypocholesterolemic drug

target revisited. Prog. Lipid Res., 2003. 42: p. 37-50.

199. Grusch, M., et al., De-regulation of the activin/follistatin system in hepatocarcinogenesis.

J. Hepatol., 2006. 45(5): p. 673-680.

200. Uetrecht, J., Dapsone and sulfapyridine, in Clinics in Dermatology, N.H. Shear, Editor.

1989, Lippincott: Philadelphia. p. 111-120.

201. Naisbitt, D.J., et al., Cellular disposition of sulphamethoxazole and its metabolites:

implications for hypersensitivity. Br. J. Pharmacol., 1999. 126: p. 1393-1407.

202. Young, J.A., L.N. Newcomer, and A.M. Keller, Aminoglutethimide-induced bone

marrow injury: report of a case and review of the literature. Cancer, 1984. 54: p. 1731-

1733.

Page 223: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

202

203. Roychowdhury, S., et al., Characterization of the formation and localization of

sulfamethoxazole and dapsone-associated drug-protein adducts in human epidermal

keratinocytes. J. Pharmacol. Exp. Ther., 2005. 314(1): p. 43-52.

204. Lu, W., X. Li, and J.P. Uetrecht, Changes in gene expression induced by carbamazepine

and phenytoin: testing the Danger Hypothesis. J. Immunotox., 2008. 5(2): p. 107-113.

205. Seguin, B., et al., Gene expression profiling in a model of D-penicillamine-induced

autoimmunity in the Brown Norway rat: predictive value of early signs of danger. Chem.

Res. Toxicol., 2005. 18(8): p. 1193-202.

206. Mustacich, D. and G. Powis, Thioredoxin reductase. Biochem. J., 2000. 346: p. 1-8.

207. Nguyen, T., P. Nioi, and C.B. Pickett, The Nrf2-antioxidant response element signaling

pathway and its activation by oxidative stress. J. Biol. Chem., 2009. 284(20): p. 13291-

13295.

208. Nagel, G.A., H.E. Wander, and H.C. Blossey, Phase II study of aminoglutethimide and

medroxyprogesterone acetate in the treatment of patients with advanced breast cancer.

Cancer Res. (Suppl.), 1982. 42: p. 3442s-3444s.

209. Sanderson, J.P., et al., Sulfamethoxazole and its metabolite nitroso sulfamethoxazole

stimulate dendritic cell costimulatory signaling. J. Immunol., 2007. 178: p. 5533-5542.

210. Khan, F.D., et al., Effect of arylhydroxylamine metbaolites of sulfamethoxazole and

dapsone on stress signal expression in human keratinocytes. J. Pharmacol. Exp. Ther.,

2007. 323: p. 771-777.

Page 224: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

203

211. Bowen, D.G., G.W. McCaughan, and P. Bertolino, Intrahepatic immunity: a tale of two

sites? Trends Immunol., 2005. 26: p. 512-517.

212. Ng, W. and J. Uetrecht, Changes in gene expression induced by aromatic amine drugs:

testing the danger hypothesis. J. Immunotox., 2013. 10: p. 178-191.

213. Waring, J.F., et al., Microarray analysis of hepatotoxins in vitro reveals a correlation

between gene expression profiles and mechanisms of toxicity. Toxicol. Lett., 2001. 120:

p. 359-368.

214. Cocconi, G., First generation aromatase inhibitors-aminoglutethimide and

testololactone. Breast Cancer Res.Treat., 1994. 30(1): p. 57-80.

215. Murray, M., E. Cantrill, and G.C. Farrell, Induction of cytochrome P450 2B1 in rat liver

by the aromatase inhibitor aminoglutethimide. J. Pharmacol. Exp. Ther., 1993. 265: p.

477-481.

216. Ng, W. and J. Uetrecht, Effect of aminoglutethimide on neutrophils in rats: implications

for idiosyncratic drug-induced blood dyscrasias Chem. Res. Toxicol., 2013. 26: p. 1272-

1281.

217. Metushi, I.G., T. Nakagawa, and J. Uetrecht, Direct oxidation and covalent binding of

isoniazid to rodent liver and human hepatic microsomes: humans are more like mice than

rats. Chem. Res. Toxicol., 2012. 25: p. 2567-2576.

Page 225: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

204

218. Goldkind, L. and L. Laine, A systematic review of NSAIDs withdrawn from the market

due to hepatotoxicity: lessons learned from the bromfenac experience.

Pharmacoepidemiology and Drug Safety, 2006. 15(4): p. 213-220.

219. Crispe, I.N., et al., The liver as a site of T-cell apoptosis: graveyard, or killing field?

Immunol. Rev., 2000. 174: p. 47-62.

220. Hayden, M.S., A.P. West, and S. Ghosh, NF-κB and the immune response. Oncogene,

2006. 25: p. 6758-6780.

221. Dutta, J., et al., Current insights into the regulation of programmed cell death by NF-κB.

Oncogene, 2006. 25: p. 6800-6816.

