2,3,7,8-tetrachlorodibenzo-p-dioxin-inducible poly(adp … › bitstream › 1807 › 65686 › 1...

221
2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP-RIBOSE) POLYMERASE IS A MONO-ADP- RIBOSYLTRANSFERASE AND A LIGAND-INDUCED REPRESSOR OF AHR TRANSACTIVATION by Laura Mariko MacPherson A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Department of Pharmacology and Toxicology University of Toronto © Copyright by Laura Mariko MacPherson (2014)

Upload: others

Post on 10-Jun-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP-RIBOSE) POLYMERASE IS A MONO-ADP-RIBOSYLTRANSFERASE AND A LIGAND-INDUCED

REPRESSOR OF AHR TRANSACTIVATION

by

Laura Mariko MacPherson

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

Department of Pharmacology and Toxicology University of Toronto

© Copyright by Laura Mariko MacPherson (2014)

Page 2: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

ii

2,3,7,8-tetrachlorodibenzo-p-dioxin-inducible poly(ADP-ribose)

polymerase is a mono-ADP-ribosyltransferase and a ligand-

induced repressor of AHR transactivation

Laura Mariko MacPherson

Doctor of Philosophy

Department of Pharmacology and Toxicology University of Toronto

2014

Abstract

2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)-inducible poly(ADP-ribose) polymerase

(TiPARP/ARTD14) is a member of the ARTD family and is regulated by the aryl hydrocarbon

receptor (AHR); however, little is known about TiPARP function. In this study we examined the

catalytic function of TiPARP and determined its role in AHR transactivation. We observed that

TiPARP exhibited auto-mono-ADP-ribosyltransferase activity and ribosylated core histones.

RNAi-mediated knockdown of TiPARP in T47D breast cancer cells increased TCDD-dependent

cytochrome P450 1A1 (CYP1A1) and CYP1B1 mRNA expression and recruitment of AHR to

both genes. Overexpression of TiPARP reduced AHR-dependent increases in CYP1A1-reporter

gene activity, which was restored by overexpression of AHR, but not ARNT. Deletion and

mutagenesis studies showed that TiPARP-mediated inhibition of AHR required the zinc finger

and catalytic domains. TiPARP and AHR co-localized in the nucleus, directly interacted and

both were recruited to CYP1A1 in response to TCDD. Overexpression of TiPARP enhanced

whereas RNAi-mediated knockdown of TiPARP reduced TCDD-dependent AHR proteolytic

degradation. TCDD-dependent induction of AHR target genes was also enhanced in Tiparp-/-

mouse embryonic fibroblasts compared to wildtype controls. Moreover, livers excised from

Page 3: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

iii

TCDD-treated Tiparp-/- mice displayed significantly greater AHR target gene expression

compared with wildtype or heterozygous mice. Comparison of TiPARP to a known negative

regulator of AHR, AHR repressor (AHRR), revealed that TiPARP and AHRR had some notable

similarities but also differences between their mechanisms of repression. Similar to TiPARP,

AHRR was recruited to AHR target regulatory regions in response to TCDD and its

overexpression repressed reporter gene activity. However unlike TiPARP, knockdown of the

AHRR did not affect AHR transactivation or its proteasomal degradation. Despite some

mechanistic similarities, our data suggest that TiPARP and AHRR independently repress AHR

transactivation. Overall, our findings show that TiPARP is a mono-ADP-ribosyltransferase and a

transcriptional repressor of AHR, revealing a novel negative feedback loop controlling AHR

function.

Page 4: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

iv

Acknowledgments There are many people who deserve recognition for their contributions to this thesis.

I would like to express my sincere gratitude to my supervisor, Dr. Jason Matthews for giving me

the opportunity to be a graduate student in his laboratory once again. It has been a pleasure to

work under his guidance and mentorship throughout the years. I am thankful for all that I have

learned from him and all the opportunities he has provided me during this experience.

I would also like to express my gratitude to my supervisory committee: Dr. Denis Grant, Dr.

Peter McPherson and Dr. Ali Salahpour for sharing their expertise in this work. All of their

suggestions were most helpful and I value their input tremendously.

Next, I would like to mention all of my wonderful Matthews’ lab members: Dr. Shaimaa Ahmed,

Sharanya Rajendra, Alvin Gomez, Debbie Bott and Laura Tamblyn for all their technical support

and insights. An extra special thanks to Shaimaa for always being an excellent resource in the

lab and the best coffee break buddy.

I would like to acknowledge the Canadian Institutes of Health Research, Ontario Graduate

Scholarship Program and the University of Toronto for financial support of this research.

Finally, last but certainly not least, I would like to thank my parents; Douglas and Patricia

MacPherson, my brothers; Gordon and Cameron and my lovely husband Tom for their

unconditional support and encouragment during this entire experience.

Page 5: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

v

Table of Contents Acknowledgments.......................................................................................................................... iv 

Table of Contents............................................................................................................................ v 

List of Tables .................................................................................................................................. x 

List of Figures ................................................................................................................................ xi 

List of Abbreviations ................................................................................................................... xiv 

Chapter 1: Introduction ................................................................................................................... 1 

1  Statement of Research Problem ................................................................................................. 1 

2  ADP-ribosylation Reactions....................................................................................................... 2 

2.1  Historical Overview............................................................................................................ 2 

2.2  NAD+ Metabolism .............................................................................................................. 3 

2.3  Mono-ADP-ribosylation ..................................................................................................... 7 

2.3.1  Mono-ADP-ribosyltransferase families .................................................................. 7 

2.3.2  ADP-ribosylhydrolases ........................................................................................... 9 

2.4  Poly-ADP-ribosylation ..................................................................................................... 10 

2.4.1  ADP-ribosyltransferase diphtheria toxin-like (ARTD) family ............................. 10 

2.4.2  Poly(ADP-ribose) Catabolism .............................................................................. 22 

2.5  Regulation of chromatin structure and transcription by ARTDs ...................................... 24 

2.5.1  Modulation of chromatin structure by ARTD1 and poly(ADP-ribose)................ 24 

2.5.2  Enhancer-binding of actions of ARTD1 ............................................................... 25 

2.5.3  Promoter-binding and transcriptional co-regulation by ARTD1 .......................... 26 

2.5.4  Transcriptional regulation by other ARTDs ......................................................... 27 

3  Aryl hydrocarbon receptor ....................................................................................................... 28 

3.1  Structure............................................................................................................................ 29 

3.2  AHR ligands...................................................................................................................... 30 

Page 6: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

vi

3.2.1  Exogenous AHR ligands....................................................................................... 31 

3.2.2  Naturally occurring AHR ligands ......................................................................... 32 

3.3  AHR-mediated toxicity..................................................................................................... 33 

3.4  AHR transactivation.......................................................................................................... 34 

3.5  Negative regulation of AHR transactivation..................................................................... 36 

3.5.1  Proteolytic Degradation of AHR .......................................................................... 37 

3.5.2  Aryl hydrocarbon receptor repressor (AHRR) ..................................................... 38 

3.5.3  Induction of a labile AHR degradation promoting factor (ADPF) ....................... 40 

3.5.4  In vivo models used to study negative regulation of AHR ................................... 41 

3.6  AHR target genes.............................................................................................................. 43 

3.6.1  TCDD-inducible poly(ADP-ribose) polymerase (TiPARP): a novel AHR target ..................................................................................................................... 44 

Chapter 2: Aims of the present study............................................................................................ 48 

Chapter 3: Materials and Methods................................................................................................ 49 

4  Materials................................................................................................................................... 49 

4.1  Chemicals and biological reagents.................................................................................... 49 

4.2  Plasticware ........................................................................................................................ 50 

4.3  Instruments........................................................................................................................ 50 

5  Methods.................................................................................................................................... 51 

5.1  Maintenance of HuH7, COS7, TREx-FLP-IN 293, T47D, NCI-N87 and MCF7 cell lines ................................................................................................................................... 51 

5.2  Generation of Expression Constructs................................................................................ 52 

5.3  Site-directed mutagenesis ................................................................................................. 54 

5.4  Transient transfection of HuH7 cells for RNA expression, Western blot and ChIP assays ................................................................................................................................ 56 

5.5  Transient transfection of T47D cells for Western blot ..................................................... 56 

5.6  Transient transfection of HuH7 cells and reporter gene assays........................................ 57 

Page 7: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

vii

5.7  Transient transfection and RNAi ...................................................................................... 57 

5.8  mRNA time-course ........................................................................................................... 58 

5.9  Tiparp-null mice and generation of mouse embryonic fibroblasts (MEF) ....................... 58 

5.10 RNA isolation and cDNA synthesis................................................................................. 59 

5.10.1  Mammalian cells ................................................................................................... 59 

5.10.2  Mouse liver ........................................................................................................... 60 

5.10.3  cDNA synthesis .................................................................................................... 60 

5.11 Chromatin Immunoprecipitation (ChIP)........................................................................... 61 

5.12 Western blot ...................................................................................................................... 63 

5.13 Co-immunoprecipitation ................................................................................................... 64 

5.14 In vitro translation and 35S PARP assays ......................................................................... 65 

5.15  Indirect immunofluorescence........................................................................................... 65 

5.16 GST purification............................................................................................................... 66 

5.17 32P-NAD+ auto-ADP-ribosylation and hetero-ADP-ribosylation assays ......................... 67 

5.18  Statistical Analysis ........................................................................................................... 67 

Chapter 4: Results ......................................................................................................................... 69 

6  TiPARP is a mono-ADP-ribosyltransferase............................................................................. 69 

6.1  Auto-ADP-ribosylation..................................................................................................... 69 

6.2  Hetero-ADP-ribosylation.................................................................................................. 76 

7  Identification of TiPARP as a novel negative regulator of AHR............................................. 79 

7.1  Temporal induction of TiPARP mRNA by TCDD........................................................... 79 

7.2  TiPARP knockdown increased TCDD-induced AHR transactivation ............................. 81 

7.3  Overexpression of TiPARP repressed AHR-mediated transactivation ............................ 88 

7.4  Tiparp knockout increased TCDD-dependent AHR target transactivation and AHR protein levels ..................................................................................................................... 93 

7.5  Tiparp knockout increased AHR-mediated gene expression in vivo................................ 96 

Page 8: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

viii

8  Understanding the mechanism of TiPARP-mediated repression of AHR-mediated transactivation .......................................................................................................................... 99 

8.1  TiPARP selective repression of AHR activity.................................................................. 99 

8.2  Ectopic AHR but not ARNT prevented TiPARP-mediated inhibition of AHR transactivation ................................................................................................................. 101 

8.3  Identification of a putative AHR repressor domain of TiPARP..................................... 103 

8.4  Repression of AHR by TiPARP required zinc finger function....................................... 112 

8.5  Repression of AHR transactivation by TiPARP required catalytic domain and function ........................................................................................................................... 115 

8.6  TiPARP interaction with AHR ....................................................................................... 120 

8.7  Identification of nuclear localization signal of N-terminus of TiPARP. ........................ 125 

8.8  TiPARP is a labile factor that negatively regulates AHR transactivation ...................... 128 

9  Comparison of TiPARP- and AHRR-mediated negative regulation of AHR transactivation132 

9.1  Kinetics of AHRR induction by TCDD.......................................................................... 133 

9.2  AHRR knockdown did not affect AHR transactivation or protein levels ...................... 136 

9.3  AHRR overexpression repressed AHR-regulated reporter gene activity ....................... 141 

9.4  Expression patterns of the AHRR................................................................................... 144 

Chapter 5: Discussion ................................................................................................................. 146 

10 TiPARP is a mono-ADP-ribosyltransferase........................................................................... 146 

11 TiPARP is a novel negative regulator of AHR transactivation.............................................. 148 

11.1 Potential co-repression by ADP-ribosylation of AHR................................................... 150 

11.2 Potential targeting of histones and other co-regulatory factors...................................... 151 

11.3 TiPARP is a candidate AHR degradation promoting factor (ADPF) ............................ 152 

11.3.1  TiPARP and the nuclear proteasome .................................................................. 153 

11.3.2  Connecting nuclear localization with repression of AHR transactivation.......... 154 

11.4 Identification of an AHR repressive domain.................................................................. 155 

12 Comparing TiPARP- with AHRR-mediated transrepression ................................................ 156 

Page 9: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

ix

12.1 Gene expression and recruitment.................................................................................... 156 

12.2 Overexpression of ARNT and co-regulators.................................................................. 157 

12.3  Subcellular localization.................................................................................................. 157 

12.4  Knockdown and knockout models ................................................................................. 158 

12.5  Potential overlap of TiPARP and the AHRR repressive pathways................................ 159 

13 TiPARP is an in vivo negative regulator of AHR transactivation.......................................... 160 

13.1 Role of TiPARP in TCDD-mediated toxicity................................................................. 161 

13.1.1 Purported effects of TiPARP on the NAD+ pool and SIRT activity.................... 161 

14 TiPARP-mediated regulation of other transcription factors .................................................. 163 

15 Limitations and future considerations.................................................................................... 163 

15.1 TiPARP antibody ............................................................................................................ 163 

15.2  ARTD catalytic activity assays ...................................................................................... 164 

16 Overall significance of research............................................................................................. 165 

References................................................................................................................................... 167 

Copyright Acknowledgements.................................................................................................... 204 

Page 10: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

x

List of Tables Table 1. Organization and enzymatic activities of mammalian ARTD family. ........................... 12 

Table 2. Transcription factors co-regulated by ARTD1. .............................................................. 26 

Table 3. Partial list of proteins and co-regulator proteins that interact with AHR/ARNT to

modulate transcription. ................................................................................................................. 35 

Table 4. PCR primers used for generation of TiPARP and AHRR inserts................................... 54 

Table 5. PCR primers used for site-directed mutagenesis of TiPARP. ........................................ 55 

Table 6. List of qPCR primers used in RNA expression assays. .................................................. 60 

Table 7. List of qPCR primers used for ChIP assays.................................................................... 62 

Table 8. Frequency of co-localization of GFP-TiPARP and AHR in COS7 cells ..................... 120 

Page 11: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

xi

List of Figures Figure 1. Schematic representation of the mono-ADP-ribosylation reaction................................. 5 

Figure 2. Metabolism of poly(ADP-ribose).................................................................................... 6 

Figure 3. Domain architecture of the 18 human ARTD family members. ................................... 13 

Figure 4. ARTD1 domain architecture. ........................................................................................ 15 

Figure 5. Schematic of the modular domains of three bHLH-PAS family members. .................. 30 

Figure 6. Selected exogenous AHR ligands. ................................................................................ 32 

Figure 7. General mechanism of AHR transactivation. ................................................................ 36 

Figure 8. Three proposed mechanisms of negative regulation of AHR transactivation. .............. 37 

Figure 9. Functional domain structure of TCDD-inducible poly(ADP-ribose) polymerase. ....... 46 

Figure 10. TiPARP is a mono-ADP-ribosyltransferase with auto-ribosylation activity. ............. 72 

Figure 11. Putative auto-ribosylation domain of TiPARP is between residues 275 and 328. ...... 74 

Figure 12. TiPARP is a mono-ADP-ribosyltransferase with heteroribosylation activity............. 78 

Figure 13. Temporal induction of TiPARP mRNA in T47D and Hepa1c1c7 cell lines treated with

TCDD............................................................................................................................................ 80 

Figure 14. TiPARP knockdown in NCI-N87, HuH7 and T47D cells increased TCDD-induced

AHR transactivation...................................................................................................................... 83 

Figure 15. TiPARP knockdown increased TCDD-induced AHR transactivation. ....................... 85 

Figure 16. TiPARP knockdown increased AHR/ARNT recruitment to CYP1A1 and CYP1B1

regulatory regions and reduced TCDD-induced degradation of AHR. ........................................ 87 

Figure 17. TiPARP overexpression reduced AHR-regulated reporter gene activity.................... 90 

Page 12: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

xii

Figure 18. Overexpression of TiPARP repressed TCDD-induced AHR transactivation. ............ 92 

Figure 19. Tiparp knockout fibroblasts exhibited increased TCDD-induced AHR transactivation.

....................................................................................................................................................... 95 

Figure 20. Female Tiparp-/- mice demonstrated increased TCDD-induced AHR transactivation.97 

Figure 21. Male Tiparp-/- mice demonstrated increased TCDD-induced AHR transactivation. .. 98 

Figure 22. CYP1A1-luc reporter gene activity was preferentially inhibited by TiPARP. ......... 100 

Figure 23. Overexpression of AHR but not ARNT prevented TiPARP-mediated inhibition of

AHR transactivation.................................................................................................................... 102 

Figure 24. In vitro expression of 35S-labelled N-terminal TiPARP truncations. ........................ 104 

Figure 25. Overexpression of N-terminal TiPARP truncations on AHR-regulated reporter gene

activity......................................................................................................................................... 106 

Figure 26. Overexpression of N-terminal GFP-tagged TiPARP truncations on AHR-regulated

reporter gene activity. ................................................................................................................. 108 

Figure 27. Overexpression of GFP-tagged TiPARP N-terminal truncations. ............................ 109 

Figure 28. Effects of TiPARP double alanine point mutants on AHR-regulated reporter gene

activity......................................................................................................................................... 111 

Figure 29. Repression of AHR transactivation required zinc finger function. ........................... 114 

Figure 30. Repression of AHR transactivation required C-terminal catalytic domain............... 117 

Figure 31. Repression of AHR transactivation required catalytic function................................ 119 

Figure 32. TiPARP interacts with AHR in the nucleus. ............................................................. 122 

Figure 33. TiPARP truncation and point mutant interaction with AHR..................................... 124 

Figure 34. Identification of putative nuclear localization signal (NLS) within TiPARP N-

terminus....................................................................................................................................... 127 

Page 13: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

xiii

Figure 35. TiPARP is a labile negative regulator of AHR. ........................................................ 131 

Figure 36. AHRR is induced by TCDD and recruited to AHR-regulated genes. ....................... 135 

Figure 37. Effects of AHRR knockdown and TiPARP knockdown on AHR transactivation.... 139 

Figure 38. AHRR, AHR and ARNT protein levels following AHRR (A) and TiPARP (B)

knockdown.................................................................................................................................. 140 

Figure 39. Overexpression of AHRRΔ8 repressed AHR-regulated reporter gene activity. ....... 143 

Figure 40. Expression of N-terminal and C-terminal GFP-tagged AHRR in MCF7, HuH7 and

COS7 cells. ................................................................................................................................. 145 

Figure 41. Proposed model of repression of AHR transactivation by TiPARP.......................... 149 

Figure 42. Updated schematic of the structure of TiPARP. ....................................................... 155 

Page 14: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

xiv

List of Abbreviations 3MC 3-methylcholanthrene ADP adenosine diphosphate ADPF AHR degradation promoting factor ARTD ADP-ribosyltransferase diphtheria toxin-like AHR aryl hydrocarbon receptor AHRE aryl hydrocarbon receptor response element AHRR aryl hydrocarbon receptor repressor ARNT aryl hydrocarbon receptor nuclear translocator ARH ADP-ribosylhydrolase B[a]P benzo[a]pyrene BAL B-aggressive lymphoma bHLH-PAS basic Helix-Loop-Helix-Period, Aryl hydrocarbon receptor nuclear

translocator Single minded BRCT BRCA1-carboxy terminus-like Brg-1 brahma-related gene 1 CA-AHR constitutively active aryl hydrocarbon receptor ChIP chromatin immunoprecipitation CYP1A1 cytochrome P450 1A1-human CYP1B1 cytochrome P450 1B1-human DBD DNA-binding domain e-MART ecto-mono-ADP-ribosyltransferase ERα estrogen receptor α ERβ estrogen receptor β FICZ 6-formalindolo-[3,2-b]carbazole G6Pase glucose-6-phosphatase GFP green fluorescent protein GST glutathione S-transferase HAH halogenated aromatic hydrocarbon HDAC histone deacetylase HES1 hairy and enhancer of split 1 Hsp90 90 kDa heat shock protein IC3 indole-3-carbinol ICZ indolo[3,2-b]carbazole LBD ligand binding domain MART mono-ADP-ribosyltransferase MEF mouse embryonic fibroblast NAD+ nicotinamide adenine dinucleotide NASH non-alcoholic steatohepatitis NCoA nuclear co-activator NCoR nuclear receptor co-repressor NES nuclear export signal NF-κB nuclear factor kappa-light-chain-enhancer of activated B-cells NLS nuclear localization signal NQO1 NAD(P)H: quinone oxidoreductase 1 NFE2L2 nuclear factor (erythroid-derived 2)-like 2

Page 15: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

xv

PAR poly(ADP-ribose) PARG poly(ADP-ribose) glycohydrolase PARP poly(ADP-ribose) polymerase PAH polycyclic aromatic hydrocarbon PCB polychlorinated biphenyl PCG1α peroxisome proliferator-activated receptor gamma co-activator 1 alpha PEPCK phosphoenolypyruvate carboxykinase RIP140 140 kDa receptor interacting protein siRNA small inhibitory RNA SHP short heterodimer partner SIRT sirtuin SMRT silencing mediator of retinoid and thyroid hormone receptor Sp1 stimulating protein 1 TAD transcriptional activation domain TCDD 2,3,7,8-tetrachlorodibenzo-p-dioxin TIF2 translation initiation factor 2 TiPARP TCDD-inducible poly(ADP-ribose) polymerase qPCR quantitative real-time PCR XAP2 hepatitis B virus X-associated protein 2 WWE tryptophan-tryptophan-glutamate ZAP zinc finger antiviral protein Zn zinc finger

Page 16: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

1

Chapter 1: Introduction 1 Statement of Research Problem 2,3,7,8-tetracholorodibenzo-p-dioxin (TCDD)-inducible poly(ADP-ribose) polymerase

(TiPARP) is a member of the diphtheria-like ADP-ribosyltransferase (ARTD) family of

nicotinamide adenine dinucleotide (NAD+)-catabolizing enzymes, formerly known as the

poly(ADP-ribose) polymerase (PARP) family. ARTD family members are best known for their

ability to catalyze ADP-ribosylation reactions, a process that uses NAD+ as a substrate to add

ADP-ribose onto target proteins. ADP-ribosylation has been implicated in numerous cellular

processes including DNA damage responses, inflammation, apoptosis and transcriptional

regulation. The function of ARTD1 (also known as PARP1), the founding ARTD, is well

established; however, the potential function(s) of the majority of other ARTD members are just

starting to be understood.

Currently, little is known about TiPARP/ARTD14 (will be referred to as TiPARP within this

thesis). As a class 2 ARTD, TiPARP is predicted to be a mono-ADP-ribosyltransferase (MART);

however, this has not been experimentally determined. As its name suggests TiPARP gene

expression is induced by TCDD. The ligand-activated transcription factor, the aryl hydrocarbon

receptor (AHR), is activated by TCDD to regulate its target genes. TiPARP expression is

regulated by numerous AHR ligands in both in vitro and in vivo models, demonstrating it to be

an important AHR regulated gene. Despite being an important AHR target gene, the potential

function of TiPARP in AHR-mediated transcriptional regulation has not been investigated.

While the activation of AHR transcription has been extensively studied, its negative regulation

has not received equal attention and is poorly understood. Proposed mechanisms of the negative

regulation of AHR include the targeted proteasomal degradation of activated-AHR and induction

of repressive factors such as the AHR repressor (AHRR) to down-regulate AHR-mediated

transactivation; however, these models are controversial and not fully understood.

In the present study we set out to characterize TiPARP and determine its role in AHR

transactivation.

Page 17: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

2

2 ADP-ribosylation Reactions

2.1 Historical Overview The presence of poly(ADP-ribose) was first described fifty years ago by Chambon et al. (1963)

who discovered that the addition of nicotinamide adenine dinucleotide (NAD+) to hen liver

extracts stimulated the synthesis of polyadenylic acid which was later identified as poly(ADP-

ribose) (Chambon, et al., 1963). This discovery launched the research of poly(ADP-ribose)

metabolism. The structure of poly(ADP-ribose) (PAR) was later determined by three

independent laboratories (Doly and Mandel, 1967; Reeder, et al., 1967; Sugimura, et al., 1967).

The enzyme responsible for synthesizing PAR was termed poly(ADP-ribose) polymerase or

PARP. The gene encoding PARP (also referred to as poly-ADP-synthase or poly-ADP-

ribosyltransferase) was isolated in the late 1980s (Alkhatib, et al., 1987; Kurosaki, et al., 1987;

Uchida, et al., 1987). Today, this gene is known as PARP1 and for many years PARP1 was

thought to be the only enzyme with ADP-ribosylation activity in mammalian cells. Fifteen years

later, five different genes encoding PARP enzymes were identified indicating PARP1 belongs to

a family of enzymes (Ame, et al., 1999; Johansson, 1999; Kaminker, et al., 2001; Kickhoefer, et

al., 1999; Smith, et al., 1998). To date, eighteen human PARP genes have been cloned or

described based on exhaustive searches of non-redundant protein databases using the PARP1

catalytic domain. However, the PAR synthesizing activity of only a few of these putative PARPs

has been verified (Aguiar, et al., 2005; Aguiar, et al., 2000; Ame, et al., 2004; Ma, et al., 2001;

Otto, et al., 2005; Yu, et al., 2004).

Not long after the structure of PAR was determined, mono-ADP-ribosylation reactions were

discovered during studies of bacterial toxins such as diphtheria, pertussis and cholera toxins

(Corda and Di Girolamo, 2003; Di Girolamo, et al., 2005; Holbourn, et al., 2006). These

enzymes were found to be mono-ADP-ribosyltransferases (MARTs) (Gill, et al., 1969; Honjo, et

al., 1968). Subsequently, extracellular MARTs or e-MARTs (or ecto-MARTs) were identified in

mammalian cells and their enzymatic activities characterized (Haag and Koch-Nolte, 1998;

Okazaki and Moss, 1996; Okazaki and Moss, 1999; Seman, et al., 2004). Several reports have

predicted distinct families of mono-ADP-ribosylating enzymes with no obvious sequence

similarity to e-MARTs (Corda and Di Girolamo, 2002; Corda and Di Girolamo, 2003; Seman, et

al., 2004). Members of the sirtuin family (SIRTs) of NAD+-dependent histone deacetylases were

Page 18: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

3

reported to contain mono-ADP-ribosyltransferase activities and represent a potential family of

intracellular mammalian MARTs (Frye, 1999; Garcia-Salcedo, et al., 2003; Haigis, et al., 2006;

Liszt, et al., 2005; Tanny, et al., 1999). However, more recently twelve PARP family members

have been predicted to be mono-ADP-ribosylating rather than poly-ADP-ribosylating enzymes,

based on subtle differences within their catalytic domains (Hottiger, et al., 2010; Kleine, et al.,

2008). These twelve PARP-like MARTs represent potential candidates of a family of

intracellular MART and only recently are their enzymatic activities being verified. A new

nomenclature for the PARP family has been proposed that reflects their true transferase rather

than polymerase activity, renaming it to ADP-ribosyltransferase diphtheria toxin-like family or

ARTD that takes into account their resemblance to bacterial mono-ADP-ribosylating enzymes

(Hottiger, et al., 2010).

2.2 NAD+ Metabolism In eukaryotes, NAD+ is best known for its essential role as a coenzyme/transmitter molecule in

bioenergetic processes such as cellular metabolism and respiration (Magni, et al., 2004; Ziegler,

2000). The synthesis of adenosine triphosphate (ATP) and balance of redox potential are directly

dependent on cellular NAD+ levels. NAD+ serves as an electron acceptor and its reduced form

NADH as an electron donor in reactions catalyzed by enzymes of the mitochondrial electron

transport chain leading to the generation of ATP through oxidative phosphorylation (Pollak, et

al., 2007; Rich, 2003).

