Enzyme-specific Inhibition of human transferases by Dinitronaphthalene derivatives

by

Hilary Groom

A Thesis presented to The University of Guelph

In partial fulfilment of requirements for the degree of Master of Science in Molecular and Cellular Biology

Guelph, Ontario, Canada

© Hilary Groom, August, 2014

Abstract

Enzyme-Specific Inhibition of Human Glutathione Transferases by Dinitronaphthalene Derivatives

M. Hilary Groom Advisor: University of Guelph, 2014 Dr. P. David Josephy

The enzyme glutathione transferase (GST) superfamily catalyzes the detoxication of many xenobiotics and byproducts of aerobic metabolism. GSTs catalyze the conjugation of glutathione with reactive functional groups of xenobiotics. Overexpression of GSTs can contribute to drug and apoptosis resistance in cells and this has motivated research to develop GST inhibitors as adjuvants to .

1-Chloro-2,4-dinitrobenzene (CDNB) is one of the most commonly used substrates for

GST activity assays. This project focuses on the interactions of naphthalene congeners of CDNB with recombinant human GST (Alpha, Mu, and Pi classes). Several dinitronaphthalene derivatives were found to be GST inhibitors. 1-Chloro-2,4-dinitronaphthalene was not a GST substrate, but inhibited the CDNB conjugation activity of GST enzymes. The mechanism of inhibition occurs by formation of a dead-end intermediate Meisenheimer complex in the enzyme active site. Significant mutagenic and toxic effects of these dinitronaphthalene derivatives were observedby conducting Salmonella Ames assays.

iii

Acknowledgements

I would like to thank the many supporters who have assisted me with my project and offered me personal support while at the University of Guelph. First of all, I would like to thank my advisor, Dr. Josephy. He has provided excellent knowledge and guidance during my research project, and has greatly enriched my experience as a graduate student by his participation with the lab group. I am grateful for all the experiences and conferences I had the fortune to attend during my MSc.

I would like to thank my fellow lab mates. Leah McGurn, who was invaluable for helping me adapt to an unfamiliar school and Minah Hwang, who mentored me on unfamiliar lab equipment. As well as current lab mates, Kaja and Reema, I thank for their moral support. I appreciate my parents, for their moral support and for being proud of my work and my sister and pet dogs, I would like to thank for reminding me to have fun.

Lastly I would like to acknowledge the great mentors I encountered during my undergrad at UOIT. These professors and T.A. include Dr. Strap, Dr. Desaulniers, Dr. Bardin, and PhD- graduate, Tim. They have provided a great foundation of knowledge that has assisted me during my graduate work and they have helped me build my confidence to pursue graduate studies.

iv

Table of contents

Acknowledgements ...... iii i. List of Tables ...... vi ii. List of Figures ...... vii A. Introduction ...... 1 i. Glutathione transferases ...... 1 a. Glutathione ...... 3 b. GST conjugation and the mercapturic acid pathway ...... 4 c. Human glutathione transferase ...... 5 d. Catalytic mechanism of cytosolic GSTs ...... 6 e. Human cytosolic glutathione transferases ...... 7 f. Model substrates of human glutathione transferases ...... 10 g. Substrate specificity of human glutathione transferases ...... 13 ii. Inhibitors of glutathione transferase ...... 14 a. Applications of glutathione transferase inhibitors ...... 16 iii. Nitronaphthalenes ...... 20 a. Genotoxicity of nitroarenes ...... 21 b. Salmonella Ames test ...... 23 c. Genotoxicityof nitronaphthalenes ...... 25 d. Metabolism of 1-nitronaphthalene in mammals ...... 25 B. Study Summary ...... 26 C. Materials and Methods ...... 27 i. Chemicals ...... 27 ii.Media and Agar ...... 27 iii. Synthesis of dinitronaphthalene derivatives and GSH-CDNN conjugate ...... 28 a. Analytical analysis of DNNSG conjugate ...... 30 iv. Recombinant human GST expressing strains ...... 30 a. Recombinant GST purification ...... 30 v. Enzyme kinetics instrumentation ...... 33

vi. Inhibitor IC50 measurement ...... 33 vii.Characterization of inhibition mechanism (kinetics) ...... 34

v

viii. Characterization of Meisenheimer complexes ...... 34 a. Reversibility of Meisenheimer complex formation ...... 35 ix. Mutagenicity and Genotoxicity: Salmonella Ames Assay ...... 35 b. Salmonella YG1024 ...... 36 c. Salmonella mutagenicity assay protocol ...... 36 d. NICE automated colony counting ...... 36 D. Results and Discussion ...... 37 i. Synthesis of CDNN-GSH adduct...... 37 ii. Protein purification ...... 38 iii. Identification of GST inhibitors ...... 39 iv. Potency of GST inhibition by dinitronaphthalene derivatives ...... 41 a. Inhibition of GSTMu class enzymes by dinitronaphthalene derivatives ...... 44 b. Inhibition of GST Alpha class enzymes by dinitronaphthalene derivatives ...... 46 c. Inhibition of GST Pi class enzyme by dinitronaphthalene derivatives ...... 48 d. Other GST enzyme specific inhibitors ...... 49 v. Inhibition mechanism of DNN inhibitors ...... 52 vi. Meisenheimer complex formation in the GSTM2 active site ...... 53 a. Meisenheimer complex stability in the GST active site ...... 55 b. Meisenheimer complex formation by other GST inhibitors ...... 57 c. Possible intramolecular Meisenheimer complex formation of DNNSG ...... 59 vii. Salmonella Mutagenicity Assay ...... 60 viii. Biological functions of GSTM2-2...... 63 E. Conclusions ...... 66 F. References ...... 69 G. Appendices ...... 81 i. Numerical data ...... 81 a. GST IC50 Inhibition values ...... 81 b. Double reciprocal plots ...... 90 c. Salmonella mutagenicity assay ...... 92 ii. Supplementary graphs ...... 95 a. IC50 curves ...... 95 b. Double reciprocal graphs ...... 99

vi

i. List of Tables Page #

Table 1 Dinitronaphthalene derivatives investigated for GST inhibition………………29

Table 2 GST molecular weight and extinction coefficient values……...... 32

Table 3 IC50 values of dinitronaphthalene inhibitors towards GSTs………………...…41

Table 4 Absorption peak maxima of dinitronaphthalene inhibitors in buffer and in complex with GSTM2-2…………………..………………………………..….54

vii

ii. List of Figures Page #

Figure 1 GST substrates……………………………………………………………2

Figure 2 Glutathione………………………………………………………………..3

Figure 3 The four enzymatic steps of the mercapturic acid pathway……………....4

Figure 4 Human GSTP1-1 in complex with S-(2,4-dinitrobenzene)glutathione (DNBSG)……………………………………………………………….....9

Figure5 DCNB conjugation with GSH to yield CNBSG…………………….....11

Figure6 SNAr reaction of CDNB and formation of 1-(S-glutathionyl)-2,4,6- trinitrocyclohexadienate dead end Meisenheimer complex…………...... 13

Figure7 GST substrate CDNB and novel GST inhibitor CDNN…………..……14

Figure 8 Regulation of JNK signaling pathway by GSTP1-1……………………18

Figure9 Metabolic activation of nitroarenes and nucleophilic (NU) attack of nitrenium ion…………………………………………………………...... 23

Figure10 ESI-LC-MS and HPLC-UV vis analysis of DNNSG…………………....38

Figure11 SDS-PAGE of purified GSTs, stained with Brilliant Blue R-250…….....39

Figure 12 IC50 curve of CDNN inhibition of GSTM2-2 activity towards 0.5 mM CDNB, 1 mM GSH………………………………………………..……..42

Figure 13 GSTM2-2 in complex with S-methylglutathione……………...…………43

Figure 14 GST inhibitors, DNNSG and DNBSG………………………………...... 45

Figure 15 GSTA1-1 in complex with S-hexylglutathione………………………...... 47

Figure 16 GSTM2-2 specific inhibitor, 6-(7-nitro-2,1,3-benzoxadiazol-4- ylthio)hexanol (NBDHEX)……………………………………………....49

Figure 17 GSTMu class specific inhibitor, γ-glutamyl-(S-benzylcysteiny1)-β-alanine and GSTP1-1 specific inhibitor γ-glutamyl-(S-benzylcysteinyl)-(R)-(-)- phenylglycine…………………………………………………...... 50

Figure 18 GSTM2-2 inhibitors, (A) genistein; (B) kaempferol; (C) quercetin……..51

viii

Figure 19 Double reciprocal plot of DNNOEt inhibition of GSTM2-2 activity….….53

Figure 20 Difference absorption spectra of dinitronaphthalene inhibitors in the active site of GSTM2-2…………………………………………………………...55

Figure 21 CDNB substrate analogue 1,3,5-trinitrobenzene………………………...... 57

Figure 22 Mutagenicityof GST inhibitors (dinitronaphthalene derivatives) towards SalmonellastrainYG1024…………………………………...... 62

ix

List of Abbreviations

AA Aristolochic Acid ACN Acetonitrile ASK1 Apoptosis signal-regulating kinase 1 CDNB 1-Chloro-2,4-dnitrobenzene CDNN 1-Chloro-2,4-dinitronaphthalene CYP450 Cytochrome P450 monooxygenase dcmA Dichloromethane dehalogenase CDNB 1,2-Dichloro-4-nitrobenzene CNPSG S-(2-Chloro-4-nitro-phenyl)glutathione DMSO Dimethyl sulfoxide DNPSG S-(2,4-Dinitrophenyl)glutathione DNN 4-Dinitronaphthalene DNNOH 2,4-Dinitronaphthol DNNOMe 1-Methoxy-2,4-dinitronaphthalene DNNSG S-(2,4-Dinitronaphthalenyl)glutathione EDTA Ethylenediaminetetraacetic acid EtOH Ethanol ESI Electrospray ionization EWG Electron-withdrawing group GSH Glutathione GST Glutathione transferase HPLC High performance liquid chromatography IPTG Isopropyl β-D-1-thiogalactopyranoside JNK c-Jun N-terminal kinase LB Lysogeny broth MAPEG Membrane-Associated Proteins in Eicosanoid and Glutathione metabolism MAPK Mitogen-activated protein kinase MeOH Methanol MS Mass spectrometry NAT N-Acetyltransferase NBDHEX S-2,4-Dinitrobenzylglutathione-(6-(7-nitro-2,1,3-benzoxadiazol-4-ylthio)hexanol NICE NIST's Integrated Colony Enumerator 1-NN 1-Nitronaphthalene PMSF Phenylmethanesulfonylfluoride SNP Single nucleotide polymorphism SR Sarcoplasmic reticulum SDS-PAGE Sodium dodecyl sulfate polyacrylamide gel electrophoresis STAT Signal Transducer and Activator of Transcription TNB 1,3,5-Trinitrobenzene TLK-199 γ-Glutamyl-(S-benzylcysteinyl)-(R)-(-)-phenylglycine 2TY 2x Tryptone-yeast media UV-vis. Ultraviolet-visible

1

A. Introduction

i. Glutathione transferases

Members of the glutathione transferase (GST, EC 2.5.1.18) enzyme superfamily catalyze the covalent conjugation of glutathione (GSH) to electrophilic substrates. GST substrates include reactive xenobiotics (compounds foreign to an organism’s biochemistry) and endogenous reactive metabolites of aerobic respiration. Inert xenobiotics that lack electronegative functional groups can become GST substrates via oxygenation catalyzed by cytochrome P450 enzymes.

Exogenous sources of GST substrates includeenvironmental pollutants, pharmaceuticals, pesticidesand secondary plant metabolites. Endogenous sources of GST substrates include oxidized lipids and oxidized catecholamines.

GSH conjugation inactivates reactive electrophilic molecules (with the exception of small bifunctional haloalkanes, which are further activated by GSH conjugation)(Guengerich et al.,

1987)by replacing the electronegative functional group witha relatively inertbond to the sulfur atom of glutathione. Inactivation of quinones occurs by addition reactions with glutathione, which reduce quinones andinhibit further redox cycling of quinones. GSTs catalyze covalent conjugation of GSH to an electrophilic second substrate via the cysteine thiolate of GSH.

Conjugation reactions catalyzed by GSTs include the following:epoxide/aziridine addition, nucleophilic aromatic substitution, alkyl halide substitution, Michael addition, diphenyl ether addition-reduction and quinone addition-reduction, to name a few (Hayes et al., 2005)(Figure

1).Additional detoxifying functions of GSTs include peroxidase activity towards organic peroxides(Di Ilio et al., 1986). GST activity is important to detoxication and aerobic metabolism. Organisms that possess GST enzymes include aerobic bacteria, protozoans and all

2 eukaryotes (plants, insects, reptiles, fish, mammals etc.). Glutathione transferases are absent only in Archaea bacteria and anaerobic bacteria(Allocati et al., 2009).

Figure 1: Typical GST substrates: A. benzo[a]pyrene diol epoxide, B. 1-chloro-2,4- dinitronaphthalene, C. chlorambucil, D. 4-hydroxynonenal, E. fluorodifen, F. aminochrome GST enzymes also catalyzing biosynthesis reactions and regulate activity of other proteins by non-covalent interactions. GST-catalyzed biosynthesis reactions include GSH dependent isomerization of steroids and eicosanoidderivatives (Bruns et al., 1999;Johansson and

3

Mannervik, 2001;Habig et al., 1974a;Edwards and Knox, 1956). GST-protein interactions include binding to and regulating the activity of receptors. Receptors regulated by GST interactions are involved in cell proliferative pathways and intracellular calcium release(Adler et al., 1999;Abdellatif et al., 2007;Cho et al., 2001).

The research conducted in this thesis focuses on the inhibitor specificity of dinitronaphthalene (DNN) derivatives towards human GSTs. The applications of GST inhibitorswill be discussed in relation to modulating GST-protein interactions.

a. Glutathione

Figure2: Glutathione (γ-glutamylcysteinylglycine) Glutathione (γ-glutamylcysteinylglycine)is an endogenously synthesized tripeptide. It is comprised of a glutamateγ-peptide-bonded to a cysteinylglycine dipeptide (Figure 2)(Hopkins and Harris, 1929). In the GST active site, the cysteine thiol is ionized to a thiolate and attacks the electrophilic center of the second substrate. GSH conjugationreplaces the electrophilic center of the co-substrate with a relatively inert carbon-sulfur bond. As a result of GSH conjugation, the conjugate is less reactive and less potentially damaging than the electrophile. GSH conjugation alsolimits transmembrane mobility by adding polarity to the co-substrate. The γ-glutamate of

4

GSH is zwitterionic at physiological pH, and the added polarity limits solubility in lipid- membranes.

b. GST conjugation and the mercapturic acid pathway

The role of GSH conjugation in xenobiotic metabolism is well established as the initial step of the excretory mercapturic acid pathway(Habig et al., 1974b). The mercapturic acid pathway follows GSH conjugation with stepwise hydrolysis of the GSH tripeptide, catalyzed by

γ-glutamyltranspeptidase and cysteinyl-glycine dipeptidase(Figure 3).The hydrolyzed glutamate and glycine amino acids are recycled. The cysteine conjugate is acetylated by N- acetyltransferase 8(Veiga-da-Cunha et al., 2010). The mercapturic acid (N-acetyl cysteine conjugate) is excreted in the urine or bile(Habig et al., 1974b;Goldberg, 1980;Veiga-da-Cunha et al., 2010).

Figure 3: The four enzymatic steps of the mercapturic acid pathway; R-X indicates electrophilic group of xenobiotic

5

c. Human glutathione transferase

The human glutathione transferase superfamily is divided into three sub-families, whose family names are loosely based on their intracellular location. These families arethe cytosolic

(or soluble) GSTs, the mitochondrial GSTs, and membrane-associated proteins involved in eicosanoid and glutathione metabolism (MAPEG). The human mitochondrial GST family has only one member, GST Kappa(Ladner et al., 2004). Members of the human MAPEG family include the following enzymes, 5-lipoxygenase activating protein, leukotriene C4 synthase, microsomal prostaglandin E2 synthases -1 and -2 (mPGES), mGST1, mGST2 and mGST3(Martinez et al., 2008). The human cytosolic GST family is the largest GST family in humans. It contains seven sub-classes with one to five members per class. These cytosolic GST classes are Alpha (GSTA), Mu (GSTM), Pi (GSTP), Theta (GSTT), Omega (GSTO), Sigma

(GSTS) and Zeta (GSTZ). Excluding deletion polymorphisms in the population, there are in total 17 confirmed functional human cytosolic GSTs(Nebert and Vasiliou, 2004). Cytosolic

GSTs are catalytically active as dimers and thusare denoted as GSTX#-#. Cytosolic GSTsare capable of forming hetero-dimers within classes,e.g. GSTA1-2(Mannervik et al., 2005).

The human GST superfamily is polyphyletic. The ability of each GST family to conjugate GSH to co-substrates has developed by convergent evolution. The conserved GSH binding domain of GSTs assumes a thioredoxin-like fold. The thioredoxin fold of GST Kappa is distinct from cytosolic GSTs; it shares greater sequence identity and three dimensional structure with bacterial disulfide oxidoreductase (DsbA)(Ladner et al., 2004).MAPEG GSTs also lack to cytosolic GSTs and GSTKappa, suggesting their function as a glutathione transferase enzyme is convergent(Hayes and Pulford, 1995). Across families of GSTs, there is evidence for convergent functions (aside from drug metabolism), such as the ability to catalyze

6

the GSH dependent isomerization of PGH2 (prostaglandin H2) to PGE2. In the human liver this is catalyzed by MAPEG GSTs, microsomal PGE2 synthases -1 and -2. However, in the human brain, this reaction is primarily catalyzed by cytosolic GSTM2-2 and GSTM3-3(Beuckmann et al., 2000).

A feature of the cytosolic GST class is catalytic redundancy. Many cytosolic GSTs catalyze GSH conjugation reactions with common electrophilic substrates. It would be desirable to identify cytosolic GST enzyme specific substrates and inhibitors to determine the preferred substrates of individual GSTs.The DNN derivatives examined in this study were tested towards cytosolic GSTs. A goal of this study was to use the DNN derivatives to identify GST enzyme specific activity or inhibitor binding.

d. Catalytic mechanism of cytosolic GSTs

The catalytic mechanism of GSH conjugation by GSTs is the reduction thecysteine thiol pKa of enzyme bound GSH toapproximately physiological pH. Glutathione ionization can be monitored by UV vis spectroscopy; the glutathione cysteine thiolate absorbs at 239 nm.

Outside of the GST active site, the cysteine thiol of glutathione possesses a pKa of 9.2. In the rat

GSTM2-2 active site, the pKa of GSH is reduced to about 6.6 (the pKa decrease varies by GST)

(Graminski et al., 1989a). The pKa of rGSTM2-2 bound GSH was determined by titrating the rGSTM2-2-GSH complex over a pH range of 5 to 8 and monitoring glutathione thiolate formation at 239 nm.In cytosolic GSTs, a catalytically conserved serine, cysteine or tyrosine residue in the GSH binding site is essential for pKa reduction(Kong et al., 1992b). Previous studies conducted site-directed mutagenesis to change the catalytically conserved Tyr7 residue in recombinant human GSTP1-1 to a phenylalanine. The Y7F variantspecific activity towards co- substrate CDNB (1-chloro-2,4-dinitronaphthalene) was about 1% of wildtype(Kong et al.,

7

1992a). The Km values of CDNB and GSH were unaffected, indicating altered substrate binding was not a contributing factor to decreased enzyme activity. Activity of the Y7F variant was partially restored by increasing the pH of buffer above 8.0, confirming the essential role of the conserved Tyr residue in reducing GSH pKa(Kong et al., 1992b).

In rat GSTA1-1,the catalytic Tyr9 residue also possesses a reduced pKa (of the Tyr9 phenol) in the GST active site, relative to solution. A reductionof pKa from pH 10 to pH 7 was observed(Atkins et al., 1993). The ionized tyrosine deprotonates the GSH cysteine(Angelucci et al., 2005). The reduced pKa of the Tyr residue also alters the hydrogen-bonding equilibrium of the phenol hydrogen, which better stabilizes the GSH thiolate(Atkins et al., 1993). Reduction of the conserved Tyr residue pKa in rat GSTA1-1 is dependent on pressure exerted from the C- terminal helices of the GST enzyme. Atkins et al.,1997 monitored the pressure-dependent fluorescenceemission shifts of Tyr9 between phenolic (em. max. 305 nm) and tyrosinate (em. max. 345 nm) states in the Trp-null variant of rat GSTA1-1. Rat GSTA1-1 possesses one non- essential tryptophan in its protein sequence and fewer Tyr residues (eighttotal). This allows for the monitoring of Tyr ionization by fluorescence, without interference of tryptophan chromophores. Both pKa and ionization of Tyr9 decreased linearly with increasing active site pressure between 0 and 120 KPa. Decrease in active site solvation (-33mL/mol) accompanied pressure increase. The authors suggestedthat compaction of the GST active site and loss of solvation by second substrate ligand binding co-occur with GSH ionization(Atkins et al., 1993).

e. Human cytosolic glutathione transferases

Cytosolic GSTs are the most abundant GST family, found in all aerobic organisms(Frova,

2006). The progenitor of cytosolic GSTs is thought to have arisen from the

8

Methylbacteriumdichloromethane dehalogenase (dcmA). dcmA uses a glutathione cofactor to convert dichloromethane to formaldehyde + 2 Cl-ions,allowingMethylbacterium to use dichloromethane as a sole carbon source(La Roche and Leisinger, 1990). DcmA shows high sequence similarity to GSTTheta, a GST abundant in invertebrates and vertebrates. It is hypothesized that other GST classes derive from class Theta(da Fonseca et al., 2010).