222. Dissanayake, D., et al., Nuclear factor-kB1 controls the functional maturation of

dendritic cells and prevents the activation of autoreactive T cells. Nat. Med., 2011. 17: p.

1663-1667.

223. Uetrecht, J., N-oxidation of drugs associated with idiosyncratic drug reactions. Drug

Met. Rev., 2002. 34(3): p. 651-65.

224. Cocconi, G., First generation aromatase inhibitors--aminoglutethimide and

testololactone. Breast Cancer Res. Treat., 1994. 30(1): p. 57-80.

225. Harris, A.L., et al., Agranulocytosis associated with aminoglutethimide: pharmacological

and marrow studies. Brit. J. Cancer, 1986. 54(1): p. 119-22.

Page 226: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

205

226. Goss, P.E., M. Jarman, and L.J. Griggs, Metabolism of aminoglutethimide in humans:

quantification and clinical relevance of induced metabolism. Brit. J. Cancer, 1985. 51(2):

p. 259-62.

227. Seago, A., et al., Identification of hydroxylaminoglutethimide as induced urinary

metabolites of aminoglutethimide in C57BL/6 mice, in Biological oxidation of nitrogen in

organic molecules, J.W. Forrod and L.A. Damani, Editors. 1985, Ellis Horwood Ltd.:

Chichester. p. 149-1563.

228. Kiorpelidou, E., et al., IL-8 release from human neutrophils cultured with pro-haptenic

chemical sensitizers. Chem. Res. Toxicol., 2012. 25: p. 2054-2056.

229. Ali, H., et al., Comparison of in vitro and in vivo haemotoxic effects of aminoglutethimide

and glutethimide. Toxicol. in Vitro, 1990. 4: p. 381-83.

230. Nicholls, P.J., et al., Aminoglutethimide: absorption, physiological disposition and

pharmacokinetics, in Aminoglutethimide as an aromatase inhibitor in the treatment of

cancer: an international symposium held during the 13th International Congress of

Chemotherapy, G.A. Nagel and R.J. Santen, Editors. 1984, Hans Huber, Berne,

Switzerland: Vienna, Austria.

231. Lonning, P.E., et al., Single-dose and steady-state pharmacokinetics of

aminoglutethimide. Clin. Pharmacokinet., 1985. 10: p. 353-364.

232. Adam, A.M., I.D. Bradbrook, and H.J. Rogers, High-performance liquid

chromatographic assay for simultaneous estimation of aminoglutethimide and

Page 227: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

206

acethylamidoglutethimide in biological fluids. Cancer Chemother. Pharmacol., 1985. 15:

p. 176-178.

233. Saad, A., et al., Differential analysis of rat bone marrow by flow cytometry. Comp.

Haematol. Int., 2001. 10: p. 97-101.

234. Murray, F.T., et al., Serum aminoglutethimide levels: studies of serum half-life,

clearance, and patient compliance. J. Clin. Pharmacol., 1979. 11-12(704-711).

235. Uetrecht, J., et al., Metabolism of dapsone to a hydroxylamine by human neutrophils and

mononuclear cells. J. Pharmacol. Exp. Ther., 1988. 245: p. 274-279.

236. Rubin, R.L. and J.T. Cumutte, Metabolism of procainamide to the cytotoxic

hydroxylamine by neutrophils activated in vitro. J. Clin. Invest., 1989. 83: p. 1336-1343.

237. Robba, C., G. Mazzocchi, and G.G. Nussdorfer, Effects of a prolonged treatment with

aminoglutethimide on the zona fasciculata of rat adrenal cortex: a morphometric

investigation. Cell Tissue Res., 1987. 248: p. 519-525.

238. Loffler, S., et al., Clozapine mobilizes CD34+ hematopoietic stem and progenitor cells

and increases plasma concentration of interleukin 6 in patients with schizophrenia. J.

Clin. Psychopharmacol., 2010. 30: p. 591-595.

239. Iverson, S., et al., Effect of clozapine on neutrophil kinetics in rabbits. Chem. Res.

Toxicol., 2010. 23(7): p. 1184-91.

240. Mantovani, A., et al., Neutrophils in the activation and regulation of innate and adaptive

immunity. Nat. Rev. Immunol., 2011. 11: p. 519-531.

Page 228: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

207

241. Uetrecht, J., Immune-mediated adverse drug reactions. Chem. Res. Toxicol. , 2009. 22:

p. 24-34.

242. Watkins, P.B., Idiosyncratic liver injury: challenges and approaches. Toxicol. Pathol.,

2005. 33: p. 1-5.

243. Siest, G., et al., Transcription factor and drug-metabolizing enzyme gene expression in

lymphocytes from healthy human subjects. Drug Metab. Dispos., 2008. 36: p. 182-189.

244. Kiorpelidou, E., et al., IL-8 release from human neutrophils cultured with pro-haptenic

chemical sensitizers. Chem. Res. Toxicol., 2012. 25: p. 2054-2056.