In addition to its vital role in energy metabolism, NAD+ is utilized by ADP-ribosyltransferases to

transfer ADP-ribose to a substrate protein or to itself (Hassa, et al., 2006). In mono-ADP-

ribosylation reactions, NAD+ is hydrolyzed by mono-ADP-ribosyltransferases and the ADP-

ribose moiety is transferred to an acceptor protein (the enzyme itself or a target protein) and the

vitamin nicotinamide is released (Figure 1). Poly-ADP-ribosylation is similar to mono-ADP-

ribosylation but also includes an additional elongation reaction where the ADP-ribose units are

linked by glycosidic ribose (1’-2’) ribose bonds creating long, irregular, branched polymers of

ADP-ribose (Desmarais, et al., 1991; Kawaichi, et al., 1981; Mendoza-Alvarez and Alvarez-

Gonzalez, 2004; Miwa, et al., 1981) (Figure 2). Polymers can reach 200-400 ADP-ribose units

in vitro and in vivo and long polymers are branched with a frequency of approximately one

branch point per linear section of 20-50 ADP-ribose units (Alvarez-Gonzalez and Jacobson,

Page 19: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

4

1987; Juarez-Salinas, et al., 1982; Juarez-Salinas, et al., 1983; Kanai, et al., 1982; Miwa, et al.,

1981). Poly-ADP-ribosylation is proposed to be the most important factor affecting the

maintenance of the NAD+ pool in cells (Elliott and Rechsteiner, 1975; Rechsteiner, et al., 1976;

Williams, et al., 1985). The catabolism of NAD+ in mammalian cells occurs primarily through

poly-ADP-ribosylation reactions (Berger, et al., 2004). Activation of poly-ADP-ribosylation

reactions, coinciding with increased poly-ADP-ribose polymer generation depletes intracellular

NAD+ levels to 10-20% of their normal levels within 5-15 min upon initiation of stimulus

(Goodwin, et al., 1978; Skidmore, et al., 1979). Cellular NAD+ depletion consequently results in

ATP depletion, since NAD+ is an essential coenzyme/transmitter for the generation of ATP

through oxidative phosphorylation (Berger, 1985). Because NAD+ connects bioenergetics to

ADP-ribosylation reactions and cellular NAD+ levels have important physiological consequences

in the regulation of multiple cellular processes, the control of NAD+-dependent enzymes must be

tightly regulated.

Page 20: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

5

Figure 1. Schematic representation of the mono-ADP-ribosylation reaction.

The chemical structure of the mono-ADP-ribosylation reaction to an acceptor protein is shown.

Mono-ADP-ribosyltransferase (MART) catalyzes the addition of ADP-ribose to an acceptor

protein using nicotinamide adenine dinucleotide (NAD+) as a precursor resulting in the release of

nicotinamide. ADP-ribosylhydrolases (ARH) remove the ADP-ribose moiety from an acceptor

protein.

Page 21: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

6

Figure 2. Metabolism of poly(ADP-ribose).

PARPs hydrolyse NAD+ and catalyze the successive addition of ADP-ribose units to acceptor

proteins. Poly-ADP-ribose size and complexity is indicated by the x and y labels that represent

values from 0 to >200. Poly-ADP-ribose-glycohydrolase (PARG) and ADP-ribosylhydrolase 3

(ARH3) can both hydrolyze poly-ADP-ribose at the indicated positions. ADE; adenine, NAM;

nicotinamide, Rib; ribose. Modified from (Hakme, et al., 2008).

Page 22: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

7

2.3 Mono-ADP-ribosylation Mono-ADP-ribosylation of proteins is a phylogentically ancient and reversible posttranslational

modification. This modification was originally identified as an important aspect of bacterial

physiopathology, catalyzed by several toxins including diphtheria, pertussis and cholera toxins

(Corda and Di Girolamo, 2003; Di Girolamo, et al., 2005; Holbourn, et al., 2006). So far, six

subclasses of bacterial MARTs have been identified based on amino acid sequence identity

(Hassa, et al., 2006). The amino acid residues of host cell acceptor proteins modified by bacterial

MARTs include: arginine, asparagine, glutamate, aspartate, cysteine and modified histidine

(diphthamide) (Corda and Di Girolamo, 2002; Corda and Di Girolamo, 2003; Okazaki and Moss,

1996; Okazaki and Moss, 1999). Diphtheria toxin can be used as an example of the specificity of

mono-ADP-ribosylation and the important physiological consequences of a single residue

modification by a MART. Diphtheria toxin, is an exotoxin secreted by the bacterium

Corynebacterium diphtheria, which mono-ADP-ribosylates host cell elongation factor 2 (eEF2)

at diphthamide residue 699 blocking the interaction of eEF2 with other factors involved in

translation and effectively preventing host cell protein synthesis (Collier, 2001).

Mono-ADP-ribosylation reactions in multicellular eukaryotes were later discovered. Endogenous

MART activities that modify arginine, glutamate, cysteine, phosphoserine and potentially

aspartate and asparagine on acceptor proteins have been detected in mammalian cells (Corda and

Di Girolamo, 2003; Okazaki and Moss, 1996; Ord and Stocken, 1977; Seman, et al., 2004; Smith

and Stocken, 1975). Intracellular mono-ADP-ribosylation in mammalian cells has been

suggested to play important roles in processes such as: regulation of intracellular signalling

cascades, transcriptional regulation, unfolded protein response, DNA repair, insulin secretion,

immunity and cellular differentiation and proliferation (Corda and Di Girolamo, 2003; Di

Girolamo, et al., 2005; Feijs, et al., 2013; Koch-Nolte, et al., 2008).

2.3.1 Mono-ADP-ribosyltransferase families

2.3.1.1 ecto-MARTs (e-MARTs)

Human and mouse e-MARTs that have been identified are glycosylphosphatidylinositol-

anchored surface proteins or secretory proteins (Glowacki, et al., 2002; Koch-Nolte, et al., 2008;

Okazaki and Moss, 1999). E-MARTs transfer ADP-ribose to target proteins within the

Page 23: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

8

extracellular compartment and are involved in cell communication (Koch-Nolte, et al., 2008).

Five e-MARTs (e-MART-1 to 5) have been identified in mammalian cells. Humans only express

four e-MARTs due to a defective e-MART-2 gene while mice express six e-MARTs as a result of

a duplication of the e-MART-2 gene (Seman, et al., 2004). e-MART-1, -2a, -2b and -5 transfer

ADP-ribose to arginine residues of extracellular target proteins on cell surfaces, or circulating in

body fluids (Corda and Di Girolamo, 2003; Glowacki, et al., 2002; Seman, et al., 2004). E-

MART-3 is expressed on the surface of unstimulated human monocytes whereas e-MART-4 is

only expressed on the cell surface once monocytes are stimulated by lipopolysaccharide

(Grahnert, et al., 2002). Cell surface mono-ADP-ribosylated proteins on human monocytes are

covalently modified at cysteine residues, suggesting that e-MART-3 and -4 may be cysteine-

specific e-MARTs (Grahnert, et al., 2002).

2.3.1.2 Sirtuins

Silent information regulator SIR2-like proteins (sirtuins or SIRTs) are a conserved family of

NAD+-dependent protein deacetylases (SIRT1 to 7) present in eukaryotes, prokaryotes and

archaea (Blander and Guarente, 2004; Denu, 2005; Grubisha, et al., 2005). The SIRT family

regulates a broad range of cellular processes including development, metabolism, DNA repair,

DNA recombination, cellular differentiation, transcriptional regulation, apoptosis and lifespan

(Blander and Guarente, 2004; Denu, 2005; Grubisha, et al., 2005). Most mammalian SIRTs are

reported to exhibit in vivo and in vitro histone deacetylation activity (Blander and Guarente,

2004; Denu, 2005; Michishita, et al., 2005). Of the seven mammalian SIRTs, SIRTs 1, 2, 3 and 5

have been shown to have NAD+-dependent deacetylase activities in vitro (Blander and Guarente,

2004; Denu, 2005; Michishita, et al., 2005). SIRTs also have non-histone substrates and not all

mammalian SIRTs localize to the nucleus.

The NAD+-dependent deacetylation reaction is an unusual mechanism performed by SIRTs

where NAD+ is hydrolyzed into nicotinamide and ADP-ribose (Abdellatif, 2012). The ADP-

ribose moiety is essential for the deacetylation reaction and serves as the acceptor of the acetyl

group removed from the target protein forming the intermediates 2’-O-AADP-ribose and 3’-O-

AADP-ribose (Smith and Denu, 2006; Zhao, et al., 2003; Zhao, et al., 2004). Several reports

have suggested mammalian SIRT family members to contain mono-ADP-ribosyltransferase

activities. SIRTs 1, 2 and 6 were reported to transfer mono-ADP-ribose to bovine serum albumin

Page 24: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

9

and histones in vitro (Frye, 1999; Mostoslavsky, et al., 2006; Tanny, et al., 1999). SIRT6 has

auto-mono-ADP-ribosylation (or self ribosylation) but does not possess deacetylase activity

indicating enzymatic mechanisms of each SIRT may differ (Liszt, et al., 2005). Currently, the

potential mono-ADP-ribosyltransferase activities of the different SIRT family members has not

been fully examined and further investigation is required to determine if indeed the SIRTs are a

class of bona fide intracellular mono-ADP-ribosyltransferases (Denu, 2005).

2.3.2 ADP-ribosylhydrolases

ADP-ribosylhydrolases (ARH) reverse mono-ADP-ribosylation reactions by catalyzing the

removal of ADP-ribose from modified proteins (Moss, et al., 1992). The best characterized

ADP-ribosylhydrolase is ADP-ribosylhydrolase 1 (ARH1), which specifically hydrolyzes ADP-

ribose-arginine bonds, leading to the release of free mono-ADP-ribose and regeneration of the

guanidino group of arginine (Moss, et al., 1986; Takada, et al., 1993). In mammalian cells, most

ARH1 activities are cytosolic, although some are also located on the cell surface (Hassa, et al.,

2006). Besides ARH1, additional unidentified proteins exhibiting ADP-ribosyl lyase activities

have been reported, which are thought to function on glutamate, lysine or cysteine residues,

releasing a deoxy form of ADP-ribose (Oka, et al., 1984; Okayama, et al., 1978; Tanuma and

Endo, 1990). In silico screening revealed two ARH-like genes, ARH2 and ARH3 homologous to

the human ARH1 gene (Glowacki, et al., 2002). ARH2 does not appear to possess hydrolase

activity, but ARH3 was shown to hydrolyze poly(ADP-ribose) but not mono-ADP-ribose (see

section 2.4.2.2)(Oka, et al., 2006).

Recently, macrodomain-containing proteins were discovered to catalyze the hydrolysis of mono-

ADP-ribose from modified proteins (Jankevicius, et al., 2013; Rosenthal, et al., 2013; Sharifi, et

al., 2013). Macrodomains are a family of evolutionarily conserved proteins that bind mono- or

poly-ADP-ribose, polyadenylation or O-acetyl-ADP-ribose (Han, et al., 2011; Karras, et al.,

2005; Kim, et al., 2012; Neuvonen and Ahola, 2009). Human proteins MacroD1, MacroD2 and

C6orf130 hydrolyzed mono-ADP-ribose from modified proteins at acidic residues (Jankevicius,

et al., 2013; Rosenthal, et al., 2013). MacroD-like hydrolases inactivated MART target proteins

by removing mono-ADP-ribose and rendering them inactive and thus form the functional

antagonists of intracellular MARTs (Rosenthal, et al., 2013). These findings define

Page 25: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

10

macrodomain-containing proteins as a novel group of specific mono-ADP-ribosylhydrolases and

establish mono-ADP-ribosylation as a fully reversible posttranslational modification.

2.4 Poly-ADP-ribosylation Poly-ADP-ribosylation was discovered in multicellular eukaryotes and appears to be less widely

used compared to mono-ADP-ribosylation (Hassa, et al., 2006). Poly-ADP-ribosylation is a

rapid and reversible posttranslational modification in which linear and branched sequences of

ADP-ribose units are covalently attached onto acceptor proteins (Otto, et al., 2005). DNA strand

breakage was long considered to be the main trigger of poly-ADP-ribosylation of proteins;

however, more recent studies suggest poly-ADP-ribosylation is important in a much broader

spectrum of cellular and molecular functions including regulation of chromatin structure,

transcriptional regulation, stress response, inflammation, cell proliferation and differentiation,

apoptosis and maintenance of genome stability (Hakme, et al., 2008; Kraus, 2008; Krishnakumar

and Kraus, 2010b; Luo and Kraus, 2012). Poly-ADP-ribosylation reactions are mediated by

some members of the ARTD family that catalyze the transfer of ADP-ribose units from NAD+ to

form a polymer of ADP-ribose units onto glutamate, aspartate and lysine residues of acceptor

proteins (Hakme, et al., 2008). ARTD1 (formerly known as PARP1), the founding and best-

studied ARTD family member, was first thought to be the only enzyme in mammalian cells to

generate poly(ADP-ribose) polymers. However in recent years, in silico studies have identified

seventeen other proteins that share a catalytic PARP domain with PARP1 and that may contain

poly-ADP-ribosylation activity (Ame, et al., 2004; Otto, et al., 2005).

2.4.1 ADP-ribosyltransferase diphtheria toxin-like (ARTD) family

2.4.1.1 Structural features and nomenclature

ARTDs are a family of ADP-ribosyltransferases that consists of eighteen members all encoded

by different genes. All ARTD family members contain a conserved catalytic domain

homologous to the ARTD1 catalytic domain (Ame, et al., 2004). Besides the catalytic domain, a

wide spectrum of modular domains is present in the different ARTD family members (Figure 3),

suggesting that these members are involved diverse physiological responses (Kleine, et al.,

2008). The catalytic domain is located at the C-terminus of the protein (except in ARTD4) other

modular domains are involved in DNA or RNA binding, protein-protein interactions or cell

Page 26: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

11

signalling (Schreiber, et al., 2006). The catalytic domain binds NAD+ via a protein fold termed

the ‘PARP signature’ that shares homology with bacterial mono-ADP-ribosylating exotoxins,

such as diphtheria toxin (Gibson and Kraus, 2012). Based on both structural homology of the

diphtheria toxin in complex with NAD+ and in silico characterization of ARTDs, histidine and

tyrosine (H862 and Y896 in ARTD1) are critical for the proper positioning the ribose moiety of

NAD+ for ADP-ribosylation (Bell and Eisenberg, 1996; Otto, et al., 2005). In poly-ADP-

ribosylating ARTDs, a catalytic glutamate (E988 in ARTD1) is essential for the catalysis of

additional ADP-ribose units or poly-ADP-ribose chain elongation onto acceptor residues

(Marsischky, et al., 1995; Rolli, et al., 1997; Ruf, et al., 1998). These three residues, histidine-

tyrosine-glutamate, make up the catalytic ‘HYE’ triad that appears in all bona fide poly-ADP-

ribosylating ARTDs (Gibson and Kraus, 2012). The absence of the catalytic glutamate of the

ARTD catalytic domain restricts activity to mono-ADP-ribosylation (Kleine, et al., 2008;

Marsischky, et al., 1995; Rolli, et al., 1997; Ruf, et al., 1998). Other structural features such as

the loop length between β-4 and β-5 sheets of the catalytic core have been correlated with lack of

poly-ADP-ribosylation activity (Han and Tainer, 2002; Kleine, et al., 2008; Ruf, et al., 1998;

Sun, et al., 2004). Because the HYE triad of ARTDs is most similar to that of diphtheria toxin,

PARPs were renamed diphtheria-like ADP-ribosyltransferases, abbreviated ARTDs, which better

reflects their transferase rather than polymerase activity (Hottiger, et al., 2010). Of the eighteen

ARTD family members only six contain the catalytic glutamate in the HYE triad, while in the

remaining twelve the catalytic glutamate is replaced by isoleucine, leucine, threonine, valine or

tyrosine residues, suggesting that they will exhibit mono- rather than poly-ADP-

ribosyltransferase activity (Otto, et al., 2005). ARTD9 and ARTD13 lack both the catalytic

glutamate as well as the histidine of the HYE triad and are predicted to be catalytically inactive

(Otto, et al., 2005). Thus ARTDs can be divided into three classes based on their catalytic

activities (Table 1). Class 1 ARTDs are poly-ADP-ribosyltransferases, class 2 are mono-ADP-

ribosyltransferases and class 3 are catalytically inactive.

Page 27: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

12

Table 1. Organization and enzymatic activities of mammalian ARTD family.

Class Transferase Name

PARP Name Alternative Name

Triad Motif Enzymatic Activity*

1 ARTD1 PARP1 HYE P & B 1 ARTD2 PARP2 HYE P & B 1 ARTD3 PARP3 HYE M (P predicted) 1 ARTD4 PARP4 VPARP HYE P 1 ARTD5 PARP5a Tankyase 1 HYE P & O 1 ARTD6 PARP5b Tankyase 2 HYE P & O 2 ARTD7 PARP15 BAL3 HYL M 2 ARTD8 PARP14 BAL2 HYL M 2 ARTD10 PARP10 HYI M 2 ARTD11 PARP11 HYI M (predicted) 2 ARTD12 PARP12 ZC3HDC1 HYI M 2 ARTD14 PARP7 TiPARP HYI M (predicted) 2 ARTD15 PARP16 HYI M 2 ARTD16 PARP8 HYI M (predicted) 2 ARTD17 PARP6 HYY M (predicted) 2 ARTD18 TPT1 HHV M (predicted) 3 ARTD9 PARP9 BAL1 QYT I 3 ARTD13 PARP13 ZAP YYV I

*Known or predicted enzymatic activity: mono- (M), oligo- (O) or poly-ADP-ribosylation (P) or

branching (B) or inactive (I). (Adapted from (Gibson and Kraus, 2012))

Page 28: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

13

Figure 3. Domain architecture of the 18 human ARTD family members.

Schematic comparison of the domain architecture of the human ARTD (PARP) family (adapted

from (Hottiger, et al., 2010)). The following domains are indicated: The ART domain is the

catalytic core required for ART activity. Within each ART domain, the region that is

homologous to the PARP signature (residues 859–908 of ARTD1) as well as the equivalent of

the ARTD1 catalytic E988 is shaded. The PARP regulatory domain (PRD) might be involved in

regulation of the ADP-ribose branching activity. The WGR domain named after a conserved

central motif (W-G-R) is also found in a variety of polyA polymerases and in proteins of

unknown function. The BRCT domain (BRCA1 carboxy-terminal domain) is found within many

DNA damage repair and cell cycle checkpoint proteins. The sterile alpha motif (SAM), a domain

found within many proteins, can mediate homo- or hetero-dimerization. The ankyrin repeat

domains (ARD) mediate protein–protein interactions. The vault protein interalphatrypsin (VIT)

Page 29: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

14

and von Willebrand type A (vWA) domains are conserved domains found in all inter-alpha-

trypsin inhibitor family members. Both of these domains are presumed to mediate protein–

protein interactions. The WWE domain is named after three conserved residues (W-W-E), and is

predicted to mediate specific protein–protein interactions. The Macro domains can serve as

ADPr or O-acetyl-ADP-ribose binding module. ZFD: zinc finger domains. SAP:

SAF/Acinus/PIAS-DNA-binding domain, MVP-ID: Major-vault particle interaction domain,

NLS: nuclear localization signal. CLS: centriole-localization signal. HPS: histidine-proline-

serine region. RRM is an RNA-binding/recognition motif. UIM: ubiquitin interaction motif.

MVP-ID: M-vault particle interaction domain. TPH: Ti-PARP homologous domain. GRD:

glycine-rich domain. TMD: transmembrane domain.

2.4.1.1.1 Class 1 ARTDs

ARTD1/PARP1 is involved in a number of cellular activities including DNA repair, apoptosis,

chromatin remodeling and transcriptional regulation (D'Amours, et al., 1999; Kim, et al., 2005;

Kraus and Lis, 2003; Schreiber, et al., 2006; Tulin and Spradling, 2003). ARTD1 accounts for

approximately 85-90% of mammalian cell poly-ADP-ribosylation activity (Shieh, et al., 1998).

ARTD1 is a nuclear enzyme that is characterized mainly by three modular domains: an N-

terminal DNA binding domain consisting of three zinc finger domains, an automodification

domain and a C-terminal catalytic ADP-ribosyltransferase domain containing the conserved

HYE motif (Figure 4)(D'Amours, et al., 1999). The DNA binding domain plays a critical role in

the recognition of DNA strand aberrations and concurrent activation of ARTD1 (Benjamin and

Gill, 1980; D'Amours, et al., 1999; Ikejima, et al., 1990). The automodification domain is

comprised of a BRCA1-carboxy terminus-like module that mediates several protein-ARTD1 and

DNA-ARTD1 interactions and is also auto-ADP-ribosylation domain (de Murcia and Menissier

de Murcia, 1994; de Murcia, et al., 1994; Mazen, et al., 1989; Schreiber, et al., 1995; Smith,

2001). ARTD1 is best known for its function in the base excision repair pathway during DNA

damage (D'Amours, et al., 1999; de Murcia and Menissier de Murcia, 1994). ARTD1 is a DNA

nick-sensor that is activated by single- and double-strand DNA breaks, DNA hairpins, cruciform

DNA and stable DNA loops (Durkacz, et al., 1980; Lonskaya, et al., 2005; Purnell, et al., 1980).

DNA binding by ARTD1 leads to a conformational change of the enzyme, followed by extensive

auto-ADP-ribosylation and hetero-ADP-ribosylation of histones H1 and H2B (Langelier, et al.,

Page 30: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

15

2012; Poirier, et al., 1982). ARTD1-dependent ADP-ribosylation has two effects. One, the

poly(ADP-ribose) chains displace histones leading to decondensation of chromatin (Poirier, et

al., 1982; Zahradka and Ebisuzaki, 1982) and two, the poly(ADP-ribose) chains function as a

binding platform for DNA repair enzymes (Dantzer, et al., 2000; El-Khamisy, et al., 2003;

Leppard, et al., 2003). The accumulation of negative charges added to ARTD1 through auto-

ADP-ribosylation leads to an electrostatic repulsion from the DNA and subsequent loss of

activity (D'Amours, et al., 1999; Schreiber, et al., 2006). Poly(ADP-ribose) chains are then

degraded through hydrolysis by poly(ADP-ribose) glycohydrolase (PARG) (D'Amours, et al.,

1999; Zahradka and Ebisuzaki, 1982). Apart from its role in DNA repair ARTD1 is also

involved in apoptosis, inflammation and transcriptional regulation (Altmeyer, et al., 2010; Kraus,

2008; Kraus and Lis, 2003; Liaudet, et al., 2002).

Figure 4. ARTD1 domain architecture.

ARTD1 uses NAD+ to form polymers of ADP-ribose onto acceptor proteins or to itself

(automodification). Major domains include: DNA-binding domain, a nuclear localization signal

(NLS), an automodification domain containing a BRCT (BRCA1-carboxy terminus-like) motif

and a catalytic ADP-ribosyltransferase domain containing the conserved ‘PARP signature’

NAD+ fold. Adapted from (David, et al., 2009).

ARTD2/PARP2 was discovered as a result of the residual DNA-dependent PARP activity in

embryonic fibroblast derived from Artd1-deficient mice (Ame, et al., 1999; Shieh, et al., 1998).

Page 31: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

16

The ARTD2 catalytic domain displays the highest sequence homology (69% similarity) to that of

ARTD1 (Ame, et al., 2004). Similar to ARTD1, ARTD2 is a nuclear protein that is activated by

DNA nicks leading to its auto-ADP-ribosylation and synthesis of long branched chains of

poly(ADP-ribose) (Ame, et al., 1999; Schreiber, et al., 2002). Both Artd1- and Artd2-deficient

mice display sensitivity to DNA damaging agents associated with defects in DNA damage

processing and cell cycle progression but to different extents. Artd1-deficient mice and cell lines

are more sensitive to high-dose ionizing radiation and alkylating agents, whereas Artd2-deficient

mice display hyper-radiosensitivity to low-dose irradiation, demonstrating that ARTD1 and

ARTD2 functions are complementary but do not fully overlap (Chalmers, et al., 2004). Double

mutant Artd1- and Artd2-deficient mice are embryonically lethal highlighting a critical role in

early embryogenesis and underlying partial functional redundancy in maintaining genome

integrity (Menissier de Murcia, et al., 2003; Schreiber, et al., 2002). In addition to its role in

maintenance of genomic stability and DNA repair, ARTD2 has been implicated in various

differentiation processes, including spermatogenesis, adipogenesis and immune cell development

(Bai, et al., 2007; Dantzer, et al., 2006; Robert, et al., 2009; Yelamos, et al., 2006).

ARTD3/PARP3 was initially identified as a core component of the centrosome preferentially

located at the daughter centriole throughout the cell cycle (Augustin, et al., 2003). ARTD3

overexpression interfered with the G1/S phase cell cycle progression and it was described to

interact with ARTD1 at the centrosome (Augustin, et al., 2003). A later study disputed the

centrosomal localization of ARTD3 and suggested that ARTD3 localized to the nucleus and

associated with polycomb group proteins involved in gene silencing and DNA repair networks

including DNA protein kinases, DNA ligase III and IV, Ku70 and Ku80 and ARTD1 (Boehler, et

al., 2011; Rouleau, et al., 2007). ARTD3-depleted cells and Artd1-/- and Artd3-/- double knockout

mice showed increased hypersensitivity to ionizing radiation suggesting a functional synergy of

ARTD3 and ARTD1 in the cellular response to DNA damage (Boehler, et al., 2011). These

studies suggested that ARTD3 is an important partner in the maintenance of genomic stability.

Auto-ADP-ribosylation and hetero-ADP-ribosylation activities of ARTD3 were initially

described as mono-ADP-ribosyltransferase activity (Loseva, et al., 2010); however, a later report

has described ARTD3 to possess poly-ADP-ribosyltransferase activity (Boehler, et al., 2011).

ARTD3 was described to poly-ADP-ribosylate the mitotic factor NuMA directly and indirectly

through ARTD5, suggesting that ARTD3 was required for mitotic spindle integrity during

Page 32: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

17

mitosis (Boehler, et al., 2011). Collectively, these reports implicate ARTD3 in the maintenance

of genomic integrity, mitotic spindle integrity and transcriptional repression.

ARTD4/PARP4/VPARP is the largest member of the ARTD family and was originally

identified as a component of mammalian cytoplasmic ribonucleoprotein complexes called vault

particles that have been proposed to be involved in multidrug resistance of human tumours and

to function in intracellular transport (Kickhoefer, et al., 1999). ARTD4 associates with two

essential proteins of the vault particle, major vault protein (MVP) and telomerase-associated

protein (TEP1) (Kickhoefer, et al., 1999). ARTD4 is also present in the nucleus and at the

mitotic spindle suggesting that it may play multiple roles not yet identified (Kickhoefer, et al.,

1999). The structure of ARTD4 is unusual in that it is the only ARTD member to have its

catalytic domain located at the N-terminal portion of the protein. Despite this unique feature,

ARTD4 is catalytically active and poly-ADP-ribosylates MVP and as well as itself (Kickhoefer,

et al., 1999).