There are seven human cytosolic GSTs classes, and within each class there are one to five members. Different members on each class have arisen by gene duplication. of each class are proximally located on different . Within GST classes,amino acid sequence identity is between 40 to <90%, and across classes <25%. Excluding polymorphic gene deletions, human cytosolic GST members are as follows: GSTA(1 to 5), GSTK1, GSTM(1 to 5), GSTO(1 and 2), GSTP1, GSTT(1 and 2), GSTZ1(Hayes et al., 2005). Heterozygous and homozygous gene deletions are reported in the population for GSTM1 and GSTT1. Most notably, the GSTM1*0-1*0 double gene deletion occurs at frequencies between 27%-53%, depending on population demographic. There are also reports of GSTM1 gene duplication(Garte et al., 2001). Cytosolic GST genes also display protein sequencepolymorphismsthat result in altered enzyme activities.

Cytosolic GSTs possess a conserved topology. For a three-dimensional structure of human GSTP1-1, see Figure 4(Oakley et al., 1999). All cytosolic GSTs must form homodimers

(or heterodimers within classes) for catalytic activity to occur. Each protomer is about 25 kDa in size and possesses two distinct domains, an N-terminal and a C-terminal domain. Within each domain, a portion of the active site residues is found. A highly conserved GSH binding site (G- site) is located in the N-terminus.Electrophilic second substrate binding occurs at the variable hydrophobic binding site (H-site) located in the C-terminus(Mannervik et al., 1985).

9

Combined,the G-site and H-site make up one active site per monomer. Monomer dimerization occurs at a dimer interface motif located in both the N-terminal and C-terminal domains.

A. B.

C. D. E.

Figure 4: Human GSTP1-1 in complex with S-(2,4-dinitrophenyl)glutathione (DNPSG); A. GSTP1-1 dimer B. GSTP1 monomer C. GSTP1 N-terminal domain (R3-76), D. GSTP1 C-terminal domain (R84-210), E. space filling model (green) and stick model of DNPSG; alpha helices are shown in red and beta sheets in blue (PDB ID 18GS). The N-terminal domain of cytosolic GSTs assumes a thioredoxin-like-fold tertiary structure, with a β1α1β2α2β3β4α3 motif. The thioredoxin fold is a conserved fold observed in other GSH binding-enzyme superfamilies, including GSH reductases, GSH peroxidases, and the disulfide oxidoreductase DsbA-like superfamily(Martin, 1995). Within the G-site, a conserved series of polar residues, the SNAIL/TRAIL motif, stabilizes GSH binding via interactions with

10 the polar functional groups of the GSH peptide backbone. The C-terminal domain is comprised entirely of α-helices (and turns) and forms the variable H-site(Mannervik, 1985).

Cytosolic GSTs are further classified into two groups, based on the conserved catalytic residue. As previously discussed, aconserved Tyr, Ser or Cys is located in the N-terminal G-site that is critical for GSH ionization and the conjugation activity of GSTs. One group is the catalytic Tyr residue in ‘Y-type’ GSTs. The conserved Tyr residue islocated in β-sheet 1. The other group is the catalytic Cys or Ser residue in ‘C/S-type’ GSTs. The conserved Cys/Ser residueis located in α-helix 1(Atkinson and Babbitt, 2009). Y-type GSTs include the classesAlpha, Mu, Pi and Sigma, all of which are found in humans. C/S-type GSTs include classes that are found in humans(Theta, Omega and Zeta) and classes that are not(Phi, Tau and

Beta). Y-type GSTs are the more recently diverged of the cytosolic GSTs; they are primarily found in mammals,whereas the C/S-type is abundant in prokaryotes and eukaryotes. Y-type

GSTs are the only enzyme group to possess a thioredoxin-like fold with a catalytic Tyr residue(Atkinson and Babbitt, 2009). The recombinant human cytosolic GST enzymes examined in this study are all Y-type GSTs of classes Alpha, Mu and Pi. In humans, they are the most widely expressed classes of GSTs and are the primary GST contributors to drug metabolism(Hayes et al., 2005).

f. Model substrates of human glutathione transferases

Mercapturic acids have been identifiedin urine as far back as the 1800s, when Baumann

(1879) identified sulfur-containing metabolites in canine urine after the animals werefed bromo-, chloro-,or iodobenzene-treated dog food. Halogenated arenes have since served as popular substrates for monitoring GST activity. TheUV active chromophores of aryl substratesallowfor product formation to be monitored by UV vis. spectrophotometry. GSH conjugation to anaryl

11 chromophore is accompanied by an absorption shift to a longer wavelength. This is due to the delocalization of electrons from the GSH sulfur atom into the aryl chromophore. Thus, product formation can be differentiated from the parental compound absorption.Glutathione transferase activity was first monitored in partially purified soluble rat liver extract using the substrate 3,4- dichloro-nitrobenzene (DCNB)(Booth et al., 1961)(Figure 5).DCNB peak absorption maximum is at 270 nm. GSH conjugation yields S-(2-chloro-4-nitro-phenyl-glutathione (CNPSG), which has an absorption peak maximum at 340 nm.

Figure 5: DCNB conjugation with GSH to yield CNPSG(Figure adapted from Booth et al., 1961)

1-Chloro-2,4-dinitrobenzene (CDNB) has spectral properties similar to DCNB and has replaced DCNB asthe standard model GST substrate. It was previously observed that partially purified GST fractions (from rat liver) had higher specific activities for CDNB, and CDNB activity could be detected in fractions that lacked considerable DCNB activity(Habig et al.,

1974b). CDNB has been found not to be a substrate for a few GSTs, including human

12

GSTTheta(Jemth and Mannervik, 1997). Despite this, CDNB is still considered the standard substrate for monitoring GST activity. All GSTs examined in this study accept CDNB as a substrate.

CDNB conjugation to GSH in the GST active site follows a nucleophilic aromatic substitution (SNAr) mechanism (Figure 6).During the SNAr reaction, a catalytic intermediate

‘Meisenheimer complex’/σ-complex forms in the GST active site. For compounds which lack a leaving group, such as trinitrobenzene, the GST SNAr reaction does not go to completion, and a dead end Meisenheimer complex forms in the GST active site (Graminski et al., 1989b) (Figure

6). Dead-end Meisenheimer complexes stay dormant in the GST active site. Due to their stability in the GST active site,dead-end Meisenheimer complexes of conjugated arenes can be observed by UV vis spectrometry. They also occupy the active site and inhibit GST catalytic activity in a competitive manner. The Meisenheimer complex shows distinct spectral properties from the start compounds and GSH conjugation production product; thus, they can be observed in the GST active site when the reaction does not go to completion. The TNB-GSH

Meisenheimer complex, 1-(S-glutathionyl)-2,4,6-trinitro-cyclohexadienate,is a competitive GST inhibitor.

13

Figure 6: A. CDNB SNAr reaction andB.formation of 1-(S-glutathionyl)-2,4,6-trinitro- cyclohexadienate dead-end Meisenheimer complex g. Substrate specificity of human glutathione transferases

Early work attempted to categorize GSTs based on substrate specificity for certain functional groups,e.g. glutathione-S-aryltransferase, glutathione-S-alkyltransferase, etc.(Boyland and Chasseaud, 1969). However,it was soon realizedthat there isno clear functional group specificity of GSTenzymes. Substrate promiscuity varies among GST. Some enzymes have high specific activities for certain functional groups,e.g. GSTA4-4 and unsaturated carbonyls

(Michael acceptors), GSTA2-2 and peroxides. Other GSTs have less distinct characteristic substrate profiles. Predicting substrates for specific GSTs is difficult(Hayes et al., 2005).

Identifying GST-specific substrates and inhibitors would be useful in characterizing human GST activities.

A primary goal of this study was to identify cytosolic GST specific substrates/ inhibitors.

The tactic was to take the universal substrate, CDNB, and change the aryl framework to a larger ring, naphthalene. This CDNB analogueis 1-chloro-2,4-dinitro-naphthalene (CDNN)(Figure

14

7).It was anticipated that by increasing substrate size (aryl ring size), a change in enzyme specificity would be observed. As a preliminary result, replacing the benzyl ring of CDNB with naphthalene produced (surprisingly) a GST inhibitor. The following study describes the identification and characterization of a novel group of GST inhibitors (possible applications are discussed in section ii). As a complementary study to the identification of dinitronaphthalene inhibitors, the mutagenic potentials of these compounds were also investigated by conducting standard Salmonella mutagenicity (Ames-test) assays. This was investigated becausenitroarenes are well-known mutagens (discussed in section iii).

Figure 7: A. Universal GST substrate CDNB; B. Novel GST inhibitor CDNN

ii. Inhibitors of glutathione transferase

The wide range of GST inhibitors reflects the broad substrate specificities of these enzymes. An additional site of inactivation of GSTs further increases the range of possible GST inhibitor structures. Allosteric inhibition of GSTs occurs by inhibitor binding to the GST dimer interface, as well as competitive inhibition at the active site. The GST dimer interface partially overlaps with the GST active site(Oakley et al., 1999).GSTs are known to bind large aliphatic

15 structures at the dimer interface. Theseinclude hormones, steroids, polyaromatic dyes, antibiotics, bilirubin and heme(Litwack et al., 1971). The non-substrate binding function of

GSTs was first attributed to the unidentifiedprotein termed ‘Ligandin’. Ligandin was later identified as an Alpha class GST(Habig et al., 1974a). Non-competitive GST inhibitors that bind to the dimer interface include iodocyanine green (azo-dye), 8-anilinonaphthalene sulfonate, bilirubin, 3,6-dibromosulfophthalein (dye) andcefalothin (antibiotic)(Ketley et al., 1975).

Natural product inhibitors of GSTsincludethe phenolic antioxidants,curcumin and ellagic acid. Both are mixed or non-competitive GST inhibitors(Hayeshi et al., 2007). Plant derived flavonoids kaemferol, quercetin and genistein inhibit GSTs, but the mechanism of inhibition is uncharacterized. Other plant secondary metabolites that are non-competitive inhibitors of GSTs include naphthala-quinones,diosphyrin and geshoidin(Hayeshi et al., 2004).

Synthetic inhibitors, designed specifically for GST inhibition, are discussed in detail in the Results sections. They include the competitive inhibitor NBDHEX (6-(7-nitro-2,1,3- benzoxadiazol-4-ylthio)hexanol), which forms a Meisenheimer complex in the GST active site(Ricci et al., 2005). Bivalent inhibitors have been developed which possess linkers that bridge the dimer interface and bind to both GST dimer active sites(Mahajan et al., 2006).Lastly, product mimic inhibitors-GSH conjugates have been investigated. They include GSH conjugateswith modified GSH peptides. They are referred to as ‘peptidomimetic inhibitors’ because they possess the GSH tripeptide,and they include the inhibitor γ-glutamyl-S-

(benzyl)cysteinyl-R-phenylglycine-diethylester(TLK199, ezatiostat®)(O'Brien et al., 1999).

16

a. Applications of glutathione transferase inhibitors

GST inhibitors have beeninvestigated for their ability to sensitize cancer cells to chemotherapeutics. Many chemotherapeutics possess functional groups that serve as GST substrates. Most are the DNA alkylating nitrogenmustards which possess halogenated alkyl functional groups(Hayes and Pulford, 1995;Hayes et al., 2005). GSTs are commonly overexpressed in human cancer tissues, including breast, colon, lung and stomach(Howie et al.,

1990). GSTP1-1 is the most commonly observedoverexpressed GST in cancer tissue. An early investigated adjuvant GSTP1-1 inhibitor was ethacrynic acid. Ethacrynic acid reached phase I human clinical trials(O'Dwyer et al., 1991). However, ethacrynic acid is a renal diuretic. Its associated renal toxicity has since prompted investigation of alternate GST inhibitors.Resistanceto chemotherapeutics also arises in which over express GST even when the drugs are not GST substrates. A second mechanism of GST-mediated chemotherapy resistance has been recently discovered: the function of GSTs as regulators of cell cycle progression.

In recent years, new functions of cytosolic GSTs have been discovered that are distinct from xenobiotic metabolism. These functions involve regulation ofkinases and regulation of receptoractivity. Specific GST-kinase interactions occur with MAPK (mitogen activated protein kinases) proteins, JNKs and ASK1. GST-receptor interactions occur withsarcoplasmic reticulum ryanodine receptors RyR1 (skeletal muscle specific) and RyR2 (cardiac muscle specific). Some of these functions have become new targets for discovery of GST inhibitors. Most research efforts are towards inhibition of GSTP1-1. GSTP1-1 binds to and inhibits JNKsthat down regulate cell proliferation(Adler et al., 1999;Wang et al., 2001).

17

JNK (c-Jun NH2 terminal kinase) proteins are a sub-family of mitogen activated protein kinases (MAPKs). There are three JNK isoforms; JNK1, JNK2 and JNK3, with different splice variants(Davis, 2000). JNK activity is dependent on phosphorylation by other upstream kinases.

These upstream kinases respond to stress,including stimuli such as cytokines (e.g., Tumor necrosis factor alpha, TNFα) and environmental stressors (e.g., , heat stress, UV light etc.)(Weston and Davis, 2002).Downstream targets of JNK phosphorylation regulate cell proliferation and apoptosis. They include a collection of transcription factors, including c-Jun, and non-transcription factors(Liu and Lin, 2005).

JNK effects on cell growth/death are isoform, splice variant and tissue specific. Jnk1-/- jnk2-/-mice are resistant to apoptosis induced by UV radiation, anisomycin (MAPK-JNK agonist), TNFα, and methylmethanesulfonate (alkylating agent)(Tournier et al., 2000).In cell models, prolonged JNK activation is required for TNFα and UV induced apoptosis(Liu and Lin,

2005). An example of a JNK downstream effector is the phosphorylation of the transcription factor c-Jun. C-Jun is activated by phosphorylation by JNKs at residues Ser63/73 and Thr91/93. c-Jun phosphorylation at residue Ser63/Ser73 has been observed to be essential for induced apoptosis in a number of neural and cancer cell lines(Watson et al., 1998;Madeo et al.,

2010).Recently, it has been discovered that GSTP1-1 binds to and inhibits activity of

JNKs(Adler et al., 1999). GSTP1-1 binding to and inhibiting JNK is thought to be an underlying mechanism by which GSTP1-1 over-expression contributesto drug and apoptosis resistance of cancer cells(Figure 8). This is in addition to the function of GSTP1-1 as an enzyme that deactivates chemotherapeutics(Townsend and Tew, 2003).

18

Extracellular stress Intracellular stress

• UV light, heat & • ER stress osmotic shock • Cytokines (TNFα) •TLR ligands ASK1

MKK4/ GST MKK7 P1-1 Inhibitor

GST JNK JNK JNK P1-1

GST P1-1 ROS GST c-jun P1-1 GST P1-1 No Apoptosis Apoptosis

Figure 8: Activation of JNK signaling pathway by stressors and downstream regulation of JNK activity by GSTP1-1.Figure adapted from(Simic et al., 2009). GSTP1-1 inhibits phosphorylation and activation of JNK1α2 and JNK2 under non-stress conditions(Adler et al., 1999;De Luca et al., 2012). GSTP1-1 binds to JNK2 via the GST C- terminal domain. However GSH binding to the GSTP1-1 N-terminal domain is able to inhibit

JNK interaction(Wang et al., 2001). GSTP1-1 dissociation from JNK also occurs by GSTP1-1 oligomerization. Under oxidative stress GSTP1-1 polymerizes by the formation of covalent cystine bonds between monomer units(De Luca et al., 2012).As a result of GSTP1-1 dissociation, the JNK pathway is able to transmit pro-apoptotic cascades from up-stream

MAPKs(Adler et al., 1999; Wang et al., 2001). Thus, cellular levels of GSHand oxidative stress regulate GSTP1 binding to JNK2. The use of GST inhibitors has also been investigated as a method of modifying GSTP1-JNK interactions. The GSTP1-1 inhibitor NBDHEX blocks

19

GSTP1-JNK interactions(Turella et al., 2005). NBDHEX is effective at sensitizing cancer cell lines to apoptosis bypreventing GSTP1 association with JNKs in several cancer cell lines(Turella et al., 2005).

The interaction of GSTP1 with JNKs is complicated, when considering that there are three classes of JNKs, with different splice variants, whose functions antagonize each other(De Luca et al., 2012). In GSTP1-/- mice, total white blood cell count is elevated 2-fold (other blood cell types remain unchanged) due to increased myeloproliferation(Gate et al., 2004). This occurs, in part, due to increased phosphorylation/activation of STAT (Signal Transducer and Activator of

Transcription) proteins by JNKs in the absence of GSTP1-1. The GSTP1 inhibitor TLK199

(generic name; ezatiostat·HCl, brand name; Telintra©) stimulates bone marrow myelo- proliferation in wildtype mice (expressing GSTP1-1)(Ruscoe et al., 2001). TLK199 has reached phase 3 clinical trials for the treatment of myelodysplastic syndrome, a disease state in which insufficient granulocyte precursor cells (mature cells; basophils, eosinophils, neutrophils and mast cells) are produced in bone marrowand severe chronic neutropenia, a disease state in which insufficient neutrophils are produced. For results of phase 1 clinical trials see (Raza et al., 2009) and for ongoing trials, see http://www.telintra.us/trials_main.html.

A kinase upstream of JNK is the ASK1 (Apoptosis signal-regulating kinase 1), also referred to as MAP3K5. ASK1 activity regulates the JNK and p38 signaling pathways. GSTM1-1 has recently been identified as a regulator of ASK1 activity.Overexpression of ASK1in COS7

(Fibroblast-like Kidney Cells) induces apoptotic cell death, and a dominant-negative mutant of

ASK1 prevents TNFα-induced apoptosis(Ichijo et al., 1997). Under non-stress conditions,

GSTM1-1 binds to ASK1. GSTM1-1 inhibits ASK1 by preventing ASK1 oligomerization, a requirement for ASK1 phosphorylation activity(Cho et al., 2001). Overexpression of GSTM1-

20

1inhibits heat-shock activation of ASK1(in Chinese hamster CCL39 and human HeLa cell lines)(Cho et al., 2001).In HEK 293 (human embryonic kidney cells), inhibition of ASK1 activation by GSTM1-1 inhibits downstream JNK signaling cascades and activation of downstream transcription factors(Dorion et al., 2002).

Less research has been done on modulating GSTM1-1 regulation of ASK1;emphasis has been placed on GSTP1-1•JNK interactions. It is also interesting to note that a prevalent

GSTM1-/- null genotype persists in human populations;its relation to ASK1 activity (and downstream effects) has not been thoroughly investigated. What has emerged from these findings on GST-MAPK interactions is the proposal of a new function for cytosolic GSTs, as endogenous regulators of the stress response-JNK signaling pathway(Cho et al., 2001).These newly identified functions of GSTs have gained the most attention for pharmaceutical applications of GST inhibitors and will be discussed as a possible application of the dinitronaphthalene GST inhibitorsidentified in this study.

iii. Nitronaphthalenes

The compounds in this study examined for GST inhibition are all C1-substituted 2,4- dinitronaphthalene derivatives. They include the following: 1-chloro-2,4-dinitronaphthalene

(CDNN) has applications in chemical synthesis and was previous used as a fungicide(Zander and

Hutzinger, 1970). 2,4-Dinitronaphthol (DNNOH) is used as a biological stain, and is also referred to as ‘Martius Yellow’. 1-Methoxy-2,4-dinitronaphthalene (DNNOMe) and 1-ethoxy-

2,4-dinitronaphthalene (DNNOEt) were previously known(Terrier, 2013). S-(2,4-

Dinitronaphthalene)glutathione (DNNSG) was synthesized specifically for this project. 1,3-

Dinitronaphthalene (1,3-DNN) is the only compound investigated in this study that is a documented environmental pollutant. This is the first study to address possible interactions of

21 this pollutant as a GST inhibitor. The sections below describe the genotoxicity of nitroarene compounds and the Ames mutagenicity assay. The Ames assay was used in this study to measure the potential mutagenic effects of the dinitronaphthalene GST inhibitors listed above.