245. Klebanoff, S.J., Myeloperoxidase: friend or foe. J. Leukocyte Biol., 2005. 77: p. 598-625.

246. Uetrecht, J., The role of leukocyte-generated reactive metabolites in the pathogenesis of

idiosyncratic drug reactions. Drug Metab. Rev., 1992. 24: p. 299-366.

247. Mackay, I.R., Hepatoimmunology: a perspective. Immunol. Cell. Biol., 2002. 80: p. 36-

44.

248. Uetrecht, J.P., Drug metabolism by leukocytes, its role in drug-induced lupus and other

idiosyncratic drug reactions. CRC Crit. Rev, Toxicol., 1990. 20: p. 213-235.

249. Uetrecht, J., N. Zahid, and R. Rubin, Metabolism of procainamide to a hydroxylamine by

human neutrophils and mononuclear leukocytes. Chem. Res. Toxicol., 1988. 1: p. 74-78.

250. Naisbitt, D.J., et al., Immunological prinicples of adverse drug reactions Drug Safety,

2000. 23: p. 483-507.

Page 229: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

208

251. Li, J. and J. Uetrecht, The Danger Hypothesis applied to idiosyncratic drug reactions, in

Adverse Drug Reactions, J. Uetrecht, Editor. 2010, Springer-Verlag: Berlin. p. 473-509.

252. Tobler, A. and H.P. Koeffler, Myeloperoxidase: localization, structure, and function, in

Blood Cell Biochemistry 3: Lymphocytes and Granulocytes, J.R. Harris, Editor. 1991,

Plenum Press: New York. p. 255-288.

253. Cocconi, G., First generation aromatase inhibitors--aminoglutethimide and

testololactone. Breast Cancer Res.Treat., 1994. 30(1): p. 57-80.

254. Choquet-Kastylevsky, G., et al., Increased levels of interleukin 5 are associated with the

generation of eosinophilia in drug-induced hypersensitivity syndrome. Brit. J. Dermatol.,

1998. 139: p. 1026-1032.

255. Langereis, J.D., et al., GM-CSF and TNF modulate protein expression of human

neutrophils visualized by fluorescence two dimensional difference gel electrophoresis.

Cytokine, 2011. 56: p. 422-429.

256. Murray, F.T., et al., Serum aminoglutethimide levels: studies of serum half-life,

clearance, and patient compliance. J. Clin. Pharmacol., 1979. 11-12: p. 704-711.

257. Khader, S.A., S.L. Gaffen, and J.K. Kolls, Th17 cells at the crossroads of innate and

adaptive immunity against infectious diseases at the mucosa. Mucosal Immunol., 2009.

2: p. 403-411.

Page 230: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

209

258. Zhu, X. and J. Uetrecht, A novel TH17-type cell is rapidly increased in the liver in

response to acetaminophen-induced liver injury: Th17 cells and the innate immune

response. J. Immunotox., 2012. Oct 8 [Epub ahead of print].

259. Holt, M.P., L. Cheng, and C. Ju, Identification and characterization of infiltrating

macrophages in acetaminophen-induced liver injury. J. Leukoc. Biol., 2008. 84: p. 1410-

1421.

260. Sunahara, K.K.S., P. Sannomiya, and J.O. Martins, Briefs in insulin and innate immune

response. Cell. Physiol. Biochem., 2012. 29: p. 1-8.

261. Harris, A.L., et al., Agranulocytosis associated with aminoglutethimide: pharmacological

and marrow studies. British Journal of Cancer, 1986. 54(1): p. 119-22.

262. Coleman, M.D., L.F. Khalaf, and P.J. Nicholls, Aminoglutethimide-induced leucopenia in

a mouse model: effects of metabolic and structural determinates. Environmental

Toxicology and Pharmacology, 2003. 15: p. 27-32.

263. Alvir, J.M., J.A. Lieberman, and A.Z. Safferman, Do white-cell count spikes predict

agranulocytosis in clozapine recipients? Psychopharmacol. Bull., 1995. 31: p. 311-314.

264. Iverson, S.L., In vitro and in vivo investigations into idiosyncratic drug reactions: the

role of reactive metabolites produced by the target tissue in terbinafine-induced

cholestatic hepatitis and antipsychotic-induced agranulocytosis, in Pharmacuetical

Sciences. 2002, University of Toronto: Toronto, ON Canada.

Page 231: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

210

265. Iverson, S., et al., Effect of clozapine on neutrophil kinetics in rabbits. Chemical

Research in Toxicology, 2010. 23(7): p. 1184-91.

Page 232: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

211

APPENDIX 1

Supplementary Material from Chapter 2: Changes in Gene Expression Induced by

Aromatic Amine Drugs

Table A1.1. SMX-induced hepatic gene changes at 12 hr (fold-change between treated and

controls).