ARTD5/PARP5a/Tankyrase 1 and ARTD6/PARP5b/Tankyrase 2 are two closely related

ARTD family members that share 83% sequence identity (Kaminker, et al., 2001; Kuimov, et

al., 2001; Smith, et al., 1998). ARTD5 and ARTD6 have been implicated in a diverse range of

functions including telomere maintenance, WNT signalling, mitosis and mediation of insulin

stimulated glucose uptake (Chi and Lodish, 2000; Cook, et al., 2002; Dynek and Smith, 2004;

Guo, et al., 2012; Hsiao and Smith, 2008; Huang, et al., 2009; Smith and de Lange, 2000; Smith,

et al., 1998). ARTD5 was first discovered as a factor that regulated telomere length by binding

the negative regulator of telomere length telomeric repeat binding factor 1 (TRF1) (Smith, et al.,

1998). ARTD5 was originally named tankyrase 1 due to its interaction with TRF1, the presence

of 24 consecutive ankyrin repeats which make up the major portion of the enzyme, and the

presence of ARTD catalytic domain (Smith, et al., 1998). ARTD5 catalyzed auto-poly-ADP-

ribosylation and poly-ADP-ribosylation of TRF1, implicating ARTD5 in telomere elongation

(Smith, et al., 1998). A careful analysis of ARTD5 auto-ADP-ribosylation revealed that ARTD5

synthesized ADP-ribose polymers with an average length of 20 ADP-ribose units but polymers

lacked branching (Rippmann, et al., 2002). ARTD6 was also reported to associate with and poly-

ADP-ribosylate TRF1, indicating a potential redundant role of ARTD5 and ARTD6 in telomere

regulation (Cook, et al., 2002; Kaminker, et al., 2001). ARTD6 also associates with ARTD5 and

both enzymes share most of their protein partners including insulin-responsive aminopeptidase

Page 33: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

18

(IRAP), NuMA and 182 kDa tankyrase-binding protein (TAB182) (Kaminker, et al., 2001;

Sbodio and Chi, 2002; Sbodio, et al., 2002; Seimiya and Smith, 2002). Artd5- and Artd6-

deficient mice do not display observed defects in development; however, the Artd5-/- and Artd6-/-

double knockout was embryonical lethal (Chiang, et al., 2008).

2.4.1.1.2 Class 2 ARTDs

ARTD7/PARP15/BAL3 and ARTD8/PARP14/BAL2 were originally identified as two genes

closely related to ARTD9/PARP9/BAL1 (B-aggressive lymphoma 1, see section 2.4.1.1.3)

(Aguiar, et al., 2005; Aguiar, et al., 2000). The encoded proteins both contain N-terminal

macrodomains and a C-terminal PARP catalytic domain (Aguiar, et al., 2005). Macrodomains

are protein domains known to bind mono- and poly-ADP-ribose (Forst, et al., 2013; Han and

Tainer, 2002; Kleine and Luscher, 2009; Welsby, et al., 2012). Recently, macrodomains 2 and 3

of ARTD8 were reported to recognize and read mono-ADP-ribosylated ARTD10 and substrates

of ARTD10 (Forst, et al., 2013). Both ARTD7 and ARTD8 demonstrated auto-mono-ADP-

ribosylation activity (Aguiar, et al., 2005). ARTD7 was reported to have a transcriptionally

repressive function through its N-terminal macrodomains and its auto-ADP-ribosylation activity

was suggested to counteract the repressive effect of the macrodomains (Aguiar, et al., 2005).

Whereas ARTD7 remains poorly characterized, ARTD8 is better understood and has been

implicated in STAT6 (Signal Transducer and Activator of Transcription 6)-dependent

transcriptional control and cytokine-regulated control of cellular metabolism (Cho, et al., 2011;

Goenka, et al., 2007). ARTD8 potentiated interleukin 4 (IL4)-induced STAT6 transactivation via

its macrodomains and catalytic activity (Goenka and Boothby, 2006; Goenka, et al., 2007).

ARTD8 was also recently identified as a downstream effector of the Jun N-terminal kinase 2

(JNK2)-dependent pro-survival signal by binding to and inhibiting JNK1 pro-apoptotic activity

promoting the survival of myeloma cells (Barbarulo, et al., 2012).

ARTD10/PARP10 is the founding member of the class 2 ARTDs and was discovered through in

silico screening ARTD family members (Ame, et al., 2004). The domain structure of ARTD10,

with the exception of the catalytic domain is unique from the other ARTDs. The N-terminus

contains a putative RNA recognition motif (RRM) followed by a glycine-rich region, which is

present in other RNA-binding proteins (Burd and Dreyfuss, 1994). The C-terminus contains a

functional nuclear export signal (NES) within its glutamate-rich region and two functional

Page 34: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

19

ubiquitin-interaction motifs (UIMs) (Yu, et al., 2005). ARTD10 is a MART that auto-ADP-

ribosylates itself as well as each of the four core histones (Kleine, et al., 2008; Yu, et al., 2005).

ARTD10 was also reported to be an interacting partner of the oncoprotein MYC (Yu, et al.,

2005). MYC/HA-RAS- and E1A/HA-RAS-dependent transformation of rat embryo fibroblasts

was inhibited by ARTD10 (Yu, et al., 2005). Inhibition was independent of ARTD10 catalytic

activity and neither MYC nor its heterodimerization partner MAX were ADP-ribosylated by

ARTD10 (Yu, et al., 2005). Recently, ARTD10 has also been reported to be a regulator of the

NF-κB pathway by mono-ADP-ribosylating NF-κB essential modulator, reducing its poly-

ubiquitination and activation of NF-κB (Verheugd, et al., 2013). Unlike the regulation of MYC,

the regulation of NF-κB was dependent on ARTD10 catalytic activity and the UIMs (Verheugd,

et al., 2013).

ARTD12/PARP12/ZC3HDC1 has been implicated in inflammatory processes in relation to

various pathogens. During infection by the alphavirus, Venezuelan equine encephalitis virus

(VEEV), the long isoform (containing the catalytic domain) of ARTD12 was up-regulated

(Atasheva, et al., 2012). ARTD12 exhibited inhibitor effects on replication of VEEV as well as

other alphaviruses and RNA viruses (Atasheva, et al., 2012). Interferon stimulation up-regulated

ARTD12 gene expression to counteract infections which was suggested as being a cellular

defence against invading viral pathogens, although this was not mechanistically determined

(Schoggins, et al., 2011). In addition to its role in viral pathogen infection inhibition, ARTD12

has been implicated in post-transcriptional gene expression (Leung, et al., 2011). In that study,

overexpressed ARTD12 assembled in cytoplasmic mRNA-protein complexes called stress

granules and alleviated microRNA (miRNA)-mediated repression (Leung, et al., 2011).

ARTD12 exhibited auto-mono-ADP-ribosylation activity and associated with both long and

short isoforms of ARTD13 (short isoform is missing catalytic domain), ARTD5 and ARTD7

within stress granules (Leung, et al., 2011).

ARTD14/PARP7/TiPARP (will be referred to as TiPARP) was first identified as a 2,3,7,8-

tetrachlorodibenzo-p-dioxin (TCDD or dioxin)-induced gene under the control of TCDD-

activated aryl hydrocarbon receptor (AHR) (Ma, et al., 2001). TiPARP is broadly expressed in

murine tissues, suggesting that TiPARP may be involved in diverse physiological functions (Ma,

et al., 2001). In its initial study, TiPARP exhibited enzymatic ADP-ribosylation capacity but it

Page 35: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

20

was not determined whether it undergoes auto- or hetero-ADP-ribosylation and whether this was

mono-ADP-ribosylation activity (Ma, et al., 2001). A more recent report has identified TiPARP

as a mediator the TCDD-induced suppression of hepatic glucose production (Diani-Moore, et al.,

2010). This study proposed the TCDD-induced expression of TiPARP contributed to NAD+

depletion by TiPARP catalytic activity, which decreased SIRT1 activity leading to decreased

gluconeogenic gene expression (Diani-Moore, et al., 2010). TCDD-induced TiPARP gene

expression and NAD+ depletion has also been implicated in the deactivation of SIRT3 which

leads to decreased superoxide dismutase 2 activity resulting in heightened sensitivity to non-

alcoholic steatohepatitis (He, et al., 2013). Both proposed mechanisms of TiPARP-mediated

NAD+ depletion were derived from observation that TCDD induces TiPARP gene expression

and decreases NAD+ levels. The enzymatic activity of TiPARP was not properly investigated in

these studies and neither study confirmed the mono-ADP-ribosyltransferase activity of TiPARP

and/or if this was required for the observed outcomes.

ARTD15/PARP16 was the first ARTD member found to associate with the endoplasmic

reticulum (ER) (Di Paola, et al., 2012). ARTD15 was the only ARTD member to contain a

predicted C-terminal transmembrane domain and was found to be an ER-tailed anchored enzyme

that associated with the nuclear transport factor karyopherin-β1/importin-β1 (Kapβ1) at the ER

and the nuclear envelope (Di Paola, et al., 2012). Kapβ1 was selectively mono-ADP-ribosylated

by ARTD15 and this modification was hypothesized to control nuclear transport (Di Paola, et al.,

2012). ARTD15 also demonstrated auto-mono-ADP-ribosylation that could be inhibited with

canonical PARP inhibitors (Di Paola, et al., 2012; Karlberg, et al., 2012). Mono-ADP-

ribosylation was reported to not occur at arginine, glutamate or cysteine residues and could not

be reverted by ARH1 or ARH3 (Di Paola, et al., 2012). A recent report has described that

ARTD15 was required for activation of the ER stress response kinases PERK and IREα during

the unfolded protein response (UPR), an ER stress signal that ultimately leads to apoptosis (Hetz,

2012; Jwa and Chang, 2012). During ER stress, ARTD15 was found to auto-ADP-ribosylate

itself and hetero-ADP-ribosylate PERK and IREα (Jwa and Chang, 2012). ADP-ribosylation of

PERK and IREα increased their kinase activities and the endoribonuclease activity of IREα,

which appeared necessary for proper execution of the UPR (Jwa and Chang, 2012).

Page 36: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

21

ARTD17/PARP6 was recently described as a negative regulator of cell proliferation and

proposed tumour suppressor involved in colorectal cancer development (Tuncel, et al., 2012).

ARTD17 overexpression was reported to arrest cells in S-phase and was dependent on the

presence of the catalytic domain (Tuncel, et al., 2012). The catalytic activity of ARTD17 has not

been evaluated and its potential activity in other cellular functions has not been fully determined.

ARTD11/PARP11, ARTD16/PARP8 and ARTD18/TPT1 (the single member of the NAD+-

dependent tRNA 2’-phosphotransferase family) have no known domains outside of their

catalytic domains (with the exception of the single WWE domain of ARTD11) and their

functions have not been determined.

2.4.1.1.3 Class 3 ARTDs

ARTD9/PARP9/BAL1 is a nucleocytoplasmic shuttling protein that has been identified as a

risk-related gene product in aggressive diffuse large B-cell lymphoma (DLBCL) (Aguiar, et al.,

2000; Juszczynski, et al., 2006). ARTD9 contains two prototypical macrodomains within the N-

terminus, which can bind mono- and poly-ADP-ribose (Karras, et al., 2005). ARTD9 was

described to possess transcriptional repressive activity that was dependent on interaction through

its macrodomains but independent of its catalytic activity (Aguiar, et al., 2005). Overexpression

of ARTD9 promoted lymphocyte migration, indicating a tumour-promoting role in high-risk

DLBCL (Aguiar, et al., 2000). Promotion of lymphocyte migration by ARTD9 has been

suggested through modulation of interferon gamma (IFNγ) signalling-related gene expression

(Juszczynski, et al., 2006). ARTD9 was identified as a novel co-repressor of transcription of

interferon response factor 1 (IRF1), a tumour suppressor (Camicia, et al., 2013). ARTD9 directly

interacted with STAT1β (signal transducer and activator of transcription 1 isoform β) to inhibit

IRF1 expression, repressing the anti-proliferative and pro-apoptotic INFγ-STAT1-IRF1-p53

complex (Camicia, et al., 2013). ARTD9 has also been linked to the DNA damage response

pathway (Yan, et al., 2013). In response to DNA strand breaks, ARTD9 and its partner BBAP

(B-lymphoma and BAL-associated protein), an E3 ligase were recruited to DNA damage sites

and co-localized with ARTD1 and its product poly(ADP-ribose) (Yan, et al., 2013). ARTD9 and

BBAP at DNA damage sites mediate the specific recruitment of the adaptor protein RAP80 and

checkpoint mediators 53BP1 and BRCA1 through BBAP-mediated ubiquitylation, which limits

early and delayed DNA damage and enhances cellular viability (Yan, et al., 2013).

Page 37: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

22

ARTD13/PARP13/ZAP/ZC3HAV1 is a type 1 interferon-inducible host factor that regulates

viral RNA transcripts and was initially identified as zinc finger antiviral protein (ZAP) in a

screen for host factors that confer resistance to the retrovirus murine leukemia virus (MLV)

infection (Bick, et al., 2003; Gao, et al., 2002; Muller, et al., 2007). ARTD13 antiviral activities

were later confirmed in other retroviruses including HIV-1 and other RNA virus families,

including alphaviruses and filoviruses (Bick, et al., 2003; Muller, et al., 2007; Zhu, et al., 2011).

ARTD13 binds directly to specific viral mRNAs through its N-terminal zinc finger domains and

recruits cellular mRNA degradation factors to promote degradation of the target viral mRNA and

the inability of the virus to replicate efficiently (Guo, et al., 2004; Guo, et al., 2007; Zhu, et al.,

2011; Zhu and Gao, 2008). These antiviral properties are not due to ADP-ribosylation since full-

length ARTD13 is not catalytically active nor is its short isoform lacking the PARP catalytic

domain (Kleine, et al., 2008; Leung, et al., 2011). Both isoforms of ARTD13 were recently

reported to localize to cytoplasmic stress granules along with ARTD5, ARTD7 and ARTD12

(Leung, et al., 2011). Interestingly, when overexpressed, both isoforms decreased miRNA-

mediated silencing (Leung, et al., 2011).

2.4.2 Poly(ADP-ribose) Catabolism

2.4.2.1 Poly(ADP-ribose) glycohydrolase (PARG)

The removal of poly(ADP-ribose) chains from modified proteins is catalyzed by poly-ADP-

ribose glycohydrolase (PARG), which is the primary enzyme responsible for poly(ADP-ribose)

degradation in vivo (Alvarez-Gonzalez and Althaus, 1989; Davidovic, et al., 2001; Jonsson, et

al., 1988). PARG has both exo- and endoglycosidase activities that hydrolyze the glycosidic

bond between ADP-ribose units that generate large amounts of free ADP-ribose (Desnoyers, et

al., 1995; Heeres and Hergenrother, 2007; Oka, et al., 1984). The structure of a bacterial PARG

revealed that the PARG catalytic domain is a distant member of the macrodomain family

suggesting a potential role in the regulation by poly-ADP-ribosylation (Slade, et al., 2011). The

single mammalian PARG gene encodes several PARG proteins that localize to various cellular

compartments that also contain ARTDs (Meyer-Ficca, et al., 2004). Four PARG isoforms have

been identified so far, but the two predominant isoforms include PARG-110/111 (110 kDa), the

full-length and very active nuclear isoform and PARG-59/60 (65 kDa), a shorter isoform

(Brochu, et al., 1994; Meyer-Ficca, et al., 2004). PARG-102 (102 kDa), a splice variant that

lacks exon 1 and PARG-99 (99 kDa), a splice variant that lacks exons 1 and 2 are cytoplasmic

Page 38: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

23

but their roles are not fully understood (Meyer-Ficca, et al., 2004). Because of its low abundance

in eukaryotic cells and its extreme sensitivity to proteases, PARG has been difficult to study

(Bonicalzi, et al., 2005). Furthermore, a detailed analysis of the biological role of PARG has

been precluded by the lack of specific and selective inhibitors (Schreiber, et al., 2006).

Much of the current understanding of the biological role of PARG is largely based on gene

disruption studies and data obtained from Parg-deficient mice (Burkle and Virag, 2013).

Targeted deletion of exon 4 resulted in complete deletion of all PARG isoforms in Parg-deficient

mice (Koh, et al., 2004). Mice lacking all PARG isoforms arrest during development at

embryonic day 3.5 as a result of apoptosis that was induced by poly(ADP-ribose) accumulation

(Koh, et al., 2004). In vitro gene knockdown studies have revealed that PARG, like ARTD1, is

required for efficient DNA repair of single- and double-strand breaks and mitotic spindle

checkpoints since knockdown of either PARG or ARTD1 sensitized cells to apoptotic cell death

following DNA damage (Ame, et al., 2009; Erdelyi, et al., 2009; Fisher, et al., 2007; Keil, et al.,

2006). Mice that are deficient of full-length PARG (PARG110/111, Parg110-deficient) are

viable with no obvious phenotype and normal poly(ADP-ribose) metabolism (Cortes, et al.,

2004). However, Parg110-deficient mice are hypersensitive to ionizing radiation and alkylating

agents, similar to Artd1-deficient mice (Cortes, et al., 2004). Data obtained from Parg110-

deficient embryonic fibroblasts are similar to those reported for ARTD1 inhibition/knockout

studies, including defects in DNA repair, genomic instability and chromosomal aberrations (Min,

et al., 2010). These observations support a model in which a balance between poly(ADP-ribose)

synthesis and degradation is necessary for efficient repair of DNA strand breaks and cell

survival.

2.4.2.2 ADP-ribosylhydrolase 3 (ARH3)

ADP-ribosylhydrolase 3 (ARH3) is a member of the family of dinitrogenase reductase-activating

glycohydrolase-related proteins and shares very little similarity with the PARG catalytic domain

sequence (Mueller-Dieckmann, et al., 2006; Oka, et al., 2006). ARH3 localizes to the

mitochondria but is also present in the nucleus and cytosol (Niere, et al., 2008; Oka, et al., 2006).

ARH3 catalyzes the hydrolysis of both poly(ADP-ribose) and O-acetyl-ADP-ribose (OAADPR),

a metabolite that results from the activity of sirtuins (Imai and Guarente, 2010; Oka, et al., 2006;

Ono, et al., 2006). ARH3 was shown to degrade poly(ADP-ribose) in vitro and in cell but its

Page 39: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

24

specific activity was significantly less than that of PARG (Niere, et al., 2008; Oka, et al., 2006).

Embryonic fibroblasts from Arh3-deficient mice lacked most of the mitochondrial poly(ADP-

ribose) degrading activity and ARH3 appears to be the only identified enzyme to degrade PAR in

the mitochondrial matrix (Niere, et al., 2012). Because of its localization to the cytosol and

nucleus as well as its poly(ADP-ribose) degrading activity, ARH3 has also been proposed to

assist PARG in degrading poly(ADP-ribose) in the nucleus and cytosol (Niere, et al., 2012).

2.5 Regulation of chromatin structure and transcription by ARTDs

The disruption of chromatin structure by ADP-ribosylation of histones and destabilization of

nucleosomes were some of the earliest characterized effects of ARTD1 (Huletsky, et al., 1989;

Mathis and Althaus, 1987; Poirier, et al., 1982). The ability of ARTD1 to modulate chromatin

structure and function underlies its contribution to transcriptional regulation (Krishnakumar and

Kraus, 2010b). Early studies reported that ARTD1 activity is preferentially associated with

regions of chromatin that are actively transcribed (De Lucia, et al., 1996; Levy-Wilson, 1981;

Mullins, et al., 1977). ARTD1 was identified as a factor that could increase specificity of RNA

polymerase transcriptional initiation, supporting ARTD activity to directly regulate transcription

(Matsui, et al., 1980; Slattery, et al., 1983). Transcriptional regulation by ARTD activity has

been proposed to occur by at least two mechanisms that are not mutually exclusive: (1)

modification of histones to alter chromatin structure and (2) functioning as part of the regulatory

region (enhancer/promoter) binding complexes in conjunction with other DNA-binding factors

and co-activators (Kraus and Lis, 2003).

2.5.1 Modulation of chromatin structure by ARTD1 and poly(ADP-ribose)

Poly-ADP-ribosylation of chromatin protein can cause profound effects on nucleosome

architecture and on the stability of individual nucleosomes (D'Amours, et al., 1999). Histones H1

and H2B show the most poly-ADP-ribosylation in vivo and are the preferred targets of ARTD1

in vitro, although all histones are modified to some extent (Adamietz and Rudolph, 1984;

D'Amours, et al., 1999; Huletsky, et al., 1989; Poirier, et al., 1982). As seen directly by electron

microscopy, native polynucleosomes when poly-ADP-ribosylated by purified ARTD1 led to

complete decondensation and mimicked the effects of linker histone H1 depletion from higher-

order chromatin (Poirier, et al., 1982). Further studies suggested polyanionic poly(ADP-ribose),

Page 40: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

25

attached to target proteins or as a free polymer acts as an attractive core histone binding matrix

that can act as a histone acceptor to further destabilize nucleosomes (Mathis and Althaus, 1987;

Realini and Althaus, 1992). Moreover, ARTD1-dependent accumulation of poly(ADP-ribose) at

decondensed, transcriptionally active loci in native chromatin has been reported (Tulin and

Spradling, 2003). Collectively, these data support a model of transcriptional regulation whereby

ARTD1 promotes the decondensation of chromatin by causing the disassembly of nucleosomes

through poly-ADP-ribosylation of H1 and core histones, causing destabilization of nucleosomes

(D'Amours, et al., 1999; Kraus and Lis, 2003; Rouleau, et al., 2004; Tulin and Spradling, 2003).

More recent biochemical studies have demonstrated that in the absence of NAD+ or significant

auto-ADP-ribosylation, ARTD1 binds to nucleosomes and promotes the condensation of

chromatin by bringing together neighbouring nucleosomes (Kim, et al., 2004; Wacker, et al.,

2007). Addition of NAD+, which led to considerable auto-ribosylation of ARTD1, promoted the

release of ARTD1 from chromatin, leading to nucleosome decondensation and the restoration of

transcription (Kim, et al., 2004; Wacker, et al., 2007). These results suggest a new NAD+-

dependent model of modulation of chromatin structure and transcription whereby ARTD1 can

regulate chromatin structure without modifying histones or promoting disassembly of

nucleosomes (Kim, et al., 2004; Wacker, et al., 2007).

2.5.2 Enhancer-binding of actions of ARTD1

In addition to modulating transcription through alterations in chromatin structure, ARTD1 can

regulate transcription by functioning as a direct enhancer-binding factor. Many of the initial

studies describing the effects of ARTD on transcriptional regulation of target genes focused on

the binding of ARTD1 to specific DNA sequences or structures (Akiyama, et al., 2001; Butler

and Ordahl, 1999; Huang, et al., 2004; Nirodi, et al., 2001; Zhang, et al., 2002). Although no

consensus DNA binding sequence for ARTD1 has been established, the ability of ARTD1 to

bind DNA in a sequence-specific manner has been demonstrated where ARTD1 preferentially

bound to DNA oligonucleotides containing the 5’-TGTTG-3’ nucleotide sequence motif (Huang,

et al., 2004; Simbulan-Rosenthal, et al., 2003). Two independent studies demonstrated ARTD1

binds to specific sequences immediately upstream of chemokine (CXC motif) ligand 1 (CXCL1)

and in the first intron of B-cell lymphoma 6 (BCL6) to repress transcription (Ambrose, et al.,

2007; Amiri, et al., 2006).

Page 41: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

26

2.5.3 Promoter-binding and transcriptional co-regulation by ARTD1

Genomic studies have revealed that ARTD1 localizes to as many as 90% of expressed Pol II-

transcribed promoters in vitro which suggested that it plays a role in promoting the formation of

chromatin structures that are receptive to transcription (Krishnakumar, et al., 2008). The

enrichment of ARTD1 at these promoters correlated with the depletion of linker histone H1, and

a high ARTD1/H1 ratio specified genes that are actively transcribed (Krishnakumar, et al.,

2008). These findings indicated that ARTD1 localized to sites of ongoing transcription, where it

may exert stimulatory or inhibitory effects (Krishnakumar, et al., 2008). Roles of ARTD1 as a

promoter-specific co-regulator (either co-activator or co-repressor) of different sequence specific

DNA-binding transcriptional factors have been reported (Table 2). In some cases, the catalytic

activity of ARTD1 is required, while in other cases it is not (Ju, et al., 2004; Olabisi, et al., 2008;

Zaniolo, et al., 2007; Zhang, et al., 2013). These studies highlight the diverse mechanisms of

ARTD1 co-regulation, which are likely to vary in a gene- or promoter-specific manner.

Table 2. Transcription factors co-regulated by ARTD1.

Transcription factor

Co-regulatory effect

Catalytic function required?

Reference

AP2 stimulatory n/d (Kannan, et al., 1999) b-MYC stimulatory no (Cervellera and Sala, 2000) Elk-1 stimulatory yes (Cohen-Armon, et al., 2007) ERα stimulatory yes (Zhang, et al., 2013) HES1 stimulatory yes (Ju, et al., 2004)

HTLV Tax1 stimulatory no (Anderson, et al., 2000) NFAT stimulatory yes (Olabisi, et al., 2008) NF-κB stimulatory no (Hassa and Hottiger, 2002) Oct-1 stimulatory n/d (Nie, et al., 1998) p53 inhibitory yes (Mendoza-Alvarez and Alvarez-

Gonzalez, 2001) RAR stimulatory no (Pavri, et al., 2005) Sp1 inhibitory yes (Zaniolo, et al., 2007)

TEF-1 inhibitory no (Butler and Ordahl, 1999) TR/RXR inhibitory yes (Miyamoto, et al., 1999)

YY1 inhibitory n/d (Oei, et al., 1997)

AP2; activating protein 2, Elk-1; ETS domain-containing protein, ERα; estrogen receptor α,

HES1; hairy and enhancer of split 1, HTLV; human T-cell leukemia virus type 1, NFAT; nuclear

factor of activated T-cells, NF-κB; nuclear factor kappa-light-chain-enhancer of activated B-

cells, Oct-1; octamer transcription factor 1, RAR; retinoic acid receptor, Sp1; specificity protein

Page 42: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

27

1, TEF-1; transcriptional enhancer factor 1, TR/RXR; thyroid hormone receptor/retinoid X

receptor heterodimer, YY1; yin ying 1. n/d; not determined.

2.5.4 Transcriptional regulation by other ARTDs

Studies have emerged examining the regulation of transcription by other ARTD members,

including, ARTD2, ARTD8, ARTD9 and ARTD10 (Aguiar, et al., 2005; Camicia, et al., 2013;

Szanto, et al., 2012; Yu, et al., 2005).

Similar to ARTD1, ATRD2 has been demonstrated to co-regulate DNA-binding transcription

factors (Bai, et al., 2011; Bai, et al., 2007; Maeda, et al., 2006). The activity of the peroxisome

proliferator-activated receptor γ (PPARγ) was modulated by ARTD2 (Bai, et al., 2007). ARTD2

demonstrated direct interaction and co-occupancy with PPARγ/RXR heterodimers at promoter

regions of PPARγ target genes and acts as a co-activator of transcription leading to triglyceride

accumulation (Bai, et al., 2007). Estrogen receptor α (ERα) but not ERβ-dependent transcription

was decreased by ARTD2 knockdown suggesting a co-activator role of ARTD2 in ERα-

dependent transcription (Bai, et al., 2007; Szanto, et al., 2012). ARTD2 along with ARTD1 have

also been reported to serve as co-activators of thyroid transcription factor 1 (TTF1) and regulate

expression of surfactant protein B in lung (Maeda, et al., 2006). ARTD2 has been reported to be

a potent transcriptional repressor of the SIRT1 promoter and the reduction or knockout of

ARTD2 expression induced SIRT1 mRNA expression, protein and activity (Bai, et al., 2011).