Sources of nitroaromatic compounds include the exhausts of vehicle, coal, kerosene and diesel combustion, cigarette smoke and ammunition refuse.Commercial uses of dinitronaphthalenes includeammunitions and explosives and asintermediates in commercial chemical synthesis. The production of nitronaphthalenes by the German chemical company

Farbenfabriken Bayer during World War I was estimated to be105 tonnes/week. Soil samples retrieved from abandoned munitions sites in Germany had concentrations of di- and tri- nitronaphthalenes of 132 mg/kg and 676 mg/kg, respectively. Mono-nitronaphthalenes were detected at concentrations below 1 mg/kg(Bausinger et al., 2004). The authors suggested mononaphthalenes are biodegraded or subject to additional nitration, whereas polynitronaphthalenes are persistent pollutants(Bausinger et al., 2004). Dinitronaphthalenes also form in the atmosphere by spontaneous nitration of naphthalene.Gas-phase dinitration of naphthalene to form isomers 1,3-DNN, 1,2-DNN and 2,3-DNN has been reported ingaseous mixtures of NO2 and CH2Cl2(Squadrito et al., 1989). Under atmospheric conditions, nitration of naphthalene and benzene occurs by reactions with NO3 and •OH radical(Atkinson and Arey,

2007). Nitration of naphthalene is observed by reactions withgaseous N2O5 which forms fromatmospheric O3 and NO2(Pitts et al., 1985). These studies indicate that dinitronaphthalenes are persistent environmental contaminants.

a. Genotoxicity of nitroarenes

Classified class 1 carcinogens, according to IARC (International Agency for Research on

Cancer), with nitroarene structures include aristolochic acids (AA) and aza-thiopurine (nitro,

22 amine-heterocycle).AAs are secondary plant metabolites found in members of the plant family

Aristolochiaceae. AAs are nephrotoxicants and inducers of urothelial cancer; AAs are thought to be the causative agents of Balkan endemic nephropathy(Arlt et al., 2007).

DNA damage by nitroarenes may occur by two mechanisms, metabolic activation and to a lesser extent by DNA intercalation(Rosenkranz and Mermelstein, 1983). Metabolic activation of nitroarenes occurs by reduction of the nitro groups to generate a nitrenium ion or oxidation of the aryl structure to form an epoxide. The reduction of aryl nitro groups generatesa reactive nitrenium ion in several steps. The first step is metabolic nitroreduction to a hydroxyarylamine, catalyzed by several different mammalian enzymes including CYP450 reductase, xanthine oxidoreductase, aldehyde oxidase, DT-diaphorase, NADH-cytochrome b3 reductase and mitochondrial succinate dehydrogenase(Guise et al., 2007;Ueda et al., 2005).

The second step is the O-acetylation of the hydroxyarylamine by N-hydroxyarylamine-O- acetyltransferases/ N-acetyltransferases. Human N-acetylating proteins capable of activation of nitroarenes include NAT1 and NAT2(Blum et al., 1990). This generates an unstable N,O- substituted acetate group that spontaneous dissociates to generate a very electrophilic nitrenium ion (Figure 9)(Wild, 1990;Josephy and Novak, 2012). Likewise, the mutagenic effects of arylamines and arylazides are attributed to metabolic activation to nitrenium ions(Wild et al.,

1989). Other mechanisms of nitroary activation include oxidation of the aryl ring catalyzed by cytochrome P450 monooxygenases (CYP450). This generates an electrophilic epoxide ring(Gelboin, 1980). Both nitrenium ions and epoxide rings are subject to nucleophilic attack by

DNA bases or protein residues. The result is the formation of DNAand protein adducts, with mutagenicity and cellular toxicity as potential consequences.

23

Figure 9: Metabolic activation of nitroarenes and nucleophilic (NU) attack of nitrenium ion

b. Salmonella Ames test

The Salmonella mutagenicity (Ames) test was developed by Bruce Ames in 1973 as a means to quantify and compare the mutagenic potential of substances(Ames et al., 1973). The

Ames test uses Salmonellatyphimuriumstrains with mutations in the histidine biosynthesis operon. The strains are histidine auxotrophs that require supplemental histidine to grow.

Salmonella strains used in the assaydiffer by mutation types,e.g. base-pair substitution or frame- shift mutation. These mutations are located in mutation sensitive ‘hotspots’. An active mutagen could cause a mutagenic event in the hotspot of the histidine operon that reverses the histidine auxotrophy, allowing the bacterial cellto grow on minimal medium. The histidine auxotrophy is reverted and the colonies that grow on histidine minimal medium are termed ‘revertant colonies’.Histidine minimal medium is used to allow cells to undergo a few rounds of division, and for revertants to express functional His biosynthesis enzymes. Non-revertant auxotrophs divide until the limiting His is consumed. The non-revertants grow toform a ‘background lawn’, whereas revertants continue to divide to form colonies. The number of revertant colonies indicates the number of histidine-auxotrophy reverting mutagenic events. All tests are run with a background control (solvent without mutagen) to measure spontaneous reversions.

24

Ames tests are recognized worldwide by regulatory agencies and are usually the initial tests done to measure the potential mutagenic effects of chemicals/ drugs. Benefits of this bacteria based mutagenicity test include the following(Mortelmans and Zeiger, 2000). The experiment period is short; results are generated in 2 days, typically, and tests can be done in a high through-put manner. Salmonella strains with different mutation sensitivities,e.g. TA98

(frame-shift mutation sensitive) and TA100 (base-pair substitution sensitive), allow for the characterization of the mutation mechanisms of samples. Mutagen incubations prepared with animal liver S9-fractions allow for tests to include the metabolic activating effect of liver enzymes. As an alternative to animal liver extract, a wide range of Salmonella tester strains have been engineered to express bacteria and human metabolizing enzymes e.g. CYP450s, NATs,

GSTs etc.(Josephy, 2002).These strains allow for metabolic activation of samples to occur withoutliver S9 fractions.

Aderivative of SalmonellaTA98 was selected for use in the mutagenic Ames assay in this study because of its sensitivity to nitroarenes. TA98 possesses mutations which sensitize the strain to mutagens. The uvrB gene deletion eliminates DNA excision repair mechanisms. Arfa mutation results in faulty lipopolysaccharide synthesis, which makes the bacterial cell membrane morepermeable to large chemicals. Plasmid pKM101 carries the muc genes (mutagenesis; UV and chemical), which encode a DNA polymerase ‘pol R1’ (and associated DNA binding proteins) that bypasses DNAdamage sites. Salmonella TA98 is sensitive to frame-shift mutagenic events, specifically a 2-basepair deletion in a CG repeat hotspot. A TA98 derivative, strainYG1024,was used to test the mutagenicity of dinitronaphthalene derivative in this study,because of its sensitivity to nitroarene mutagenicity. Strain YG1024 overexpressesSalmonellaarylamine N-acetyltransferase/ N-hydroxyarylamine-O-

25 acetyltransferase.As mentioned earlier, this enzyme acetylates hydroxyarylamines, a step involved in nitro-group activation (Josephy et al., 1997;Watanabe et al., 1989)(Figure 9).

c. Genotoxicityof nitronaphthalenes

Mutagenicity of nitronaphthalene isomers and dinitronaphthalene regio-isomers has been reported previously with Ames strains TA98 and TA100(Rosenkranz et al., 1980). Mono nitronaphthalenes were identified as weak acting mutagens while dinitronaphthalenes weremore potent. For both mono and di-substituted naphthalenes,the mutagenic potency varied by regioisomer. For mononaphthalenes, the difference in mutagenicity has been attributed to the different DNA adductsformed by different isomers. 1-Nitronaphthalene forms adducts with the

O6atom of guanine. 2-Nitronaphthalene forms adducts with the C6 and N2 atoms of guanine and the N6atom of adenine (Kadlubar et al., 1980). Studies with 2,4-dinitronaphthol (DNNOH), an inhibitor examined in this study, showedhigher mutagenicity by frame-shift mutation towards strain TA98 vs. basepair substitution in towards strain TA100.Studies on methoxy and ethoxy substituted dinitronaphthalenes have not been conducted. As a part of the research conducted for this study, the mutagenic effects of DNN, CDNN, DNNOH, DNNOMe,

DNNOEtinSalmonellastrain YG1024 weremeasured.

d. Metabolism of 1-nitronaphthalene in mammals

Metabolism of 1-nitronaphthalene (1-NN) has been described. Glutathione conjugates are the major metabolites of 1-NN in mouse liver and lung extracts(Watt and Buckpitt, 2000). In animal tissue 1-NN is oxidized to C5-C6 and C7-C8 epoxides, which are GST substrates. GSH conjugationyields,1-nitro-7-hydroxy-8-glutathionyl-7,8-dihydronaphthalene and 1-nitro-5- hydroxy-6-glutathionyl-5,6-dihydronaphthalene(Watt and Buckpitt, 2000). The major excretory metabolitefound in mouse urine,post-administration of intravenous 1-nitronaphthalene,wasthe

26 mercapturic acid, N-acetyl-S-(hydroxy-1-nitro-dihydronaphthalene)-L-cysteine(Halladay et al.,

1999). These results confirm that epoxides of 1-nitronaphthalene are conjugated by GSTs in mammals. In the Rhesus macaque, inhaled 1-nitronaphthalene forms adducts to a wide range of cellular proteins in the brachial airway and lung, in addition to forming DNA adducts. These results confirmed that 1-nitronaphthalene is mobile across cellular membranes and forms both cytotoxic and genotoxic adducts(Lin et al., 2006).

The structures of the dinitronaphthalene compounds in this study are analogous to 1- nitronaphthalene and to the universal GST substrate CDNB. Based on structure this suggests they will bind to the GST active site as competitive inhibitors. Previous studies also indicate dinitronaphthalene compounds are mutagenic in the Salmonella Ames assay and potentially cytotoxic via cellular protein adduction. Thus, it was anticipated that the compounds in this study would also test positive for mutagenicity in the Ames test.

B. Study Summary

The following researchdescribes theinteractions of human cytosolic glutathione transferases with dinitronaphthalene derivatives. Based on the preliminaryfinding of CDNN (1- chloro-2,4-dinitronaphthalene)inhibition of GSTs, presented here are novel findings of a group of compounds that selectively inhibit human GSTs. As a complementary study, the mutagenicities of these GST inhibitors were examined with theSalmonella Ames assay.

27

C. Materials and Methods

i. Chemicals

Sources of chemicals were as follows: glutathione-agarose, IPTG (isopropyl-β-D-1- thiogalactopyranoside), ampicillin sodium salt, tetracycline·HCl, β-mercaptoethanol, Martius

Yellow (2,4-dinitronaphthol), D-biotin (99%), L-histidine·HCl·H2Oand bovine serum albumin

(BSA): Sigma-Aldrich (Oakville, ON); 1-chloro-2,4-dinitrobenzene (CDNB; 98%): Alfa Aesar

(Ward Hill, MA); glutathione (99.6%): ChemImpex (Wood Dale, IL); 2,4-dinitronaphthalene

(DNN; 100%), 1-chloronaphthalene (95.5%), 1-naphthol (100%) (Accu-Standard, Inc. (New

Haven, CT)); Brilliant Blue R-250: (Fisher Bioreagents); Difco agar, Tryptone, Yeast Extract:

Becton, Dickson Co. (Sparks, MD); Oxoid Nutrient Broth No. 2: Oxoid, Ltd. (Hampshire,

England); lysozyme: Boehringer-Mannheim (Germany).

ii. Media and Agar

2x TY medium for growth of recombinant protein-expressing bacteria was prepared with ddH2O as follows (per 1L); 5 g NaCl, 10 g yeast extract, and 16 g tryptone. Ampicillin wasprepared in ddH2O, filter sterilizedand added as appropriate (100 μg/mL).

Mediumfor the Salmonella Ames mutagenicity assay was preparedas described by(Maron and Ames, 1983).Salmonella YG1024 was grown in Oxoid nutrient broth No. 2 (2.5 g/ 100 mL).

Ampicillin (25 μg/mL) and tetracyclin (6.25 μg/mL)were prepared as previously mentioned and added as appropriate.

Minimal glucose bottom agar was prepared with the following; 20 mL 50x VB salts (50x per 1L; 10 g MgSO4·7H2O, 100 g citric acid·H2O, 500 g K2HPO4, 175g NaHNH4PO4·4H2O), 50 mL of 40% (w/v) D-glucose, 15g Difco agar, and autoclaved.

28

Top agar was prepared with 600 mg/L Difco agar, 500 mg/L NaCl and autoclaved.

Histidine- biotin mix was prepared in ddH2O, filter sterilized (per 100 mL: 9.6 mg L- histidine·HCl·H2Oand 12.36 mg D-biotin) and added to top agar for a final concentration of 50

μM, each.

iii. Synthesis of dinitronaphthalene derivatives and GSH-CDNN conjugate

The GST inhibitors examined in this study are listed above (Table 1). DNN and

DNNOH were purchased from commercial sources (see section i. Chemicals). CDNN,

DNNOMe and DNNOEt were synthesized by collaborators Moses Lee and Pravin Patil.

Synthesis of these three compounds is described elsewhere(Groom et al., 2014). 1-

Chloronaphthalene and 1-naphthol were purchased from commercial sources and also examined for GST inhibition. These two compounds were found to have no effect on GST activity and no further investigation was carried out (data not shown).

DNNSG was synthesized following a modified method by(Shiotsuki et al., 1990). 1 mM

CDNN and 1 mM GSH was dissolved in 1.2 mL EtOH. Solution was stirred at RT and 2 M

NaOH (1 mL) was added drop-wise. Immediately upon addition of base, solution colour changed from yellow to red-orange. Reaction continued at RT for 30 min and was stopped by neutralization by addition of 3 drops conc. HCl. Red-orange solids were collected by vacuum filtration and residual start compounds removed by washing (twice) with ice cold 1 mL volumes of water, ethanol and ethyl acetate. Solids were re-crystallized from 5 mL hot ethanol.

29

Table 1: Dinitronaphthalene derivatives studied for GST inhibition

R- group Name & abbreviation Source

H 1,3-dinitronaphthlene (DNN) Commercial

OH 2,4-dinitronaphth-1-ol (DNNOH) Commercial

Cl 1-chloro-2,4-dinitronaphthalene Lee1 (CDNN)

OMe 1-methoxy-2,4-dinitronaphthalene Lee1 (DNNOMe)

OEt 1-ethoxy-2,4-dinitronaphthalene Lee1 (DNNOEt)

Glutathione S-(2,4-dinitronaphthalene)- This study glutathione (DNNSG)

1Moses Lee and Pravin Patil, Hope College, Holland, MI.

30

a. Analytical analysis of DNNSG conjugate

Purity of the synthesized DNNSG conjugate was assessed by HPLC-UV analysis and

ESI-HPLC-MS. HPLC analysis was carried out on a Gilson HPLC system equipped with auto injector; model no. 234, dual pumps; model no. 322, Waters Symmetry® C18 column (5 μm;

4.6 × 150 mm), absorbance detector; model 152, and UniPoint™ software. ESI-HPLC-MS analysis was conducted using a Dionex UHPLC UltiMate® 3000 liquid chromatograph interfaced to a Bruker amaZon SL® iontrap mass spectrometer (Bruker Daltonics, Billerica, MA,

USA) at the Advanced Analysis Centre, University of Guelph. Peak absorption maxima was characterised by UV-vis spectrophotometry using a Cary 300 (Agilent Technologies©) dual beam spectrophotometer.

iv. Recombinant human GST expressing strains

Recombinant human GST-expressing E. coliXL-1 Blue strains were the kind gift of Dr.

BengtMannervik (Uppsala University, Sweden) (Mannervik et al., 1999).

a. Recombinant GST purification

Recombinant GST proteins were expressed in E. coli and purified using a previously published method, with modifications(Mukanganyama et al., 2002).AllGST expressing

E. colicultures were grown with vigorous shaking at 37°C.Overnight culture (10 mL) was grown in 2TY media with ampicillin (0.1 mg per mL). The next day, overnight culture wasdiluted 100- fold into 1L of 2TY medium with ampicillin (0.1 mg per mL).GST expression was induced at an

OD600= 0.5 with 1 mM IPTG. Induction proceeded overnight.Cultures were separated into four

250 mL centrifuge tubes. Cells were pelleted by centrifugation for 20 minat 15000 rpm, 4°C.

31

Cell pellets wereresuspended in 12 mL lysis buffer(pH 7.4; 50 mM TrisHCl (7.88 g/L),

100 mM NaCl (2.9 g/L), 1 mM EDTA (300 mg/L), 1 mM β-mercaptoethanol (70 μL/L) prepared in ddH2O and filter sterilized before use).The serine protease inhibitor phenylmethane- sulfonylfluoride (PMSF, dissolved in dimethyl sulfoxide (DMSO)) was added to a final concentration of 0.1 mM (solvent concentration, 1%). Resuspended cells were incubated with

0.25 mg/mL lysozyme for 30 min on ice,followed by sonication on ice for 5 min. Lysate was collected by centrifugation for 30 min at 14000 rpm, 4°C. Lysate was combined and a second aliquot of 0.1 mM PMSF added before further purification.

Lysate GSTspecific activity was measured using standard GST activity conditions(Habig et al., 1974b),with 1 mM CDNB and 1 mM GSH in 100 mM potassium phosphate buffer, pH

6.5. Protein concentration was determined by Bradford assay,with BSA as the reference standard(Bradford, 1976). Specific activity was calculated monitoring formation of theCDNB-

GSH conjugate; S-(2,4-dinitrobenzene)glutathione,ε=9.6 mM-1cm-1;

GST was purified from cell lysate on a column of glutathione-agarose affinity resin (2 mL bed vol.). Lysate was applied to the column and incubated with the resin for 30 min, with gentle shaking at 4 °C. To remove non-specific binding proteins, the column was washed with cold PBS until no protein was present in the eluate. Eluate protein concentration was measured by absorbance reading at 280 nm. GST was eluted with 3 mL of elution buffer, pH 9 (50 mM tris base(6.06 g/L) and50 mM GSH (15.35 g/L)). The eluate was collected and transferred todialysis tubing(Fisherbrand® regenerated cellulose MW cut-off 12-14 kDa). To remove glutathione,protein was dialyzed at 4°C indialysis buffer (100 mM potassium phosphate buffer,

32 pH 6.5, 10% glycerol (v/v), 1 mM β-mercaptoethanol (70 μL/L), 1L).Dialysis occurred overnight. The following morning, overnight buffer was replaced with fresh buffer and dialysis continued for 1-4 h. Protein samples were aliquoted into cryo-vials, frozen on dry ice and stored at -80°C until use.The GSH-agarose column was cleaned with alternating washes of

Regenerating Buffer A (100 mM tris base (15.76 g/L), 0.5 M NaCl (29.22 g/L), pH 8.5) and

Buffer B (100 mM sodium acetate (13.6 g/L), 0.5 M NaCl (29.22 g/L), pH 4.5), 5 bed vol, x3 washes. The column wasstored in 10 mL column storage buffer(2 M NaCl (117 mg/L), 1 mM

NaN3 (65 mg/L)) at 4 °C.

Purified protein concentration was measured at 280 nm using the following molecular weight and extinction coefficients, obtained from ExPASy online ProtParam calculator (based on amino acid composition) (http://web.expasy.org/protparam/, Table 2). Specific activity was calculated using methods previously described.

Table 2: GST molecular weight and extinction coefficient values

GST M.W. ε (cm-1M-1)

GSTA1 25630 20400 GSTA2 25664 20400 GSTA4 25704 17420 GSTM1 25712 40130 GSTM2 25745 39880 GSTP1 23360 29130

SDS gels were prepared with Amresco’s NEXT GEL® 10% acrylamide solution (Lot

#2642C120) and running buffer (Lot #0832C047). Gels were cast with 0.6 % (v/v) ammonium persulfate and 0.06% (v/v) tetramethylethylenediamine. Gels were stained with Coomassie stain

33

solution (45% (v/v) MeOH, 45% (v/v) ddH2O, 10% (v/v) CH3COOH, 0.25 (w/v) Brilliant Blue

R-250) and destained with the same solution lacking Brilliant Blue.

v. Enzyme kinetics instrumentation

All enzyme kinetic activity measurements and absorption spectra were conducted using a

Cary 300 dual-beam UV-visspectrophotometer. Spectrometry readings were taken in 1.5 mL quartz cuvettes or 3 mL tandem Yankeelov quartz cuvettes when indicated(Yankeelov, 1963).

vi. Inhibitor IC50measurement

Inhibition potency of dinitronaphthalene derivatives towards GST catalytic conjugation of CDNB and GSH was determined by IC50 curve analysis. For screening purposes,preliminary tests were conducted with 25 μM inhibitor and different GSTs. Inhibitor-enzyme combinations which showed appreciable inhibition with 25 μM inhibitor were characterized by IC50 analysis.

Assay conditions are as follows: purified human GST was incubated with 1 mM GSH and varying concentrations of inhibitor in 100 mM potassium phosphate buffer, pH 6.5 for 5 min at

37 °C. The incubation was transferred to a quartz cuvette. CDNB activity was initiated by addition of 0.5 mM CDNB. GST activity towards CDNB was monitored by measuring product formation at 340 nm. Control experiments, with no inhibitor, were carried out under identical condition with equal volume of solvent instead of inhibitor. Activity assays used 15 μg enzyme

(~600 nM), with the exception of high potency inhibitor-enzyme combinations, where less enzyme was used,to maintain the condition [I] >> [E]. Enzyme concentration was decreased to the lowest amount which would still yield control activity suitable for inhibition comparison.

1 μg (~40 nM) GSTM2 was assayed with CDNN, DNN, DNNOMe, DNNOEt and DNNSG; 1

μg GSTM1 with DNN and 5 μg (~200 nM) GSTA1 with DNNOH.