RefSeq Gene symbol Gene name Fold-

change

p-value

Up-regulation

NM_001009965 Eiih hepatic protein EIIH 2.86 0.0227

J03863 Sds serine dehydratase 2.15 0.0237 NM_021660 Ihpk2 inositol hexaphosphate kinase 2 1.91 0.0206

NM_030845 Cxcl1 chemokine (C-X-C motif) ligand 1 1.67 0.0476

NM_198776 MGC72973 hemoglobin, beta, adult major chain 1.67 0.1509 NM_022671 Onecut1 one cut homeobox 1 1.61 0.3665

NM_033234 Hbb hemoglobin, beta 1.54 0.1647

NM_001034125 Per1 period homolog 1 (Drosophila) 1.53 0.1517 NM_001013853 LOC287167 globin, alpha 1.52 0.1826

NM_001010946 LOC288741 huntingtin interactor protein E 1.51 0.0480

NM_012571 Got1 glutamate oxaloacetate transaminase 1 1.51 0.0035 NM_001007146 Tob2 transducer of ERBB2, 2 1.50 0.0456

Down-regulation

NM_017136 Sqle squalene epoxidase -1.51 0.0336 NM_017268 Hmgcs1 3-hydroxy-3-methylglutaryl-Coenzyme A synthase 1 -1.52 0.0098

NM_080886 Sc4mol sterol-C4-methyl oxidase-like -1.55 0.0196

NM_057107 Acsl3 acyl-CoA synthetase long-chain family member 3 -1.57 0.0374 NM_013060 Id2 inhibitor of DNA binding 2 -1.58 0.0221

NM_053380 Slc34a2 solute carrier family 34 (sodium phosphate), member 2 -1.59 0.0076

NM_001012072 Ppp1r3c protein phosphatase 1, regulatory subunit 3C -1.61 0.0383 NM_001008352 Pmvk phosphomevalonate kinase -1.62 0.0108

NM_001006995 Acat2 acetyl-Coenzyme A acetyltransferase 2 -1.63 0.0259

NM_001106882 Unc5cl unc-5 homolog C (C. elegans)-like -1.66 0.0357 NM_019237 Pcolce procollagen C-endopeptidase enhancer protein -1.67 0.0075

NM_198738 Psat1 phosphoserine aminotransferase 1 -1.67 0.0772

NM_032063 Dll1 delta-like 1 (Drosophila) -1.68 0.0029 NM_212505 Ier3 immediate early response 3 -1.69 0.0022

NM_053328 Bhlhb2 basic helix-loop-helix domain containing, class B2 -1.72 0.0969

NM_001013071 Tm7sf2 transmembrane 7 superfamily member 2 -1.76 0.0139 NM_206950 Mig12 MID1 interacting G12-like protein -1.76 0.0010

NM_145677 Slc25a25 solute carrier family 25 (mitochondrial, phosphate), member

25

-1.86 0.0162

NM_001033066 Ddhd1 DDHD domain containing 1 -1.87 0.0323

NM_031327 Cyr61 cysteine rich protein 61 -1.87 0.0025 NM_021835 Jun jun proto-oncogene -1.97 0.0019

NM_001009632 G0s2 G0/G1 switch gene 2 -2.11 0.0201

NM_019232 Sgk serum/glucocorticoid regulated kinase 1 -2.70 0.0023 NM_053769 Dusp1 dual specificity phosphatase 1 -3.58 0.0020

Data shown are gene changes ≥ 1.50 fold expression. p-values are unadjusted (n = 4).

Page 233: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

212

Table A1.2. DDS-induced hepatic gene changes at 12 hr (fold-change between treated and

controls).

RefSeq Gene symbol Gene name Fold-

change

p-value

Up-regulation

NM_013215 Akr7a3 aldo-keto reductase family 7, member A3 (aflatoxin aldehyde reductase)

5.38 2.89 x 10-7

NM_001009965 Eiih hepatic protein EIIH 4.23 0.0038

NM_001013181 Zbtb16 zinc finger and BTB domain containing 16 3.55 0.0154 NM_001009920 Yc2 glutathione S-transferase ,Yc2 subunit 3.19 7.52 x 10-7

NM_001012111 Lpin1 lipin 1 2.59 0.0200

NM_001014162 RGD1309578 acyl-CoA synthetase medium-chain family member 5 2.37 0.0329 NM_022407 Aldh1a1 aldehyde dehydrogenase family 1, member A1 2.33 0.0049

NM_058208 Socs2 suppressor of cytokine signaling 2 2.08 0.1430

NM_012603 Myc myelocytomatosis oncogene 2.00 0.0238 BC89763 Nr1i3 nuclear receptor subfamily 1, group I, member 3 1.95 0.0052

NM_145775 Nr1d1 nuclear receptor subfamily 1, group D, member 1 1.94 0.0885

NM_001025623 Vnn1 vanin 1 1.92 0.0957 NM_052798 Zfp354a zinc finger protein 354A 1.85 0.1089

NM_198776 MGC72973 hemoglobin, beta, adult major chain 1.80 0.1046

Down-regulation

NM_001012072 Ppp1r3c protein phosphatase 1, regulatory subunit 3C -1.81 0.0129

NM_053380 Slc34a2 solute carrier family 34 (sodium phosphate), member 2 -1.82 0.0014 NM_032063 Dll1 delta-like 1 (Drosophila) -1.85 0.0008