The mono-ADP-ribosyltransferase ARTD8 was reported to activate a STAT6-regulated reporter

gene, which was dependent on its macro- and catalytic domains (Goenka and Boothby, 2006;

Goenka, et al., 2007). ARTD8 acted as a transcriptional switch, where under non-stimulating

conditions ARTD8 was bound to STAT6-responsive promoters and recruited histone deacetylase

2 (HDAC2) and HDAC3 to keep genes silent (Mehrotra, et al., 2011). Upon IL4 stimulation

ARTD8 catalyzed the ADP-ribosylation of itself, HDAC2 and HDAC3 leading to their

dissociation from the promoter thereby allowing STAT6 dimers binding (Mehrotra, et al., 2011).

ARTD10 may possibly play a role in MYC-induced transcription. ARTD10 was identified as an

interacting partner of MYC and this interaction was demonstrated to occur in the nucleus

(Kleine, et al., 2012; Yu, et al., 2005). ARTD10 has also recently been reported to negatively

regulate NF-κB signalling in response to IL-1β and TNFα stimulation by inhibiting NF-κB

Page 43: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

28

target gene and protein expression (Verheugd, et al., 2013). ARTD10-mediated inhibition of NF-

κB signalling was dependent on its catalytic activity and mono-ADP-ribosylation of NF-κB

essential modulator by ARTD10 reduced its poly-ubiquitination, leading to reduced expression

of NF-κB target genes (Verheugd, et al., 2013). ARTD10 also prevented the nuclear

translocation of the NF-κB transcription factor p65 (Verheugd, et al., 2013).

Recently, ARTD9 has been identified as a novel co-repressor of transcription of IRF1, a tumour

suppressor (Camicia, et al., 2013). ARTD9 directly interacted with STAT1β through its

macrodomains mediated by trans-mono-ADP-ribosylation to inhibit IRF1 expression (Camicia,

et al., 2013). This was shown to repress the anti-proliferative and pro-apoptotic INFγ-STAT1-

IRF1-p53 complex and mediated proliferation, survival and chemo-resistance in DLBCL

(Camicia, et al., 2013).

The transcriptional regulatory roles of the remaining ARTD members, including TiPARP have

not been evaluated. The expression of TiPARP is regulated by aryl hydrocarbon receptor (AHR),

a ligand activated transcription factor, but its potential function in AHR transactivation has not

been evaluated.

3 Aryl hydrocarbon receptor The AHR is a ligand-activated transcription factor and a member of the basic-helix-loop-helix

Per (Period)-ARNT (aryl hydrocarbon receptor nuclear translocator)-SIM (single-minded)

(bHLH/PAS) family of heterodimeric transcriptional regulators (Burbach, et al., 1992). Other

members of the bHLH/PAS family of transcription factors include ARNT, hypoxia factor 1α

(HIF1α), SIM, Clock, Per, and the p160 family of co-activators (Hankinson, 1995). bHLH/PAS

family of transcription factors are involved in the control of a variety of physiological processes

which include: circadian rhythms, organ development, neurogenesis, metabolism and the stress

response to hypoxia (Crews, 1998; Fernandez-Salguero, et al., 1997; Gonzalez and Fernandez-

Salguero, 1998; Whitlock, 1999). Although AHR plays an important role in physiology it is most

well-known for mediating the toxic effects of environmental contaminants (Abbott, et al., 1995;

Hahn, 2002; Lahvis, et al., 2000; Lin, et al., 2001; Mimura, et al., 1997; Schmidt, et al., 1996;

Van den Berg, et al., 1998).

Page 44: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

29

3.1 Structure The human AHR is composed of several modular domains with distinct functions (Figure 5A).

AHR shares similar domain architecture to other bHLH/PAS proteins, with a highly conserved

N-terminal bHLH motif, adjacent PAS region, and loosely conserved C-terminal transactivation

or transrepression regions (Figure 5B and C) (Kewley, et al., 2004). The basic region of the

bHLH motif mediates DNA binding and the HLH domain is involved in dimerization (Kikuchi,

et al., 2003). Within the N-terminus of AHR is a bipartite nuclear localization signal (NLS),

which overlaps with the basic region of the bHLH (Ikuta, et al., 1998). Moreover, AHR also

contains a leucine-rich NES with helix 2 of the bHLH motif (Ikuta, et al., 1998). The PAS region

of AHR contains two tandem PAS domains designated PAS A and PAS B and both PAS

domains along with the bHLH motif contribute to dimerization with AHR obligatory interaction

partner ARNT and interaction with chaperone protein HSP90 (heat shock protein 90 kDa)

(Lindebro, et al., 1995; Pongratz, et al., 1992). The PAS A domain controls dimerization

specificity and stability, and strengthens DNA binding (Chapman-Smith, et al., 2004; Lindebro,

et al., 1995; Pongratz, et al., 1998). The PAS B domain contains the ligand binding domain and

deletion of this domain results in a constitutively active receptor (McGuire, et al., 2001).

The C-terminal half of AHR contains a modular transactivation domain (TAD) that is composed

of three distinct subdomains: acidic, glutamine-rich (Q-rich) and proline-serine-threonine-rich

(P/S/T) (Jain, et al., 1994; Rowlands, et al., 1996; Sogawa, et al., 1995). Individually, each of the

three subdomains exhibits low levels of transactivation; however, any combination of two

subdomains can synergistically activate transcription (Rowlands, et al., 1996). The Q-rich sub-

domain is critical for transcriptional activation and its deletion completely inactivates the

receptor (Kumar, et al., 2001). Additionally the Q-rich subdomain also appears to be required for

interaction with co-regulatory proteins nuclear co-activator 1 (NcoA1) and receptor interacting

protein 140 (RIP140) (Kumar and Perdew, 1999; Kumar, et al., 2001; Kumar, et al., 1999).

Page 45: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

30

Figure 5. Schematic of the modular domains of three bHLH-PAS family members.

(A) Aryl hydrocarbon receptor (AHR), (B) Aryl hydrocarbon nuclear translocator (ARNT), and

the (C) Aryl hydrocarbon receptor repressor (AHRR). bHLH (basic helix-loop-helix), PAS

(Per/ARNT/Sim), P/S/T (Proline/Serine/Threonine).

3.2 AHR ligands AHR is a promiscuous receptor that can bind and be activated by a wide variety of structurally

diverse exogenous and natural compounds that produce a spectrum of the biological and toxic

effects (Denison and Heath-Pagliuso, 1998; Denison and Nagy, 2003; Denison, et al., 2011;

Nguyen, et al., 1999; Poland and Knutson, 1982; Safe, 1990). AHR ligands can be divided into

two major categories: exogenous and naturally occurring.

Page 46: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

31

3.2.1 Exogenous AHR ligands

The majority of the high affinity AHR ligands that have been identified and characterized are

exogenous, which include planar hydrophobic halogenated aromatic hydrocarbons (HAHs) and

polycyclic aromatic hydrocarbons (PAHs) and related compounds (Figure 6) (Denison and

Heath-Pagliuso, 1998; Gillner, et al., 1993; Poland and Knutson, 1982; Safe, 1990). Many HAHs

and PAHs are environmental contaminants that are by-products of chlorine chemistry, waste

incineration, metal production and fossil fuel or wood burning (Denison and Nagy, 2003;

DeVito, et al., 1994; Hankinson, 2005; Kulkarni, et al., 2008; Safe, 1986; Van den Berg, et al.,

1998). The HAHs represent the most potent class of AHR ligands and include chlorinated and

brominated dibenzo-p-dioxins, dibenzofurans and biphenyls. HAHs are metabolically stable and

have binding affinities in the picomolar to nanomolar range (Denison and Nagy, 2003). PAHs

include: 3MC, B[a]P, benzoanthracenes and benzoflavones (Denison and Nagy, 2003; Poland

and Knutson, 1982). PAHs are metabolically less stable than HAHs and tend to have lower

binding affinities usually within the nanomolar to micromolar range. Structure-activity

relationship analysis of many HAHs and PAHs congeners has demonstrated toxicity is positively

correlated to the affinity of the congener for AHR (Kafafi, et al., 1993; Mhin, et al., 2002;

Poland and Knutson, 1982; Safe, 1986; Tuppurainen and Ruuskanen, 2000; Waller and

McKinney, 1995). The HAH, 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD, dioxin) has the

highest affinity for AHR and is the most toxic of the known AHR ligands and remains the

prototype AHR ligand for which all subsequent ligands are compared (Nebert, et al., 2004).

Page 47: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

32

Figure 6. Selected exogenous AHR ligands.

3.2.2 Naturally occurring AHR ligands

The greatest source of human exposure to AHR ligands (exogenous and naturally occurring)

comes from the diet (Denison and Nagy, 2003; Jeuken, et al., 2003). A variety of naturally

occurring dietary constituents are AHR ligands (Bjeldanes, et al., 1991; MacDonald, et al., 2001;

Wattenberg and Loub, 1978). Indole-3-carbinol (IC3) is found in cruciferous vegetables and is

hydrolyzed under the acidic conditions of the gastrointestinal tract to many products including

3,3’-diindolylmethane (DIM) and indolo[3,2-b]carbazole (ICZ) (Bjeldanes, et al., 1991; Grose

and Bjeldanes, 1992). DIM is a partial agonist for AHR while ICZ is a potent activator of AHR

activity (Broadbent and Broadbent, 1998; Chen, et al., 1996; Kleman, et al., 1994). ICZ has high

Page 48: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

33

affinity for AHR but it does not cause toxicity (Pohjanvirta, et al., 2002). Naturally occurring

AHR ligands can also be antagonists and numerous have been previously described as such.

Resveratrol (3,5,4’-trihydroxystilbene), a polyphenol present in grapes, peanuts and berries and

kaempferol (3,4’,5,7,-tetrahydroxyflavone), a flavonol found in teas, broccoli, grapefruit and

other plant sources are two potent AHR antagonists (Casper, et al., 1999; Ciolino, et al., 1999;

MacPherson and Matthews, 2010).

To date no definitive endogenous AHR ligand has been identified; however, numerous

endogenous chemicals have been identified that bind to AHR and/or activate AHR-dependent

gene expression (Denison and Nagy, 2003). Bilirubin, tryptophan metabolites, arachidonic acid

metabolite lipoxin A4, several prostaglandins, retinoids, indigo and indirubin are some examples

of endogenous compounds known to activate AHR (Adachi, et al., 2001; Bittinger, et al., 2003;

Gambone, et al., 2002; McMillan and Bradfield, 2007; Phelan, et al., 1998; Schaldach, et al.,

1999; Seidel, et al., 2001; Sugihara, et al., 2004). FICZ (6-formylindolo[3,2-b]carbazole), a

tryptophan photoproduct has been reported to be the most potent and highest affinity candidate

endogenous AHR ligand known to date, capable of competitively displacing TCDD (Fritsche, et

al., 2007; Helferich and Denison, 1991; Rannug, et al., 1987; Wincent, et al., 2009). Femtomolar

levels of FICZ were reported to sufficiently induce AHR target mRNA and enzyme activity

(Wincent, et al., 2012). Unlike TCDD, however, FICZ is a substrate for Phase I enzymes

cytochrome P450 1A1, CYP1A2 and CYP1B1 and is readily metabolized (Wincent, et al.,

2009).

3.3 AHR-mediated toxicity Most of what is known about the toxicity of TCDD and related compounds come from studies in

experimental animals, but effects are also observed in humans (Birnbaum, 1994; Birnbaum and

Tuomisto, 2000; Pohjanvirta and Tuomisto, 1994; Sewall and Lucier, 1995; Tuomisto, 2005;

White and Birnbaum, 2009). One of the most dramatic effects from a single TCDD exposure in

experimental animals is the lethal wasting syndrome that is characterized by progressive weight

loss, hypophagia and gluconeogenic suppression (Pohjanvirta and Tuomisto, 1994; Seefeld, et

al., 1984; Seefeld and Peterson, 1984; Stahl, et al., 1993). While wasting accompanies death,

wasting per se is not likely the sole cause of death since maintenance of body weight by

parenteral nutrition does not prevent mortality (Gasiewicz, et al., 1980). Lethality of exposed

Page 49: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

34

animals is delayed and does not ensue until 1-8 weeks after exposure with the time course

depending on the sensitivity of the animal (Birnbaum and Tuomisto, 2000; Poland and Knutson,

1982). A wide range of other short-term toxic effects following TCDD exposure are known,

including thymic atrophy, hepatic necrosis or hypertrophy, lipid accumulation within

hepatocytes, a variety of endocrine imbalances and immunosuppression (Birnbaum and

Tuomisto, 2000; Pohjanvirta and Tuomisto, 1994). Ahr-null mice challenged with TCDD are

resistant to TCDD-induced toxicities as well as to TCDD-induced transcriptional alterations

(Boutros, et al., 2009; Fernandez-Salguero, et al., 1995; Mimura, et al., 1997; Schmidt, et al.,

1996; Tijet, et al., 2006). Mice expressing a mutated form of AHR that is unable to translocate to

the nucleus or a DNA binding mutant are also essentially refractory to TCDD-mediated toxicity

(Bunger, et al., 2008; Bunger, et al., 2003). Furthermore, mice hypomorphic for ARNT are

phenotypically non-responsive to dioxins (Walisser, et al., 2004). Collectively, these genetic

manipulations indicate that TCDD toxicities are dependent on the transcriptional regulatory

function and DNA-binding capacity of AHR.

3.4 AHR transactivation

In its latent state, AHR is found in the cytosol where it is stably associated with a multi-protein

complex consisting of two molecules of chaperone protein HSP90, hepatitis B virus X-associated

protein 2 (XAP2) and a 23 kDa co-chaperone protein referred to as p23 (Figure 7)(Carver, et al.,

1994; Kazlauskas, et al., 1999; Ma and Whitlock, 1997; Meyer, et al., 1998; Perdew, 1988;

Pollenz, et al., 1994). HSP90 interacts with AHR via both the bHLH motif and the PAS B

domain and is an essential component of the unliganded AHR complex as its presence is

necessary for the high affinity ligand binding conformation of AHR and retention of AHR in the

cytosol by masking its NLS sequence (Antonsson, et al., 1995a; Antonsson, et al., 1995b;

Dolwick, et al., 1993; Pongratz, et al., 1992; Whitelaw, et al., 1993; Whitelaw, et al., 1995).

XAP2 has a multifunctional role which includes maintaining proper cytosolic folding of AHR,

improving chaperone complex stability, subcellular localization and ligand binding ability of

AHR and repressing AHR transcriptional activity (Kazlauskas, et al., 2000; LaPres, et al., 2000;

Meyer and Perdew, 1999; Meyer, et al., 1998; Petrulis, et al., 2000). Co-chaperone p23 stabilizes

the chaperone complex and is important for nuclear import of AHR (Kazlauskas, et al., 1999).

Page 50: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

35

Most AHR ligands are highly lipophilic and therefore can enter the cell through simple diffusion

(Delescluse, et al., 2000; Goryo, et al., 2007). Ligand binding induces a conformational change

in AHR that exposes its NLS (Henry and Gasiewicz, 2003; Ikuta, et al., 2004; Whitlock, 1999).

Liganded AHR translocates into the nucleus where it dissociates from its chaperone protein

complex and then heterodimerizes with ARNT and is converted to its high affinity DNA

conformation (Hankinson, 2005; Okey, et al., 1980; Tomita, et al., 2000). The AHR/ARNT

heterodimer recognizes a DNA element designated aryl hydrocarbon receptor response element

(AHRE) located within the regulatory regions of target genes such as cytochrome P450 1A1

(CYP1A1) (Mimura and Fujii-Kuriyama, 2003; Whitlock, 1999). The AHRE is composed of two

half-sites, TnGC and GTG recognized by AHR and ARNT, respectively (Denison, et al., 1988;

Swanson, et al., 1995). Once bound to an AHRE the AHR/ARNT complex interacts with

multiple co-regulator proteins to regulate transcription (Hankinson, 2005). A list of co-regulator

proteins known to interact and/or modulate AHR/ARNT activity is provided in Table 3.

Table 3. Partial list of proteins and co-regulator proteins that interact with AHR/ARNT to

modulate transcription.

Factors Response Co-activators NCoA1/NCoA2/NCoA3/p300/CBP, interact with AHR and/or ARNT to facilitate gene NCoA6 activation (Hankinson, 2005; Harper, et al., 2006) RIP140/NRIP enhances AHRE-driven transcription (Hankinson,

2005; Harper, et al., 2006) Brg-1/Brm1 enhances transcription mediated by AHR/ARNT

(Hankinson, 2005) Co-repressors SMRT, SHP inhibits transcriptional activity of AHR/ARNT

complex (Jepsen and Rosenfeld, 2002; Nguyen, et al., 1999)

General Transcription Factors Pol II, TBP, TFIIB, TFIIF key components of the general transcription

machinery that are directly involved in gene activation by AHR/ARNT (Hankinson, 2005; Harper, et al., 2006)

For a complete list of AHR co-regulators see reference (Hankinson, 2005).

Page 51: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

36

Figure 7. General mechanism of AHR transactivation.

Ligands bind (TCDD) to AHR in the cytosol and triggers translocation into the nucleus. AHR is

released from its chaperone complex and then binds with its dimerization partner ARNT. The

AHR/ARNT complex binds AHREs located within the regulatory region of target genes. Once

bound to the AHRE, AHR/ARNT associates with co-activators to activate transcription.

3.5 Negative regulation of AHR transactivation There are several mechanisms by which AHR transactivation can be attenuated (Figure 8). One

mechanism involves ligand-dependent degradation of AHR protein through a proteasomal

pathway (Pollenz, 2002; Wentworth, et al., 2004). The decrease in AHR protein leads to reduced

AHR-dependent transcription (Pollenz, et al., 1998). Other proposed mechanisms involve an

autoregulatory feedback loop created by induction of AHR-dependent transcriptional repressor.

The aryl hydrocarbon receptor repressor (AHRR) is induced by activated-AHR and is a proposed

transcriptional repressor of AHR (Mimura, et al., 1999). Another and more hypothetical

mechanism combines ideas from the two previous mechanisms where a yet to be identified

AHR-dependent labile factor promotes the proteolytic degradation of AHR protein and the

attenuation of transcription (Ma, 2002; Ma and Baldwin, 2002).

Page 52: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

37

Figure 8. Three proposed mechanisms of negative regulation of AHR transactivation.

(A) Targeted proteolytic degradation of AHR protein through the ubiquitin 26S proteasome

pathway. (B) Induction of the transcriptional repressor AHRR. (C) Induction of a labile factor

that promotes the proteolytic degradation of AHR.

3.5.1 Proteolytic Degradation of AHR

Numerous studies have established AHR ligand binding leads to rapid and sustained reduction in

the level of endogenous AHR protein in both in vitro and in vivo models (Pollenz, 2002). The

ligand-induced degradation of AHR has been observed in many cell lines of different tissue

lineages, although the time-course and range of maximal degradation varied amongst cell lines

(Pollenz and Buggy, 2006). AHR degradation has also been observed in vivo. For example, AHR

protein concentration was reduced in liver, spleen, thymus and lung from Sprague-Dawley rats

given an acute dose of TCDD (Pollenz, et al., 1998). The ubiquitin-proteasome pathway has

been implicated in the ligand-induced degradation of AHR due the ability of the proteasome

inhibitors, MG-132 and lactacystin, two specific inhibitors of the 26S proteasome, to completely

block degradation (Davarinos and Pollenz, 1999; Ma and Baldwin, 2000). Inhibitors of calpains,

Page 53: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

38

serine/cysteine, lysosomal and serine proteases failed to block AHR degradation (Davarinos and

Pollenz, 1999; Ma and Baldwin, 2000; Roberts and Whitelaw, 1999; Wormke, et al., 2000). In

addition, AHR degradation is inhibited in cells that contain a temperature-sensitive mutation in

the E1 ubiquitin conjugating enzyme (Ma and Baldwin, 2000). The inhibition of AHR

degradation resulted in sustained levels of AHR in the nucleus over time and significantly

increased in the concentration of DNA-bound AHR/ARNT heterodimers (Ma and Baldwin,

2000; Roberts and Whitelaw, 1999). These studies also demonstrated that AHR degradation

required the C-terminal transactivation domain and DNA-binding activity, as AHR lacking the

C-terminus and a DNA-binding mutant were more resistant to degradation than wildtype AHR

(Ma and Baldwin, 2000). TCDD induced the accumulation of high molecular weight AHR that

contained ubiquitin; however, specific ubiquitination sites are yet to be identified (Ma and

Baldwin, 2000). Taken together, these studies provide compelling evidence that AHR is

degraded via the 26S proteasome.

The ligand-induced degradation of AHR was connected to down-regulation of AHR target gene

expression by analyzing the duration and magnitude of gene expression following proteasomal

inhibition. AHR-regulated reporter gene experiments demonstrated pretreatment of cells with

MG-132 resulted in significantly greater reporter gene activity following TCDD treatment

compared to cells treated with TCDD alone (Davarinos and Pollenz, 1999). Inhibition of AHR

degradation also resulted in higher endogenous AHR target mRNA induction compared to non-

inhibited cells when exposed to TCDD (Ma and Baldwin, 2000). Reporter genes were also used

to assess whether down-regulation of endogenous AHR rendered cells non-responsive. Cell lines

propagated in the presence of TCDD for 12 days and then re-challenged with TCDD

demonstrated reductions in AHR protein and the absolute level of reporter gene activity (Pollenz,

et al., 1998). Collectively, these results correlate ligand-induced gene expression increases with

AHR degradation inhibition.

3.5.2 Aryl hydrocarbon receptor repressor (AHRR)

The aryl hydrocarbon receptor repressor (AHRR) is a bHLH-PAS protein that was discovered

because of its similarity to AHR (Mimura, et al., 1999). The AHRR shares high amino acid

identity with the AHR in the N-terminal region containing the bHLH (81%) and PAS A domain

(60%) but diverges significantly towards the C-terminus and does not contain a ligand binding

Page 54: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

39

(PAS B) and transactivation domain (Figure 5C) (Mimura, et al., 1999). Due to the lack of a

ligand binding domain, the AHRR does not bind AHR ligands and appears to function in a

ligand-independent manner (Karchner, et al., 2009). Like AHR, the AHRR can heterodimerize

with ARNT and bind AHREs through its HLH-PAS A domains and basic region respectively

(Mimura, et al., 1999). Consistent with this, the initial mechanism of AHRR-mediated repression

was hypothesized to involve competition between AHR and the AHRR for ARNT

heterodimerization and competition between AHR/ARNT and AHRR/ARNT heterodimers for

AHRE binding (Mimura, et al., 1999). AHRR/ARNT binding to AHREs do not transcriptionally

activate due to the AHRR lacking a transactivation domain (Mimura, et al., 1999). Four putative

AHREs have been identified in the 5’ proximal to the promoter of the AHRR gene, and not

surprisingly AHR ligands have been reported to induce AHRR gene expression (Baba, et al.,

2001; Evans, et al., 2005; Karchner, et al., 2002; Mimura, et al., 1999; Tsuchiya, et al., 2003;

Yamamoto, et al., 2004). Induction of the AHRR by AHR agonists and AHRR-mediated

repression of AHR transactivation implicates that the AHRR forms a negative feedback loop

with AHR.

AHRR mRNA is constitutively expressed in a variety of tissues with the testis, lung, spleen,

heart and kidneys expressing the highest levels (Bernshausen, et al., 2006; Korkalainen, et al.,

2004; Nishihashi, et al., 2006; Tsuchiya, et al., 2003; Yamamoto, et al., 2004). In some tissues

that express high AHR levels such as the liver, constitutive AHRR mRNA is low and only

becomes measurable after induction with potent AHR ligands such as TCDD, B[a]P or 3MC

(Bernshausen, et al., 2006; Korkalainen, et al., 2005; Korkalainen, et al., 2004; Mimura, et al.,

1999). Conversely, constitutively high AHRR mRNA levels have been attributed to the lack of

AHR ligand inducibility of targets in primary cells and some cell lines (Gradin, et al., 1999;

Tsuchiya, et al., 2003). Reduction of AHRR expression by siRNA can in some cases restore

inducibility of AHR targets, but low constitutive expression of AHRR does not always correlate

with high AHR target inducibility (Haarmann-Stemmann, et al., 2007; Tsuchiya, et al., 2003;

Yamamoto, et al., 2004). For instance, constitutive AHRR mRNA is decreased in the heart and

brain tissue of Ahr-/- mice (Bernshausen, et al., 2006). Additionally, B[a]P induces significant

CYP1A1 mRNA in murine kidney tissue but does not induce detectable levels of AHRR mRNA

(Bernshausen, et al., 2006). The ability of AHR to transactivate the AHRR gene and the ability

of AHRR to repress AHR activity is likely to be tissue-, cell- or context-specific.

Page 55: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

40

The AHRR has been demonstrated to be an effective transcriptional repressor of AHR activity;

however, its repressive activity on AHR transactivation has been largely based on in vitro

overexpression reporter gene experiments (Evans, et al., 2008; Evans, et al., 2005; Karchner, et

al., 2002; Karchner, et al., 2009; Mimura, et al., 1999; Oshima, et al., 2007). Overexpression

experiments using GFP-tagged AHRR were used to determine GFP-AHRR and presumably

endogenous AHRR localize primarily to the nucleus in MCF7, Hepa1c1c7 and COS7 cells

(Evans, et al., 2008; Kanno, et al., 2007; Mimura, et al., 1999). Using a series of GFP-tagged

AHRR truncations the AHRR was found to contain a functional NLS and NES and was shown to

undergo nucleocytoplasmic shuttling that favoured nuclear localization (Kanno, et al., 2007).

Overexpression experiments determined that the C-terminus of the AHRR was required for

repression, suggesting the presence of a ‘transrepression domain’ (Oshima, et al., 2007). These

observations support the hypothesis that AHRR is a nuclear transcriptional repressor. The exact

molecular mechanisms by which the AHRR represses AHR transactivation is currently subject to

debate. The initial mechanism proposed for AHRR-mediated repression involved competition

between the AHRR and AHR for heterodimerization with ARNT; however, subsequent studies

have reported that overexpression of ARNT had no effect on AHRR-mediated repression,

demonstrating that competition for ARNT is not the primary mechanism of repression (Evans, et

al., 2008; Karchner, et al., 2009). Moreover, a DNA-binding point mutant of the AHRR

effectively repressed AHR activity indicating that AHRE binding was not required for repression

(Evans, et al., 2008). Furthermore, interaction experiments demonstrated the AHRR interacted

with ARNT and AHR in a ligand-independent manner (Karchner, et al., 2009). These studies

illustrate the complexity of AHRR-mediated repression and the mechanism by which the AHRR

represses requires further clarification.