34

Activity assays were conducted with a final incubation volume of 1 mL. Inhibitors were dissolved in acetonitrile and CDNB was prepared in EtOH. Final incubation solvent content was

1% each respectively. GSH was dissolved in ddH2O. The reference cuvette held CDNB and

GSH in buffer.All activity assays were done in triplicate, with a no-inhibitor control. Enzyme activity vs.inhibitor concentration was plotted on SigmaPlot™. IC50 values were determined using SigmaPlot™ four-parameter logistic curve regression analysis.

vii. Characterization of inhibition mechanism(kinetics)

To determine the mechanism of dinitronaphthalene GST inhibition, double-reciprocal plots were constructed for selected inhibitor-enzyme combinations. GST activity assays were conducted with varying inhibitor and CDNB concentrations (as indicated); assays were done usingconditions previously described. Double reciprocal plots were made with GSTA1-1and inhibitors CDNN, DNNOH,DNNSG; andwith GSTM2-2 and inhibitors DNN, DNNOMe,

DNNOEt.

Dialysis experiments were conducted to determine whether GST inhibition was reversible. Incubations of GST A1-1 (25 μg), CDNN (25 μM) and GSH (1 mM) in 100 mM potassium phosphate buffer (1 mL), pH 6.5, were dialyzed against 1L of 100 mM potassium phosphate buffer for 24 h at 4ºC. Control incubations lacked CDNN.Activity with 0.5 mM

CDNB, 1 mM GSH was measured before and after dialysis.

viii. Characterization of Meisenheimer complexes

Characterization of inhibitor-GSHMeisenheimer complex formation in the GST active site was done following previous methods(Graminski et al., 1989b). Prior to experimentation, proteins were dialyzed overnight in 100 mM potassium phosphate buffer to remove β-

35 mercaptoethanol, a thiol-containing compound which may interfere with complex formation.Difference absorption spectra were measured by experimentation with Yankeelov tandem cuvettes. In one chamber, 50 μM GSTM2-2 and 1 mM GSH wereadded in 100 mM potassium phosphate buffer, pH 6.5. The second chamber contained 12.5 μM inhibitor dissolved in ACN 2% (v/v) (all inhibitors except DNNOH were dissolved in MeOH) in 100 mM potassium phosphate buffer, pH 6.5. The reference Yankeelov cuvette contained all components mentioned previously but with buffer added instead of GST and solvent added instead of inhibitor.

Absorption spectra(300 - 700 nm) were taken before and after mixing the chamber contents of both sample and reference cuvettes. Difference spectra were obtained by subtracting the pre- mix spectra from post-mix spectra (SigmaPlot™).

a. Reversibility of Meisenheimer complex formation

Additional dialysis experiments were conducted to monitor the reversibility of

Meisenheimer complex formation in the GSTM2-2 active site. Pre-incubations ofGSTM2-2 were dialyzed overnight in 100 mM potassium phosphate buffer to remove β-mercaptoethanol.

Absorbance spectrawere measured (300 – 700 nm) for incubations containing50 μM GSTM2-2,

1 mM GSH and 25 μM of each inhibitor:DNN, CDNN, DNNOMe and DNNOEt in 100 mM potassium phosphate buffer, pH 6.5. Incubations were dialyzed overnight using dialysis buffer(without β-mercaptoethanol) and a secondabsorption spectrum measured.

ix. Mutagenicity and Genotoxicity: Salmonella Ames Assay

The mutagenic effect and mechanism of genotoxicity of nitroarenes are well documented

(see introduction, section iii). A Salmonella mutagenicity assay (Ames test) was conducted to measure the potential mutagenic effect of dinitronaphthalene GST inhibitors: DNN, CDNN,

DNNOH, DNNOMe, and DNNOEt. DNNSG was not tested due to poor solubility.

36

b. Salmonella YG1024

The Salmonella mutagenicity assay was conducted with Salmonella strain

YG1024(Watanabe et al., 1989). YG1024 expresses high levels of Salmonella arylamine N- acetyltransferase/ N-hydroxyarylamine O-acetyltransferase, making it highly sensitive to the mutagenicity of nitroaromatic compounds(Watanabe et al., 1989;Josephy et al., 1997;Josephy et al., 2002). Freezer stocks of Salmonella YG1024 were prepared by inoculating Oxoid nutrient broth No. 2 containing ampicillin (25 μg/mL) and tetracycline (6.25 μg/mL). Cells were grown to an OD600 of 1.0. Aliquots of cells (1 mL) were mixed with 90 μL DMSO and frozen in liquid nitrogen for future use.

c. Salmonella mutagenicity assay protocol

Salmonella mutagenicity assays were carried out using standard methods(Maron and

Ames, 1983). Salmonellastrain YG1024 was grown overnight to an OD600= 0.8 - 1.0 under conditions previously described. Stock mutagens were prepared in ACN, with the exception of

DNNOH prepared in MeOH. Culture (100 μL)was incubated with sodium phosphate buffer (500

μL), pH 7.4 and mutagen solution (10 μL) at 37 °C for 30 min.After 30 min incubation,

2 mL of top agar was added to the cell-mutagen incubationsand thenplated. Colonies were counted after 48 h incubation at 37 °C. Tests were done in triplicate, twice, on two different days for a total sample size of six replicates.

d. NICE automated colony counting

Pictures of revertant plates were taken with a digital camera.NIST Integrated Colony

Enumerator (NICE) software was used to automate counting of revertant colonies(Clarke et al.,

2010). Colonies were counted in the inscribed square region ofthe circular 90 mm-diameter

37 plates (software limitation). Colony counts in the square region were corrected for the circular plate area by multiplication by(π/2).

D. Results and Discussion

i. Synthesis of CDNN-GSH adduct

CDNN is a yellow solid. Synthesis of the CDNN-GSH conjugate yielded S-(2,4- dinitronaphthalene)glutathione (DNNSG). Purified DNNSG is a red-orange solid. HPLC-

UVanalysis and ESI-HPLC-MS confirmed absence of starting material and product purity. Poor solubility of DNNSG made it unsuitable for NMR analysis.HPLC-UV analysis was conducted at the absorption peak maximum of CDNN, 390 nm, to observe for residual CDNN. Using the gradient specified in Figure 8, CDNN elutes at t = 6 min (data not shown).Analysis yielded a single peak at t = 1 min (Figure 10). No residual CDNN was observed in the purified DNNSG sample.The M+H+ and M+Na+ ion peaks of DNNSG were observed by ESI-HPLC-MS

(Figure10).UV vis. spectrum of DNNSG showed a peak absorption maximumat 370 nm and a shoulder at518 nm. The molar yield of DNNSG synthesis was 21%.

38

%

Figure 10: (Above) ESI-LC-MS of DNNSG; (Below) HPLC-UV vis analysis of DNNSG Green dashes: % ACN; red dashes: %, 1% Acetic Acid in H2O, pH 2.9. ii. Protein purification

Six human cytosolic GSTs were purified from recombinant protein-expressing E. coli, GSTs

A1, A2, A4, M1, M2 and P1(Mukanganyama et al., 2002).SDS-PAGE (sodium dodecyl sulfate polyacrylamide gel electrophoresis) of purified GSTs was done to confirm protein purity.

Protein bands were seen within the anticipated MW range for monomeric GSTs, 23.4 kDa – 25.7 kDa (Figure 11).These specific classes of cytosolic GSTs were selected for inhibitor activity

39 analysis because of their dominant roles in xenobiotic conjugation(Mannervik et al., 1985;Hayes et al., 2005). Within classes Alpha, Mu, and Pi, specific enzymes were selected for their high activity towards CDNB. (For example, GSTA3-3was not assayed. It has low specific activity towards CDNB and functions rather as a highly efficient steroid isomerase(Johansson and

Mannervik, 2002)). The specific activities of purified GSTs towards 1 mM CDNB, 1 mM GSH are listed in Table3.

250 150 100 75

50

37

25 20 15 10

Figure 11: SDS-PAGE of purified GSTs (1 μg protein/lane), stained with Brilliant Blue R-250;arrow indicates GSTP1-1, which consistently showed darker staining under the conditions described (ii. Protein purification). iii. Identification of GST inhibitors

CDNN was initially assayed as a potential substrate for GST conjugation. CDNN possesses an analogous structure to the universal GST substrate CDNB, including aryl ring, chloride (leaving group) and activating o- and p- nitro groups. These structural commonalities

40

with CDNB suggested CDNN would react with GSH (in the GST active site) following a SNAr mechanism. This was not the case for any of the GSTs tested, A1, A2, A4, M1, M2,or P1.

CDNN activity was tested over a pH range 6.5 – 7.8. Formation of the DNNSG conjugate could not be detected in any GST incubation (data not shown). Specifically, no absorption peak maximumof DNNSG was detected at 518 nm, and enzyme incubations with CDNN did not generate the anticipated red-orange colour of DNNSG. High concentrations of CDNN (> 500

µM) precipitated fromcontrol incubations containing phosphate buffer with GSH,but without

GST enzyme. In the presence of GSTA1-1, high concentrations of CDNN dissolved better. This observation suggestedthat an interaction between CDNN and GSTA1-1 was occurring. GSTA1-

1 activity towards CDNB was tested with CDNN present. A significant decrease in GSTA1-1 activity towards CDNB was seen with CDNN present, compared to incubations lacking CDNN.

Thus, CDNN was identified as an inhibitor of GSTA1-1. Following this discoveryother compounds were tested as potential GST inhibitors. To determine the structural features of

CDNN required for GST inhibition,compounds were tested that possess individual functional groups of CDNN. Potential inhibitors tested were 1-chloronaphthalene (CNN), 1-naphthol

(NOH) and 1,3-dinitronaphthalene (1,3-DNN). Only 1,3-DNN proved to be a GST inhibitor.

Thus, subsequent inhibitors examined in this studywere limited to compounds with the 2,4- dinitronaphthalene motif.

DNNOH, DNNOMe and DNNOEt were expected to inhibit GST activity. This was anticipated due to the poor leaving group in the C1 position, which would prevent SNAr completion in the GST active site. The synthesized CDNN-GSH conjugate, DNNSG, was also expected to be a GST inhibitor,because it is a GST product analogue. The following results

41 summarize the enzyme specificities, inhibition mechanisms and potencies of inhibition of

CDNN, DNNOH, 1,3-DNN, DNNOMe, DNNOEt, and DNNSGtowards GST enzymes.

iv. Potency of GST inhibition by dinitronaphthalene derivatives

A preliminary test to identify potent GST-inhibitor combinations was done. IC50 curves were constructed for GST-inhibitor combinations thatshowed appreciable inhibition at 25 μM inhibitor present. A typical IC50 curve and a summary of IC50 values are shown below (Table 3;

Figure 12).

Table 3: IC50 values (μM) of dinitronaphthalene inhibitors towards GSTs

GST A1-1 A2-2 A4-4 M1-1 M2-2 P1-1

Spec. Act1 35.4 6.3 10.4 156 495 93.8

CDNN 4.9 * * 35.7 0.72 91.2

DNNOH 0.92 76.2 * 4.1 97.6 89.7

DNNSG 9.8 * * 7.6 3.4 *

DNNOMe * * * 2.9 0.12 *

DNNOEt * * * 2.1 0.18 *

DNN * * * 1.3 1.3 *

*Indicates no inhibitory effect with 25 μM inhibitor 1Specific activity (μmol/min/mg)measured with 1 mM CDNB, 1 mM GSH; no inhibitor present (Materials & Methods)

For numerical data see section H; i. numerical data, forother IC50 curves see section; Supplementary graphs

42

Figure 12: IC50 curve of CDNN inhibition of GSTM2-2 activity towards 0.5 mM CDNB, 1 mM GSH; tests were done in triplicate, data points indicate mean ± standard error. (For additional IC50 curves, see supplementary graphs, section a.) GSTM1-1 and GSTM2-2 have high specific activities towards CDNB, the benzene congener of CDNN (Table 3). CDNN was also a potent inhibitor ofGSTM1-1 and GSTM2-2.

However, no correlation was observed between CDNB specific activity and CDNN inhibition potency amongst GST enzymes. It was anticipated that there could be a correlation between

CDNN inhibition potency and CDNB activity because they are structural congeners, but this was not the case. Thus, CDNB activity could not be used to predict CDNN inhibition preference/ potency towards different GSTs.The most potent inhibitory effects were observed with GSTM2-

2 and the O-alkyl compounds 1-methoxy- and 1-ethoxy-2,4-dinitronaphthalene.

Dinitronaphthalene derivatives were potentinhibitors towards Mu class GSTs (IC50< 5

μM). CDNN and DNNOH were potent inhibitors towards GSTA1-1. GSTMu class enzymes possess similar three-dimensional topology to GSTP1-1(Wilce and Parker, 1994), but there are significant differences in inhibition potency of dinitronaphthalene derivatives towards these two classes of GSTs. A distinguishing feature of the GST Mu family is the characteristic ‘mu-loop’, located between β-strand 2 and α-helix 2 (R 33-42)(Ji et al., 1992) (Figure 13). The mu-loop

43 extends the active site (H-site) of GST Mu enzymes. The greater size of the GST Mu active site contributes to the broad substrate specificity of GSTMu enzymes and the ability to conjugate large compounds(Wilce and Parker, 1994). For example,GSTM1-1 has high activity forlarge aryl compounds such as dibenz[a,h]anthracene-diolepoxide(Hayes et al., 2005). GSTM2-2 accepts a wide range of substrates, including oxidized catecholamines, aliphatic prostaglandins and the nitroareneCDNB(Hayes et al., 2005).

Figure 13: GSTM2-2 (ligand not shown); ‘mu-loop’ is shown in orange for each monomer; alpha helices are shown in red and beta sheets in blue (PDB entry 2AB6).

44

a. Inhibition of GSTMu class enzymes by dinitronaphthalene derivatives

The results presented in Table 3 suggest that the dinitronaphthalene structure determines

GST class potency of inhibition. Differences of inhibitor potency between GSTM1-1 and

GSTM2-2 suggest that substitution at the C1 group confers inhibitor specificity within the

GSTMu enzyme class. Mu class enzymes,GSTM2-2and GSTM1-1, share 84% amino acid sequence identity. Between GSTM1-1 and GSTM2-2 34 residues differ, only four of those residues line the substrate binding site(Hansson et al., 1999). Despite this small difference in active site residues, activity towards many substrates differs by a factor of 100 (including 1,2- dichloro-4-nitrobenzene (DCNB) and aminochrome(Hansson et al., 1999;Comstock et al., 1994).

A similar difference between GSTM1-1 and GSTM2-2 dinitronaphthalene inhibitor potency was observed. Active site residues which contribute to the substrate specificity of GSTM2-2 include a network of amino acids that form hydrogen bonds (Arg107-Asp161-Arg165-Glu164) at the base of H-site and a tripeptide sequence (Ala111-Lys112-Leu113) in the H-site interior(Hansson et al., 1999;Comstock et al., 1994). The former network form stabilizing H-bonds with the negatively charged nitro groups on DCNB and could possibly do the same with the dinitronaphthalene compounds. X-ray crystallography could determine inhibitor orientation in the GSTM2-2 active site and show whether these proposed stabilizing interactions occur with dinitronaphthalene inhibitor binding.

The IC50values of dinitronaphthalene derivatives with C1 R groups (-OMe, -OEt, -Cl) differ between GSTM1-1 and GSTM2-2 enzymes, by factors of 20 to 50 fold. Comparably,

DNN has no functional group at C1 and has equal potency for GSTM1-1 and GSTM2-2 enzymes. The presence of a C1 ether functional group on DNNOMe and DNNOEt increased potency of inhibition of DNN for GSTM2-2 but decreased potency of DNN for GSTM1-

45

1.Inhibitors with neutral R-groupswere selectively potent for GSTM2-2. This suggeststhat a functional group at the C1 position of dinitronaphthalene inhibitorsis a structural feature that affectsGSTMu inhibition potency.

Figure 14: GST inhibitors DNNSG (left) and DNBSG (right) DNNSG is a GSH conjugate that was found to inhibit GST Mu enzymes. The CDNB-

GSH conjugate product,S-(2,4-dinitrobenzene)glutathione (DNBSG), is a benzene congener of

DNNSG and inhibits GSTM1-1 and GSTM2-2(Comstock et al., 1994)(Figure 14).DNNSG is a considerably more potent inhibitor than DNBSG towards GST Mu class enzymes. IC50 values of

DNBSG are 140 μM and 250 μM for GSTM1-1 and GSTM2-2, respectively(Comstock et al.,

1994).The high IC50 of DNBSGand low IC50 of DNNSG would suggest that the dinitronaphthalene structure has higher affinity and/or is more stabilized in the GSTM1-1 and

GSM2-2 active site. Unfortunately, the inhibition potency of DNBSG towards other human

GSTs classes is not known. It would be interesting to test DNBSG inhibition of other GST classes. It may also be worth examining the inhibitory effects of the benzyl congeners of the naphthalene inhibitors identified in the study. Those would include ether-substituted

46 dinitrobenzenes e.g. 1-methoxy-2,4-dinitrobenzene. For these compounds, the intermediary

Meisenheimer complex would be less stable due to smaller ring size (discussed in results section vi), but a trade-off may be that the smaller ringhas better accessibilityto other GST active sites.

b. Inhibition of GST Alpha class enzymes by dinitronaphthalene derivatives

Weak or no inhibitory effect of dinitronaphthalene inhibitors was observed towards

GSTA2-2 and GSTA4-4, whereas DNNOH inhibition was enzyme-specific and potent towards

GSTA1-1.The low inhibition potency towards GSTA2-2 and GSTA4-4 can be related to their active site structure, specialized functions and substrate specificity. GSTA4-4 has high specific activity towards aliphatic oxidized products of lipid-peroxidation, for example 4- hydroxynonenol(Hubatsch et al., 1998). A structural characteristic specific to alpha class GSTs is the C-terminal α-9 helix that bridges over the GST alpha active site; it is thought to contribute to substratebinding andspecificity (Figure 15). The α-9 helix (R 212-222) packs into the H-site, resulting in a smaller, more hydrophobic binding site relative to other GST classes (Wilce and

Parker, 1994). The α-9 helix of GSTA1-1 is intrinsically disordered in the absence of ligand and assumes an ordered structure upon ligand binding. Conversely GSTA4-4 has a rigid C-terminal helix, that does not adjust upon ligand binding(Hou et al., 2007). A promiscuity index (J), based on catalytic efficiencies of non-substrate specific enzymes has been suggested. Catalytic activity against 30 substrates was compared between GST enzymes, GSTA1-1 and GSTA4-4. Based on higher specific activities for a greater variety of substratesGSTA1-1 was found to have a promiscuity index 3-fold greater than GSTA4-4(Nath and Atkins, 2008).

47

Figure 15: GSTA1-1 (ligand not shown), GST alpha α9 C-terminal helix is shown in purple; other alpha helices are shown in red and beta sheets in blue (PDB entry 3KTL). GSTA2-2 is distinguished from other alpha GSTs by having high (glutathione dependent) peroxidase activity. GSTA2-2 has high specific activities towards lipid hydroperoxide metabolites of linoleic acid and arachidonic acid, while having low specific activity towards aryl substrates such as CDNB(Zhao et al., 1999). Both GSTA2-2 and GSTA4-4 have preference for substrateswith structures that are very different from the dinitronaphthalene inhibitors tested in this study. The finding that dinitronaphthalene derivatives were non-inhibitory towards GSTA2-

2 and GSTA4-4 reflects these enzyme’s substrate preferences and is supportive of active site binding as the mechanism of inhibition for dinitronaphthalenes.

As a comparison, GSTA1-1 has high specific activity for large aryl structures, including the sixaromatic ring, dibenzopyrene[a,l]diolexpoxide and similarly structured polycyclicaromatic

48 compounds(Dreij et al., 2010).The inhibition potency of DNNOH towards GSTA1-1 may be related to the ionized state of DNNOH (pKa 2.12) at neutral pH.GSTA1-1 hashigh activity for epoxides(Dreij et al., 2002), which generate an oxyanion upon ring opening. It is also interesting to note that the unsubstituted dinitronaphthalene and 1-naphthol had no effect on

GSTA1-1 activity. Thus, the hydroxyl and nitrogroups on the naphthalene ring are required for potent GSTA1-1 inhibition.

c. Inhibition of GST Pi class enzyme by dinitronaphthalene derivatives

GSTP1-1 was poorly inhibited by the dinitronaphthalene derivatives investigated in this study. Only weak inhibition was observed with CDNN and DNNOH. As previously mentioned,GSTM1-1 and GSTP1-1 crystal structures show a high degree of similarity(Dirr et al.,

1999). GSTP1-1 also shows high specific activity for CDNB. These observations would suggest that the dinitronaphthalene inhibitors could have high affinity for GSTP1-1.Crystal structures

GSTP1-1 and GSTM2-2 bound to the GST inhibitor NBDHEX (discussed in the following section) show that a single residue,Ile104, in the active site of GSTP1-1, obstructs inhibitor binding in the H-site.In GSTM2-2, the equivalent residue is an alanine. As a result, the inhibitor binds with much greater affinity to GSTM2-2(Federici et al., 2009). This residue is a prime candidate for site-directedmutagenesis studies. There is a strong possibility that Ile104 affects dinitronaphthalene inhibitor binding, considering that NBDHEX possess a similar structure to the inhibitors examined. A suggested substitution would be Ile104 to an alanine or valine(Wu and Dong, 2012). The Val104 variant of GSTP1-1, denoted as allele GSTP1*b, is an abundant

SNP in the human population,relevant to differences in drug metabolism of chemotherapeutics(Lo and Ali-Osman, 2007).