NM_031678 Per2 period homolog 2 (Drosophila) -1.86 0.0074

NM_012565 Gck glucokinase -1.89 0.0450 NM_001013175 Apol3 apolipoprotein L, 3 -1.90 0.0005

NM_053328 Bhlhb2 basic helix-loop-helix domain containing, class B2 -1.93 0.0500

NM_138912 Ppp1r3b protein phosphatase 1, regulatory subunit 3B -1.96 0.0038 NM_001009632 G0s2 G0/G1 switch gene 2 -2.03 0.0263

NM_019138 Cyp7b1 cytochrome P450, family 7, subfamily b, polypeptide 1 -2.05 0.0286

NM_012703 Thrsp thyroid hormone responsive protein -2.08 0.0023 NM_001003706 LOC360228 WDNM1 homolog -2.17 0.0022

NM_021835 Jun jun proto-oncogene -2.49 0.0002

NM_031327 Cyr61 cysteine rich protein 61 -2.50 0.0001

NM_001033066 Ddhd1 DDHD domain containing 1 -2.50 0.0040

NM_206950 Mig12 MID1 interacting protein 1 -2.59 1.06 x 10-5

NM_145677 Slc25a25 solute carrier family 25 (mitochondrial, phosphate), member 25

-2.67 0.0008

NM_053769 Dusp1 dual specificity phosphatase 1 -3.01 0.0052

NM_019232 Sgk serum/glucocorticoid regulated kinase 1 -3.72 0.0003

Data shown are gene changes ≥ 1.80-fold expression. p-values are unadjusted (n = 4).

Page 234: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

213

Table A1.3. AMG-induced hepatic gene changes at 12 hr (fold-change between treated and

controls).

RefSeq Gene symbol Gene name Fold-

change

p-value

Up-regulation

NM_001009965 Eiih hepatic protein EIIH 12.94 0.0072 NM_145775 Nr1d1 nuclear receptor subfamily 1, group D, member 1 5.77 0.0291

NM_031576 Por P450 (cytochrome) oxidoreductase 4.78 0.0003

NM_133401 Abcb1a ATP-binding cassette, sub-family B, member 1A 4.50 0.0001 NM_133586 Ces2 carboxylesterase 2C 4.38 0.0051

NM_022407 Aldh1a1 aldehyde dehydrogenase family 1, member A1 4.15 0.0121

NM_001008321 Gadd45b growth arrest and DNA-damage-inducible, beta 3.81 0.0009 NM_001008363 Zfand2a zinc finger, AN1-type domain 2A 3.77 0.0076

NM_022927 Mid1 midline 1 3.71 0.0021

NM_001009474 Pir pirin (iron-binding nuclear protein) 3.67 2.98 x 10-5 NM_173295 Udpgtr2 UDP glucuronosyltransferase 2 family, polypeptide B1 3.59 0.0008

NM_144743 Ces6 carboxylesterase 2A 3.48 3.33 x 10-5

NM_013105 Cyp3a23 cytochrome P450, family 3, subfamily a, polypeptide 23/polypeptide 1

3.37 3.79 x 10-5

NM_144755 Trib3 tribbles homolog 3 (Drosophila) 2.79 0.0329

NM_012600 Me1 malic enzyme 1, NADP(+)-dependent, cytosolic 2.76 0.0076 NM_033352 Abcd2 ATP-binding cassette, sub-family D (ALD), member 2 2.57 0.0039

NM_001047090 Aph1b anterior pharynx defective 1b homolog (C. elegans) 2.56 0.0291 NM_175761 Hspca heat shock protein 90, alpha (cytosolic), class A member 1 2.52 0.0145

NM_013063 Parp1 poly (ADP-ribose) polymerase 1 2.46 0.0002

NM_001001509 Hdmcp solute carrier family 25, member 47 2.46 0.2365 NM_001007706 RGD1359191 tRNA methyltransferase 61 homolog A (S. cerevisiae) 2.40 0.1436

NM_001025623 Vnn1 vanin 1 2.35 0.3351

NM_173116 Sgpl1 sphingosine-1-phosphate lyase 1 2.34 0.0246 NM_001011901 Hsph1 heat shock 105/110 protein 1 2.34 0.0048

NM_001034090 Ephx1 epoxide hydrolase 1, microsomal 2.33 0.0009

NM_001109022 LOC368066 indolethylamine N-methyltransferase 2.32 0.1376 NM_013215 Akr7a3 aldo-keto reductase family 7, member A3 (aflatoxin

aldehyde reductase)

2.23 0.0305

NM_199103 Ikbkg inhibitor of kappa light polypeptide gene enhancer in B-

cells, kinase gamma

2.23 0.0006

NM_001009920 Yc2 glutathione S-transferase Yc2 subunit 2.23 0.0066

NM_031614 Txnrd1 thioredoxin reductase 1 2.23 0.0003 NM_131906 Slco1a4 solute carrier organic anion transporter family, member 1A2 2.22 0.0702