3.5.3 Induction of a labile AHR degradation promoting factor (ADPF)

A third proposed mechanism by which AHR transactivation can be terminated is through a

hypothetical labile factor that promotes the proteasomal degradation of AHR, termed the AHR

degradation promoting factor (ADPF) (Ma, 2007; Ma and Baldwin, 2002; Ma, et al., 2000). The

existence of a potential labile factor was first proposed based on the observation that protein

synthesis inhibition of liver cells increased the AHR ligand-induced target gene expression well

above levels of AHR ligand treatment alone (Whitlock and Gelboin, 1973). Cycloheximide, a

potent protein synthesis inhibitor, enhanced the TCDD-induced CYP1A1 mRNA to levels

Page 56: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

41

greater than the sum of each compound alone, a phenomenon termed superinduction (Ma and

Baldwin, 2002; Ma, et al., 2000). Superinduction results from increased transcriptional rate and

subsequent generation of AHR target mRNA and protein but not from increased mRNA or

protein stability (Israel, et al., 1985; Israel and Whitlock, 1983). Superinduction requires

substantial (>90%) protein synthesis inhibition and functional AHR and ARNT suggesting that

the synthesis of a repressor protein, which is labile and induced during AHR target induction is

inhibited (Lusska, et al., 1992). CHX co-treatment was later found to fully inhibit the

proteasomal degradation of AHR that correlated with superinduction in a time- and dose-

dependent manner (Ma, et al., 2000). Similarly, proteasomal inhibition with MG-132 also

superinduced AHR targets following AHR agonist treatment further supporting the notion that

increasing the stability of AHR by decreasing its degradation superinduced target gene

expression (Ma and Baldwin, 2000; Ma, et al., 2000). Together these observations establish the

hypothesis that a repressor induced by agonist-activated AHR and sensitive to CHX mediates the

proteasomal degradation of AHR, this unknown factor is referred to as the ADPF (Ma, et al.,

2000). Two possible mechanisms, which are not mutually exclusive, have been proposed for the

ADPF based on results from inhibitor studies. One is an autoregulatory mechanism in which the

ADPF is an AHR target protein that is induced by AHR agonists (Ma and Baldwin, 2002).

Blockage of the induction of the ADPF by protein synthesis disrupts the autoregulation (Ma and

Baldwin, 2002). The second is the ADPF is a labile factor that requires constant protein synthesis

for function and therefore sensitive to cycloheximide (Ma and Baldwin, 2002). Despite the

evidence supporting its existence, an ADPF has not yet been identified.

3.5.4 In vivo models used to study negative regulation of AHR

3.5.4.1 Ahrr-deficient mice

Ahrr-deficient mice were generated to study the function of the AHRR in AHR transactivation in

vivo (Hosoya, et al., 2008). The AHR-dependent induction of CYP1A1 was expected to be

greater in Ahrr-deficient mice than wildtype mice following 3MC treatment; however, this

pattern of induction was highly tissue-specific and limited to only the heart, spleen and skin

(Hosoya, et al., 2008). Tissues with high AHR expression including the liver, lung, kidney and

thymus did not display significant increases in 3MC-treated Ahrr-deficient mice compared with

treated wildtype mice nor did the brain or stomach (Hosoya, et al., 2008). Constitutive

expression of CYP1A1 in Ahrr-deficient mice was not significantly different than wildtype mice

Page 57: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

42

for all the tissues examined (Hosoya, et al., 2008). Ahrr-deficient mice treated with subcutaneous

B[a]P showed a significant delay in skin tumour formation than wildtype (Hosoya, et al., 2008).

It was suggested that the increased CYP1A1 expression in the skin shifted the balance of

metabolic activity of CYP1A1 in favour of detoxification and this attributed to the delay in skin

carcinogenesis (Hosoya, et al., 2008). B[a]P-treated Ahrr-deficient mice demonstrated a slightly

lower mortality rate trend compared with treated-wildtype mice (Hosoya, et al., 2008).

3.5.4.2 Constitutively active AHR mice

Deletion of the entire PAS B domain of AHR results in a protein that does not interact with

HSP90, shows constitutive nuclear localization, binds ARNT and exhibits potent transactivation

in the absence of ligand i.e. TCDD, a constitutively active AHR (CA-AHR) (McGuire, et al.,

2001). Transgenic mice expressing CA-AHR were generated and used to study the in vivo

consequences of lack of AHR transactivation termination (Andersson, et al., 2002; Andersson, et

al., 2003; Brunnberg, et al., 2006; Brunnberg, et al., 2011; McGuire, et al., 2001; Moennikes, et

al., 2004; Wejheden, et al., 2010). Transgenic CA-AHR animals expressed CA-AHR in

lymphatic organs including thymus, spleen, lymph nodes, enriched T and B cells and bone

marrow as well as in lung, liver, muscle, skin, brain, heart, kidney and throughout the

gastrointestinal tract (Andersson, et al., 2002). All tissues that showed CA-AHR expression also

demonstrated induced CYP1A1 mRNA expression at varying levels (Andersson, et al., 2002).

Additionally, homozygous CA-AHR mice demonstrated significantly greater CYP1A1, CYP1A2

and AHRR mRNA expression in the liver compared with wildtype animals (Andersson, et al.,

2002). CYP1A1 mRNA induction in homozygous CA-AHR mice was similar to that of low to

mid-dose TCDD-treated wildtype mice in the liver, thymus and stomach (Andersson, et al.,

2002). CA-AHR mice also showed several symptoms that are similar to TCDD toxicity such as

thymic atrophy and liver enlargement (Andersson, et al., 2002). CA-AHR mice showed a

significantly reduced lifespan where mice began dying as early as 6 months and often without

symptoms (other than body weight loss a day before death), only a few animals survived after 12

months (Andersson, et al., 2002). Another distinguishing feature of CA-AHR mice was the

development of invasive tumours of the glandular stomach at an early age that correlated with

increased mortality (Andersson, et al., 2002). Other findings in these CA-AHR mice included:

decreased population of B lymphocytes in the peritoneal cavity, enlarged heart and kidneys and

Page 58: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

43

increased hepatocarcinogenesis (Andersson, et al., 2003; Brunnberg, et al., 2006; Brunnberg, et

al., 2011; Moennikes, et al., 2004).

3.6 AHR target genes The up-regulation of a battery of six enzymes involved in biotransformation of xenobiotics and

endogenous compounds in response to TCDD and related AHR ligands has been well established

and studied in several species (Nebert, 1989). This battery includes cytochrome P450 1A1

(CYP1A1), CYP1A2, glutathione S-transferase Ya (GSTYa), aldehyde-3-dehydrogenase

(ALDH-3) and NAD(P)H: quinone oxidoreductase (NQO1) (Nebert, et al., 1990; Nebert, et al.,

2000; Schrenk, 1998). A third cytochrome P450 enzyme, CYP1B1 was later identified to be up-

regulated by AHR and added to the gene battery (Savas, et al., 1994; Shen, et al., 1993; Sutter, et

al., 1994). CYP1A1, CYP1A2 and CYP1B1 are NADPH-dependent monooxygenases that

mediate phase I metabolism rendering end-products more reactive for subsequent reactions. The

other four enzymes mediate phase II metabolism by conjugating charged groups onto xenobiotics

to facilitate their elimination. Many AHR ligands are metabolized by these enzymes and the idea

of induction of these enzymes as an adaptive response to AHR ligand exposure has been

proposed (Gu, et al., 2000; Nebert, et al., 2004; Okey, 1990). The induction of these enzymes,

however, cannot explain the plethora of toxic effects that are associated with TCDD exposure.

TCDD is poorly metabolized and thus causes chronic and sustained induction of cytochrome

P450 enzymes, leading to oxidative stress, formation of reactive oxygen species, lipid

peroxidation and DNA damage (Hassoun, et al., 2002; Hassoun, et al., 2001; Nebert, et al.,

2000; Shertzer, et al., 1998; Slezak, et al., 2000).

Several different methodological approaches have been used to discover new AHR target genes,

such as microarrays, differential display, representational difference analysis and serial analysis

of gene expression (Boverhof, et al., 2006; Dong, et al., 1997; Fletcher, et al., 2005; Frericks, et

al., 2006; Frueh, et al., 2001; Hayes, et al., 2007; Kurachi, et al., 2002; Ohbayashi, et al., 2001;

Oikawa, et al., 2002; Puga, et al., 2000; Svensson and Lundberg, 2001; Tijet, et al., 2006;

Zeytun, et al., 2002). These studies revealed that AHR ligands induce the expression of a number

of genes that do not encode metabolizing enzymes. Some of these genes have been shown to be

‘true’ AHR target genes, being transcriptionally regulated via a genomic AHR pathway

involving AHRE binding within the gene regulatory region while others are affected by a non-

Page 59: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

44

AHRE binding or non-genomic pathway (Dere, et al., 2011; Lo, et al., 2011; Lo and Matthews,

2012). TiPARP is an example of a novel AHR target gene that encodes a non-xenobiotic

metabolizing enzyme and is regulated via the genomic AHR pathway (Hao, et al., 2012; Ma, et

al., 2001).

3.6.1 TCDD-inducible poly(ADP-ribose) polymerase (TiPARP): a novel AHR target

TCDD-inducible poly(ADP-ribose) polymerase (TiPARP) was first identified in mouse

hepatoma cells by differential display as a TCDD-induced mRNA (Ma, et al., 2001). The human

orthologue of the mouse Tiparp gene was later identified and contains six exons and human

TiPARP protein displayed 91.8% total amino acid identity with mouse TiPARP (Katoh and

Katoh, 2003). Induction of TiPARP gene expression by TCDD is mediated via the AHR (Ma,

2002; Ma, et al., 2001). TiPARP induction by TCDD demonstrated characteristics similar to

other AHR target genes such as superinduction in the presence of proteasomal and protein

synthesis inhibitors (Ma, 2002). Subsequent microarray studies have shown that TiPARP mRNA

is consistently up-regulated by TCDD both in vitro and in vivo (Dere, et al., 2006; Forgacs, et

al., 2013; Frericks, et al., 2006; Henry, et al., 2010; Lo, et al., 2011; Tijet, et al., 2006).

Additionally, TiPARP gene regulation by other AHR ligands both exogenous and naturally

occurring as well as AHR antagonists has been previously observed (de Waard, et al., 2008;

Hao, et al., 2012; Ito, et al., 2006; Pansoy, et al., 2010; Wahl, et al., 2008). Together these

studies support TiPARP as an AHR target gene.

3.6.1.1 AHR regulation of TiPARP expression

TiPARP induction by TCDD has been reported to be rapid reaching maximal levels after 1.5 h to

2 h of treatment leading to TiPARP sometimes being referred to as an immediate early gene

(Diani-Moore, et al., 2010; Hao, et al., 2012; Ma, 2002; Ma, et al., 2001). Currently, the

regulation of TiPARP by activated-AHR is not fully understood. Previous studies have reported

regions upstream of the TiPARP transcriptional start that are highly bound by ligand activated-

AHR suggesting the presence of an AHRE-containing regulatory region (Ahmed, et al., 2009;

Hao, et al., 2012; Lo, et al., 2011; Lo and Matthews, 2012; Pansoy, et al., 2010). There are a

total of 24 AHRE core elements (GCGTG) located 10 kb upstream of the mouse Tiparp

transcriptional start site (Hao, et al., 2012). Six of these AHREs are clustered together within a

Page 60: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

45

45 bp fragment including four tandem repeats of AHRE core elements arranged in a concatemer

configuration (Hao, et al., 2012). This AHRE concatemer was bound by activated-AHR and

when expressed as a reporter gene displayed activity, although the level of ligand-induced

activity was not substantial due to high constitutive activity (Hao, et al., 2012).

3.6.1.2 TiPARP Structure

TiPARP encodes a protein of 657 amino acid residues that contains three predicted modular

domains: a C-terminal PARP catalytic domain, a central tryptophan-tryptophan-glutamate

(WWE) domain and a single CX8CX5CX3-type zinc-finger (Figure 9)(Schreiber, et al., 2006).

The PARP catalytic domain shares 43% sequence homology with the catalytic domains of

ARTDs (Ma, et al., 2001). The catalytic triad of the catalytic domain is H-Y-I and TiPARP is

predicted to possess mono-ADP-ribosyltransferase (MART) activity (Otto, et al., 2005). During

its initial characterization, TiPARP was reported to be a catalytically active enzyme; however, its

potential mono-ADP-ribosyltransferase activities were not evaluated and remain to be

determined (Ma, et al., 2001).

The WWE domain is composed of two conserved tryptophan residues and one glutamic acid

residue and is the putative protein-protein interaction motif of TiPARP that also occurs in classes

of proteins associated with ubiquitination (Aravind, 2001). Other ARTD family members

including ARTD8, ARTD11, ARTD12 and ARTD13 contain at least one WWE domain and it

was recently reported that the WWE domain of ARTD11 binds ADP-ribose (He, et al., 2012).

The role of the TiPARP WWE domain is currently unknown.

The CCCH-type zinc finger is a putative RNA binding motif characterized by three cysteines and

one histidine, which together coordinate a zinc ion to form a local peptide structure (Blackshear,

2002; Liang, et al., 2008). ARTD13 contains four CCCH-type zinc fingers in tandem and has

demonstrated to bind specific viral mRNA sequences (Gao, et al., 2002). CCCH-type containing

proteins are best known for their RNA binding ability but they also bind single-stranded DNA

(Bai and Tolias, 1996; Collart, et al., 2005). The functionality and nucleic acid binding potential

of CCCH-type zinc finger of TiPARP have not been determined.

Page 61: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

46

Figure 9. Functional domain structure of TCDD-inducible poly(ADP-ribose) polymerase.

TiPARP/ARTD14/PARP7 contains a single CCCH-type zinc-finger, which is a putative RNA-

binding module, a centrally located WWE (tryptophan-tryptophan-glutamate) domain and a

carboxy-terminal PARP catalytic domain. The catalytic domain contains a ‘HYI’ catalytic triad,

which is found in mono-ADP-ribosyltransferases and not poly-ADP-ribose polymerases.

3.6.1.3 Potential role of TiPARP in TCDD-mediated toxicity

TiPARP has been implicated in the suppression of hepatic gluconeogenesis, a prominent feature

of the lethal wasting syndrome seen in experimental animals exposed to TCDD (Diani-Moore, et

al., 2010). Phosphoenolpyruvate carboxykinase (PEPCK) and glucose-6-phosphatase (G6Pase)

are two key gluconeogenic genes known to be down-regulated by TCDD (Fan and Rozman,

1994; Hsia and Kreamer, 1985; Stahl, et al., 1993; Weber, et al., 1991). Because TCDD

treatment of chick embryo livers reduced NAD+ levels, increased total cellular PARP activity

and TiPARP expression is induced by TCDD, it was hypothesized that TiPARP was responsible

for NAD+ and NADH depletion (Diani-Moore, et al., 2010). The depletion of NAD+ was

suggested to decrease the activity of SIRT1, which is dependent on NAD+. PEPCK and G6Pase

are transcriptionally regulated by peroxisome proliferator-activated receptor γ co-activator 1α

(PGC1α), which is activated by deacetylation by SIRT1 (Diani-Moore, et al., 2010). The

decreased SIRT1 activity would lead to decreased PGC1α activity and decrease PEPCK and

G6Pase expression levels. TiPARP overexpression resulted in decreased glucose output, NAD+

levels, PEPCK mRNA, PGC1α protein levels and increased acetylated PGC1α (Diani-Moore, et

al., 2010). Knockdown of TiPARP resulted in increased glucose output and NAD+ levels (Diani-

Page 62: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

47

Moore, et al., 2010). It appeared that TiPARP mediated the TCDD-induced suppression of

gluconeogenesis and its overexpression mimicked the effects of TCDD in a chick embryo model.

Similarly, TiPARP has been recently suggested to be involved in heightened sensitivity to non-

alcoholic steatohepatitis (NASH) through depletion of NAD+ (He, et al., 2013). Activation of

AHR has been reported to lead to spontaneous fatty liver disease, lipid peroxidation and

oxidative stress, all of which are risk factors for NASH (Dalton, et al., 2002; Lee, et al., 2010;

Lu, et al., 2011). The increase in reactive oxygen species (ROS) production in mouse liver that is

associated with activated-AHR was proposed to be a result of lowered superoxide dismutase 2

(SOD2) activity and compromised clearance of ROS (He, et al., 2013). In another elaborated

model, TiPARP gene induction by TCDD was connected to a decrease in NAD+ levels by

catalytic consumption, this in turn deactivated mitochondrial SIRT3, which reduced SOD2

deacetylation decreasing its activity (He, et al., 2013). The decreased activity of SOD2 led to

increased ROS production and increased sensitivity to NASH (He, et al., 2013).

These two studies based their proposed models of TiPARP-mediated hepatic depletion of NAD+

to TiPARP gene induction by TCDD. Both studies failed to measure increased TiPARP protein

levels following TCDD treatment or increased catalytic activity of TiPARP, making models

highly presumptuous. Additionally, ARTD1 when activated accounts for 90% cellular

poly(ADP-ribose) formation leading one to question whether TiPARP, a predicted MART could

be solely responsible for NAD+ depletion via catalytic consumption (Schreiber, et al., 2002).

Page 63: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

48

Chapter 2: Aims of the present study The ARTD superfamily consists of eighteen members and their potential activities in diverse

cellular processes are just now being explored. TiPARP, the seventh ARTD family member to be

identified, has received little attention since its discovery over ten years ago. TiPARP is better

known as an AHR target gene that is induced by a vast array of AHR ligands in both in vivo and

in vitro models. Despite being an apparently important AHR target gene, the potential role of

TiPARP in AHR-mediated transcriptional regulation is not known. Moreover, the predicted

functional domains of TiPARP also give little insight into a potential role in AHR transcriptional

regulation. The TiPARP catalytic triad motif predicts that TiPARP is a class 2 ARTD with

mono-ADP-ribosyltransferase activity rather than a class 1 poly-ADP-ribosylating ARTD;

however, this has not been experimentally confirmed. Because TiPARP is an important AHR

target and a potential mono-ADP-ribosyltransferase, we were interested in whether TiPARP and

mono-ADP-ribosylation were involved in the AHR-mediated transcription.

The repression of AHR-mediated transcription is an aspect that is less understood than

transactivation. Three mechanisms of negative regulation have been proposed: the targeted

proteasomal degradation of activated-AHR via the 26S proteasome, the AHRR and an

autoregulatory feedback loop involving an unidentified labile factor that promotes the proteolytic

degradation of AHR. Because TiPARP expression is induced by activated-AHR we examined its

potential as negative regulator of AHR transactivation.

The overall objective of this thesis research was to understand TiPARP and its potential role in

AHR transactivation. Specific aims were:

• Demonstrate TiPARP is a mono-ADP-ribosyltransferase

• Determine role of TiPARP in AHR transactivation and understand its mechanism

• Compare TiPARP- and aryl hydrocarbon receptor repressor (AHRR)-mediated regulation

of AHR transactivation

Page 64: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

49

Chapter 3: Materials and Methods 4 Materials

4.1 Chemicals and biological reagents 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) was purchased from Wellington Laboratories

(Guelph, ON, Canada). Dimethyl sulfoxide (DMSO) was purchased from Sigma-Aldrich (St.

Louis, MO). Cell culture reagents: Dulbecco’s Modified Eagle’s medium (DMEM)

supplemented with 1 g/L, high glucose (4.5 g/L glucose) DMEM, 1:1 mixture of DMEM and

Ham’s F-12 nutrient mixture (DMEM/F-12), RPMI-1640 medium, fetal bovine serum (FBS),

penicillin/streptomycin antibiotic mixture, 0.25% Trypsin-EDTA, Phosphate Buffered Saline,

were all purchased from Sigma-Aldrich (St. Louis, MO). Protein A-Agarose Fast Flow 50%

(v/v) was purchased from Sigma-Aldrich. Lipofectamine LTX with Plus reagent was purchased

from Life Technologies (Burlington, ON) and FuGENE HD tranfection reagent was purchased

from Roche (Laval, QC).

Antibodies used for chromatin immunoprecipitation (ChIP), co-immunoprecipitation (IP),

indirect immunofluorescence (IF) or Western blot (WB) experiments were: anti-AHR (H-211,

ChIP, IF, WB), anti-AHR (N-19, WB), anti-ARNT (H-172, ChIP), anti-ARNT (C-19, WB), anti-

rabbit immunoglobin (IgG, sc-2027, ChIP, IP) (all from Santa Cruz Biotechnology, Santa Cruz,

CA), anti-PARP7 (Abcam, 84664, Cambridge, MA, IF), anti-AHRR (HPA019614, ChIP, WB)

and anti-β-actin (A2228, WB) both from Sigma-Aldrich, anti-AHR (SA-210, Biomol

International Inc., Plymouth Meeting, PA), anti-GFP (632460, ChIP) and anti-GFP (JL-8, WB)

both from Clontech (Mountain View, CA). Conjugated secondary antibodies used were:

Horseradish Peroxidase (HRP)-conjugated anti-rabbit IgG whole antibody (NA934, GE

Healthcare; Buckinghamshire, UK), HRP-conjugated normal anti-mouse IgG (sc2954) and HRP-

conjugated normal anti-goat IgG (sc2741) all from Santa Cruz, and HRP-conjugated Clean Blot

IP Reagent (ThermoScientific) and Alexa 568 Fluor-conjugated anti-rabbit IgG (Life

Technologies, Burlington, ON, Canada). ECL Advance Western blotting detection system and

ECL Plus both from GE Healthcare and SuperSignal West Dura (ThermoScientific) were used to

visualize proteins for both Western blots and Co-IP experiments. For purification of PCR

products Buffer PB (Qiagen, Toronto, ON) and EZ-10 Spin Column PCR Purification Kit (Bio

Page 65: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

50

Basic Inc. Markham, ON) were used. For RNA extraction and cDNA synthesis, Aurum™ Total

RNA Mini kit (Bio-Rad, Hercules, CA) and Superscript III reverse transcriptase (Life

Technologies) were used. Pfu Ultra DNA polymerase (Agilent, Mississauga, ON) and Pfx50

DNA polymerase (Life Technologies) were used for site-directed mutagenesis. High Fidelity

PCR enzyme mix (ThermoScientific) was used for amplification of PCR products. QIAquick Gel

Extraction Kit and HiSpeed Plasmid Maxi Kit were purchased from Qiagen. SsoFast EvaGreen®

Supermix (Bio-Rad) was used for all quantitative real-time PCR reactions. ONE-glo™

Luciferase Assay System, TNT® Coupled Reticulocyte Lysate System and pGEM®-T Easy

Vector System were purchased from Promega (Madison, WI). 35S-radiolabelled L-methionine

and 32P-radiolabelled β-NAD+ were purchased from Perkin-Elmer (Woodbridge, ON). All

primers used for real-time PCR and for the generation of expression constructs were synthesized

by Integrated DNA Technologies (IDT, Coralville, IA).

4.2 Plasticware Plasticware used for cell culture was purchased from Sarstedt (Newton, CT) which include: T-

25, T-75, and T-175 tissue culture flasks, 10 cm tissue culture dishes, 6-well tissue culture plates,

and 2 ml cryovials. The 12-well and 96-well clear flat bottom plates were purchased from BD

Biosciences-Falcon (San Jose, CA). The 96-well black flat bottom plates used for luciferase

assays were purchased from Corning (Lowell, MA). The 1.5 ml (Axygen) and 1.7 ml low-

binding (Progene) microcentrifuge tubes were purchased from Ultident Scientific (St. Laurent,

QC). 96-well non-skirted PCR plates were purchased from D-Mark Biosciences (Toronto, ON).

4.3 Instruments For cell culture, HERAcell® 150 incubators (Kendro, Langenselbold, Germany) were used. The

water bath (model # 1228, Sheldon manufacturing, Cornelius, OR) was purchased from VWR

International (Plainfield, NJ). Cells were centrifuged with Centrifuge 5702 (Eppendorf,

Hamburg, Germany). All centrifuge steps in PCR Purification and RNA extraction processes

used Centrifuge 5415D (Eppendorf). Locator JR Cryo Biological Storage System (Thermolyne)

was used to freeze, and store human immortalized cells was purchased from VWR International

(Plainfeld, NJ). To count cells, a Bright-Line Hemocytometer (Hausser Scientific, Horsham, PA)

and Vistavision Light Microscope (VWR) were used. The purity and concentration of RNA and

protein were measured with an Ultrospec 2100 pro Spectrophotometer (Biochrom, Cambridge,

Page 66: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

51

England). Branson Digital Sonifier 450 (Danbury, CT) was used to sonicate cells. The rotator

used for the ChIP assay and Western blotting was from Mini LabRoller (Labnet Inc., Edison,

NJ). Real-time PCR amplification of ChIP DNA and cDNA was performed on Chromo4 Real-

Time PCR detector (Bio-Rad). A DNAEngine Peltier Thermal Cycler (Bio-Rad) was used for

site-directed mutagenesis and cDNA synthesis of extracted RNA. Protein and β-galactosidase

levels were determined using a ThermoScientific 96 well Multiskan EX Photometer (Waltham,

MA). Luciferase levels from reporter gene constructs were measured using GLO-max 96-well

microplate luminometer (Promega). Indirect immunofluorescence images were viewed with

Imager.Z1 epifluorescence microscope and Axiovision software (Zeiss, Toronto, ON) or an

Olympus Fluoview 1000 confocal microscope (Center Valley, PA). Mouse livers were

homogenized on a VWR PowerMax Advanced Homogenizing System 200 (VWR).

5 Methods

5.1 Maintenance of HuH7, COS7, TREx-FLP-IN 293, T47D, NCI-N87 and MCF7 cell lines

HuH7 human hepatoma, T47D human breast carcinoma, MCF7 human breast carcinoma and

NCI-N87 human gastric carcinoma cell lines were purchased from the American Type Culture

Collection (ATCC, Manassas, VA). TREx-FLP-IN 293 human embryonic kidney cells were

purchased from Life Technologies. COS7 African Green Monkey kidney cells were a generous

gift from Dr. Stephane Angers (Faculty of Pharmacy, University of Toronto). All cell lines were

cryopreserved and stored in liquid nitrogen. All cells were ~80% confluent before freezing, cells

were trypsinized and neutralized with fresh medium and centrifuged for 2 min at 4000 rpm. The

cell pellet was re-suspended in a 90% FBS containing 10% DMSO (v/v) solution and aliquoted

into labelled cryovials. Cryovials were immediately frozen at -80ºC and then transferred to liquid

nitrogen for long-term storage. Before use, HuH7, COS7 and TREx-FLN cells were rapidly

thawed at 37°C and transferred to a T-75 tissue culture flask containing high glucose DMEM

supplemented with 10% (v/v) FBS and 1% (v/v) penicillin/streptomycin (PEST). Once cell lines

were at least 80% confluent they were sub-cultured 1:6 to 1:12 twice a week. T47D cells were

thawed and transferred to a T-25 tissue culture flask containing DMEM/F12 medium

supplemented with 10% FBS and 1% PEST. Once cells were at least 80% confluent, cells were

sub-cultured 1:2 to 1:3 twice a week. NCI-N87 cells were thawed and transferred to T-25 tissue

Page 67: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

52

culture flask containing RPMI-1640 supplemented with 10% FBS and 1% PEST. MCF7 cells

were thawed and transferred to T-25 tissue culture flask containing low glucose DMEM

containing 10% FBS and 1% PEST. Once cells were at least 80% confluent they were sub-

cultured 1:2 to 1:3 twice a week. All cell lines were sub-cultured as follows: medium was

aspirated from culture flask and cells were washed once with 5-10 ml of sterile PBS. PBS was

aspirated and 2-4 ml of pre-warmed trypsin-EDTA was added to the flask. The flask was placed

back into incubator for 1-5 min until cells detached completely. An equal volume of pre-warmed

completed medium was added to cell suspension in flask to neutralize the activity of tryspin. Cell

suspension was vigorously pipetted up and down until homogenous then 1-4 ml of the cell

solution was transferred to a T-175 flask containing 40 ml of completed media. All cell lines

were maintained at 37°C and 5% CO2.