49

d. Other GST enzyme specific inhibitors

The majority of GST inhibition efforts have focusedon GSTP1-1, due to its documented over-expression in tumor cells and role in regulating cell proliferation pathways (see introduction, section ii). The inhibitors in this study showed low potency towards GSTP1-1, but much research is available on inhibitors designed for this enzyme, as discussed below.

NBDHEX is a potent GSTM2-2 specific inhibitor (Figure16).NBDHEX has been proven effective as a sensitizing agent to chemotherapeutics (e.g. temozolomide) in melanoma cell lines

Me501 and A375(Tentori et al., 2011). NBDHEX forms a Meisenheimer complex in the GST active site of GSTM2-2 andGSTP1-1*Val104 but not GSTP1-1*Ile104(Federici et al.,

2009).NBDHEX displays low nanomolar potency towards GSTM2-2 (IC50< 0.010 μM) and

GSTP1-1(IC50 0.8 μM)(Giorgio Ricci et al., 2005).

Figure16: GSTM2-2 specific inhibitor 6-(7-nitro-2,1,3-benzoxadiazol-4-ylthio)hexanol (NBDHEX) Crystallographic structures of NBDHEX in GST active sites revealed that NBDHEX is partially stabilized by the same interactions in GSTM2-2 and GSTP1-1, that is π-π stacking interactions of the benzene ring with residues Tyr108 (GSTP1-1) and Tyr115 (GSTM2-2).

Selective potency arises because of unobstructed binding in the GSTM2-2 active site. The

50 negative nitro group oxygens of NBDHEX are able to interact with positively charged residues deep in the GSTM2-2 active site. Deep binding of NBDHEX in the active site of GSTP1-1 is inhibited due to steric hindrance of residue Ile104(Wu and Dong, 2012). Replacing Ile104 in

GST P1-1 with alanine or valine decreases the IC50 of NBDHEX to GSTP1-1 by 4-fold.

NBDHEX shares similar structural features with DNNOMe andDNNOEt; both possess a p-nitro group, a double ring structure (in NBDHEX the second ring is an oxadiazole heterocycle) and a

C1 aliphatic chain. The C1 benzyl ring of NBDHEX forms stacking interactions with Y115 in

GSTM2-2 and the p-nitrogroup of NBDHEX interacts with positive amino acids R107 and

R165(Federici et al., 2009). These interactions may also occur with the dinitronaphthalene inhibitors discussed in this study, due to their similar structure. Crystal structures of DNN derivatives in M2-2 active sites could test this suggestion of binding interactions.

Figure17: GSTMu class specific inhibitor γ-glutamyl-(S-benzylcysteiny1)-β-alanine (left) and GSTP1-1 specific inhibitor γ-glutamyl-(S-benzylcysteinyl)-(R)-(-)-phenylglycine (TLK199, ezatiostat•HCl, Telintra®) (right) Selective GST inhibition by product mimics also arises by modification of the GSH tri- peptide. The GSH-phenylmethyl conjugate (Figure17, left)has selectivity for Mu class enzymes, with highest potencies of 22 μM and 26 μM towards GSTM1-1 and GSTM2-2, respectively(Lyttle et al., 1994). The GSTP1-1 IC50 is higher (710 μM). By modifying the GSH

51

moiety with the addition of a phenyl group to the α-C of glycine (Figure 17, right)the IC50 towards GSTP1-1 dropped to 0.42 μM and increased 2- and 7-fold for GSTM1-1 and GSTM2-2 respectively(Lyttle et al., 1994a). Crystal structure examination revealed that the additional phenyl group on the glycine moiety clashes sterically with the α-9 loop of Alpha class GSTs and with the mu-loop (R33-42) of Mu class GSTs. Within the GSTP1-1 GSH binding site,no disruption of GSH binding occurs with the addition of a phenyl group to the α-C of glycine(Oakley et al., 1997).

This engineered GSTP1-1specific inhibitor (TLK199, generic name; ezatiostat·HCl, brand name; Telintra©) has been approved by the FDA for phase 3 clinical trials. It has been granted “orphan drug” designationby the FDA for the treatment for myelodysplastic syndrome(Raza et al., 2009).Orphan drugs are pharmaceuticals specifically engineered to treat rare diseases that would otherwise not be investigated due to lack of financial incentive. The

DNNSG conjugate inhibitor investigated in this study was selective for GSTM2-2 inhibition and had no affinity for GSTP1-1. Modifications to the GSH moiety would be a reasonable consideration for future studies, with aims to modify DNNSG for GSTP1-1 inhibition.

Figure 18: GSTM2-2 inhibitors (A) genistein; (B) kaempferol; (C) quercetin

52

Natural product GST inhibitors selective for Mu class enzymes, GSTM1-1 and GSTM2-

2, include the phytoestrogen-isoflavone genistein and the flavonoids kaempferol and quercetin(Hayeshi et al., 2007) (Figure 18). The IC50valuesof genistein and kaempferol towards

GSTM1-1 and GSTM2-2 are in the single-digit micromolar range. Quercetin potency towards

GSTM1-1 and GSTM2-2 are in the mid-nanomolar range. These three compounds share a common flavone (or isoflavone) backbone. The inhibition mechanism of flavonoid inhibitors towards GSTMu enzymes were not characterized,so affinity of these compounds maybe for the

GST active site or the GST dimer interface.

v. Inhibition mechanism of DNN inhibitors

The mechanism of inhibition of DNN derivatives towards GSTs was determined by constructing double reciprocal plots of potent inhibitor-GST combinations. DNN, DNNSG,

CDNN, DNNOH, DNNOMe and DNNOEt showed kinetics characteristic of competitive inhibition (Figure 19).That is, inhibited and control activities had the same Vmax; a common intersection on the y-axis. The DNNSG is a product mimic and was anticipated to be a competitive inhibitor. Spectral evidence of inhibitor GSH intermediate formation in the GST active site (discussed in the section below) supports the characterization of dinitronaphthalene derivatives as competitive inhibitors.

53

0.03 1200 nM

600 nM

0.02 300 nM

0 nM

0.01

(nmoles/min)

o

1/V 0.00

-0.01 -5 0 5 10 15 20 1/[CDNB] mM

Figure 19: Double reciprocal plot of DNNOMe inhibition of GSTM2-2 activity; for additionalplots, see supplementary graphs section b. vi. Meisenheimer complex formation in the GSTM2 active site

GSTM2-2 was chosen for analysis of inhibitor complex formation in the GST active site due to the high potency of inhibition of dinitronaphthalene derivatives towards GSTM2-2. A characteristic absorption spectrum of the SNAr reaction intermediate‘Meisenheimer complex’was observed in the GSTM2-2 active site, using enzyme concentrations of 650 μg/mL (see introduction section (i;f), (Figure20).The absorption peak maxima of unbound inhibitor and GST bound Meisenheimercomplexes are summarized in Table4;consistent with their spectral profiles, solutions of GST, GSH and dinitronaphthalene inhibitor were red in colour. The characteristic red colour of the Meisenheimer complex had not previously been seen at low, catalytic concentrations of enzyme and is not seen in the absence of enzyme (in buffer). The minimal detectable limit of the Meisenheimer complex by UV vis. spectrometry was identified at enzyme concentrations of 25 μg/mL GSTM2-2. The Meisenheimer complexes for different

54 dinitronaphthalene derivates possess similar absorption maximums in the GST active site. This was an anticipated observation because they share a common structural chromophore: a 2,4- dinitronaphthalene conjugated system.

Table4: Absorption peak maxima of dinitronaphthalene inhibitors in buffer and in complex with GSTM2-2-GSH-inhibitor

λmax (nm) inhibitor complex CDNN 372 478 DNNOH 393, 436 - DNNSG 370, 518 - DNNOMe 295, 360 480 DNNOEt 295, 360 474 DNN 288, 371 495

Inhibitors CDNN, DNNOMe, DNNOEt, and DNN produced Meisenheimer complexes in the GSTM2-2 active site. This observation supports kinetic evidence of competitive inhibition because the Meisenheimer complex forms in the GST active site. DNNSG did not, as anticipated, form a Meisenheimer complex in the GST active site. The presence of glutathione at the C1 position of DNNSG would occupy the G-site of the enzyme and impede binding of free

GSH. Meisenheimer complex formation in GST has not been reported for GSH-conjugate inhibitors. DNNOH did not show a shift in absorption maxima in the GST active site, suggesting no Meisenheimer complex had formed. This may be due to the DNNOH hydroxyl functional group. DNNOH hydroxyl has a pKa of 2.12. Thus, it is ionized at physiological pH. The oxyanion of ionized DNNOH may repel nucleophilic attack by the glutathione

55 thiolate,preventing Meisenheimer complex formation. Crystallization of DNNOH in the active site of GST would be useful for determining the orientation of DNNOH binding.

0.4

0.3

0.2

absorption 0.1

0.0

-0.1 300 400 500 600 wavelength (nm)

CDNN DNN DNNOEt DNNOMe

Figure 20:Difference absorption spectra of dinitronaphthalene inhibitors in the active site of GSTM2-2.

a. Meisenheimer complex stability in the GST active site

Data from this studyprovides novel evidence that the structural features that stabilize

Meisenheimer complex formation in solution have the same stabilizing effect on Meisenheimer complex formation in the GST active site. For example (in MeOH), the stability of the methoxy anionMeisenheimer complex with 2,4-dinitronaphthalene is 104 times greater than the methoxy and 2,4-dinitrobenzene complex. The free energy of Meisenheimer complex formation with 2,4-

56 dinitroarenes decreases by about 30 kJ/mol between naphthalene and benzene structures(Terrier,

2013).

GSH reacts with the universal GST substrate CDNB via a Meisenheimer complex

- intermediate to form the final products,DNBSG and Cl . Along the SNAr reaction coordinate, there is an energy barrierbetween the Meisenheimer complex and breaking of the chlorine- carbon bond. The negative ΔG between the Meisenheimer intermediate and products(GSDNB and chloride) drives the reaction to completion(Ji et al., 1993). CDNN does not react with GSH in the GST active site, whereas CDNB reacts rapidly with GSH in the GST active site. This would suggest that the second fused ring of CDNN increased the stability of the CDNN-GSH-

GST Meisenheimer complex, to such an extent that the reaction does not go to completion, regardless of Cl- being a stable leaving group.Thus, the additional ring on the naphthalene inhibitors provides stability of Meisenheimer complex in solution and GST-active site. The other dinitronaphthalene derivative inhibitors examined,DNNOMe, DNNOEt, DNNOH and

DNN,do not possess good leaving groups in the C1 position, so non-reactivity in the GSH active site was expected. Conversely, CDNN was expected to react with GSH in the GST active site, but did not.

Meisenheimer complex formation with CDNN, DNNOMe, DNNOEt and DNN in the

GSTM2-2 active site was reversible by dialysis. Overnight dialysis of 50 μM GSTM2, 1 mM

GSH and 25 μM of inhibitor in 100 mM potassium phosphate buffer resulted in a loss of

Meisenheimer complex absorption (data not shown). Several attempts were made to measure the dissociation constant (Kd) of dinitronaphthalene inhibitors towards GSTM2-2. However, inhibitor binding was so potent that under conditionswhere [E]<<[I], the absorbance spectra were too weak for accurate measurement (recall that the detection limit of Meisenheimercomplex is

57 about 25 μg/mL enzyme). Dissociation of theMeisenheimer complex was observed over a 24 h time period at 37oC (dissociation did not occur at room temperature). Meisenheimer complex formation was reversible by addition of more GSH, indicating Meisenheimer complex loss was due to GSH auto-oxidation. An interestingfuture experiment to carry out would be to measure

Kd by isothermal titration calorimetry (ITC).

b. Meisenheimer complex formation by other GST inhibitors

Meisenheimer complex formation by GST inhibitors has been documented in a number of GST active sites. As previously mentioned, GST inhibitor, NBDHEX, forms a Meisenheimer complex with GSH in the active sites of GSTM2-2 and GSTP1-1*Val104(Bico et al., 1994) (see introduction, section ii). Another well-studied Meisenheimercomplex-forming GST inhibitor is

1,3,5-trinitrobenzene, a CDNB substrate analogue(Figure 21). Crystal structures of TNB complexes; have been solved for rat rGST3-3 (rGSTM1-1), rat rGST4-4 (rGSTM2-2), moth GST

(Galleria mellonella) and human GSTP1-1(Clark and Sinclair, 1988;Graminski et al.,

1989b;Prade et al., 1997). The inhibition potency of the TNB-Meisenheimer complex has been characterized towards moth GST; IC50= 10 μM(Clark and Sinclair, 1988).

Figure 21: CDNB substrate analogue 1,3,5-trinitrobenzene

58

With several enzymes for comparison, it is interesting to note that the TNB-

Meisenheimer absorption spectra differ, depending on the GST active sitein which it forms.

Difference absorption spectra of the TNB-Meisenheimer complex in the active site of rGSTM1-1 showed peak maxima at 457 and a shoulder at 520 nm(Graminski et al., 1989b). In complex with human GSTP1-1, a more pronounced shoulder peak maximum occurs at 570 nm. The authors suggest that the difference in absorption spectra occurs due to different microenvironments in the GST active site of rGTM1-1 vs. hGSTP1-1(Bico et al., 1994).

Absorption spectra of dinitronaphthalene Meisenheimer complexes have been produced for

GSTM2-2 in this study, but can be done for GSTM1-1 in future investigations. Due to active site similarities, Meisenheimer complex absorption spectra in GSTM1-1 would likely differ very little from GSTM2-2, but could be worth comparing. If other, non-human GSTs are found to be inhibited by the dinitronaphthalene compounds in this study, there would be grounds for more comparisons to be made.

A direct correlation between complex formation constant (Kf) of TNB in GSTP1-1, rGSTM1-1 and rGSTM4-4 and catalytic efficiency towards CDNB has been observed(Bico et al., 1994). The authors concluded that the ability of GSTs to stabilize the TNB Meisenheimer complex in the GST active site directly correlates to catalytic efficiency towards nucleophilic aromatic substitution of substrate CDNB. The TNB Meisenheimer complex is a dead-end intermediate in the SNAr reaction pathway and resembles CDNB. Results from my study,comparing dinitronaphthalene inhibition IC50 values and CDNB, activity did not show such a correlation. The discrepancy between CDNB activity and inhibitor potency maybe due to size exclusion of dinitronaphthalenes in the GST active site (see results section iv.,IC50 results).

59

A better comparison for this study would be to compare dinitronaphthalene Kd values and CDNB activity.

C. Possible intramolecular Meisenheimer complex formation of DNNSG

As previously mentioned the absorption spectrum of DNNSG shows a peak maximum at and a shoulder at 518 nm. This spectrum is similar to the spectra observed for the Meisenheimer complex of TNB and GSH in rGSTM1-1(Graminski et al., 1989b). This suggests the possibility of the formation of an intermolecular Meisenheimer complex with DNNSG. Such a complex could form with CDNN, the cysteine thiol of GSH and an additional nucleophile on the GSH molecule. Additional GSH nucleophiles include the carboxyl terminal of glycine and the carboxy and amino terminal of the γ-bonded glutamate. Stable Meiseneheimer complexes have been observed with 2,4-dinitronaphthalene and ethane-1,2-diol, propane-1,3-diol and butane-1,4- diol in the presence of sodium hydroxide in water and solutions of DMSO and water. Complex stability and rate of complex formation was found to decrease with increasing ring size(Crampton and Willson, 1975). The authors suggest this is due to a loss of rotational freedom, which increases with longer chains and results in a greater loss in entropy upon formation of the Meisenheimer complex. Thus, it could be more likely that a GSH peptide amide may contribute to the Meisenheimer complex, to yield a 4 or 5 atom ring.

Meisenheimer complexes involving nitroarenes form in the presence of base and a stabilizing cation such as Na+ of K+. In the case of zwitterionic Meisenheimer complexes an intramolecular substituent with a positive charge may stabilize the complex, such as an amine cation. GSH and

CDNN would be capable of forming a zwitterionic Meisenheimer complex due to the charges of the glutamate amino acid termini of GSH (at physiological pH). Crystal structures have been

60 obtained for 1,2,3-trinitrobenzene Meisenheimer complex with 1-(2,3-dihydroxypropyl)pyridine-

1-ium(Borbulevych et al., 1999). X-ray analysis of crystallized DNNSG, could be done to determine if an intermolecular Meisenheimer complex does occur in solution. Presumably this interaction would be unable to occur in the GST active site due to restricted conformation of

GSH binding to the G-site.

vii. Salmonella Mutagenicity Assay

Potential mutagenicity of the dinitronaphthalene derivatives investigated in this study was assessed by Ames assays with Salmonella strain YG1024. Strain YG1024 was selected for the assay because it has been engineered to be more sensitive to nitroarene mutagenicity(Watanabe et al., 1989) (see introduction section iii). The results of the Salmonella Ames assay are reported in Figures 22. All dinitronaphthalene derivatives tested were mutagenic towards Salmonella

YG1024, and were toxic (bactericidal effects) at higher concentrations. Toxicity was observed as a drop in revertant frequency, thinning of the background bacteria lawn and abnormal growth morphology of revertants. All of these observations are characteristics of toxic effects in the

Ames test(McGregor et al., 1984).CDNN showed the most potenttoxic effect. Toxicity was observableat a concentration of 0.3 nmole/plate. This was an anticipated effect, attributed to the reactivity of CDNN nitro groups and additional reactivity of the nitro-activated aryl-halogen center that the other compounds lack.

61

3000

2500 DNNOEt DNNOMe CDNN 2000

1500

1000

revertantsplate per 500

0 0.001 0.01 0.1 1 10 100

nmoles per plate

2500 DNN 2000 DNNOH

1500

1000

revertants per plate per revertants 500

0 0.001 0.01 0.1 1 10 100 1000

nmoles per plate Figure 22: Mutagenicities of dinitronaphthalene derivatives in Salmonellastrain YG1024; two experiments, done in triplicate, were carried out on different days. Each data point represents the mean ± standard error of all six replicates.

In previous experiments, both DNNOH and DNN have tested positive as direct-acting mutagens towards TA98 and TA100(McCoy et al., 1981). TA98 is the parental strain of

YG1024. YG1024 differs from TA98; it contains a plasmid that drives constitutive over-

62 expression inSalmonella N-hydroxyarylamine O-acetyltransferase/ N-acetyltransferase. The mutagenic response of YG1024towards DNN and DNNOH, reported in this study, is 12-72 fold greater (revertants/mole) than the response observed in TA98(McCoy et al., 1981). A similar increase in mutagen response of YG1024 compared to TA98 has been observed for CDNB.

CDNB was found to be a direct-acting mutagen towards Salmonella TA98(Summer et al., 1980;

Catterall et al., 2002). The revertant yield of strain YG1024 towards CDNB was 1000-fold greater than with TA98, and toxic effects of CDNB were observed at much lower doses

(Catterall et al., 2002). The elevated response of strain YG1024 towards DNN and DNNOH observed in this study is consistent with the activating effect of N-hydroxyarylamine O- acetyltransferase in the Salmonella Ames assay.

Mutagenicity ofnitroaryl compounds is attributed to metabolic activation of the nitro substituent, and possibly also to the ability of nitroarenes to DNA intercalate(McCoy et al.,

1981). The positive results for mutagenicity observed in this study suggest that it would be beneficial for drug development to replace the nitro groups on the dinitronaphthalene inhibitors with other, inert,electron-withdrawing (EWG) functional groups. PotentialSNAr activating

EWGs include the triflyl (F3CSO2) and nitrile (ACN) groups. A review of the rate constants (K1) of Meisenheimer complex formation(in solvent) of nitroarenes, bearingdifferent EWGs, found thatthe F3CSO2 group better stabilizes intermediate σ-complexes and is more activating than

NO2, while NO2 is more stabilizing and activating than ACN (Terrier, 2013). Replacing NO2 with F3CSO2on the dinitronaphthalene derivatives would be expected to reduce the mutagenicity, and because F3CSO2 better stabilizes Meisenheimer complex formation, they may also increase inhibition potencyby producing a more stable inhibitor intermediate. Alternatively the second aromatic ring maybe altered to increase electron delocalization capacity. Abenzofurazan,or

63 similar N/O heterocycle, in place of the second aromatic ring on the naphthalene backbone, has similar electron-withdrawing effects. From the Ames assay data provided from this study, dinitronaphthalene compounds maybesuitable lead compounds for pharmaceutical applications; they would benefit from modifications to reduce potential toxic effects.

viii. Biological functions of GSTM2-2

The GST inhibitors identified in this study showed specific, high potencyinhibition towards

GSTM2-2. Thus, it is relevant to discuss the functionsof GSTM2-2 in the body. In GSTM1-/- individuals, GSTM2 protein levels are 2-fold greater, compared to GSTM1+/+ individuals(Bhattacharjee et al., 2013). This suggests that GSTM2-2 overexpression may serve as a compensation mechanism for absence of GSTM1-1. This also suggests GSTM2-2 is of particular importance to GSTM1-/- null individuals. An interesting observation is that the reverse is not true. Unlike its closest homologue, GSTM1, GSTM2 does not occur as a null genotype in the human population and has no documented polymorphic alleles (in amino acid sequence).