NM_001005384 Osmr oncostatin M receptor 2.21 0.0501

NM_212504 Hspa1b heat shock 70kD protein 1B 2.20 0.0130 NM_178091 Insig2 insulin induced gene 2 2.20 0.0018

M58040 Tfrc transferrin receptor 2.19 0.1851

NM_024127 Gadd45a growth arrest and DNA-damage-inducible, alpha 2.19 0.1613 NM_144752 Oas1b 2-5 oligoadenylate synthetase 1B 2.14 0.0020

NM_145091 Pdp2 pyruvate dehyrogenase phosphatase catalytic subunit 2 2.09 0.0115

NM_001007687 Cndp1 carnosine dipeptidase 1 (metallopeptidase M20 family) 2.08 0.0635 NM_001010946 LOC288741 huntingtin interactor protein E 2.08 0.0782

NM_053906 Gsr glutathione reductase 2.06 0.0000

NM_001013181 Zbtb16 zinc finger and BTB domain containing 16 2.06 0.5563 NM_201424 Ugt1a3 UDP glycosyltransferase 1 family polypeptide A3 2.06 2.86 x 10-5

AF082126 Ahr aryl hydrocarbon receptor 2.05 0.0733

NM_019283 Slc3a2 solute carrier family 3 (activators of dibasic and neutral amino acid transport), member 2

2.05 0.0485

NM_001008301

Usp14

ubiquitin specific peptidase 14 (tRNA-guanine

transglycosylase) 2.04 0.0135 ENSRNOT00000056174 Usp18 ubiquitin specific peptidase 18 2.04 0.0300

NM_153312 Cyp3a2 cytochrome P450, family 3, subfamily a, polypeptide 2 2.03 0.0116

NM_001007690 Fbxo30 F-box protein 30 2.01 0.0246 NM_053612 Hspb8 heat shock protein B8 2.01 0.0036

Down-regulation

NM_031678 Per2 period homolog 2 (Drosophila) -2.00 0.1059 NM_001003706 LOC360228 WDNM1 homolog -2.00 0.1222

NM_001108118 Itga5 integrin, alpha 5 (fibronectin receptor, alpha polypeptide) -2.01 0.0205

NM_031815 Inhbe inhibin beta E -2.01 0.0132

Page 235: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

214

NM_001011985 Mknk2 MAP kinase-interacting serine/threonine kinase 2 -2.02 0.0612

NM_080782 Cdkn1a cyclin-dependent kinase inhibitor 1A -2.03 0.3319

NM_198780 Pck1 phosphoenolpyruvate carboxykinase 1 (soulable) -2.05 0.0272 NM_012941 Cyp51 cytochrome P450, subfamily 51 -2.07 0.0346

NM_001109430 LOC680883 leucine-rich repeats and transmembrane domains 2 -2.07 0.0635

NM_012508 Atp2b2 ATPase, Ca++ transporting, plasma membrane 2 -2.08 0.0006 NM_001106873 LOC301113 solute carrier family 25 (mitochondrial; phosphate), member

23

-2.11 0.0257

NM_001034081 Paqr7 progestin and adipoQ receptor family member VII -2.14 0.0071 NM_178096 Nrep neuronal regeneration related protein -2.15 0.0273

ENSRNOT00000027756 Fads2 fatty acid desaturase 2 -2.20 0.0116

NM_001006995 Acat2 acetyl-Coenzyme A acetyltransferase 2 -2.28 0.0526 BC107921 RGD1309879 family with sequence similarity 89, member A -2.31 0.0121

NM_022392 Insig1 insulin induced gene 1 -2.32 0.0272

NM_012703 Thrsp thyroid hormone responsive -2.33 0.0441 NM_198790 Rgsl2h regulator of G-protein signaling like 2 homolog (mouse) -2.34 0.0385

NM_001033066 Ddhd1 DDHD domain containing 1 -2.41 0.1299

NM_001004096 Reg4 regenerating islet-derived family, member 4 -2.47 0.0026 NM_022392 Insig1 insulin induced gene 1 -2.50 0.0298

NM_022280 Lrat lecithin-retinol acyltransferase (phosphatidylcholine-retin -2.59 0.0212

U11038 Lox lysyl oxidase -2.70 0.3845 NM_001109912 Tsc22d1 TSC22 domain family, member 1 -2.72 0.0080

NM_013144 Igfbp1 insulin-like growth factor binding protein 1 -2.73 0.5886

ENSRNOT00000024658 Slc28a2 solute carrier family 28 (sodium-coupled nucleoside transporter), member 2

-2.80 0.0076

NM_053922 Acacb acetyl-Coenzyme A carboxylase beta -2.80 0.0197 NM_001008352 Pmvk phosphomevalonate kinase -2.83 0.0063

NM_053380 Slc34a2 solute carrier family 34 (sodium phosphate), member 2 -3.02 0.0025

NM_019138 Cyp7b1 cytochrome P450, family 7, subfamily b, polypeptide 1 -3.04 0.0815 NM_001012621 Slc13a4 solute carrier family 13 (sodium/sulfate symporters),

member 4

-3.15 0.0023

NM_139258 Bmf Bcl2 modifying factor -3.25 0.0123 NM_053769 Dusp1 dual specificity phosphatase 1 -3.61 0.0712

NM_145677 Slc25a25 solute carrier family 25 (mitochondrial, phosphate), member

25

-4.05 0.0076

NM_019232 Sgk serum/glucocorticoid regulated kinase 1 -5.37 0.0064

Data shown are gene changes ≥ 2-fold expression. p-values are adjusted using FDR (n = 4).