5.2 Generation of Expression Constructs All full-length human TiPARP expression constructs, TiPARP N-terminal truncation expression

constructs and the C-terminal truncation 1-234 were created by PCR amplification of TiPARP

cDNA from pCMV6-XL4-hTiPARP (Origene, Rockwood, MD). Two-hundred micrograms of

pCMV6-XL4-hTiPARP was added to a reaction mixture containing 0.2 μM of desired forward

primer containing Kozak sequence and/or EcoRI restriction site and reverse primer containing

SalI restriction cut site (Table 4), 0.2 mM dNTPs and High Fidelity PCR enzyme mix

(ThermoScientific) and amplified using the DNAEngine Peltier Thermal Cycler (Bio-Rad). PCR

products were loaded on a 1% agarose gel and separated by gel electrophoresis. PCR products

were excised from gel and purified using QIAquick Gel Extraction Kit following manufacturer’s

instructions. One hundred nanograms of purified PCR product was ligated into 50 ng pGEM-T

Easy vector with T4 DNA ligase (New England Biolabs, Whitby, ON) in a 10 μl reaction

overnight at room temperature. Ligation reactions were transformed into 100 μl chemically

competent DH5α strain E. coli. Bacteria were plated on LB-ampicillin agar plates containing 40

μg/ml X-gal and incubated overnight at 37°C. Individual white colonies were selected and grown

in 3 ml LB-ampicillin overnight at 37°C with constant shaking. The following morning, plasmid

DNA from selected colonies was purified using GenElute Plasmid Miniprep kit (Sigma)

according to the manufacturer’s instructions. Purified plasmids were digested with EcoRI and

SalI restriction enzymes for 1 h and digestions were loaded on a 1% agarose gel and separated by

Page 68: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

53

gel electrophoresis. Inserts were excised from gel and purified using QIAquick Gel Extraction

Kit (Qiagen). Inserts were ligated into pcDNA3.1 (Life Technologies) using EcoRI and XhoI

restriction sites, pEGFP-C2 (Clontech) using EcoRI and SalI restriction sites and pGEX-4T1

(GE Healthcare) using EcoRI and XhoI restriction sites overnight at room temperature. The

following day ligations were transformed into chemically competent DH5α strain E. coli and

grown on LB-ampicillin plates for pcDNA3.1 and pGEX-4T1 ligations or LB-kanamycin plates

for pEGFP-C2 ligations overnight at 37°C. The following day individual colonies were selected

and grown in LB-ampicillin or LB-kanamycin overnight at 37°C with constant shaking. Plasmids

were purified using GenElute Plasmid Miniprep kit (Sigma) according to the manufacturer’s

instructions. pcDNA3.1 and pGEX-4T1 constructs were screened by restriction digests using

EcoRI and XhoI restriction enzymes and pEGFP-C2 constructs were digested using EcoRI and

SalI. Digests were loaded onto a 1% agarose gel and separated by gel electrophoresis. All

plasmids that screened positive were sent to The Centre for Applied Genomics DNA Sequencing

and Synthesis Facility (Medical and Related Sciences (MaRS) building, Toronto, ON, Canada)

to verify sequence and framing. Mouse Tiparp expression constructs were generated by PCR

amplification of mTiPARP cDNA from pCMV6-kan/neo-mTiPARP (Origene) using forward

primer containing EcoRI restriction site and reverse primer containing XhoI restriction site

(Table 4). Chicken TiPARP expression vectors were generated by PCR amplification of chicken

TiPARP cDNA from total Gallus gallus liver RNA (Zyagen, San Diego, CA) using forward

primer containing EcoRI restriction site and reverse primer containing SalI site (Table 4). Both

chicken and mouse TiPARP inserts were cloned into EcoRI and XhoI sites of pcDNA3.1 and

pGEX-4TI and EcoRI and SalI sites of pEGFP-C2.

pEGFP-C2-AHRR-FL, pEGFP-C2-AHRRΔ8 (delta exon 8), pAcGFP-N1-AHRR-FL and

pAcGFP-N1- AHRRΔ8 were generated by sub-cloning from pcDNA3.1-AHRR-FL and pcDNA-

AHRRΔ8, (both gracious gifts from Dr. Mark Hahn, Woods Hole Oceanographic Institute,

Woods Hole, MA). PCR products were amplified using forward primer containing EcoRI

restriction site and reverse primer containing stop codon and XhoI restriction site or without stop

codon and BamHI restriction site (Table 4). Both AHRR inserts were cloned into EcoRI and SalI

sites of pEGFP-C2 or EcoRI and BamHI sites of pAcGFP-N1 (Clontech).

Page 69: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

54

Table 4. PCR primers used for generation of TiPARP and AHRR inserts.

TiPARP EcoRI 5’ CAAA GAATTC ATGGAAATGGAAACCACCGAACC

TiPARP SalI 3’ CAAA GTCGAC TCAAATGGAAACAGTGTTACTGAC EcoRI TiPARP 33 5’ CAAA GAATTC ATCACTCCATTGAAGACTTGTTTTAAG EcoRI Kozak TiPARP 33 3’ CAAA GAATTC ACCATG

ATCACTCCATTGAAGACTTGTTTTAAG EcoRI TiPARP 53 5’ CAAA GAATTC CTGAGGTCTTTGAGGCCAATATT EcoRI Kozak TiPARP 53 5’ CAAA GAATTC ACCATG CTGAGGTCTTTGAGGCCAATATT EcoRI TiPARP 103 5’ CAAA GAATTC GCAGAAAATAATATGTCTGTTCTGAT EcoRI Kozak TiPARP 103 5’ CAAA GAATTC ACCATG

GCAGAAAATAATATGTCTGTTCTGAT EcoRI TiPARP 200 5’ CAAA GAATTC CTGGACACCAGTGATAAGCTGAGT EcoRI Kozak TiPARP 200 5’ CAAA GAATTC ACCATG

CTGGACACCAGTGATAAGCTGAGT EcoRI TiPARP 218 5’ CAAA GAATTC GCTTCCCTTGACCTCGTGTTTGA EcoRI Kozak TiPARP 218 5’ CAAA GAATTC ACCATG GCTTCCCTTGACCTCGTGTTTGA EcoRI TiPARP 225 5’ CAAA GAATTC GAGCTGGTGAACCAGTTGCAGTAC EcoRI Kozak TiPARP 225 5’ CAAA GAATTC ACCATG

GAGCTGGTGAACCAGTTGCAGTAC EcoRI TiPARP 245 5’ CAAA GAATTC GACTTTCTGCAAGGCACTTGTATTTA EcoRI Kozak TiPARP 245 5’ CAAA GAATTC ACCATG

GACTTTCTGCAAGGCACTTGTATTTA EcoRI TiPARP 275 5’ CAAA GAATTC ACTCAAAAGTGGCAGAGTGTATTC EcoRI Kozak TiPARP 275 5’ CAAA GAATTC ACCATG

ACTCAAAAGTGGCAGAGTGTATTC EcoRI TiPARP 328 5’ CAAA GAATTC CTACGAAGGCTGTCCACACCAC EcoRI Kozak TiPARP 328 5’ CAAA GAATTC ACCATG CTACGAAGGCTGTCCACACCAC EcoRI TiPARP 445 5’ CAAA GAATTC TCAGCAAACTTTTACCCTGAAACTT EcoRI Kozak TiPARP 445 5’ CAAA GAATTC ACCATG

TCAGCAAACTTTTACCCTGAAACTT SalI TiPARP 234 3` CAAA GTCGAC TCAAGTGTGGTACTGCAACTGGTTCAC EcoRI mouse TiPARP 5’ CAAA GAATTC ATGGAAGTGGAAACCACTGAACC XhoI mouse TiPARP 3’ CAAA CTCGAG TCAAATGGAAACAGTGTTACTGACT EcoRI chicken TiPARP 5’ CAAA GAATTC ATGGACACCGAAGCAAAGCCA SalI chicken TiPARP 3’ CAAA GTCGAC TCAAATGGAGACAGTGTTGCTA EcoRI hAHRR 5’ CAAA GAATTC ATGCCGAGGACGATGATCCCGCC XhoI hAHRR stop 3’ CAAA CTCGAG CTATGGCAGGAATGTGCACCCC BamHI AHRR no stop 3’ CAAA GGATCC CG TGGCAGGAATGTGCACCCCGAA

5.3 Site-directed mutagenesis Catalytic, zinc finger, nuclear localization point mutants, double alanine point mutants and C-

terminal 1-448 truncation were generated by site-directed mutagenesis. Fifty nanograms of

pcDNA-TiPARP, pEGFP-C2-TiPARP and pGEX-4T1-TiPARP was added to a reaction mixture

containing 0.2 mM dNTPs, 0.2 μM forward and reverse primers containing desired mutation

Page 70: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

55

(Table 5), Pfu Ultra DNA polymerase (Agilent) or Pfx50 DNA polyermase (Life Technologies)

with 0.5 mM MgCl2 and amplified using the DNAEngine Peltier Thermal Cycler (Bio-Rad).

PCR products were digested by 1 μl DpnI restriction enzyme (New England BioLabs) for 1 h at

37°C and 10 μl of digested PCR product was transformed into chemically competent DH5α E.

coli. Bacteria were plated on LB-ampicillin or LB-kanamycin agar plates and incubated

overnight at 37°C. The following day colonies were selected and grown overnight at 37°C under

constant agitation. The following morning plasmid DNA from selected colonies was purified

using GenElute Plasmid Miniprep kit (Sigma) according to the manufacturer’s instructions.

Purified plasmid DNA was sent to The Centre for Applied Genomics DNA Sequencing and

Synthesis Facility (Medical and Related Sciences (MaRS) building, Toronto, ON, Canada) to

verify mutations.

Table 5. PCR primers used for site-directed mutagenesis of TiPARP.

TiPARP K41A 5’ ATTGAAGACTTGTTTTGCGAAAAAGGATCAGAAAAGA TiPARP K41A 3’ TCTTTTCTGATCCTTTTTCGCAAAACAAGTCTTCAAT TiPARP K41A; K42A; K43A 5’

CCATTGAAGACTTGTTTTGCGGCAGCGGATCAGAAAAGATTGGG

TiPARP K41A; K42A; K43A 3’

CCCAATCTTTTCTGATCCGCTGCCGCAAAACAAGTCTTCAATGG

TiPARP C243A 5’ AACGGAATTGAAATTGCCATGGACTTTCTGCAAGG TiPARP C243A 3’ CCTTGCAGAAAGTCCATGGCAATTTCAATTCCGTT TiPARP C251A 5’ CTTTCTGCAAGGCACTGCTATTTATGGCAGGGATT TiPARP C251A 3’ AATCCCTGCCATAAATAGCAGTGCCTTGCAGAAAG TiPARP C257A 5’ TATTTATGGCAGGGATGCTTTGAAGCACCACACTGT TiPARP C257A 3’ ACAGTGTGGTGCTTCAAAGCATCCCTGCCATAAATA TiPARP H532A 5’ AGACATTTATTTGCTGGAACATCCCAGG TiPARP H532A 3’ CCTGGGATGTTCCAGCAAATAAATGTCT TiPARP Y564A 5’ TTGGACAAGGCAGTGCTTTTGCAAAGAAGG TiPARP Y564A 3’ CCTTCTTTGCAAAAGCACTGCCTTGTCCAA TiPARP I631A 5’ TTTCTTTGAGCCTCAGGCTTTTGTCATTTTTAATG TiPARP I631A 3’ CATTAAAAATGACAAAAGCCTGAGGCTCAAAGAAA TiPARP I631E 5’ TTTCTTTGAGCCTCAGGAATTTGTCATTTTTAATGATG TiPARP I631E 3’ CATCATTAAAAATGACAAATTCCTGAGGCTCAAAGAAA TiPARP 449 stop 5’ TTCAGCAAACTTTTAGCCTGAAACTTGG TiPARP 449 stop 3’ CCAAGTTTCAGGCTAAAAGTTTGCTGAA A218; S219A 5’ AGGACAAAAGTGAAGAGGCTGCCCTTGACCTCGTGTTTGAGCTG A218; S219A 3’ CAGCTCAAACACGAGGTCAAGGGCAGCCTCTTCACTTTTGTCCT L220A; D221A 5’ ACAAAAGTGAAGAGGCTTCCGCTGCCCTCGTGTTTGAGCTGGTGAA L220A; D221A 3’ TTCACCAGCTCAAACACGAGGGCAGCGGAAGCCTCTTCACTTTTGT L222A;V223A 5’ GTGAAGAGGCTTCCCTTGACGCCGCGTTTGAGCTGGTGAACCAGTT

Page 71: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

56

L222A; V223A 3’ AACTGGTTCACCAGCTCAAACGCGGCGTCAAGGGAAGCCTCTTCAC F224A; E225A 5’ AGGCTTCCCTTGACCTCGTGGCTGCGCTGGTGAACCAGTTGCAGTA F224A; E225A 3’ TACTGCAACTGGTTCACCAGCGCAGCCACGAGGTCAAGGGAAGCCT L226A; V227A 5’ CCCTTGACCTCGTGTTTGAGGCGGCGAACCAGTTGCAGTACCACAC L226A; V227A 3’ GTGTGGTACTGCAACTGGTTCGCCGCCTCAAACACGAGGTCAAGGG N228A; Q229A 5’ ACCTCGTGTTTGAGCTGGTGGCCGCGTTGCAGTACCACACTCACCA N228A; Q229A 3’ TGGTGAGTGTGGTACTGCAACGCGGCCACCAGCTCAAACACGAGGT L230A; Q231A 5’ TGTTTGAGCTGGTGAACCAGGCGGCGTACCACACTCACCAAGAGAA L230A; Q231A 3’ TTCTCTTGGTGAGTGTGGTACGCCGCCTGGTTCACCAGCTCAAACA Y232A; H233A 5’ AGCTGGTGAACCAGTTGCAGGCCGCCACTCACCAAGAGAACGGAAT Y232A; H233A 3’ ATTCCGTTCTCTTGGTGAGTGGCGGCCTGCAACTGGTTCACCAGCT T234A; H235A 5’ TGAACCAGTTGCAGTACCACGCTGCCCAAGAGAACGGAATTGAAAT T234A; H235A 3’ ATTTCAATTCCGTTCTCTTGGGCAGCGTGGTACTGCAACTGGTTCA Q236A; E237A 5’ AGTTGCAGTACCACACTCACGCAGCGAACGGAATTGAAATTTGCAT Q236A; E237A 3’ ATGCAAATTTCAATTCCGTTCGCTGCGTGAGTGTGGTACTGCAACT N238A; G239 5’ AGTACCACACTCACCAAGAGGCCGGAATTGAAATTTGCATGGAC N238A; G239 3’ GTCCATGCAAATTTCAATTCCGGCCTCTTGGTGAGTGTGGTACT I240A;E241A 5’ ACACTCACCAAGAGAACGGAGCTGCAATTTGCATGGACTTTCTGCA I240A;E241A 3’ TGCAGAAAGTCCATGCAAATTGCAGCTCCGTTCTCTTGGTGAGTGT

5.4 Transient transfection of HuH7 cells for RNA expression, Western blot and ChIP assays

HuH7 cells were seeded in six-well plates (35 mm diameter) at a density of 2.0 x 105 cells/well.

Following 24 h cells were transfected with 3 μg pcDNA or pcDNA-TiPARP or pEGFP or

pEGFP-C2-TiPARP using 6 μl Lipofectamine LTX and 6 μl Plus reagent. Medium was changed

the following day and cells were treated with 10 nM TCDD or DMSO (0.1%) for 6 h and RNA

isolation was performed (see below). For ChIP assays transfected cells were treated with 10 nM

TCDD or DMSO for 45 min and cells were harvested for ChIP assays (see below). For Western

blots, HuH7 were transfected with 2 μg pEGFP-TiPARP or pEGFP-TiPARP point mutant or

truncation construct with 4 μl Lipofectamine LTX + 4 μl Plus reagent. Following 24 h medium

was changed and proteins were overexpressed for another 24 h and cells were prepared for

protein extraction.

5.5 Transient transfection of T47D cells for Western blot T47D cells were plated in six-well plates at a density of 3.0 x 105 cells/well. Following 24 h cells

were transfected with 3 μg pcDNA or pcDNA-TiPARP using 6 μl Lipofectamine LTX and 6 μl

Page 72: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

57

Plus reagent. Medium was changed the following day and cells were treated for 0-24 h with 10

nM TCDD or DMSO and cells were harvested for protein extraction.

5.6 Transient transfection of HuH7 cells and reporter gene assays

HuH7 cells were plated in 12-well plates (22 mm diameter) at a density of 1.2 x 105 cells/well.

The following day cells were transfected with 50-1500 ng pcDNA-TiPARP and/or 250-750 ng

pRc-CMV-hAHR (a generous gift from Dr. Patricia Harper) and/or 250-750 ng pcDNA4-

hARNT (a generous gift from Dr. Kevin Gardner) or 200 ng pcDNA-mTiparp, pcDNA-

chTiPARP, 200-1500 ng pCMV-XL-ARTD1 or pCMV-XL-ARTD12 (Origene) or 200 ng

pcDNA-TiPARP truncation or point mutant plasmids or 5-100 ng pcDNA-AHRRΔ8 or 500-100

ng pcDNA-AHRR-FL and 200 ng pCYP1A1-luc or pGL3-CYP1B1-luc plasmid using 2-4 μl

Lipofectamine LTX and 2-4 μl Plus reagent. All reactions included 50 ng of pCH110-β-Gal (GE

Healthcare) to normalize for transfection efficiency. pcDNA3.1 was used as carrier DNA for all

transfection experiments. After 24 h cells were treated with TCDD or DMSO for 24 h and cells

were washed once with 1 ml PBS and lysed with 250 μl 1X Passive Lysis Buffer (Promega) for

15 min with gentle rocking. Twenty-five microliters of lysate were loaded in duplicate into a 96-

well black flat bottom plate and 25 μl ONE-Glo luciferase substrate (Promega) were added to

each well and luciferase activity was determined using a GLO-max luminometer (Promega). For

β-galactosidase activity 10 μl of lysate were loaded into a 96-well clear flat bottom plate in

duplicate and 100 μl β-galactosidase buffer (0.6 M Na2HPO4, 0.04 M NaH2PO4, 0.01 M KCl, 1

mM MgSO4, 100 mM β-mercaptoethanol) and 2.5 mM ONPG (ortho-Nitrophenyl-β-galactoside)

were added to lysate and plate was incubated at 37ºC for 30 min and reaction was stopped

addition of 160 mM Na2HCO3 to reaction and 420 nm absorbance was measured using

Multiskan EX Photometer (ThermoScientific). Luciferase activity was normalized to

corresponding β-galactosidase activity.

5.7 Transient transfection and RNAi T47D and NCI-N87 cells were plated in 6-well plates at a density of 3.0 x 105 cells/well. HuH7

cells were plated in in 6-well plates at a density of 2.0 x 105 cells/well. MCF7 cells were plated

in 6-well plates at a density of 3.5 x 105 cells/well. Twenty-four hours after plating T47D, HuH7

and NCI-N87 cells were transfected with 100 nM SMARTpool: ON-TARGETplus TiPARP

Page 73: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

58

siRNAs (siTiPARP; L-013948-00) or ON-TARGETplus Non-targeting Pool (NTP; D-0011810-

10) siRNAs with 4 μl DharmaFECT1 (Dharmacon, Lafayette, CO). T47D cells were transfected

with 100 nM ON-TARGETplus TiPARP (siTiPARP-12; J-013948-12, siTiPARP-13; J-013948-

13) siRNAs or ON-TARGETplus non-targeting siRNA #2 (NT; D-0018100-02) (Dharmacon)

with 4 μl DharmaFECT1. MCF7 cells were transfected with 50 nM MISSION AHRR siRNAs

(siAHRR1; SASI_Hs02_00353810, siAHRR2; SASI_Hs01_00213384, Sigma-Aldrich) or NT2

(Dharmacon) with 2 μl DharmaFECT1. Medium of all cell lines was changed 24 h following

transfection. TiPARP was knocked down for 48 h post-transfection and dosed with 10 nM

TCDD or DMSO for 1.5, 3 h and 24 h and cells were harvested for Western blot, ChIP and RNA

expression assays. AHRR was knocked down for 24 h post-transfection and dosed with 10 nM

TCDD or DMSO for 24 h and cells were harvested for protein and RNA expression assays.

5.8 mRNA time-course T47D cells were seeded in 6-well plates at a density of 3.0 x 105 cells/well. Following 24 h cells

were treated with 1 μg/ml actinomycin D (Sigma) for 1 h to 24 h or 10 nM TCDD for 0.25 h to

24 h. MCF7 cells were seeded in 6-well plates at a density of 3.5 x 105 cells/well. Following 24 h

cells were treated with 10 nM TCDD or DMSO for 0.25 h to 24 h.

5.9 Tiparp-null mice and generation of mouse embryonic fibroblasts (MEF)

Tiparp-/- mice (strain B6;129S4-TiparpGt(ROSA)79Sor) were purchased from Jackson Laboratory

(Bar Harbor, ME) and housed and bred at the Division of Comparative Medicine (University of

Toronto) under the management of our lab technician, Debbie Brenneman. Six to eight week old

male and female wildtype (Tiparp+/+), knockout (Tiparp-/-) and heterozygous (Tiparp+/-) mice

were treated with a single dose of 30 μg/kg b.w. TCDD or corn oil (control) by intraperitoneal

injection. Mice were sacrificed after 6 h of treatment and livers excised and snap frozen in liquid

nitrogen and transferred to a -80ºC freezer. Four mice of each group were tested.

Wildtype, Tiparp-deficient (Tiparp-/-) and Tiparp+/- fibroblasts were prepared from E14.5

embryos derived from matings of mice heterozygous for disruption of the Tiparp allele. Primary

fibroblasts from wildtype, Tiparp-/- and Tiparp+/- siblings were immortalized at passage 2 after

transfection with Simian virus large T antigen (SV40gp6) in pSG5 (Stratagene) and a puromycin

Page 74: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

59

resistance plasmid, and selected in puromycin-containing medium. The genotypes of the MEFs

were verified by PCR. Once immortalized, all MEF cell lines were plated into 10 cm dishes at a

density of 2.0 million cells/dish. Twenty-four hours after plating cells were treated with 10 nM

TCDD or DMSO for 45 min for ChIP assays and 24 h for Western blots. For RNA expression

experiments MEFs were seeded in 6-well plates at a density of 1.5 x 105 cells/well and treated 24

h following plating with 10 nM TCDD or DMSO for 6 h or 24 h or pre-treated with 10 μg/ml

cycloheximide (CHX, Sigma) for 1 h followed by 10 nM TCDD for 6 h or pre-treated with 10

μM MG132 (Sigma) for 30 min followed by 10 nM TCDD for 4 h. Following all treatments cells

were harvested for RNA isolation.

5.10 RNA isolation and cDNA synthesis

5.10.1 Mammalian cells For RNA isolation of all transfected and/or treated cells, medium was aspirated and cells were

washed once with PBS. Three hundred and fifty microliters of Total RNA Lysis Solution (Bio-

Rad) containing 1% β-mercaptoethanol was added to each well to lyse cells. Following 10 min

cell lysates were scraped and transferred to microcentrifuge tubes. An equal volume of 70%

ethanol was added to each tube and was mixed by gentle pipetting to bind RNA. RNA was

purified using Aurum™ Total RNA Mini Kit (Bio-Rad) following the manufacturer’s protocol.

Briefly, RNA was bound to spin columns by centrifugation at 13,000 rpm for 30 s. The column

membranes were washed once with Total RNA Wash Low Stringency solution and 80 μl DNAse

I solution (DNAse I stock solution diluted 1:40 in DNAse I Dilution Solution) was added

directly to column membranes and allowed to digest for 15 min at room temperature. DNAse I

was removed by addition of Total RNA Wash High Stringency solution followed by

centrifugation at 13,000 rpm for 30 s. Column membranes were then washed with Total RNA

Wash Low Stringency solution and centrifuged at 13 000 rpm for 1 min. Columns were dried by

centrifugation at 13,000 rpm for 2 min prior to elution. To elute, columns were transferred to

new microcentrifuge tubes and 40 μl of Total RNA Elution Solution was placed directly onto

column membrane and allowed to incubate for 1 min. Tubes were centrifuged at 13,000 rpm for

2 min. RNA concentration was measured using the spectrophotometer (Biochrom) and all

samples were set to 50 ng/μl concentrations by addition of DNAase/RNAse-free distilled water.

Page 75: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

60

5.10.2 Mouse liver

Approximately 40 mg of snap frozen liver was thawed on ice and re-suspended in 700 μl Total

RNA Lysis Solution (Bio-Rad) containing 1% β-mercaptoethanol and homogenized for 20 s

using a VWR PowerMax AHS homogenizer. Homogenates were centrifuged at 13,000 rpm for 3

min and supernatant was transferred to a new microcentrifuge tube. An equal volume of 70%

ethanol was added to supernatant and mixed by gentle pipetting until homogenous. RNA was

purified using Aurum™ Total RNA Mini Kit (Bio-Rad) following the manufacturer’s protocol.

5.10.3 cDNA synthesis

For cDNA synthesis, 500 ng of extracted RNA was reversed transcribed using 2500 U

SuperScriptIII (Life Technologies) in a total volume of 20 μl containing 2.5 μM random

hexamers and 1 μM dNTPs. cDNA was synthesized for 1 h at 50°C then heated to 70°C for 15

min to inactivate enzyme. After synthesis cDNA samples were diluted with 60 μl

DNAse/RNAse-free distilled water. Real-time PCR (qPCR) was performed on 1 μl of cDNA

synthesis sample using SsoFast EvaGreen® SYBR Supermix (Bio-Rad) and primers to amplify

CYP1A1, CYP1B1, TiPARP, AHRR, NFE2L2, NQO1 and β-actin mRNAs (Table 6) using

Chromo4 Real-Time PCR detector (Bio-Rad) and analyzed using Opticon Monitor 3 (Bio-Rad).

All target gene transcripts were normalized to β-actin mRNA and run in triplicate. Fold change

was calculated using the difference between cycle threshold (Ct) value of treatment and control

using the formula comparative method. ΔCt sample - ΔCt control = ΔΔCt. The expression level

of each gene was then calculated using 2-ΔΔCt and compared to DMSO or 0 h treatment (=1). PCR

was performed using specific primers (Table 6). Cycling conditions consisted of an initial

denaturation step of 95°C for 3 min followed by 45 cycles of 95 °C for 15 s, and the 60°C

annealing temperature for 30 s.