The conservationof cytosolic GSTs, despite catalytic redundancy towards many substrates,is partially attributed to theiracquired non-catalytic functions(da Fonseca et al., 2010). As with

GSTP1-1, recent research has identified important function of GSTM2-2 aside from drug metabolism, particularly in brain and in cardiac muscle tissues. These areas of function are consistent with the tissue distribution profile of GSTM2-2 in humans. GSTM2-2 expression pattern is unique amongst cytosolic GSTs,due to its absence in liver tissue. It is found in highest concentrations in the brain, testis, heart and skeletal muscle(Rowe et al., 1997;Abdellatif et al.,

2007).

In cardiac muscle (and, less so, skeletal muscle) GSTM2-2 has been shown to modulate the activity of ryanodine receptors in the sarcoplasmic reticulum (SR).GSTM2-2 interacts with

64 its C-terminal domain on the luminal side of the cardiac ryanodine receptor (RyR2), in a non- covalent manner(Hewawasam et al., 2010). The GSTM2 C-terminal domain has Kd= 37 μM towards RyR2, well above the concentration of GSTM2-2 detected in sheep cardiac tissue (93

μM)(Liu et al., 2009). GSTM2-2 allosterically binds to andinhibits Ca2+ release from the SR via interactions with RyR2(Abdellatif et al., 2007).Ca2+ release into cardiocyte cytoplasm is required for cardiac muscle contraction to occur. The function of GSTM2-2 inhibition of RyR2 is thought to be protective in nature. Regulation of RyR2 prevents overload of Ca2+ release and indirectly regulates muscle contraction.Cessation of Ca2+ release from the SR is required for cardiac muscle repolarisation to occur. Synthetic GSTM2-2 peptides are being investigated as a possible treatment for heart failure(Liu et al., 2012).

In the brain, GSTM2-2 has been identified as the primary prostaglandin E2 (PGE2) synthase(Beuckmann et al., 2000b). Prostaglandins are a class of eicosanoid, chemical messengers derived from arachadonic acid. PGE2 is synthesized by the GSH dependent isomerization of PGH2. In the liver this reaction is catalyzed by microsomal PGE2 synthase-1 and -2, members of a different GST family, the MAPEG family (membrane associated proteins involved in eicosanoid and glutathione metabolism). When researchers tried to identify the

PGE2 synthase in the human brain, they found that activity was localized to the cytosolic fraction

(not the microsomal). By sequencing the proteins with PGE2 synthase activity, researchers confirmed that the enzymes that catalyze the isomerization of PGH2 were GSTs M2-2 and M3-

3(Beuckmann et al., 2000b). Invitro activity analysis found that GSTM2-2 has a specific activity of282 nmol/min/mgtowards PGH2 isomerization. Specific activity of mPGE2 synthase-1 is 13x greater than GSTM2-2 (4 mmol/min per mg pro),but, nonetheless,the mRNA of mPGE2synthase-

1 and -2 is not found in the brain(Murakami et al., 2002). This finding suggeststhat GSTs M2-2

65

and M3-3 are primarily responsible for PGE2 synthesis in the brain.Aside from the identification of GSTM2-2 and GSTM3-3 as PGE2 synthases in the brain, no studies have further elaborated this finding.Microdialysis administration of PGE2into regions of the brain of the Rhesus

Macaque found PGE2 activity localized to the pre-optic area and tuberomammillary region of posterior hypothalamus. PGE2 was found to regulate wakefulness, heart rate, and body temperature in these areas(Onoe et al., 1992). How GSTM2-2 may contribute to the homeostasis of these bodily processes is unknown.

A second, poorly studied function of GSTM2-2 in the brain is the metabolism of oxidized catecholamines. In vitro, GSTM2-2 has high specific activities towards dopachrome, aminochrome and noradrenochrome. These compounds are oxidized ortho-quinones of the catecholamine-neurotransmitters, L-DOPA, dopamine and noradrenalin, respectively. Specific activities of other cytosolic GSTs for oxidized catecholamines are <1 μmol/min/mg protein(Segura-Aguilar et al., 1997). It is suggested that GSH-catecholamine conjugates are metabolized to cysteine conjugates and are either incorporated into neuromelanin or excreted in cerebral spinal fluid(Munoz et al., 2012).GSH conjugation reduces the oxidized cathecholamine quinone and the resultant conjugate is resistant to re-oxidation(Dagnino-Subiabre et al., 2000).

GSH conjugation prevents redox cycling of oxidized catecholamines and the generation of reactive oxygen species (ROS). ROS generation and catecholamine redox cycling are neurotoxic mechanisms associated with neurodegeneration(Sofia Baez et al., 1997). The high activity of

GSTM2-2 towards oxidized catecholamines suggests that GSTM2-2 plays a neural protective function in the brain, particularly in dopaminergic neurons. However, the role of GSTM2-2 in neurodegenerative diseases is not well characterized. The proposed functions of GSTM2-2 in the body, aside from drug metabolism, have plenty of room for further investigation.

66

E. Conclusions

Human cytosolic glutathione transferase (GST) enzymes contribute to the detoxication of cytotoxic and mutagenic electrophilic molecules. Detoxication occurs by catalyzing the conjugation of endogenously produced glutathione to co-substrates. Like other drug metabolizing enzymes,cytosolicGSTs have a remarkable capacity to accept a wide range of co- substrates, with different structures, in one active site. An initial goal of thisstudy was to identify specific GST enzyme substrates, to better determine the substrate specificity of these enzymes.

The first trial substrate was a naphthalene congener of the universal GST substrate 1-chloro-2,4- dinitrobenzene (CDNB), 1-chloro-2,4-dinitronaphthalene (CDNN).

Prior to experimentation, it was hypothesized that human GST enzymes M2-2, M1-1 and A1-

1 would have highest activity towards CDNN. This suggestion was based on high specificity constants of these GSTs towards CDNB and the ability of these enzymes to fit large aryl structures in their respective active sites. However, CDNN proved not be a GST substrate for six recombinant GSTs, A1-1, A2-2, A4-4, M1-1, M2-2 and P1-1. Instead, CDNN inhibits GST activity towards CDNB. Subsequent studies identified the dinitronaphthalene structure of

CDNN to be necessary for GST inhibition. Several additional, novel GST inhibitors were identified. The structures were selected by replacing the CDNN chloride group with functional groups with poor leaving capacity (e.g. methoxy).

Dinitronaphthalene derivatives werespecific and potent inhibitors towards GSTMu class enzymes M1-1 and M2-2; some potent inhibition was observed towards GSTA1-1. Inhibitor selectivity towards GSTs M1-1, M2-2 and A1-1 was consistent with a previous hypothesis that suggested the same enzymes would be active towards CDNN. The inhibition mechanism of dinitronaphthalene inhibitors was characterized as being competitive towards GSTM2-2, results

67 consistent with observations of a stable Meisenheimer complex in the GST active site. Stable

Meisenheimer complex inhibitors have previously been observed in the GST active site but not of a compound (CDNN) which has anticipated activity as a substrate.

GST inhibitors have been investigatedin other studies as chemotherapy adjuvants. The majority of GST inhibitor research has focused on GSTP1-1, because of its ability to metabolize chemotherapeutic drugs and its role as a modulator of cell proliferation/ apoptosis. The inhibitors in this study were selective for GSTM2-2, an enzyme with cardio-protective and potentially neuro-protective functions. Inhibiting this particular GST would likely not have the same beneficial effect as would GSTP1-1 inhibition. The dinitronaphthalene GST inhibitors were also confirmed to be mutagenic and cytotoxic in the Salmonella Ames assay. This wasan anticipated result, attributed to bioactivation of the nitro group substituents.These observations suggest structural modifications of the dinitronaphthalene inhibitors would be necessary for pharmaceutical applications. These include modification to reduce inhibitor toxicity and modification of inhibitors for selectivity towards GSTP1-1.

Changes in dinitronaphthalenestructures to be addressed in future studies could include replacing the nitro groups with other electron withdrawing groups e.g.trifyl (F3CSO2) or nitrile (-

C≡N). GSTP1-1 selectivity has been documented for GST inhibitor product mimics, such as the

S-(2,4-dinitronaphthalene)glutathione (DNNSG) inhibitor examined in this study. Modification of the glutathione moiety, by addition of an α-glycine phenyl group,could improve DNNSG potency towards GSTP1-1. Benzene congeners of naphthalene inhibitors 1-methoxy/1-ethoxy-

2,4-dinitrobenzene may also function as GST inhibitors, with less enzyme specificity due to smaller ring size. Lastly, previous studies have shown a significant difference in inhibitor

68 binding and affinity to GSTP1 allelic variants Ile*104 vs. Val*104, indicating both should be examined during GST inhibitor development.

This study has identified a novel class of human GST inhibitors and characterized the potential mutagenicity/ toxicity of these inhibitors by Salmonella Ames assay. These inhibitors represent possible lead compounds for GST inhibitor pharmaceuticals. Suggestions have been made, based on results from this study and others, to modify these dinitronaphthalene inhibitors.

These modifications would be useful in augmenting the dinitronaphthalenes for potential pharmaceutical applications.

69

F. References

Abdellatif, Y., D.Liu, E.M.Gallant, P.W.Gage, P.G.Board, and A.F.Dulhunty. 2007. The Mu class glutathione transferase is abundant in striated muscle and is an isoform-specific regulator of ryanodine receptor calcium channels. Cell Calcium. 41:429-440.

Adler, V., Z.Yin, S.Y.Fuchs, M.Benezra, L.Rosario, K.D.Tew, M.R.Pincus, M.Sardana, C.J.Henderson, C.R.Wolf, R.J.Davis, and Z.Ronai. 1999. Regulation of JNK signaling by GSTp. EMBO J. 18:1321-1334.

Allocati, N., L.Federici, M.Masulli, and C.Di Ilio. 2009. Glutathione transferases in bacteria. FEBS J. 276:58-75.

Ames, B.N., F.D.Lee, and W.E.Durston. 1973. An improved bacterial test system for the detection and classification of mutagens and carcinogens. Proc. Natl. Acad. Sci. U.S.A. 70:782- 786.

Angelucci, F., P.Baiocco, M.Brunori, L.Gourlay, V.Morea, and A.Bellelli. 2005. Insights into the catalytic mechanism of glutathione S-transferase: the lesson from Schistosoma haematobium. Structure. 13:1241-1246.

Arlt, V.M., M.Stiborova, B.J.vom, M.L.Simoes, G.M.Lord, J.L.Nortier, M.Hollstein, D.H.Phillips, and H.H.Schmeiser. 2007. Aristolochic acid mutagenesis: molecular clues to the aetiology of Balkan endemic nephropathy-associated urothelial cancer. Carcinogenesis. 28:2253- 2261. Atkins, W.M., R.W.Wang, A.W.Bird, D.J.Newton, and A.Y.Lu. 1993. The catalytic mechanism of glutathione S-transferase (GST). Spectroscopic determination of the pKa of Tyr-9 in rat alpha 1-1 GST.J. Biol. Chem. 268:19188-19191. Atkinson, H.J. and P.C.Babbitt. 2009. Glutathione transferases are structural and functional outliers in the thioredoxin fold. Biochemistry. 48:11108-11116.

Atkinson, R. and J.Arey. 2007. Mechanisms of the gas-phase reactions of aromatic hydrocarbons and PAHs with OH and NO3 radicals. Polycyclic Aromat. Compd. 27:15-40.

Baez, S., J.Segura-Aguilar, M.Widersten, A.-S.Johansson, and B.Mannervik. 1997. Glutathione transferases catalyse the detoxication of oxidized metabolites(o-quinones) of catecholamines and may serve as an antioxidant system preventing degenerativecellular processes. Biochem.J. 324:25-28.

Bausinger, T., U.Dehner, and J.Preuss. 2004. Determination of mono-, di- and trinitronaphthalenes in soil samples contaminated by explosives. Chemosphere. 57:821-829.

70

Beuckmann, C.T., K.Fujimori, Y.Urade, and O.Hayaishi. 2000. Identification of mu-class glutathione transferases M2-2 and M3-3 as cytosolic prostaglandin E synthases in the human brain. Neurochem. Res. 25:733-738.

Bhattacharjee, P., S.Paul, M.Banerjee, D.Patra, P.Banerjee, N.Ghoshal, A.Bandyopadhyay, and A.K.Giri. 2013. Functional compensation of glutathione S-transferase M1 (GSTM1) null by another GST superfamily member, GSTM2. Sci. Rep. 3:2704.

Bico, P., C.Y.Chen, M.Jones, J.Erhardt, and H.Dirr. 1994. Class pi glutathione S-transferase: Meisenheimer complex formation. Biochem. Mol. Biol. Int. 33:887-892.

Blum, M., D.M.Grant, W.McBride, M.Heim, and U.A.Meyer. 1990. Human arylamine N- acetyltransferase genes: isolation, chromosomal localization, and functional expression. DNA Cell Biol. 9:193-203.

Booth, J., E.Boyland, and P.Sims. 1961. An enzyme from rat liver catalysing conjugations with glutathione. Biochem. J.79:516-524.

Borbulevych, O.Y., O.V.Shishkin, and V.N.Knyazev. 1999. A zwitterionic Meisenheimer complex of 2,4,6-trinitrobenzene. Acta. Cryst. 55:1704-1706. Boyland, E. and L.F.Chasseaud. 1969. Glutathione S-aralkyltransferase. Biochem. J. 115:985- 991. Bradford, M.M. 1976. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding.Anal.Biochem. 72:248-254.

Bruns, C.M., I.Hubatsch, M.Ridderstrom, B.Mannervik, and J.A.Tainer. 1999. Human glutathione transferase A4-4 crystal structures and mutagenesis reveal the basis of high catalytic efficiency with toxic lipid peroxidation products. J. Mol. Biol. 288:427-439.

Catterall, F., L.J.King, and C.Ioannides. 2002. Mutagenic activity of the glutathione S- transferase substrate 1-chloro-2,4-dinitrobenzene (CDNB) in the Salmonella mutagenicity assay. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 540:119-124.

Cho, S.G., Y.H.Lee, H.S.Park, K.Ryoo, K.W.Kang, J.Park, S.J.Eom, M.J.Kim, T.S.Chang, S.Y.Choi, J.Shim, Y.Kim, M.S.Dong, M.J.Lee, S.G.Kim, H.Ichijo, and E.J.Choi. 2001. Glutathione S-transferase mu modulates the stress-activated signals by suppressing apoptosis signal-regulating kinase 1. J. Biol. Chem. 276:12749-12755.

Clark, A.G. and M.Sinclair. 1988. The Meisenheimer complex of glutathione and trinitrobenzene. A potent inhibitor of the glutathione S-transferase from Galleria mellonella.Biochem Pharmacol. 37:259-263.

71

Clarke, M.L., R.L.Burton, A.N.Hill, M.Litorja, M.H.Nahm, and J.Hwang. 2010. Low-cost, high- throughput, automated counting of bacterial colonies. Cytometry. 77:790-797.

Comstock, K.E., M.Widersten, X.Y.Hao, W.D.Henner, and B.Mannervik. 1994. A comparison of the enzymatic and physicochemical properties of human glutathione transferase M4-4 and three other human Mu class enzymes. Arch. Biochem. Biophys. 311:487-495. da Fonseca, R.R., W.E.Johnson, S.J.O'Brien, V.Vasconcelos, and A.Antunes. 2010. Molecular evolution and the role of oxidative stress in the expansion and functional diversification of cytosolic glutathione transferases. BMC.Evol. Biol. 10:281.

Dagnino-Subiabre, A., B.K.Cassels, S.Baez, A.S.Johansson, B.Mannervik, and J.Segura-Aguilar. 2000. Glutathione transferase M2-2 catalyzes conjugation of dopamine and dopa o-quinones. Biochem.Biophys. Res. Commun. 274:32-36.

Davis, R.J. 2000. Signal transduction by the JNK group of MAP kinases. Cell 103:239-252.

De Luca, A., L.Federici, M.De Canio, L.Stella, and A.M.Caccuri. 2012. New insights into the mechanism of JNK1 inhibition by glutathione transferase P1-1. Biochemistry. 51:7304-7312.

Di Ilio, C., P.Sacchetta, B.M.Lo, A.M.Caccuri, and G.Federici. 1986. Selenium independent activity associated with cationic forms of glutathione transferase in human heart. J. Mol. Cell Cardiol. 18:983-991.

Dirr, H., P.Reinemer, and R.Huber. 1999. X-ray crystal structures of cytosolic glutathione S- transferases. Implications for protein architecture, substrate recognition and catalytic function.Eur. J. Biochem. 220:641-661.

Dorion, S., H.Lambert, and J.Landry. 2002. Activation of the p38 signaling pathway by heat shock involves the dissociation of glutathione S-transferase Mu from Ask1. J. Biol. Chem. 277:30792-30797.

Dreij, K., K.Sundberg, B.Jernstrom, A.S.Johansson, B.Mannervik, and A.Seidel. 2010. Inactivation of carcinogenic diol epoxides of dibenzo[a,l]pyrene (dibenzo[def,p]chrysene) by human alpha class glutathione transferases. Polycyclic Aromat. Compd. 22:823-829.

Dreij, K., K.Sundberg, A.S.Johansson, E.Nordling, A.Seidel, B.Persson, B.Mannervik, and B.Jernstrom. 2002. Catalytic activities of human Alpha class glutathione transferases toward carcinogenic dibenzo[a,l]pyrene diol epoxides. Chem. Res. Toxicol. 15:825-831.

Edwards, S.W. and W.E.Knox. 1956. Homogentisate metabolism: the isomerization of maleylacetoacetate by an enzyme which requires glutathione. J. Biol. Chem. 220:79-91.

72

Federici, L., C.Lo Sterzo, P.Silvia, M.Adele, F.Scaloni, G.Federici, and A.M.Caccuri. 2009. Structural basis for the binding of the anticancer compound 6-(7-nitro-2,1,3-benzoxadiazol-4- ylthio)hexanol to human glutathione S-transferases. Cancer Res. 69:8025-8034.

Frova, C. 2006. Glutathione transferases in the genomics era: new insights and perspectives. Biomol. Eng. 23:149-169.

Garte, S., L.Gaspari, A.K.Alexandrie, C.Ambrosone, H.Autrup, J.L.Autrup, H.Baranova, L.Bathum, S.Benhamou, E.Taioli.et al., 2001. Metabolic gene polymorphism frequencies in control populations.Cancer Epidemiol.Biomarkers Prev. 10:1239-1248.

Gate, L., R.S.Majumdar, A.Lunk, and K.D.Tew. 2004. Increased myeloproliferation in glutathione S-transferase pi-deficient mice is associated with a deregulation of JNK and Janus kinase/STAT pathways. J. Biol. Chem. 279:8608-8616.

Gelboin, H.V. 1980. Benzo[a]pyrene metabolism, activation and carcinogenesis: role and regulation of mixed-function oxidases and related enzymes. Physiol Rev. 60:1107-1166.

Ricci, G., F.De Maria, G.Antonini, P.Turella, A.Bullo, L.Stella, G.Filomeni, G.Federici, and A.M.Caccuri. 2005. 7-Nitro-2,1,3-benzoxadiazole derivatives, a new class of suicide inhibitors for glutathione S-transferases. Mechanism of action of potential anticancer drugs.J. Biol. Chem. 280:26397-26405.

Goldberg, D.M. 1980. Structural, functional, and clinical aspects of gamma- glutamyltransferase.CRC Crit Rev. Clin. Lab Sci. 12:1-58.

Graminski, G.F., Y.Kubo, and R.N.Armstrong.1989a. Spectroscopic and kinetic evidence for the thiolate anion of glutathione at the active site of glutathione S-transferase.Biochemistry 28:3562- 3568.

Graminski, G.F., P.H.Zhang, M.A.Sesay, H.L.Ammon, and R.N.Armstrong. 1989b. Formation of the 1-(S-glutathionyl)-2,4,6-trinitrocyclohexadienate anion at the active site of glutathione S- transferase: evidence for enzymic stabilization of sigma-complex intermediates in nucleophilic aromatic substitution reactions. Biochemistry 28:6252-6258.

Groom, H., M.Lee, P.Patil, and P.D.Josephy. 2014. Inhibition of human glutathione transferases by dinitronaphthalene derivatives. Arch. Biochem. Biophys. 555-556:71-76.

Guengerich, F.P., L.A.Peterson, J.L.Cmarik, N.Koga, and P.B.Inskeep. 1987. Activation of dihaloalkanes by glutathione conjugation and formation of DNA adducts. Environ. Health Perspect. 76:15-18.

Guise, C.P., A.T.Wang, A.Theil, D.J.Bridewell, W.R.Wilson, and A.V.Patterson. 2007. Identification of human reductases that activate the dinitrobenzamide mustard prodrug PR-104A:

73 a role for NADPH:cytochrome P450 oxidoreductase under hypoxia. Biochem.Pharmacol. 74:810-820.