Page 236: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

215

APPENDIX 2

In Vitro Metabolism of AMG by Neutrophils to a Hydroxylated Metabolite

A2.1 BACKGROUND

Aminoglutethimide (AMG) has been found to decrease the number of circulating

neutrophils in mice, but it has the opposite effect on neutrophils in rats. This discrepancy could

be caused by a species difference in the metabolism of AMG because N-hydroxyl AMG was

observed in the urine of humans and mice [1, 2], but not in rats [3]. However, lack of detection

of the N-hydroxyl AMG metabolite does not mean that it does not form, and localized formation

of reactive metabolites in immune cells, which may not make a significant contribution to

urinary metabolites, may provide a greater stimulus for immune activation. To determine the

metabolic capacity of oxidative enzymes within the neutrophil to directly metabolize AMG,

peripheral blood neutrophils from rats were isolated and incubated with AMG. For this study

neutrophils from mice were tested in addition to rat neutrophils to determine whether there was a

species difference in AMG metabolism to the N-hydroxyl metabolite. Ascorbic acid was added

to the incubation to prevent the oxidation of N-hydroxyl AMG to the nitroso AMG metabolite.

Page 237: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

216

A2.2 MATERIALS AND METHODS

A2.2.1 Animals. Male Brown Norway rats (176-200 g) and male C57BL/6 mice (7-9

weeks old) were purchased from Charles River (Montreal, QC). Rodents were housed doubly

under standard conditions with automatic watering, 12:12 hr light/dark cycle at 22°C. Food and

water were provided ad libitum and all animals were given standard rodent chow. Animals were

acclimatized for one week before the start of experiments. The experiment protocol was

approved by and performed in accordance with the University of Toronto Faculties of Medicine

and Pharmacy Animal Care Committee.

A2.2.2 Neutrophil Isolation from Blood. For isolation of peripheral blood neutrophils,

blood (5 mL) was collected terminally, after CO2 asphyxiation, through the central vein from rats

or pooled from the venous sinus of mice into heparin coated syringes. Equal parts of 3% dextran

(in 0.9% NaCl) were mixed with the blood in 15 mL tubes and incubated at room temperature for

18 min. The straw-coloured upper layer was collected and centrifuged at 300g for 5 min. The

pellet was washed once more with Minimum Essential Media (MEM-α; Invitrogen, Life

Technologies) and the leukocytes were resuspended in a final volume of 2 mL MEM-α.

Isolated leukocytes were carefully overlaid on a Percoll gradient (80%, 65%, 55%; GE

Healthcare) and centrifuged at 400g for 30 min at 4°C. The supernatant was removed from the

65% layer and the neutrophils were collected from the 80/65% interface. The cells were washed

once with PBS and then lysed with 2 mL red blood cell (RBC) lysis buffer (155 mM NH4Cl, 10

mM KHCO3, and 0.1 mM EDTA) for 5 min and the reaction was stopped by adding 10 mL PBS.

Cells were washed again before resuspension in 2 mL Hank’s Balanced Salt Solution (HBSS)

Page 238: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

217

and cell counting using trypan blue and Countess Automated Cell Counter (Invitrogen, Life

Technologies).

A2.2.3. AMG Metabolism by Activated Neutrophils. The protocol to determine the

metabolism of AMG by activated neutrophils was performed similar to Uetrecht et al. [4].

Briefly, a reaction mixture of 2.5x106 neutrophils (in 250 µL HBSS), 1 mM ascorbate, 40 ng/mL

phorbol 12-myristate 13-acetate (PMA; 5 µL in DMSO), and 100 µM AMG (10 µL solution in

DMSO) was made and incubated at 37 °C in a shaking water bath for 10 min. The mixture was

then centrifuged at 13000g for 2 min and 10 µL of the supernatant was directly injected into an

LC system (SLC-10A, Shimadzu) coupled to a MS detection with ionspray (API 3000, PE Sciex)

to look for AMG metabolites using an Ultracarb 5 µ ODS (30) 100x2.0 mm column

(Phenomenex, Torrance, CA), mobile phase of water-methanol (80:20, v/v), and a flow rate of

0.2 mL/min. The supernatant was analyzed by selective ion monitoring for specific m/z ratios

that corresponded to AMG (m/z 232.5-233.5) and the N-hydroxy-AMG (m/z 248.5-249.5)

metabolite. A separate total ion scan was also performed to determine whether other metabolic

products formed.