Table 6. List of qPCR primers used in RNA expression assays.

hCYP1A1 mRNA 5' TGGTCTCCCTTCTCTACACTCTTGT hCYP1A1 mRNA 3' ATTTTCCCTATTACATTAAATCAATGGTTCT hCYP1B1 mRNA 5’ CCTATGTCCTGGCCTTCCTTT hCYP1B1 mRNA 3’ TGAGGAATAGTGACAGGCACAAA TiPARP mRNA 5' AGGGACCACTTTGGATGGAGAGAGT TiPARP mRNA 3' TCAGACCCCGAGAGTTGGCTT hAHRR mRNA 5’ CCCCGCCCTTGGAGACAGGA

Page 76: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

61

hAHRR mRNA 3’ AGTACTCGGTGGGCGTGCCT

β-actin mRNA 5’ CAGAGCAAGAGAGGCATCCT

β-actin mRNA 3’ GGTCTCAAACATGATCTGGGTC mCYP1A1 mRNA 5’ CGTTATGACCATGATGACCAAGA mCYP1A1 mRNA 3’ TCCCCAAACTCATTGCTCAGAT mCYP1B1 mRNA 5’ CCAGATCCCGCTGCTCTACA mCYP1B1 mRNA 3’ TGGACTGTCTGCACTAAGGCTG mTiPARP mRNA 5’ CCTTCAACCAAGAACAAAGCATCT mTiPARP mRNA 3’ GGATGTCTGTATTTTGTGGCTGAAT mAHRR mRNA 5’ CTGCCCAGGTACTCTGAACC mAHRR mRNA 3’ ACTGTCCACAAAGCCTGACC mNFE2L2 mRNA 5` GTGACGGTGGCAGCATGCCT mNFE2L2 mRNA 3` TGCCTTCAGCGTGCTTCGGG mNQO1 mRNA 5` GCAGGATTTGCCTACACATATGC mNQO1 mRNA 3` AGTGGTGATAGAAAGCAAGGTCTTC

5.11 Chromatin Immunoprecipitation (ChIP) For ChIP assays, all three MEF cell lines were plated in 10 cm dishes at a density of 2.0 million

cells/dish. MCF7 cells were plated in 10 cm dishes at a density of 3.5 million cells/dish. Cells

were grown for 24 h and treated with 10 nM TCDD or DMSO for 45 min. Treated HuH7 cells

overexpressing EGFP-TiPARP or EGFP and treated T47D cells with TiPARP knockdown (from

sections 5.4 and 5.7 respectively) were used for ChIP assays. Following treatment protein-DNA

complexes were cross-linked with 1% formaldehyde (Sigma) for 10 min and then quenched with

125 mM glycine for 5 min. The medium was aspirated and cells were washed twice with 2 ml

PBS. Five hundred microliters of PBS + 0.1% (v/v) Tween 20 (PBS/T) was added to each well

and cells were scraped and centrifuged for 5 min at 7 500 rpm at 4°C. Supernatant was aspirated

off and cells were re-suspended in ice cold 1 ml Buffer A (10 mM HEPES [pH 7.9], 1.5 mM

MgCl2, 10 mM KCl, 0.05% NP-40) containing 1X protease inhibitor cocktail (PIC, Sigma) and

rotated at 4ºC for 15 min to remove cytosol. Nuclei were pelleted by centrifugation at 7,500 rpm

for 5 min and supernatant containing cytosolic fraction was aspirated. Nuclei were re-suspended

in 400 μL TSE I buffer (50 mM Tris-Base [pH 8.0], 150 mM NaCl, 1 mM EDTA, 1% Triton X-

100) containing 1X PIC and incubated on ice for 10 min. Nuclei were sonicated on ice for 10

times for 10 sec each time, resulting in the shearing of chromatin to an average size of ~500 bp.

Soluble chromatin was collected by centrifugation at 13 000 rpm for 10 min at 4°C and

supernatant transferred to a new microcentrifuge tube. A volume of 30 μl of Protein A-Agarose

(50% slurry) was added to chromatin and incubated at 4°C under gentle rotation for 2 h. An

Page 77: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

62

aliquot of chromatin was stored at -20°C for the 5% total input sample. Agarose beads were

pelleted by centrifugation and chromatin was aliquoted into fresh microcentrifuge tubes

containing 1 μg of antibody, 0.5 μg/μl BSA and 0.05 μg/μl salmon sperm DNA in TSE I buffer.

Chromatin was immunoprecipitated overnight at 4°C under gentle rotation. The following

morning Protein A-Agarose (25 μl of 50% slurry) was added and chromatin was incubated for

another 1.5 hours at 4°C under gentle rotation. Protein A beads were pelleted by centrifugation

and washed successively for 10 min in 3 X 1 ml of TSE I, 1 ml of TSE II (20 mM Tris-Base [pH

8.0], 500 mM NaCl, 2 mM EDTA, 1% Triton X-100, 0.1% SDS), 1 ml of LiCl buffer (20 mM

Tris-Base [pH 8.0], 250 mM LiCl, 1 mM EDTA, 1% NP-40, 1% Na-deoxycholate), and 2 X 1 ml

of TE (10 mM Tris-Base [pH 8.0], 1 mM EDTA). Protein-DNA complexes were eluted in 110 µl

of elution buffer (TE + 1% SDS) for 30 min under constant rotation, and the cross-links were

reversed by overnight incubation at 65°C. The following morning, eluted DNA was bound using

5X volume of Buffer PB (Qiagen) and purified using EZ-10 Spin Column PCR Purification Kit

(Bio Basic) according to manufacturer’s instructions with slight modification. Briefly, bound

DNA was added to column and centrifuged at 11 000 rpm for 1 min. Columns were washed

twice with wash solution (containing 95% ethanol) and dried by centrifugation. DNA was eluted

from column using 100 μl of elution buffer. ChIP DNA was quantified by qPCR using SsoFast

EvaGreen® SYBR Supermix (Bio-Rad) and primers (Table 7) to amplify human CYP1A1 and

CYP1B1 enhancer regions and mouse Cyp1a1 and Cyp1b1 enhancer regions and downstream

negative binding regions of each gene. ChIP DNA samples were run in triplicate on Chromo4

Real-Time PCR detector (Bio-Rad) and analyzed using Opticon Monitor 3 (Bio-Rad). Results

were normalized to 5% total input and reported as percent recruitment relative to 100% total

input.

Table 7. List of qPCR primers used for ChIP assays.

hCYP1A1 enhancer 5’ AGGCGTGGACCGAAAATG hCYP1A1 enhancer 3’ CTAGGTCTGCGTGTGGCTTCT hCYP1B1 enhancer 5’ ATATGACTGGAGCCGACTTTCC hCYP1B1 enhancer 3’ GGCGAACTTTATCGGGTTGA hCYP1A1 5 782 bp downstream 5’ TGGTCTCCCTTCTCTACACTCTTGT hCYP1A1 5 782 bp downstream 3’ ATTTTCCCTATTACATTAAATCAATGGTTCT hCYP1B1 8 437 bp downstream 5’ AGTACATTCTGGGGAAACAACA hCYP1B1 8 437 bp downstream 3’ TTTTGGTAATGGTGTCCCAGT mCyp1a1 enhancer 5’ GAGGATGGAGCAGGCTTACG mCyp1a1 enhancer 3’ GGGCTACAAAGGGTGATGCTT

Page 78: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

63

mCyp1b1 enh XRE 5’ rt GGAGCCGACTTCCCAGAAG mCyp1b1 enh XRE 3’ rt AACAAACGGTTGGGTTGAAAAG mCyp1a1 12 791 bp downstream 5’ CGTTATGACCATGATGACCAAGA mCyp1a1 12 791 bp downstream 3’ TCCCCAAACTCATTGCTCAGAT mCyp1b1 7 576 bp downstream 5’ GTGCGGCAAAAGCATGTCTC mCyp1b1 7 576 bp downstream 3’ GGGGAAAAGCAACGTTCTGAC

5.12 Western blot

All transfected and/or treated cells were washed once with PBS and scraped in 350-750 μl

PBS/T and cells were pelleted by centrifugation at 7 500 rpm for 5 min at 4ºC and supernatants

were aspirated and pellets were frozen at -80ºC. For detection of AHR, cell pellets were thawed

on ice and re-suspended in 200 μl cold Cell Lysis Buffer (50 mM HEPES [pH 7.5], 10%

glycerol, 100 mM KCl, 2 mM EDTA, 0.1% NP-40) containing 1X PIC and 2 mM DTT. Cell

suspensions were incubated on ice for 10 min then sonicated for 15 s at 30% amplitude to lyse

cells. Cell lysates were placed back on ice and incubated for 10 min and soluble proteins were

quantified by Bradford assays. Fifty micrograms of cell lysates were resolved by SDS-8%-PAGE

and transferred to a nitrocellulose blotting membrane (Pall Life Sciences) in 25 mM Tris base

(pH 8.3) containing 19.2 mM glycine and 20% (v/v) methanol. The membranes were blocked in

2% fat-free milk PBS/T for 2 h at room temperature with constant rocking and then

immunoblotted with a 1:5000 dilution of anti-AHR (H-211, Santa Cruz) or 1:10 000 AHR (SA-

210, Biomol International Inc.) antibody in 2% fat-free milk PBS/T overnight at 4ºC with

constant rocking. The membrane was then washed with PBS/T for 30 min and then incubated

with ECL Anti-rabbit IgG, Horseradish Peroxidase linked antibody (NA934, GE Healthcare) for

1 h at room temperature with constant rocking. Proteins were visualized using ECL Select

Western Blotting Detection System (GE Healthcare) according to the manufacturer’s

instructions. The membranes were exposed to an autoradiography film for 1 to 5 min. β-actin

was used as a loading control. AHR protein levels were normalized to β-actin levels and

quantified using ImageJ analysis software (NIH).

For detection of AHRR, ARNT or GFP-TiPARP fusions, frozen cell pellets were thawed on ice

and then re-suspended in 100 μl cold TENG buffer (50 mM Tris-Base [pH 8.0], 2 mM EDTA,

150 mM NaCl, 5% glycerol, 1% NP-40) containing 2.5X PIC and 2 mM DTT. Cell suspensions

were incubated on ice for 10 min and then frozen at -80ºC for 10 min to lyse cells. Lysates were

thawed on ice for 10 min and then centrifuged at 13 000 rpm for 10 min at 4ºC to pellet cell

Page 79: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

64

debris. Supernatants were collected and transferred to new microcentrifuge tubes and solubilized

proteins were quantified by Bradford assays. For AHRR, ARNT and GFP blots 200 μg whole

cell lysates were resolved on by SDS-8%-PAGE transferred to a nitrocellulose blotting

membrane as described above. For AHRR immunoblots, membranes were blocked for 2 h at

room temperature in 2% fat-free milk TNT buffer (100 mM Tris-Base [pH 7.5], 150 mM NaCl,

0.1% (v/v) Tween 20) and then immunobloted with 1:2000 anti-AHRR antibody (HPA019614,

Sigma) in 2% milk TNT overnight at 4ºC with constant rocking. For ARNT immunoblots

membranes were blocked with 5% milk TNT for 2 h at room temperature then immunoblotted

with 1:200 anti-ARNT antibody (C-19, Santa Cruz) in 5% milk TNT overnight at 4ºC with

constant rocking. The membranes were washed with TNT for 30 min and then membranes were

incubated with ECL Anti-rabbit IgG, Horseradish Peroxidase linked antibody (GE Healthcare)

for AHRR immunoblots or HRP-conjugated normal anti-goat IgG (sc2741, Santa Cruz) for

ARNT immunoblots for 1 h at room temperature with constant rocking. Proteins were visualized

using SuperSignal West Dura (Thermo Scientific) and ECL Plus Western Blotting Detection

System (GE Healthcare) for AHRR and ARNT immunoblots respectively, according to the

manufacturers’ instructions. The membranes were exposed to an autoradiography film for 1 to 5

min. For GFP immunoblots membranes were blocked in 5% fat-free milk PBS/T for 2 h and

immunoblotted for GFP-fusion proteins using 1:2000 anti-EGFP antibody (JL-8, Clontech) in

2.5% milk PBS/T overnight at 4ºC with constant rocking. The following morning membranes

were washed with PBS/T for 30 min and membranes were incubated with HRP-conjugated

normal anti-mouse IgG (sc2954) in 2.5% milk PBS/T for 1 h at room temperature with constant

rocking. Proteins were visualized using SuperSignal West Dura (Thermo Scientific) according to

manufacturer’s instructions. Membranes were exposed to autoradiography film for 1 to 5 min.

5.13 Co-immunoprecipitation TREx-FLP-IN 293 cells were seeded in 6-well plates at a density of 4.0 x 105 cells/well were

transfected with 650 ng pRc-CMV-AHR, pRc-CMV-AHR-1-425, pCMV-AHR-TAD (464-848)

(all generous gifts from Dr. Patricia Harper), 200 ng pcDNA4-hARNT and 650 ng pEGFP-

TiPARP-FL or 650 ng pEGFP-TiPARP truncation with 3 μl FuGENE HD (Roche) according to

the manufacturer’s instructions. Twenty-four hours following transfection cells were treated with

10 nM TCDD or DMSO for 1.5 h and cells were scraped in PBS/T and centrifuged at 7 500 rpm

for 5 min at 4ºC. Whole cell extracts were prepared by two freeze-thaw cycles using 200 μl ice

Page 80: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

65

cold TENG buffer containing 1X PIC and 2 mM DTT as described above. Two micrograms of

anti-AHR antibody (H-211) or anti-GFP (632460, Clontech) 1:200 dilution were incubated with

whole cell extracts for 1 h in 4ºC with constant rotation followed by incubation with protein-A

beads for 1 h. Beads were washed five times with NP-40 buffer (20 mM Tris Base [pH 8.0], 150

mM NaCl, 10% glycerol, 1% NP-40, 2 mM EDTA) and eluted in 1X sample buffer, separated by

SDS-PAGE and transferred to nitrocellulose membrane. Membranes were exposed to anti-GFP

(JL-8, Clontech), anti-AHR (H-211) or anti-AHR (N-19) antibodies. All membranes were then

washed and incubated with the appropriate secondary antibodies for 1 h at room temperature.

After washing, bands were visualized using SuperSignal West Dura substrate or ECL Select or

ECL Plus.

5.14 In vitro translation and 35S PARP assays All pcDNA-TiPARP constructs were in vitro translated and labelled with 35S-methionine using

the TNT® Coupled Reticulocyte Lysate System according to manufacturer’s instructions to

verify expression. Briefly, one microgram of pcDNA-TiPARP construct was added to rabbit

reticulocyte lysate containing 1X TNT buffer, 1 mM amino acid mixture minus methionine, 40 U

RNAseOUT (Invitrogen), T7 RNA polymerase and 35S-methionine (Perkin-Elmer) and reactions

were incubated at 30ºC for 90 min. Five percent of each reaction was diluted in TENG buffer

and 1X loading dye was added to each tube. Proteins were resolved by SDS-8%-12% PAGE and

gels were incubated in KODAK™ ENLIGHTNING™ Rapid Autoradiography Enhancer

(Perkin-Elmer) for 30 min with gentle rocking. Gels were then incubated in gel drying solution

(10% (v/v) ethanol and 5% acetic acid) for 20 min with gentle rocking and gels were dried at

80ºC for 2 h. Proteins were visualized by autoradiography.

35S PARP assays were carried out in a 30 μl reaction volume at 30ºC in 1X PARP buffer (5 mM

Tris Base [pH 8.0], 0.2 mM DTT, 4 mM MgCl2, 1% BSA) or 500 μM β-NAD+ (Sigma) or 1 μg

activated DNA (Sigma) with 5% in vitro translated proteins. Reactions were stopped with 1X

sample buffer and resolved by SDS-8%-12% PAGE.

5.15 Indirect immunofluorescence Glass cover slips (VWR No. 1, 22 mm x 22 mm) were sterilized in 70% ethanol for 20 min in the

cell culture hood. Cover slips were transferred to 6-well cell culture plates and allowed to dry

Page 81: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

66

against well wall for 15 min. HuH7 and COS7 cells were seeded onto sterile cover slips at a

density of 2.0 x 105 cells/well and MCF7 cells were seeded onto cover slips at a density of 2.0 x

105 cells/well and cells were allowed to adhere overnight. Cells were transfected with 700 ng

pEGFP-TiPARP constructs and/or 500 ng pRc-CMV-AHR and 300 ng pcDNA4-ARNT with 3

μl Lipofectamine LTX or 100 ng pEGFP-AHRR constructs or 500 ng pAcGFP-AHRR

constructs and all proteins were allowed to overexpress for 24 h. Cells co-transfected with

pEGFP-TiPARP, pRc-CMV-AHR and pcDNA4-ARNT were treated with 10 nM TCDD for 1.5

h. Cover slips were washed once with PBS and fixed in 4% paraformaldehyde for 10 min. Cells

were PBS washed three times for 5 min each and then incubated with PBS + 0.4% Triton-X in

PBS for 20 min to permeabilize nuclei. Cover slips were washed three times with PBS for 5 min

and mounted with Vectashield containing 4,6-diamidino-2-phenylindole (DAPI) (Vector

Laboratories).

For cells co-overexpressed with GFP-TiPARP constructs, AHR and ARNT coverslips were

blocked in 1% goat serum/0.2% Triton-X in PBS for 1 h and incubated with 1:100 anti-AHR

antibody (H-211, Santa Cruz) in 1% goat serum/0.2% Triton-X in PBS overnight at 4ºC. The

following morning cover slips were then incubated with Alexa 568-conjugated secondary

antibody (Life Technologies) for 45 min and mounted with Vectashield containing DAPI.

For detection of endogenous TiPARP, HuH7 cells were fixed with methanol/acetone for 20

min/1 min at -20ºC. Cells were permeabilized with 0.4% Triton-X in PBS for 20 min and then

blocked with 1% goat serum/0.2% Triton-X in PBS for 1 h. Cover slips were incubated

overnight with anti-TiPARP (84664, Abcam) diluted 1:100 in 1% goat serum/0.2% Triton-X in

PBS overnight at 4ºC. The following morning cover slips were then incubated with Alexa 568-

conjugated secondary antibody for 45 min and mounted with Vectashield containing DAPI.

All images were acquired using an Imager.Z1 epifluorescence microscope and Axiovision

software (Zeiss) or an Olympus Fluoview 1000 confocal microscope.

5.16 GST purification pGEX-4T1-TiPARP constructs were transformed into BL21 stain E. coli. The following day

single colonies were selected and cultured overnight in 10 ml LB. Overnight cultures were

transferred to 500 ml LB-ampicillin and cultured at 30ºC under constant shaking. Once cultures

Page 82: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

67

reached OD600= 0.8-1.0 GST fusion protein expression was induced with 0.1 mM isopropyl-β-D-

thio-galactoside (IPTG, BioShop) for 3-4 h at 37ºC with shaking. Bacteria were pelleted at 8 000

rpm for 10 min at 4ºC and frozen overnight at -20ºC. Pellets were thawed on ice and re-

suspended in 10 ml GST lysis buffer (20 mM Tris Base [pH 7.6], 10 mM MgCl2, 1 mM MnCl2,

1 mM EDTA, 150 mM NaCl, 5% Glycerol) containing 1X PIC and then incubated with 1 mg/ml

lysozyme (BioShop), 1 μg/ml DNAse I (Life Technologies), 10 mg/ml RNAse A (BioShop) for

15 min at room temperature. Extracts were sonicated for 2 min at 45% amplitude and rotated at

4ºC for 30 min. Extracts were centrifuged at 14 000 rpm for 15 min and soluble protein were

transferred to pre-cooled 50 ml conical tubes. One milliliter of equilibrated Glutathione

Sepharose 4 Fast Flow beads (50% slurry in GST lysis buffer) were added to extracts and rotated

overnight at 4ºC. The following morning beads were loading into 20 ml Econo-Column

Chromatography Columns (Bio-Rad). Beads were consecutively washed at 4ºC with cell lysis

buffer (20X bed volume) 3X and NETN buffer (50 mM Tris Base [pH 7.6], 100 mM NaCl, 1

mM MgCl2, 10% glycerol, 0.5% NP-40) 3X. GST fusion proteins were eluted from beads in 2 ml

10 mM glutathione for 30 min and concentrated and purified using Amicon Ultra-0.5 ml

Centrifugal Filters 50 kDa cut off (Millipore).

5.17 32P-NAD+ auto-ADP-ribosylation and hetero-ADP-ribosylation assays

Purified ARTD1 was purchased from Trevigen (Gaithersburg, MD). Auto-ribosylation assays

were carried out in a 30 μl reaction volume at 30ºC in 1X PARP buffer (5 mM Tris Base [pH

8.0], 0.2 mM DTT, 4 mM MgCl2, 0.1 μg/ml BSA), 2 μCi 32P-NAD+ (Perkin-Elmer,

Woodbridge, ON) and/or 500 μM β-NAD+ (Sigma), 1 μg activated DNA (Sigma) or RNA.

Approximately 5 μg of GST fusion partially purified protein or 1 μg ARTD1 were used. For

hetero-ribosylation experiments 8 μg of core histones (H2A, H2B, H3, H4) (New England

Biolabs) were added to the reaction. Reactions were incubated for 30 min and stopped by

addition of 1X sample buffer, separated by SDS-PAGE and visualized by autoradiography.

5.18 Statistical Analysis All real-time PCR results are expressed as means ± standard error of the means (SEM).

Statistical analysis was calculated using GraphPad Prism statistical software (San Diego, CA).

Page 83: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

68

One-way Analysis of Variance (ANOVA) followed by Tukey’s multiple comparison tests were

used to assess statistical significance (P<0.05).

Page 84: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

69

Chapter 4: Results 6 TiPARP is a mono-ADP-ribosyltransferase

6.1 Auto-ADP-ribosylation The catalytic activities of ARTDs are dependent on a well-conserved glutamate residue within

the catalytic active site of the HYE triad sequence (Holbourn, et al., 2006; Marsischky, et al.,

1995). Sequence alignment of the ARTD catalytic core sequences of the 18 ARTD family

members showed that H532, Y564 and I631 of TiPARP are equivalent to the ARTD catalytic

HYE triad. Since TiPARP has an isoleucine in place of the catalytic glutamate and a shortened

connecting loop, it is predicted to possess MART rather than poly-ADP-ribosyltransferase

activity (Kleine, et al., 2008); however, this has not been experimentally determined. To examine

the catalytic activity of TiPARP, we compared the enzymatic properties of TiPARP to that of

ARTD1 using an auto-ADP-ribosylation assay. ARTD1 demonstrated auto-ADP-ribosylation

with the incorporation of 32P-NAD+ and a marked shift in mobility and smearing pattern which

was indicative of poly(ADP-ribose) polymer formation (Figure 10A). Partially purified GST-

TiPARP also showed evidence of auto-modification but no apparent change in mobility,

suggesting that TiPARP exhibits mono-ADP-ribosyltransferase rather than poly(ADP-ribose)

polymer synthase activity (Figure 10B). This activity was competed away with unlabelled β-

NAD+. We cannot exclude the possibility that TiPARP generates short oligo-ADP-ribose units

since they would not alter the migration of the protein. Because ARTD1/PARP1 activity is

stimulated by its interaction with DNA (D'Amours, et al., 1999) and TiPARP contained a zinc-

finger domain with putative RNA binding capacity (Kelly, et al., 2007), we tested whether the

presence of DNA or RNA modified TiPARP auto-ribosylation activity. Addition of activated

DNA or DNAse-treated total RNA did not alter the migration pattern of GST-TiPARP (Figure

10B). We next tested mutants of the TiPARP catalytic triad. GST-TiPARP-H532A and GST-

TiPARP-Y564A did not show any evidence of auto-ribosylation or shift in mobility (Figure

10B). GST-TiPARP-I631A demonstrated similar auto-ribosylation activity to that of wildtype,

suggesting that this residue was not required for enzymatic activity (Figure 10B).

Next we evaluated the catalytic activity of mouse and chicken (Gallus gallus) TiPARP as GST

fusion proteins. GST-mouse TiPARP (GST-mTiPARP) demonstrated auto-ribosylation that

Page 85: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

70

could be mono-ADP-ribosyltransferase activity due to the lack of shift in mobility (Figure 10B).

GST-chicken TiPARP (GST-chTiPARP) demonstrated weak auto-ribosylation activity compared

to human, despite the loading of similar amounts of protein (Figure 10B). Addition of activated

DNA or DNAse-treated RNA did not alter the migration pattern or auto-ribosylation ability of

GST-mTiPARP or GST-chTiPARP (Figure 10C).

To determine if mutation of isoleucine 631 to a glutamate would convert TiPARP MART

activity to poly-ADP-ribosytransferase activity, we analyzed the auto-ribosylation activity of

GST-TiPARP-I631E mutant. The auto-ribosylation activity of GST-TiPARP-I631E was similar

to that of wildtype (Figure 10D), exhibiting MART but not polymer synthase activity.

To further examine TiPARP catalytic activity we used an in vitro 35S-labelled protein shift assay,

which allowed us to directly compare mobility shift of modified and unmodified ARTD proteins.

In vitro translated 35S-labelled ARTD1 (35S-ARTD1) incubated with β-NAD+ (and activated

DNA) resulted in a marked shift in its mobility compared to unmodified 35S-ARTD1 in the

absence of β-NAD+ (Figure 11A). Conversely, only a small change in mobility was observed for 35S-TiPARP when incubated with β-NAD+, which may represent mono-ADP-ribosylation of

multiple residues or short oligo-ADP-ribose units. None of the 35S-labelled TiPARP catalytic

point mutants demonstrated altered mobility compared to wildtype TiPARP (Figure 11B). The

absence of a modest mobility shift of 35S-labelled TiPARP-I631A in the presence of β-NAD+ is

in contrast to results obtained from 32P-NAD+ PARP assays with GST-TiPARP I631A (Figure

11B). This discrepancy could be due to differences in sensitivity between the two assays. The

I631A mutant may have reduced auto-ADP-ribosylation activity that could be below the

threshold of sensitivity of the 35S protein shift assay. A similar mono-ribosylation pattern was

observed for mouse 35S-TiPARP but not chicken 35S-TiPARP (Figure 11C). 32P-NAD+ catalytic

assays performed on GST-chTiPARP demonstrated very weak auto-ribosylation that may not

have been detected by the 35S protein shift assay. Taken together, these data revealed that

TiPARP was a mono-ADP-ribosyltransferase.

We next used the 35S-labelled in vitro protein shift assay to assess the catalytic activity of the N-

terminal truncations of TiPARP. Deletion mutants from amino acid residue 33 to residue 275

showed a slight migration or shift comparable to full-length TiPARP when incubated with β-

Page 86: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

71

NAD+ (Figure 11D). However, this shift was not apparent in N-terminal deletion mutants 328-

657 and 445-657, suggesting that the auto-modification domain of TiPARP is between residues

275 to 328.

Page 87: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

72

Figure 10. TiPARP is a mono-ADP-ribosyltransferase with auto-ribosylation activity.

(A) Auto-ADP-ribosylation of ARTD1, purified ARTD1 was incubated with 32P-NAD+ and 1 μg

activated DNA then analyzed the SDS-PAGE and autoradiography (B) Auto-ADP-ribosylation

Page 88: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

73

of GST-TiPARP, GST-TiPARP catalytic mutants (GST-TiPARP-H532A, –Y564A and –I631A),

GST-mouse TiPARP (GST-mTiPARP) and GST-chicken TiPARP (GST-chTiPARP). Auto-

ribosylation reactions were carried out with semi-purified GST-TiPARP fusion proteins and 32P-

NAD+ or co-incubated 500 μM β-NAD+ or 1 μg activated DNA or RNA. Automodified GST-

TiPARP fusion proteins were analyzed by SDS-PAGE and autoradiography. (C) Auto-ADP-

ribosylation of mouse TiPARP (mTiPARP) and chicken TiPARP (chTiPARP) incubated with

DNA or RNA. Mouse GST-tagged TiPARP (GST-mTiPARP) and chicken GST-tagged TiPARP

(GST-chTiPARP) were incubated with 32P-NAD+ and activated DNA or DNAse-treated RNA

and analyzed by SDS-PAGE and autoradiography. (D) Comparison of auto-ADP-ribosylation of

GST-TiPARP-I631E (GST-I631E) with wildtype GST-TiPARP (GST-wt). All figures represent

data from three independent experiments. Closed arrow denotes mobility shifted ARTD1.