Habig, W.H., M.J.Pabst, G.Fleischner, Z.Gatmaitan, I.M.Arias, and W.B.Jakoby. 1974a. The identity of glutathione S-transferase B with ligandin, a major binding protein of liver. Proc. Natl. Acad. Sci. U.S.A. 71:3879-3882.

Habig, W.H., M.J.Pabst, and W.B.Jakoby.1974b.Glutathione S-transferases.The first enzymatic step in mercapturic acid formation.J. Biol. Chem. 249:7130-7139.

Halladay, J.S., J.M.Sauer, and I.G.Sipes. 1999. Metabolism and disposition of [(14)C]1- nitronaphthalene in male Sprague-Dawley rats. Drug Metab. Dispos. 27:1456-1465.

Hansson, L.O., R.Bolton-Grob, M.Widersten, and B.Mannervik. 1999. Structural determinants in domain II of human glutathione transferase M2-2 govern the characteristic activities with aminochrome, 2-cyano-1,3-dimethyl-1-nitrosoguanidine, and 1,2-dichloro-4-nitrobenzene. Protein Sci. 8:2742-2750.

Hayes, J.D., J.U.Flanagan, and I.R.Jowsey. 2005. Glutathione transferases. Annu. Rev. Pharmacol. Toxicol. 45:51-88.

Hayes, J.D. and D.J.Pulford. 1995. The glutathione S-transferase supergene family: regulation of GST and the contribution of the isoenzymes to cancer chemoprotection and drug resistance. Crit Rev. Biochem. Mol. Biol. 30:445-600.

Hayeshi, R., S.Mukanganyama, B.Hazra, B.Abegaz, and J.A.Hasler. 2004. The interaction of selected natural products with human recombinant glutathione transferases. Phytother. Res. 18:877-883.

Hayeshi, R., I.Mutingwende, W.Mavengere, V.Masiyanise, and S.Mukanganyama. 2007. The inhibition of human glutathione S-transferases activity by plant polyphenolic compounds ellagic acid and curcumin. Food Chem. Toxicol. 45:286-295.

Hewawasam, R., D.Liu, M.G.Casarotto, A.F.Dulhunty, and P.G.Board. 2010. The structure of the C-terminal helical bundle in glutathione transferase M2-2 determines its ability to inhibit the cardiac ryanodine receptor. Biochem.Pharmacol. 80:381-388.

Hopkins, F. and L.Harris. 1929. On Glutathione: A Reinvestigation. J. Biol. Chem. 84:269-320.

Hou, L., M.T.Honaker, L.M.Shireman, L.M.Balogh, A.G.Roberts, K.C.Ng, A.Nath, and W.M.Atkins. 2007. Functional promiscuity correlates with conformational heterogeneity in A- class glutathione S-transferases. J. Biol. Chem. 282:23264-23274.

74

Howie, A.F., L.M.Forrester, M.J.Glancey, J.J.Schlager, G.Powis, G.J.Beckett, J.D.Hayes, and C.R.Wolf. 1990. Glutathione S-transferase and glutathione peroxidase expression in normal and tumour human tissues. Carcinogenesis. 11:451-458.

Hubatsch, I., M.Ridderstrom, and B.Mannervik. 1998. Human glutathione transferase A4-4: an alpha class enzyme with high catalytic efficiency in the conjugation of 4-hydroxynonenal and other genotoxic products of lipid peroxidation. Biochem. J. 330 ( Pt 1):175-179.

Ichijo, H., E.Nishida, K.Irie, P.ten Dijke, M.Saitoh, T.Moriguchi, M.Takagi, K.Matsumoto, K.Miyazono, and Y.Gotoh. 1997. Induction of apoptosis by ASK1, a mammalian MAPKKK that activates SAPK/JNK and p38 signaling pathways. Science 275:90-94.

Jemth, P. and B.Mannervik. 1997. Kinetic characterization of recombinant human glutathione transferase T1-1, a polymorphic detoxication enzyme. Arch. Biochem. Biophys. 348:247-254.

Ji, X., R.N.Armstrong, and G.L.Gilliland. 1993. Snapshots along the reaction coordinate of an

SNAr reaction catalyzed by glutathione transferase. Biochemistry. 32:12949-12954.

Ji, X., P.Zhang, R.N.Armstrong, and G.L.Gilliland. 1992. The three-dimensional structure of a glutathione S-transferase from the Mu gene class. Structural analysis of the binary complex of isoenzyme 3-3 and glutathione at 2.2-Å resolution.Biochemistry. 31:10169-10184.

Johansson, A.S. and B.Mannervik. 2001. Human glutathione transferase A3-3, a highly efficient catalyst of double-bond isomerization in the biosynthetic pathway of steroid hormones. J. Biol. Chem. 276:33061-33065.

Johansson, A.S. and B.Mannervik. 2002. Active-site residues governing high steroid isomerase activity in human glutathione transferase A3-3. J. Biol. Chem. 277:16648-16654.

Josephy, P.D. 2002.Genetically-engineered bacteria expressing human enzymes and their use in the study of mutagens and mutagenesis.Toxicology 181-182:255-260.

Josephy, P.D., P.Gruz, and T.Nohmi. 1997. Recent advances in the construction of bacterial genotoxicity assays. Mutat. Res. 386:1-23.

Josephy, P.D. and M.Novak. 2013.Reactive electrophilic metabolites of aromatic amine and amide carcinogens.Front. Biosci.5:341-359.

Josephy, P.D., J.E.Summerscales, L.S.DeBruin, C.Schlaeger, and J.Ho. 2002. N- Hydroxyarylamine O-acetyltransferase-deficient Escherichia coli strains are resistant to the mutagenicity of nitro compounds. Biol. Chem. 383:977-982.

Kadlubar F.F., L.E.Unruh, Beland F.A., K.M.Straub, and F.E.Evans. 1980. In vitro reaction of the carcinogen, N-hydroxy-2-naphthylamine, with DNA at the C-8 and N2 atoms of guanine and at the N6 atom of adenine. Carcinogenesis. 1:139-150.

75

Ketley, J.N., W.H.Habig, and W.B.Jakoby. 1975. Binding of nonsubstrate ligands to the glutathione S-transferases. J. Biol. Chem. 250:8670-8673.

Kong, K.H., M.Nishida, H.Inoue, and K.Takahashi. 1992a. Tyrosine-7 is an essential residue for the catalytic activity of human class PI glutathione S-transferase: chemical modification and site- directed mutagenesis studies. Biochem.Biophys. Res. Commun. 182:1122-1129.

Kong, K.H., K.Takasu, H.Inoue, and K.Takahashi. 1992b. Tyrosine-7 in human class Pi glutathione S-transferase is important for lowering the pKa of the thiol group of glutathione in the enzyme-glutathione complex. Biochem.Biophys. Res. Commun. 184:194-197.

La Roche, S.D. and T.Leisinger. 1990. Sequence analysis and expression of the bacterial dichloromethane dehalogenase structural gene, a member of the glutathione S-transferase supergene family. J. Bacteriol. 172:164-171.

Ladner, J.E., J.F.Parsons, C.L.Rife, G.L.Gilliland, and R.N.Armstrong. 2004. Parallel evolutionary pathways for glutathione transferases: structure and mechanism of the mitochondrial class kappa enzyme rGSTK1-1. Biochemistry. 43:352-361.

Lin, C.Y., B.C.Boland, Y.J.Lee, M.R.Salemi, D.Morin, L.A.Miller, C.G.Plopper, and A.R.Buckpitt. 2006. Identification of proteins adducted by reactive metabolites of naphthalene and 1-nitronaphthalene in dissected airways of rhesus macaques. Proteomics. 6:972-982.

Litwack, G., B.Ketterer, and I.M.Arias. 1971. Ligandin: a hepatic protein which binds steroids, bilirubin, carcinogens and a number of exogenous organic anions. Nature. 234:466-467.

Liu, D., R.Hewawasam, Y.Karunasekara, M.G.Casarotto, A.F.Dulhunty, and P.G.Board. 2012. The inhibitory glutathione transferase M2-2 binding site is located in divergent region 3 of the cardiac ryanodine receptor. Biochem.Pharmacol. 83:1523-1529.

Liu, D., R.Hewawasam, S.M.Pace, E.M.Gallant, M.G.Casarotto, A.F.Dulhunty, and P.G.Board. 2009. Dissection of the inhibition of cardiac ryanodine receptors by human glutathione transferase GSTM2-2. Biochem.Pharmacol. 77:1181-1193.

Liu, J. and A.Lin. 2005. Role of JNK activation in apoptosis: A double-edged sword. Cell Res. 15:36-42.

Lo, H.W. and F.Ali-Osman. 2007. Genetic polymorphism and function of glutathione S- transferases in tumor drug resistance. Curr.Opin.Pharmacol.7:367-374.

Lyttle, M.H., M.D.Hocker, H.C.Hui, C.G.Caldwell, A.T.Decius, A.Engqvist-Goldstein, J.E.Flatgaard, and K.E.Bauer. 1994. Isozyme-specific glutathione-S-transferase inhibitors: design and synthesis. J. Med. Chem. 37:189-194.

76

Madeo, A., M.Vinciguerra, R.Lappano, M.Galgani, A.Gasperi-Campani, M.Maggiolini, and A.M.Musti. 2010. c-Jun activation is required for 4-hydroxytamoxifen-induced cell death in breast cancer cells. Oncogene 29:978-991.

Mahajan, S.S., L.Hou, C.Doneanu, R.Paranji, D.Maeda, J.Zebala, and W.M.Atkins. 2006. Optimization of bivalent glutathione S-transferase inhibitors by combinatorial linker design. J. Am. Chem. Soc. 128:8615-8625.

Mannervik, B. 1985.The isoenzymes of glutathione transferase.Adv.Enzymol. Relat Areas Mol. Biol. 57:357-417.

Mannervik, B., P.Alin, C.Guthenberg, H.Jensson, M.K.Tahir, M.Warholm, and H.Jornvall. 1985. Identification of three classes of cytosolic glutathione transferase common to several mammalian species: correlation between structural data and enzymatic properties. Proc. Natl. Acad. Sci. USA82:7202-7206.

Mannervik, B., P.G.Board, J.D.Hayes, I.Listowsky, and W.R.Pearson. 2005. Nomenclature for mammalian soluble glutathione transferases. Methods Enzymol. 401:1-8.

Mannervik, B., R.Bolton-Grob, A-S.Johansson. 1999. Use of silent mutations in cDNA encoding human glutathionetransferase M2-2 for optimized expression in Escherichia coli. ProteinExpress. Purif. 17:105-122

Maron, D.M. and B.N.Ames. 1983. Revised methods for the Salmonella mutagenicity test. Mutat. Res. 113:173-215.

Martin, J.L. 1995. Thioredoxin-a fold for all reasons.Structure. 3:245-250.

Martinez, M.D., S.Eshaghi, and P.Nordlund. 2008. Catalysis within the lipid bilayer-structure and mechanism of the MAPEG family of integral membrane proteins. Curr.Opin.Struct. Biol. 18:442-449.

McCoy, E.C., E.J.Rosenkranz, L.A.Petrullo, H.S.Rosenkranz, and R.Mermelstein. 1981. Structural basis of the mutagenicity in bacteria of nitrated naphthalene and derivatives. Environ. Mutagen. 3:499-511.

McGregor, D., R.D.Prentice, M.McConville, Y.J.Lee, and W.J.Caspary. 1984. Reduced mutant yield at high doses in the Salmonella/activation assay: the cause is not always toxicity. Environ. Mutagen. 6:545-557.

Mortelmans, K. and E.Zeiger. 2000. The Ames Salmonella/microsome mutagenicity assay. Mutat. Res. 455:29-60.

77

Mukanganyama, S., M.Widersten, Y.S.Naik, B.Mannervik, and J.A.Hasler. 2002. Inhibition of glutathione S-transferases by antimalarial drugs possible implications for circumventing anticancer drug resistance. Int. J.Cancer. 97:700-705.

Munoz, P., S.Huenchuguala, I.Paris, and J.Segura-Aguilar. 2012. Dopamine oxidation and autophagy. Parkinsons Dis. 2012:920953.

Murakami, M., Y.Nakatani, T.Tanioka, and I.Kudo. 2002. Prostaglandin E synthase. Prostaglandins Other Lipid Mediat. 68-69:383-399.

Nath, A. and W.M.Atkins. 2008. A quantitative index of substrate promiscuity. Biochemistry. 47:157-166.

Nebert, D.W. and V.Vasiliou. 2004. Analysis of the glutathione S-transferase (GST) gene family. Hum. Genomics 1:460-4.64.

O'Brien, M.L., B.Vulevic, S.Freer, J.Boyd, H.Shen, and K.D.Tew. 1999. Glutathione peptidomimetic drug modulator of multidrug resistance-associated protein. J. Pharmacol. Exp. Ther. 291:1348-1355.

O'Dwyer, P.J., F.LaCreta, S.Nash, P.W.Tinsley, R.Schilder, M.L.Clapper, K.D.Tew, L.Panting, S.Litwin, R.L.Comis, et al. 1991. Phase I study of thiotepa in combination with the glutathione transferase inhibitor ethacrynic acid. Cancer Res. 51:6059-6065.

Oakley, A.J., M.Bello, A.Battistoni, G.Ricci, J.Rossjohn, H.O.Villar, and M.W.Parker. 1997. The structures of human glutathione transferase P1-1 in complex with glutathione and various inhibitors at high resolution. J. Mol. Biol. 274:84-100.

Oakley, A.J., B.M.Lo, M.Nuccetelli, A.P.Mazzetti, and M.W.Parker. 1999. The ligandin (non- substrate) binding site of human Pi class glutathione transferase is located in the electrophile binding site (H-site). J. Mol. Biol. 291:913-926.

Onoe, H., Y.Watanabe, K.Ono, Y.Koyama, and O.Hayaishi. 1992. Prostaglandin E2 exerts an awaking effect in the posterior hypothalamus at a site distinct from that mediating its febrile action in the anterior hypothalamus. J. Neurosci. 12:2715-2725.

Pitts, J., R.Atkinson, J.A.Sweetman, and B.Zielinska. 1985. The gas-phase reaction of naphthalene with N2O5 to form nitronaphthalenes. Atmos. Environ.19:701-705.

Prade, L., R.Huber, T.H.Manoharan, W.E.Fahl, and W.Reuter. 1997. Structures of class pi glutathione S-transferase from human placenta in complex with substrate, transition-state analogue and inhibitor. Structure. 5:1287-1295.

Raza, A., N.Galili, S.Smith, J.Godwin, J.Lancet, M.Melchert, M.Jones, J.G.Keck, L.Meng, G.L.Brown, and A.List. 2009. Phase 1 multicenter dose-escalation study of ezatiostat

78 hydrochloride (TLK199 tablets), a novel glutathione analog prodrug, in patients with myelodysplastic syndrome. Blood 113:6533-6540.

Ricci, G., F.De Maria, G.Antonini, P.Turella, A.Bullo, L.Stella, G.Filomeni, G.Federici, and A.M.Caccuri. 2005. 7-Nitro-2,1,3-benzoxadiazole derivatives, a new class of suicide inhibitors for glutathione S-transferases. Mechanism of action of potential anticancer drugs.J. Biol. Chem. 280:26397-26405.

Rosenkranz, H.S., E.C.McCoy, D.R.Sanders, M.Butler, D.K.Kiriazides, and R.Mermelstein. 1980. Nitropyrenes: isolation, identification, and reduction of mutagenic impurities in carbon black and toners. Science. 209:1039-1043.

Rosenkranz, H.S. and R.Mermelstein. 1983. Mutagenicity and genotoxicity of nitroarenes. All nitro-containing chemicals were not created equal. Mutat. Res. 114:217-267.

Rowe, J.D., E.Nieves, and I.Listowsky. 1997. Subunit diversity and tissue distribution of human glutathione S-transferases: interpretations based on electrospray ionization-MS and peptide sequence-specific antisera. Biochem. J. 325:481-486.

Ruscoe, J.E., L.A.Rosario, T.Wang, L.Gate, P.Arifoglu, C.R.Wolf, C.J.Henderson, Z.Ronai, and K.D.Tew. 2001. Pharmacologic or genetic manipulation of glutathione S-transferase P1-1 (GSTpi) influences cell proliferation pathways. J. Pharmacol. Exp. Ther. 298:339-345.

Segura-Aguilar, J., S.Baez, M.Widersten, C.J.Welch, and B.Mannervik. 1997. Human class Mu glutathione transferases, in particular isoenzyme M2-2, catalyze detoxication of the dopamine metabolite aminochrome. J. Biol. Chem. 272:5727-5731.

Shiotsuki, T., A.Koiso, and M.Eto. 1990. Inhibition of glutathione transferase by S-benzyl glutathione analogous to the conjugate of saligenin cyclic phosphate. Pesticide Biochem. Physiol. 37:121-129.

Simic, T., A.Savic-Radojevic, M.Pljesa-Ercegovac, M.Matic, and J.Mimic-Oka. 2009. Glutathione S-transferases in kidney and urinary bladder tumors. Nat. Rev. Urol. 6:281-289.

Squadrito, G.L., F.R.Fronczek, D.F.Church, and W.A.Pryor. 1989. Free-radical nitration of naphthalene with nitrogen dioxide in carbon tetrachloride and implications for environmental nitrations. J. Org. Chem.54:548-552.

Summer, K.H., W.Goggelmann, and H.Greim. 1980. Glutathione and glutathione S-transferases in the Salmonella mammalian-microsome mutagenicity test. Mutat. Res. 70:269-278.

Tentori, L., A.S.Dorio, E.Mazzon, A.Muzi, A.Sau, S.Cuzzocrea, P.Vernole, G.Federici, A.M.Caccuri, and G.Graziani. 2011. The glutathione transferase inhibitor 6-(7-nitro-2,1,3- benzoxadiazol-4-ylthio)hexanol (NBDHEX) increases temozolomide efficacy against malignant melanoma. Eur. J. Cancer. 47:1219-1230.

79

Terrier, F. 2013.Modern Nucleophilic Aromatic Substitution. Wiley-VCH, New York.

Tournier, C., P.Hess, D.D.Yang, J.Xu, T.K.Turner, A.Nimnual, D.Bar-Sagi, S.N.Jones, R.A.Flavell, and R.J.Davis. 2000. Requirement of JNK for stress-induced activation of the cytochrome c-mediated death pathway. Science. 288:870-874.

Townsend, D.M. and K.D.Tew. 2003. The role of glutathione-S-transferase in anti-cancer drug resistance. Oncogene. 22:7369-7375.

Turella, P., C.Cerella, G.Filomeni, A.Bullo, F.De Maria, L.Ghibelli, M.R.Ciriolo, M.Cianfriglia, M.Mattei, G.Federici, G.Ricci, and A.M.Caccuri. 2005. Proapoptotic activity of new glutathione S-transferase inhibitors. Cancer Res. 65:3751-3761.

Ueda, O., K.Sugihara, S.Ohta, and S.Kitamura. 2005. Involvement of molybdenum hydroxylases in reductive metabolism of nitro polycyclic aromatic hydrocarbons in mammalian skin. Drug Metab Dispos. 33:1312-1318.

Veiga-da-Cunha, M., D.Tyteca, V.Stroobant, P.J.Courtoy, F.R.Opperdoes, and S.E.Van. 2010. Molecular identification of NAT8 as the enzyme that acetylates cysteine S-conjugates to mercapturic acids. J. Biol. Chem. 285:18888-18898.

Wang, T., A.Pinar, R.Ze'ev, and D.T.Kenneth. 2001. Glutathione S-transferase P1–1 (GSTP1–1) inhibits c-Jun N-terminal kinase (JNK1) signaling through interaction with the C terminus. J. Biol. Chem. 276:20999-21003.

Watanabe, M., M.Ishidate, and T.Nohmi. 1989. A sensitive method for the detection of mutagenic nitroarenes: construction of nitroreductase-overproducing derivatives of Salmonella typhimurium strains TA98 and TA100. Mutat. Res. 216:211-220.

Watson, A., A.Eilers, D.Lallemand, J.Kyriakis, L.L.Rubin, and J.Ham. 1998. Phosphorylation of c-Jun is necessary for apoptosis induced by survival signal withdrawal in cerebellar granule neurons. J. Neurosci. 18:751-762.

Watt, K.C. and A.R.Buckpitt. 2000. Species differences in the regio- and stereoselectivity of 1- nitronaphthalene metabolism. Drug Metab Dispos. 28:376-378.

Weston, C.R. and R.J.Davis. 2002. The JNK signal transduction pathway. Curr.Opin.Genet. Dev. 12:14-21.

Wilce, M.C. and M.W.Parker. 1994. Structure and function of glutathione S-transferases. Biochim.Biophys. Acta. 1205:1-18.

Wild, D. 1990. A novel pathway to the ultimate mutagens of aromatic amino and nitro compounds. Environ. Health Perspect. 88:27-31.

80

Wild, D., A.Dirr, I.Fasshauer, and D.Henschler. 1989. Photolysis of arylazides and generation of highly electrophilic DNA-binding and mutagenic intermediates. Carcinogenesis. 10:335-341.

Wu, B. and D.Dong. 2012. Human cytosolic glutathione transferases: structure, function, and drug discovery. Trends Pharmacol. Sci. 33:656-668.