Page 239: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

218

A2.3 RESULTS AND DISCUSSION

Using LC-MS in the selective ion monitoring (SIM) mode, peaks were observed for

AMG (m/z 233) and a hydroxylated AMG (m/z 249) both at a retention time of approximately

5.7-5.9 min in both rats and mice. The similarity in retention times is not surprising, as the

differences in polarity between AMG and a hydroxylated AMG metabolite are fairly small. A

separate total ion scan of the supernatant found no other substantial metabolic peaks.

Figure A2.1. Metabolism of AMG by rat and mouse neutrophils. Selected ion monitoring

was performed for AMG by scanning m/z 232.5-233.5 and for a hydroxylated AMG metabolite

scanning m/z 248.5-249.5.

Page 240: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

219

The differences in the magnitude of the formation of the hydroxylated AMG metabolite

between rats and mice were very subtle with a slightly higher formation observed in the mice

than rat (Figure A2.1). Interestingly, neutrophils were also able to form the hydroxylated AMG

metabolite in the absence of PMA stimulation. Although, less N-hydroxyl AMG was expected to

form in the absence of PMA, it is likely that the low levels of N-hydroxyl AMG observed

without stimulation were due to some activation of neutrophils through the isolation process on a

Percoll gradient or metabolism by other oxidizing enzymes within neutrophils.

Figure A2.2. Mass fragmentation of (A) AMG and (B) product of m/z 249.0.

A

B

Page 241: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

220

Further investigation into the hydroxylated metabolite through LC-MS/MS analysis

found that the fragmentation pattern was quite different from AMG (Figure A2.2). Nevertheless,

although these findings suggest that AMG has been hydroxylated by neutrophils, the exact

location of the hydroxyl group is unknown and hydroxylation of the heterocyclic ring is possible

as well (Figure A2.3). If the metabolic product is N-hydroxyl AMG we would expect to see the

loss of water and a m/z of 231. Although, we do see this fragment, the magnitude is low and we

would need to confirm; however, an authentic standard of N-OH AMG is not available.

Figure A2.3. Potential structures for m/z 249 hydroxylation of AMG.

These findings indicate that endogenous enzymes within rodent neutrophils are able to

oxidize AMG to a common metabolite associated with aromatic amine toxicity and the

differences between rats and mice in this metabolic capacity are quite subtle. Although we are

unable to confirm the identity of the metabolite, these findings may play a significant role in

AMG-induced agranulocytosis.

Page 242: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

221

APPENDIX 3

Effect of SMX and DDS on Peripheral Blood Neutrophils

A

24 480

5000

10000

15000

20000Control

SMX

DDS

Time (h)

To

tal W

BC

s

(x 1

03

cell

s/m

l b

loo

d)

B

24 480

500

1000

1500

2000

2500Control

SMX

DDS

Time (h)

Neu

tro

ph

ils

(x10

3/

ml

blo

od

)

Figure A3.1 Peripheral blood changes induced by SMX and DDS. Male Brown Norway rats

were treated with 150 mg/kg/day SMX or 20 mg/kg/day DDS and blood was taken from the tail

vein to determine the WBC differential using Wright-Giemsa staining after 24 and 48 h. No

Page 243: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

222

change was observed for both drugs in either (A) the total WBC count or (B) the number of

peripheral blood neutrophils. Values are mean ± SEM, n=4. None of the results are statistically

significant by two-way ANOVA.

Page 244: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

223

APPENDIX 4

Haematological Effects of AMG in Mice

0 7 14 210

500

1000

1500

2000

4000

6000

8000

10000Total WBCs

Lymphocytes

Neutrophils

Days

Cells (

x10

3/m

l b

loo

d)

Figure A4.1 Changes in peripheral blood WBCs induced by AMG in mice. Male C57BL/6

mice were treated with 75 mg/kg/day AMG through oral gavage in 0.5% methylcellulose

vehicle. Blood was taken at various time-points through the venous sinus of mice and the WBC

differential was determined through Wright-Giemsa staining. A decreasing trend was observed in

the total WBCs and lymphocytes with the most dramatic decrease found in the neutrophils.

Values are mean ± SEM, n=3.

Page 245: INVESTIGATION OF THE MECHANISMS OF AROMATIC AMINE

224

APPENDICES REFERENCES

1. Goss, P.E., M. Jarman, and L.J. Griggs, Metabolism of aminoglutethimide in humans:

quantification and clinical relevance of induced metabolism. Brit. J. Cancer, 1985. 51:

p. 259-62.

2. Seago, A., et al., Identification of hydroxylaminoglutethimide as induced urinary

metabolites of aminoglutethimide in C57BL/6 mice, in Biological Oxidation of Nitrogen

in Organic Molecules, J.W. Forrod and L.A. Damani, Editors. 1985, Ellis Horwood Ltd.:

Chichester. p. 149-1563.

3. Egger, H., et al., Metabolism of aminoglutethimide in the rat. Drug Met. Dispos., 1982.

10: p. 405-412.

4. Uetrecht, J., et al., Metabolism of dapsone to a hydroxylamine by human neutrophils and

mononuclear cells. J. Pharmacol. Exp. Ther., 1988. 245: p. 274-279.