Yankeelov, J.A. 1963. Split-compartment mixing cells for difference spectroscopy.Anal.Biochem. 6:287-289.

Zander, M. and O.Hutzinger. 1970. Phosphorimetry of chloro- and nitro-aromatic fungicides. Bull. Environ. Contam Toxicol. 5:565-568.

Zhao, T., S.S.Singhal, J.T.Piper, J.Cheng, U.Pandya, J.Clark-Wronski, S.Awasthi, and Y.C.Awasthi. 1999. The role of human glutathione S-transferases hGSTA1-1 and hGSTA2-2 in protection against oxidative stress. Arch. Biochem. Biophys. 367:216-224

81

G. Appendices i. Numerical data a. GST IC50 Inhibition values

CDNN inhibition of GSTM1 (15 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[CDNN] Mean Vo Std. Err. µM (nmoles/min)

500 0.0 0.0 300 30.7 7.3 150 28.0 1.5 130 38.3 6.3 100 41.3 2.9 50 111.3 15.7 30 304.0 23.2 10 321.3 31.6 5 408.0 9.0 3 408.0 23.2 1 425.7 8.7 0.5 430.3 12.7 0 444.6 6.7 IC50 (µM) 35.6 5.8

DNNOH inhibition of GSTM1 (15 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNNOH] Mean Vo Std. Err. µM (nmoles/min)

500 6.3 0.3 300 17.0 0.0 150 30.3 1.3 130 34.3 2.7 100 42.3 3.7 50 72.0 3.8 30 94.7 6.2 10 251.3 4.3 5 246.7 8.8 3 382.3 17.3 1 408.7 22.6 0 445 6.7 IC50 (µM) 4.1 0.8

82

DNN inhibition of GSTM1 (1 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNN] Mean Vo Std. Err. µM (nmoles/min)

50 9.4 0.6 10 17.7 0.6 5 22.2 0.7 2.5 29.2 1.2 1.0 37.5 2.1 0.75 43.4 1.5 0.50 50.7 1.8 0.10 58.7 0.4 0.05 61.8 2.5 0.0 62.6 0.78 IC50 (µM) 1.3 0.2

DNNOMe inhibition of GSTM1 (15 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNNOMe] Mean Std. Err. µM Vo(nmoles/min)

2000 1.4 1.4 1000 60.7 5.4 500 71.3 4.1 250 100.0 2.3 100 152.7 9.4 50 232.7 9.1 25 345.7 9.8 10 373.3 23.2 5 434.3 17.3 2.5 529.7 17.3 1 661.7 7.9 0.5 746.3 17.3 0.25 885.0 15.1 0.1 1024.7 17.3 0 1033.5 9.7 IC50 (µM) 2.9 0.8

83

DNNOEt inhibition of GSTM1 (15 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNNOEt] Mean Vo Std. Err. µM (nmoles/min)

2000 43.3 6.8 1000 118.0 7.0 500 177.3 10.3 250 246.3 13.7 50 328.3 5.3 25 329.7 4.7 10 397.0 10.1 5 427.3 10.3 1 651.0 15.0 0.5 720.3 37.8 0.25 833.0 30.0 0.1 1015.7 30.3 0 1033.5 9.7 IC50 (µM) 2.1 0.9

DNNSG inhibition of GSTM1 (15 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNNSG] Mean Vo Std. Err. µM (nmoles/min)

150 21.0 0.0 125 35.3 4.1 100 51.0 6.2 50 65.0 7.5 25 72.0 3.8 10 93.7 6.1 5 121.3 2.3 1 184.8 14.8 0 191.5 2.6 IC50 (µM) 7.6 2.7

84

CDNN inhibition of GSTM2 (1 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[CDNN] Mean Vo Std. Err. µM (nmoles/min)

100 2.8 1.4 50 5.6 0.7 10 7.3 0.6 5 15.6 0.6 1 31.3 1.2 0.5 36.8 1.4 0.1 47.2 3.0 0.05 63.2 3.5 0 64.1 1.7 IC50 (nM) 724.4 260.8

DNNOH inhibition of GSTM2 (1 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNNOH] Mean Vo Std. Err. µM (nmoles/min)

1500 2.3 1.2 1000 20.0 2.1 500 47.7 3.0 300 67.0 5.9 100 86.0 9.3 50 117.0 3.2 10 131.7 3.3 1 180.3 1.7 0 179.9 9.6 IC50 (µM) 97.6 --

85

DNN inhibition of GSTM2 (1 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNN] Mean Vo Std. Err. µM (nmoles/min)

100 1.6 0.3 50 4.9 0.3 10 8.8 0.8 5 20.2 1.4 2.5 29.5 1.9 0.75 41.7 1.2 0.5 48.6 0.7 0.1 57.6 1.8 0.05 63.9 1.9 0.01 72.7 2.3 0 79.8 0.6 IC50 (µM) 1.2 0.3

DNNOMe inhibition of GSTM2 (1 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNNOMe] Mean Vo Std. Err. nM (nmoles/min)

5000 4.9 1.5 1000 18.2 1.3 500 23.6 1.9 250 27.8 0.7 75 38.5 1.0 50 43.4 0.7 10 61.5 0.6 5 66.3 2.3 1 66.7 2.1 0 72.5 2.2 IC50 (nM) 119 40

86

DNNOEt inhibition of GSTM2 (1 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNNOEt] Mean Vo Std. Err. nM (nmoles/min)

5000 4.9 0.3 1000 16.3 1.3 500 25.0 1.4 250 39.6 0.6 100 48.6 1.6 50 61.8 3.6 10 71.5 4.5 5 81.6 3.4 1 88.2 3.9 0 90.2 1.4 IC50 (nM) 178 49

DNNSG inhibition of GSTM2 (1 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNNSG] Mean Vo Std. Err. µM (nmoles/min)

50 3.8 0.4 10 11.1 0.9 5 16.3 0.4 2.5 19.8 0.6 1 22.6 0.3 0.1 32.6 1.8 0.05 37.2 2.4 0 37.5 0.9 IC50 (µM) 3.4 1.9

87

CDNN inhibition of GSTA1 (15 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[CDNN] Mean Vo Std. Err. µM (nmoles/min)

1000 0 0 500 5.2 3.5 100 8.2 3.2 50 30.6 2.9 10 51.0 1.8 1 120 3.6 0 152 9.2 IC50 (µM) 4.9 1.3

DNNOH inhibition of GSTA1 (5 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNNOH] Mean Vo Std. Err. µM (nmoles/min)

50 5.0 0.4 10 9.4 0.6 5 17.7 1.1 2.5 22.4 1.0 1 30.4 0.9 0.75 37.4 2.3 0.5 45.2 3.5 0.1 55.7 1.6 0.05 60.8 1.7 0 67.7 3.0 IC50 (µM) 0.9 0.2

88

DNNSG inhibition of GSTA1 (15 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNNSG] Mean Vo Std. Err. µM (nmoles/min)

100 24.0 3.8 50 26.0 0.0 25 50.7 12.3 10 97.0 8.1 2.5 150.3 5.7 0.5 164.7 8.7 0.1 186.7 11.2 0 189.9 6.9 IC50 (µM) 3.4 1.9

DNNOH inhibition of GSTA2 (15 µg) activity towards 1 mM GSH, 0.5 mM CDNB

[DNNOH] Mean Vo Std. Err. µM (nmoles/min)

1000 0 0.0 500 13.0 2.0 300 19.7 2.4 150 28.0 1.5 100 33.0 1.2 50 51.3 2.3 10 58.0 4.2 1 79.7 3.3 0 77.6 4.4 IC50 (µM) 76.2 --

89

CDNN inhibition of GSTP1 (15 µg) activity towards 1 mM GSH, 0.5 mM CDNB

1 [CDNN] Mean Vo Std. Err. µM (nmoles/min)

1000 35.7 2.6 500 49.3 17.2 300 57.7 7.8 100 221.0 13.0 30 344.7 10.5 10 377.7 9.4 5 439.7 27.8 0 429.9 9.1 IC50 (µM) 91.2 19.4 1Indicates inhibitor solubility limit reached before 100% enzyme inhibition

DNNOH inhibition of GSTP1 (15 µg) activity towards 1 mM GSH, 0.5 mM CDNB

1 [DNNOH] Mean Vo Std. Err. mM (nmoles/min)

1.5 81.3 3.3 1 79.0 1.0 0.5 86.3 3.8 0.3 125.0 17.9 0.1 244.3 15.1 0.03 326.0 19.9 0.01 380.7 9.3 0 429.9 9.1 IC50 (µM) 89.7 18.0 1Indicates inhibitor solubility limit reached before 100% enzyme inhibition

90

b. Double reciprocal plots

A1 inhibition by CDNN

CDNB (mM) Control Act. 2 µM CDNN 4 µM CDNN 8 µM CDNN

Enzyme activity in (nmoles/min)

0 0 0 0 0 0.1 50 +/- 3 39 +/- 7 29 +/- 2 25 +/- 2 0.3 86 +/- 8 70 +/- 4 43 +/- 1 34 +/- 4 0.5 104 +/- 4 78 +/- 8 50 +/- 8 42 +/- 5 0.7 119 +/- 6 88 +/- 8 59 +/- 2 47 +/- 3 1.0 145 +/- 4 97 +/- 7 82 +/- 8 54 +/- 2 2.0 168 +/- 10 115 +/- 6 97 +/- 5 116 +/- 17

A1 inhibition by DNNOH

CDNB (mM) Control Act. 1 µM DNNOH 2 µM DNNOH 4 µM DNNOH

Enzyme activity in (nmoles/min)

0 0 0 0 0 0.1 50 +/- 3 34 +/- 2 24 +/- 2 44 +/- 3 0.3 86 +/- 8 50 +/- 2 41 +/- 5 73 +/- 3 0.5 104 +/- 4 59 +/- 4 43 +/- 4 95 +/- 1 0.7 119 +/- 6 72 +/- 9 52 +/- 8 113 +/- 2 1.0 145 +/- 4 91 +/- 2 76 +/- 4 139 +/- 3 2.0 168 +/- 10 111 +/- 9 102 +/- 15 157 +/- 8

91

A1 inhibition by DNNSG

CDNB (mM) Control Act. 5 µM DNNSG 10 µM DNNSG 20 µM DNNSG

Enzyme activity in (nmoles/min)

0 0 0 0 0 0.3 116 +/- 6 70 +/- 8 42 +/- 7 27 +/- 5 0.5 120 +/- 8 77 +/- 8 55 +/- 6 39 +/- 9 0.7 122 +/- 9 93 +/- 6 68 +/- 5 51 +/- 9 1.0 137 +/- 4 97 +/- 7 85 +/- 10 53 +/- 3 2.0 142 +/- 8 115 +/- 6 84 +/- 4 76 +/- 7

M2 inhibition by DNN

CDNB (mM) Control 1.5 µM DNN 3 µM DNN 6 µM DNN

Enzyme activity in (nmoles/min)

0 0 0 0 0 0.05 278 +/- 7 104 +/- 8 78 +/- 5 49 +/- 4 0.1 686 +/- 31 229 +/- 16 188 +/- 12 61 +/- 2 0.3 1233 +/- 17 400 +/- 23 315 +/- 14 260 +/- 6 0.5 1545 +/- 76 538 +/- 46 495 +/- 40 356 +/- 23 0.7 1632 +/- 46 755 +/- 26 550 +/- 435 391 +/- 25 1.0 1788 +/- 17 842 +/- 23 703 +/- 15 573 +/- 30

M2 inhibition by DNNOMe

CDNB (mM) Control 0.3 µM 0.6 µM 1.2 µM DNNOMe DNNOMe DNNOMe

Enzyme activity in (nmoles/min)

0 0 0 0 0 0.05 330 +/- 4 167 +/- 6 73 +/- 10 41 +/- 5 0.1 528 +/- 8 267 +/- 4 177 +/- 30 97 +/- 7 0.3 1238 +/- 25 764 +/- 46 521 +/- 30 188 +/- 6 0.5 1189 +/- 10 1250 +/- 60 885 +/- 30 356 +/- 9 0.7 1956 +/- 6 1163 +/- 17 885 +/- 30 555 +/- 18

92

M2 inhibition by DNNOEt

CDNB (mM) Control 0.3 µM 0.6 µM 1.2 µM DNNOEt DNNOEt DNNOEt

Enzyme activity in (nmoles/min)

0 0 0 0 0 0.05 330 +/- 4 158 +/- 6 76 +/- 2 44 +/- 2 0.1 528 +/- 8 265 +/- 11 170 +/- 9 87 +/- 7 0.3 1238 +/- 25 755 +/- 40 278 +/- 18 187 +/- 10 0.5 1189 +/- 10 1181 +/- 17 668 +/- 23 339 +/- 26 0.7 1956 +/- 6 1111 +/- 87 903 +/- 46 608 +/- 17

c. Salmonella Mutagenicity Assay

Salmonella YG1024 mutagenicity response to CDNN

CDNN CDNN Rev./plate +/- Std. Err. Rev./nmole (ng/plate) (pmoles)

300 1200 75 5 44 100 400 321 22 748 50 200 211 6 945 30 120 119 12 808 10 40 44 6 550 3 12 35 6 1083 0 0 22 2 0

n = 6 assays, from 2 independent, triplicate experiments

93

Salmonella YG1024 mutagenicity response to DNNOH

DNNOH DNNOH Rev./plate Std. Err. Rev./ng (µg/plate) (nmoles)

50 213.5 1631 7 8 30 128 2045 33 16 10 42.7 1652 49 38 7 29.9 1430 39 48 5 21.4 840 50 39 3 12.8 353 22 27 1 4.3 162 23 36 0.5 2.1 29 2 10 0.1 0.4 15 1 15 0 0 9 1 0

n = 6 assays, from 2 independent, triplicate experiments

Salmonella YG1024 mutagenicity response to DNN

DNN DNN (ng) Rev./plate +/- Std. Err. Rev./ng (µg/plate)

300 1374 353 78 0.24 100 458 1341 137 3 30 137 1518 59 11 10 45.8 1057 56 26 5 23 829 40 35 3 13.7 510 51 35 1 4.6 200 17 38 0.3 0.14 88 19 46 0.1 0.5 68 19 88 0 0 24 6 0.24

n = 6 assays, from 2 independent, triplicate experiments

94

Salmonella YG1024 mutagenicity response to DNNOMe

DNNOMe DNNOMe Rev./plate +/- Std. Err. Rev/nmole (µg/plate) (nmoles)

30 120 1714 76 14 10 40 1738 62 43 3 12 2564 113 212 1 4 1574 67 388 0.3 1.2 715 36 577 0.1 0.4 318 16 738 0.03 0.12 122 10 825 0.01 0.04 58 8 875 0 0 23 2 0

n = 6 assays, from 2 independent, triplicate experiments

Salmonella YG1024 mutagenicity response to DNNOEt

DNNOEt DNNOEt Rev./plate +/- Std. Err. Rev./mole (µg/plate) (nmoles)

30 114 1463 95 13 10 38 2036 75 53 3 11.4 2318 64 201 2 7.6 1629 23 212 1 3.8 882 42 227 0.3 1.14 715 14 609 0.1 0.76 225 11 268 0.03 0.14 152 21 936 0.01 0.08 48 8 338 0 0 21 3 0

n = 6 assays, from 2 independent, triplicate experiments

95

ii. Supplementary graphs

a. IC50 curves

M2-2 IC50 curves

80 60 IC50 : 0.72 +/- 0.26 M IC50 : 1.3 +/- 0.3 M

60 40

40

(nmols/min)

(nmols/min)

o

o

V

V 20 20

0 0 0 100 1000 10000 100000 0 0 10 100 1000 10000 100000 [CDNN] nM [DNN] nM Inhibition of GSTM2-2 by CDNN Inhibition of GSTM2-2 by DNN

80 100

IC50 : 0.12 +/- 0.04 M IC50 : 0.18 +/-0.05 M 80 60

60 40

(nmoles/min)

(nmoles/min) 40

o

o

V

V 20 20

0 0 0 0 1 10 100 1000 10000 0 0 1 10 100 1000 10000 [DNNOMe] nM [DNNOEt] nM Inhibition of GSTM2-2 by DNNOMe Inhibition of GSTM2-2 by DNNOEt

96

40 

IC : 97.6 M IC50 : 3.4 +/- 1.9 M 50 30 

20 

(nmols/min)

(nmoles/min)

o

o

V

V 10 

0  0 0 100 1000 10000 100000      [DNNSG] nM [DNNOH] M Inhibition of GSTM2-2 by DNNSG Inhibition of GSTM2-2 by DNNOH

GSTM1-1 IC50 curves

500 70 IC50 : 35.6 +/- 5.8 M IC50 : 1.34 +/- 0.16 M 400 60

50 300 40

200 30

(nmoles/min)

Vo (nmoles/min) Vo

o V 20 100 10 0 0 0 1 10 100 1000 0 0 100 1000 10000 100000 1000000 [CDNN] M [DNN] nM Inhibition of GSTM1-1 by CDNN Inhibition of GSTM1-1by DNN

1200 1200

IC50 : 2.9 +/- 0.8 uM IC50 : 2.1 +/- 0.9 uM 1000 1000

800 800

600 600

nmoles/min

nmoles/min

o

o

V V 400 400

200 200

0 0 0 0 0.10 1 10 100 1000 10000 0 0 0.10 1 10 100 1000 [DNNOMe] M [DNNOEt] M Inhibition of GSTM1-1by DNNOMe Inhibition of GSTM1-1 by DNNOEt

97

500 220 200 IC50 : 7.6 +/- 2.7 M IC50 : 4.1 +/- 0.8 M 180 400 160 140 300 120 100

200 (nmoles/min)

o 80

V

Vo (nmoles/min) Vo 60 100 40 20 0 0 0 0 1 10 100 1000 0 1 10 100 [DNNOH] uM [DNNSG] M Inhibition of GSTM1-1 by DNNOH Inhibition of GSTM1-1 by DNNSG

GSTA1-1 IC50 curves

80 200

IC50 : 0.92 +/- 0.2 M IC50 : 9.8 +/- 2.7 M

60 150

40 100

(nmoles/min)

o

V

Vo (nmoles/min) Vo 20 50

0 0 0 0 100 1000 10000 100000 0 0 0 1 10 100 [DNNOH] nM [DNNSG] uM Inhibition of GSTA1-1 by DNNOH Inhibition of GSTA1-1 by DNNSG

180 160 IC50 : 4.9 +/- 1.3 140 120 100 80 60

Vo (nmoles/min) Vo 40 20 0

0 0 1 10 100 1000 10000 [CDNN] M Inhibition of GSTA1-1 by CDNN

98

GSTA2-2 IC50 curves

80 IC50 : 76.2 M

60

40

Vo (nmoles/min) Vo

20

0 0 0 1 10 100 1000 [DNNOH] M Inhibition of GSTA2-2 by DNNOH

GSTP1-1 IC50 curves

500 500

IC50 : 91.2 +/- 19 M IC50 : 89.7 +/- 18 M 400 400

300 300

200 (nmoles/min) 200

o

V

Vo (nmoles/min) Vo

100 100

0 0 0.00 0.01 0.10 1.00 10.00 0.00 0.01 0.10 1.00 10.00 [CDNN] mM [DNNOH] mM Inhibition of GSTP1-1 by CDNN Inhibition of GSTP1-1 by DNNOH

99 b. Double reciprocal graphs

0.03 1200 nM DNNOMe

600 nM DNNOMe

0.02 300 nM DNNOMe

0 nM DNNOMe

0.01

(nmoles/min)

o

1/V 0.00

-0.01 -5 0 5 10 15 20 1/[CDNB] mM Double reciprocal plot of GSTM2 -2 inhibition by DNNOMe

0.03 1200 nM DNNOEt

600 nM DNNOEt

0.02 300 nM DNNOEt

0 DNNOEt

0.01

(nmoles/min)

o

1/V 0.00

-0.01 -5 0 5 10 15 20 25 1/[CDNB] mM

Double reciprocal plot of GSTM2 -2 inhibition by DNNOEt

100

0.030

M DNN 0.025 6 M DNN

0.020 3 M DNN 1.5 M DNN 0.015

(nmoles/min)

o 0.010

1/V

0.005

0.000

-5 0 5 10 15 20 25 1/[CDNB] M

Double reciprocal plot of GSTM2 -2 inhibition by DNN

0.05

20 M DNNSG

0.04 10 M DNNSG

0 M DNNSG

5 M DNNSG 0.03

(nmoles/min) o 0.02

1/V

0.01

0.00 -1 0 1 2 3 4 5 1/ [CDNB] mM

Double reciprocal plot of GSTA1-1 inhibition by DNNSG

101

0.030

2 M CDNN

0.025 4 M CDNN

8 M CDNN

0.020 M CDNN

0.015

(nmoles/min)

o

1/ V 0.010

0.005

0.000 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 1/ [CDNB] mM Double reciprocal plot of GSTA1-1 inhibition by CDNN

0.05 1 M DNNOH

0.04 2 M DNNOH

4 M DNNOH

0.03 0 M DNNOH

0.02

(nmoles/min)

o 0.01

1/ V

0.00

-0.01 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 1/[CDNB] mM Double reciprocal plot of GSTA1-1 inhibition by DNNOH