Page 89: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

74

Figure 11. Putative auto-ribosylation domain of TiPARP is between residues 275 and 328.

(A) Mobility shift comparisons of modified and unmodified 35S-labelled ARTD1 and TiPARP.

In vitro translated 35S-labelled ARTD1 and TiPARP were incubated with or without 500 μM β-

NAD+, activated DNA or RNA and analyzed by SDS-PAGE and autoradiography. (B) TiPARP

Page 90: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

75

catalytic point mutants are inactive. (C) Auto-ribosylation of mouse TiPARP and chicken

TiPARP. In vitro translated 35S-labelled TiPARP, catalytic point mutants, mouse and chicken

TiPARP were incubated with or without 500 μM β-NAD+ and analyzed by SDS-PAGE and

autoradiography. (D) TiPARP automodification domain is between residues 275 and 328. In

vitro translated 35S-labelled TiPARP and N-terminal truncations were incubated with or without

500 μM β-NAD+ and analyzed by SDS-PAGE and autoradiography. These figures represent data

from three independent experiments. Closed arrow denotes shifted ARTD1.

Page 91: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

76

6.2 Hetero-ADP-ribosylation

We then compared the ability of GST-TiPARP and ARTD1 to hetero-ribosylate core histones. In

agreement with a previous report, ARTD1 hetero-ADP-ribosylated each of the core histones

(Figure 12A)(Messner, et al.). GST-TiPARP also mono-ADP-ribosylated each of the core

histones as evidenced by the lack of a marked mobility shift in molecular weight of the histones

(Figure 12B). No ADP-ribosylation of the core histones was observed in the presence of the

catalytic mutant GST-TiPARP-H532A (Figure 12C). Similar to wildtype, GST-TiPARP-I631A

also ribosylated core histones in the presence of 32P-NAD+ (Figure 12D). These results

demonstrated TiPARP hetero-ribosylation activity was mono-ADP-ribosyltransferase.

Page 92: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

77

Page 93: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

78

Figure 12. TiPARP is a mono-ADP-ribosyltransferase with heteroribosylation activity.

(A) Hetero-ribosylation of core histones by ARTD1. Purified ARTD1 was incubated with 2 μCi 32P-NAD+, individual or mixed core histones (core) and hetero-ribosylation was analyzed by

SDS-PAGE followed by autoradiography. (B) Hetero-ribosylation of core histones by GST-

TiPARP. GST-TiPARP was incubated with 32P-NAD+ and histones H2A, H2B, H3 or H4 and

analyzed by SDS-PAGE and autoradiography. (C) Catalytic mutant GST-TiPARP-H532A did

not hetero-ribosylation core histones. GST-TiPARP-H532A (H532A) or wildtype GST-TiPARP

(wt) were incubated with mixed core histones and hetero-ribosylation was determined by SDS-

PAGE followed by autoradiography. (D) Hetero-ribosylation of core histones by TiPARP

catalytic mutant I631A. GST-TiPARP-I631A was incubated with 32P-NAD+ and histones H2A,

H2B, H3 or H4 and analyzed by SDS-PAGE and autoradiography.

Page 94: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

79

7 Identification of TiPARP as a novel negative regulator of AHR

7.1 Temporal induction of TiPARP mRNA by TCDD TiPARP was first reported to be a TCDD-responsive gene that was regulated by AHR in mouse

hepatoma Hepa1c1c7 cell line (Ma, et al., 2001); however, little is known about the regulation of

TiPARP in human cell lines. With this in mind, we determined the temporal changes in TiPARP

expression after treatment of T47D human breast carcinoma cells with TCDD. The results were

compared to those obtained for CYP1A1. TCDD-dependent increases in TiPARP mRNA levels

peaked at 1.5 h, whereas CYP1A1 mRNA levels increased over the 24 h time course (Figure

13A). Similarly, TiPARP mRNA induction peaked at 1.5 h TCDD treatment in Hepa1c1c7 cells

(Figure 13B) (Ma, 2002). However, the TiPARP mRNA declined less rapidly compared to

T47D cells, suggesting possible cell type differences in TiPARP regulation or mRNA stability.

Similar kinetic differences between TCDD-induced TiPARP and CYP1A expression levels were

observed in chick embryo hepatocytes (Diani-Moore, et al., 2010), demonstrating that these

regulatory differences between TiPARP and CYP1A genes are conserved across species.

Page 95: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

80

Figure 13. Temporal induction of TiPARP mRNA in T47D and Hepa1c1c7 cell lines

treated with TCDD.

(A) Temporal analysis of TCDD-induced TiPARP and CYP1A1 mRNA expression levels. T47D

cells were treated with 10 nM TCDD at the times indicated. (B) Temporal analysis of TCDD-

induced TiPARP mRNA expression levels in Hepa1c1c7 cells. Hepa1c1c7 cells were treated

with 10 nM TCDD for the times indicated. Data are presented as means ± S.E.M. of three

independent replicates.

Page 96: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

81

7.2 TiPARP knockdown increased TCDD-induced AHR transactivation

To test whether TiPARP modulates AHR activity, we used RNAi-mediated knockdown of

TiPARP using pooled siTiPARP siRNA (containing four different siTiPARP sequences) in

human gastric carcinoma NCI-N87 cells, human hepatoma HuH7 cells and human breast

carcinoma T47D cells and examined TCDD-dependent changes in CYP1A1 and CYP1B1

mRNA levels. TiPARP mRNA levels were reduced to approximately 25-38% compared to

control cells from the cell lines with a non-targeting pool (NTP) of siRNA (Figure 14A).

Endogenous TiPARP protein levels could not be assessed by immunoblot due to the lack of a

suitable antibody. RNAi-mediated knockdown of TiPARP resulted in significantly greater

TCDD-dependent induction of CYP1A1 and CYP1B1 mRNA expression levels compared to

NTP control cells after 24 h exposure (Figure 14B and C). Next we verified these results by

using single siTiPARP sequences (siTiPARP12 and siTiPARP13) to knockdown TiPARP in

T47D cells. TiPARP mRNA levels were reduced to approximately 23% and 34% compared to

non-targeting (NT) control cells using siTiPARP12 and siTiPARP13 sequences, respectively

(Figure 15A). Due to the lack of suitable antibody to detect endogenous TiPARP protein, we

instead co-transfected GFP-tagged TiPARP with siTiPARP sequences to demonstrate

knockdown of overexpressed GFP-TiPARP protein with anti-GFP antibody (Figure 15B).

Single-strand RNAi-mediated knockdown of TiPARP resulted in significantly greater TCDD-

dependent induction of CYP1A1 and CYP1B1 mRNA expression levels compared to NT control

cells after 24 h exposure which is in agreement with that we observed with pooled siTiPARP

knockdown (Figure 15C). We next tested how soon after treatment TCDD-induced gene expression

was increased in T47D cells following 48 h knockdown of TiPARP. T47D cells were treated with TCDD

for 1.5 h and 3 h. Increased TCDD-dependent induction of CYP1A1 and CYP1B1 mRNA levels

following TiPARP knockdown was observed after 3 h but not 1.5 h after treatment (Figure 15D

and E).

To examine the TiPARP-dependent mechanism of AHR repression we performed ChIP assays in

T47D cells after RNAi-mediated TiPARP knockdown of TiPARP and treatment with TCDD for

1 h. In agreement with increased CYP1A1 and CYP1B1 mRNA levels, knockdown of TiPARP

resulted in significantly greater recruitment of AHR and ARNT to the CYP1A1 and CYP1B1

enhancer regions compared to the NT control (Figure 16A and C). No significant differences in

Page 97: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

82

AHR or ARNT recruitment to downstream genomic control regions for both genes were

observed (Figure 16B and D).

We next determined the effect of TiPARP knockdown on ligand-induced AHR protein

degradation. Cells with TiPARP knocked down trended toward increased constitutive AHR

protein levels (Figure 16E). TiPARP knockdown cells treated with TCDD for 24 h demonstrated

reduced AHR degradation compared with TCDD-treated non-targeting cells, which is consistent

with elevated AHR-regulated gene expression and AHR recruitment results (Figure 16E). These

data suggested TiPARP mechanism of AHR repression to potentially involve reduced proteolytic

degradation of AHR.

Page 98: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

83

Figure 14. TiPARP knockdown in NCI-N87, HuH7 and T47D cells increased TCDD-

induced AHR transactivation.

Cell lines were transfected with siRNA pool (containing four different siRNAs) targeting against

TiPARP for 48 h. Cells were then treated with 10 nM TCDD and DMSO (control) for 24 h and

RNA was isolated and reverse transcribed as described in the Materials and Methods. Changes in

(A) TiPARP; (B) CYP1A1; and (C) CYP1B1 mRNA expression was then determined using

qPCR. Data were normalized to non-targeting pool (NTP) DMSO (control) and to β-actin levels

of each cell line. Results shown are means ± S.E.M for at least four independent experiments.

mRNA expression levels for each cell line significantly (P < 0.05) different than NTP transfected

cells are denoted with an asterisk.

Page 99: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

84

Page 100: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

85

Figure 15. TiPARP knockdown increased TCDD-induced AHR transactivation.

(A) TiPARP mRNA expression levels in T47D cells following 48 h knockdown with single

siRNAs against TiPARP (siTP12 and siTP13) and non-targeting control (NT). Expression levels

significantly (P < 0.05) lower than NT were denoted with asterisks. (B) Western blot analysis of

GFP-TiPARP protein overexpression following RNAi-mediated TiPARP knockdown in TREx-

FLP-IN 293 cells. (C) TiPARP knockdown increased TCDD-induced CYP1A1 and CYP1B1

mRNA expression levels in T47D cells following 24 h treatment. Gene expression levels

significantly (P < 0.05) greater than NT were denoted with asterisks. (D) TiPARP knockdown

did not affect CYP1A1 and CYP1B1 mRNA levels following 1.5 h TCDD treatment. Asterisks

denote TiPARP mRNA levels significantly lower than NT. (E) TiPARP increased TCDD-

induced CYP1A1 and CYP1B1 mRNA levels following 3 h treatment. Gene expression levels

significantly (P < 0.05) different than NT were denoted with asterisks.

Page 101: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

86

Page 102: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

87

Figure 16. TiPARP knockdown increased AHR/ARNT recruitment to CYP1A1 and

CYP1B1 regulatory regions and reduced TCDD-induced degradation of AHR.

TiPARP knockdown increased TCDD-induced AHR and ARNT recruitment to CYP1A1 (A-B)

and CYP1B1 (C-D). AHR and ARNT to CYP1A1 regulatory region (A), CYP1A1 distal

downstream region (B), CYP1B1 regulatory region (C) and CYP1B1 distal downstream region

(D). T47D cells with TiPARP knocked down were treated with 10 nM TCDD for 1 h and

harvested for ChIP assays. Percent recruitment significantly greater (P < 0.05) than NT was

denoted with an asterisk. (E) TiPARP knockdown reduced TCDD-induced AHR degradation in

T47D cells. T47D cells following TiPARP knockdown were treated with TCDD (T) or DMSO

(D) for 24 h and AHR protein expression was determined by Western blot. Left panel;

representative Western blot from three independent experiments, right panel; Quantification of

AHR protein levels, densitometry was performed using ImageJ analysis software. AHR protein

levels were normalized to pcDNA-transfected and DMSO treated cells. Asterisks denoted AHR

protein levels significantly (P < 0.05) greater than treatment-matched non-targeting-transfected

cells. All data were presented as means ± S.E.M for three independent experiments. Statistical

analysis was determined by two-tailed Student’s t-test.

Page 103: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

88

7.3 Overexpression of TiPARP repressed AHR-mediated transactivation

Because the RNAi experiments suggested that TiPARP might act as a negative regulator of AHR

transactivation, we investigated the ability of TiPARP overexpression to repress AHR-dependent

gene expression in HuH7 cells transiently transfected with a CYP1A1-regulated luciferase

reporter plasmid (CYP1A1-luc). Transient transfection of increasing amounts of TiPARP

resulted in a dose-dependent reduction of TCDD-induced reporter gene activity (Figure 17A).

Similar results were observed using a CYP1B1-regulated luciferase reporter plasmid (CYP1B1-

luc) (Figure 17B). To examine whether TiPARP-mediated repression was limited to CYP1

regulatory regions we used an artificial 3XAHRE-luc reporter (3X repeat of Hairy and Enhancer

of Split-1 (HES1) AHRE sequence) and determined TCDD-induced reporter gene activity.

Overexpression of increasing amounts of TiPARP dose-dependently repressed 3XAHRE-luc

reporter gene activity, which was similar to the CYP1-luc reporter constructs (Figure 17C).

Transfection of up to one microgram of the catalytic inactive mutant TiPARP-H532A plasmid

did not repress CYP1A1-luc reporter gene activity but rather demonstrated dominant negative

activity, indicating catalytically active TiPARP is required for repression (Figure 17D). We next

tested mouse TiPARP, which shares 92.8% sequence homology with human TiPARP and

demonstrated it to repress induction of CYP1A1-luc activity similar to human TiPARP (Figure

17E). Chicken TiPARP, which only exhibits 68.5% sequence homology with human TiPARP

did not repress reporter gene activity. The lack of repression of chicken TiPARP may have been

due the use of mammalian cells and/or reporter gene in our model. Similar findings were

observed after transfection with GFP-tagged TiPARP fusion proteins (Figure 17F). Transfection

of equal amounts of GFP-TiPARP plasmid DNA resulted in similar protein expression levels of

human and mouse GFP-tagged TiPARP, whereas higher protein levels of chicken TiPARP were

observed (Figure 17G). Collectively these data showed that TiPARP was a negative regulator of

AHR transactivation, but this effect may exhibit species-specificity.

To examine if TiPARP overexpression inhibited endogenous CYP1A1 and CYP1B1 mRNA

expression, we transiently transfected HuH7 cells with TiPARP, treated them with TCDD for 24

h prior to isolating RNA and performing qPCR. We observed a significant repression of TCDD-

dependent increases in CYP1A1 and CYP1B1 mRNA levels in cells overexpressing TiPARP,

supporting our reporter gene assay results (Figure 18A and B).

Page 104: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

89

We then determined if TiPARP was recruited to AHR target gene enhancer regions using ChIP

assays. We transiently transfected HuH7 cells with GFP-tagged TiPARP and ChIP assays were

performed with an anti-GFP antibody. HuH7 cells transfected with GFP-TiPARP revealed a

small but significant TCDD-dependent increase in GFP-TiPARP recruitment to CYP1A1 after 45

min of TCDD treatment (Figure 18C), but not to a downstream chromatin region that was not

bound by AHR (Figure 18D). Interestingly, reduced AHR occupancy at CYP1A1 was also

observed in cells overexpressing GFP-TiPARP, which supported the increased AHR and ARNT

recruitment we observed with TiPARP knockdown (Figure 18C). We next examined if TiPARP

overexpression affected AHR protein levels. Time course studies showed a reduction in AHR

protein levels in cells overexpressing TiPARP in the presence or absence of TCDD (Figure

18E), which was in support of the increased AHR protein levels observed after RNAi-mediated

knockdown of TiPARP (Figure 16E).

Page 105: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

90

Figure 17. TiPARP overexpression reduced AHR-regulated reporter gene activity.

(A-C) Concentration-dependent reduction of TCDD-induced CYP1A1-luc, CYP1B1-luc and

3xAHRE/HES1-luc reporter gene activity with TiPARP overexpression. (D) Dose-response of

TiPARP catalytic mutant H532A in HuH7 cells. (E-F) Mammalian TiPARP inhibited TCDD-

induced CYP1A1-luc reporter gene activity. Cells were co-overexpressed with reporter gene and

increasing amounts of pcDNA-TiPARP (50 ng to 200 ng), pcDNA-TiPARP-H532A (500 ng to

1000 ng), 200 ng pcDNA-chicken TiPARP (chTiPARP), 200 ng pcDNA-mouse TiPARP

(mTiPARP), pEGFP-hTiPARP, pEGFP-chTiPARP or pEGFP-mTiPARP. All transfected cells

were treated with TCDD or DMSO for 24 h and reporter gene activity was determined. Values

Page 106: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

91

were normalized to 100% TCDD treatment. Data are represented as means ± S.E.M. of three

independent experiments. Statistical significance was analyzed by one-way ANOVA and

Tukey’s multiple comparisons test. Reporter gene activity significantly different than (P < 0.05)

TCDD-treated empty vector was denoted with an asterisk. (G) Western blot of overexpressed

GFP-hTiPARP, GFP-chTiPARP and GFP-mTiPARP in HuH7 cells.

Page 107: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

92

Figure 18. Overexpression of TiPARP repressed TCDD-induced AHR transactivation.

(A-B) TiPARP overexpression repressed TCDD induction of CYP1A1 (A) and CYP1B1 (B)

mRNA in HuH7 cells. Asterisks denote mRNA expression significantly lower (P < 0.05, two-

tailed Student’s t-test) than treatment-matched empty vector transfected cells. (C) Overexpressed

GFP-TiPARP recruitment to CYP1A1 regulatory region was induced by TCDD and repressed

AHR recruitment. (D) Recruitment of GFP-TiPARP and AHR to CYP1A1 distal downstream

region. HuH7 cells were transfected with pEGFP empty vector or pEGFP-TiPARP and dosed

with TCDD for 45 min and cell were harvested for ChIP assays. Percent recruitment of GFP-

TiPARP of significantly greater (P < 0.05, two-tailed Student’s t-test) than transfection-matched

DMSO-treated cells was denoted with an asterisk. Percent recruitment of AHR significantly

lower (P < 0.05, two-tailed Student’s t-test) than treatment-matched GFP-transfected cells is

denoted with a pound sign. Data shown are representative of three independent experiments. (E)

TiPARP overexpression increased TCDD-induced AHR proteasomal degradation. T47D cells

Page 108: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

93

were transfected with vector or pcDNA-TiPARP and dosed with TCDD for the indicated times

and AHR protein levels were determined by Western blot. Left panel; representative Western

blots from three independent experiments. Right panel; quantification of AHR protein levels,

densitometry was performed using ImageJ analysis software (NIH). Asterisks denote AHR

protein levels significantly (P < 0.05, two-tailed Student’s t-test) lower than treatment-matched

vector-transfected cells.

7.4 Tiparp knockout increased TCDD-dependent AHR target transactivation and AHR protein levels

To investigate the impact of TiPARP loss on AHR transactivation and protein expression we

created immortalized mouse embryonic fibroblasts (MEFs) from wildtype, Tiparp-/- and

Tiparp+/- siblings (strain B6;129S4). As expected, TiPARP mRNA levels were induced in

wildtype and Tiparp+/- cells after 24 h treatment with TCDD, but no TiPARP was detected in

Tiparp-/- cells (Figure 19A). TCDD treatment of Tiparp-/- MEFs resulted in greater increases in

CYP1A1, CYP1B1 and AHRR mRNA levels compared to wildtype and Tiparp+/- cells, while

GAPDH mRNA levels were unaffected by TCDD treatment or by genotype (Figure 19A and

B). Ectopic expression of TiPARP into Tiparp-/- MEFs reduced TCDD-dependent AHR

transactivation compared to vector control (Figure 19C). ChIP assays revealed significantly

greater TCDD-induced AHR recruitment to Cyp1a1 and Cyp1b1 in Tiparp-/- compared to

wildtype or Tiparp+/- cells (Figure 19D and E). Because we observed increased AHR

transactivation in Tiparp-/- compared to wildtype or Tiparp+/- MEFs, we examined the relative

AHR protein levels and ligand-induced AHR degradation in the different MEF lines. Tiparp-/-

cells had higher constitutive AHR protein levels compared to wildtype and Tiparp+/- cells

(Figure 19F). TCDD-induced degradation of AHR was reduced in Tiparp-/- compared to

wildtype and Tiparp+/- cells (Figure 19F). TCDD treatment for 24 h caused a 72% and 81%

reduction in AHR protein levels in wildtype and Tiparp+/- cells, respectively, but only a 53%

reduction in Tiparp-/- MEFs (Figure 19F). In contrast to increased AHR protein levels, small but

significant TCDD-dependent decreases in AHR mRNA levels were detected in Tiparp-/- cells

(Figure 19G).

Page 109: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

94

Page 110: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

95

Figure 19. Tiparp knockout fibroblasts exhibited increased TCDD-induced AHR

transactivation.

(A) Wildtype (Tiparp+/+), Tiparp-/- and Tiparp+/- immortalized mouse embryonic fibroblast

(MEF) lines were treated with 10 nM TCDD for 24 h and gene expression determined. Gene

expression was normalized relative to DMSO treated wildtype cells. Gene expression levels

significantly (P < 0.05) different than treatment-matched wildtype cells are denoted with an

asterisk, and gene expression levels significantly (P < 0.05) different than treatment-matched

Tiparp+/- (+/-) cells are denoted with a pound sign. (B) Expression of GAPDH mRNA in

immortalized mouse embryonic fibroblast lines. (C) Genetic complementation of Tiparp-/- MEFs

with mouse TiPARP overexpression. Tiparp-/- MEFs were transfected with 2 μg pcDNA-

mTiPARP or pcDNA (vector). Following 24 h transfected cells were treated with TCDD for 24 h

and CYP1A1, CYP1B1 and AHRR mRNA levels were determined. Data presented were

representative of means ± S.E.M. of three independent experiments. mRNA expression

significantly (P<0.05) different than vector-transfected cells was denoted with an asterisk. (D-E)

Tiparp-/- cells exhibit increased TCDD-induced AHR recruitment to Cyp1a1 and Cyp1b1. (D)

AHR recruitment to Cyp1a1 regulatory region and distal downstream control region. (E) AHR

recruitment to Cyp1b1 regulatory region and distal downstream control region. MEF lines were

treated with 10 nM TCDD for 45 min and harvested for ChIP assays. Percent recruitment

significantly greater (P < 0.05) than treatment-matched wildtype cells was denoted with an

asterisk, and percent recruitment significantly greater (P < 0.05) than treatment-matched

Tiparp+/- (+/-) cells was denoted with a pound sign. (F) Tiparp-/- fibroblasts have reduced

TCDD-induced AHR degradation. MEF lines were treated with TCDD (T) or DMSO (D) for 24

h and AHR protein expression was determined by Western blot. Left panel; representative

Western blot from three independent experiments. Right panel; quantification of AHR protein

levels, densitometry was performed using ImageJ analysis software (NIH). Asterisks denoted

AHR protein levels significantly (P < 0.05, two-tailed Student’s t-test) greater than treatment-

matched wildtype cells. Pound sign denotes AHR protein levels significantly (P < 0.05, two-

tailed Student’s t-test) less than genotype-matched DMSO-treated cells. (G) AHR mRNA

expression levels of MEF cell lines treated with DMSO or TCDD for 24 h. Asterisks denoted

AHR mRNA levels significantly (P < 0.05, two-tailed Student’s t-test) greater than treatment-

matched wildtype cells. Pound sign denoted AHR mRNA levels significantly (P < 0.05, two-

tailed Student’s t-test) less than genotype-matched DMSO-treated cells.

Page 111: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

96

7.5 Tiparp knockout increased AHR-mediated gene expression in vivo

To determine if TiPARP was a negative regulator of AHR transactivation in vivo we treated male

and female 6-8 week old Tiparp+/+, Tiparp+/- and Tiparp-/- mice (strain B6;129S4-

TiparpGt(ROSA)79Sor) with 30 μg/kg b.w. TCDD or corn oil (vehicle control) for 6 h. Mice were

sacrificed, livers excised and AHR-regulated gene expression was determined. As expected

TiPARP expression was detected in hepatic tissues from wildtype and Tiparp+/- mice but not in

Tiparp-/- mice (Figures 20A and 21A). Following 6 h TCDD treatment both female and male

Tiparp-/- mice demonstrated significantly greater AHR target gene CYP1A1, CYP1B1, AHRR,

nuclear factor (erythroid-derived 2)-like 2 (NFE2L2) and NAD(P)H dehydrogenase (quinone) 1

(NQO1) mRNA expression compared to wildtype mice (Figures 20B-F and Figure 21B-F).

These data supported our TiPARP knockdown and Tiparp-/- MEF data and indicated that

TiPARP is a negative regulator of AHR transactivation in vivo.

Page 112: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

97

Figure 20. Female Tiparp-/- mice demonstrated increased TCDD-induced AHR

transactivation.

Six week old female Tiparp+/+, Tiparp+/- and Tiparp-/- mice were treated with corn oil (CO) or 30

μg/kg b.w. TCDD (TCDD) for 6 h and livers were excised, RNA was isolated and mRNA

reversed transcribed. Gene expression was normalized relative to CO-treated wildtype mice.

Each bar represented means ± S.E.M. from four animals. Data were analyzed by One-way

ANOVA and Tukey’s multiple comparisons test. Gene expression levels significantly (P < 0.05)

different than treatment-matched wildtype mice were denoted with an asterisk and gene

expression levels significantly (P < 0.05) different than treatment-matched Tiparp+/- (+/-) mice

were denoted with a pound sign.

Page 113: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

98

Figure 21. Male Tiparp-/- mice demonstrated increased TCDD-induced AHR

transactivation.

Six week old male Tiparp+/+, Tiparp+/- and Tiparp-/- mice were treated with corn oil (CO) or 30

μg/kg b.w. TCDD (TCDD) for 6 h and livers were excised, RNA was isolated and mRNA

reversed transcribed. Gene expression was normalized relative CO-treated wildtype mice. Each

bar represented the means ± S.E.M. from four animals. Data were analyzed by One-way

ANOVA and Tukey’s multiple comparisons test. Gene expression levels significantly (P < 0.05)

different than treatment matched wildtype mice were denoted with an asterisk and gene

expression levels significantly (P < 0.05) different than treatment- matched Tiparp+/- (+/-) mice

were denoted with a pound sign.

Page 114: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

99

8 Understanding the mechanism of TiPARP-mediated repression of AHR transactivation

8.1 TiPARP selective repression of AHR activity Because ARTD1 and other ARTD family members can function as transcriptional repressors or

activators (Kraus, 2008), we examined the ability of ARTD1 and ARTD12 to modulate AHR

transactivation. ARTD1 was selected since it is the most studied ARTD, whereas ARTD12 was

chosen because it shared the greatest sequence homology (46.8%) with TiPARP compared to

other members of the ARTD family (Schreiber, et al., 2006). We co-transfected 200 ng and 1500

ng of ARTD1 and ARTD12 with CYP1A1-luc into HuH7 cells and determined reporter gene

activity after 24 h TCDD treatment. Transfection of HuH7 cells with 200 ng of TiPARP caused a

~50% reduction in TCDD-dependent induction of CYP1A1-luc reporter gene activity, whereas

transfection with the same amount of ARTD1 or ARTD12 had no effect (Figure 22). Increasing

the amount of transfected plasmid DNA to 1500 ng caused a greater than 85% reduction in

CYP1A1-luc reporter gene activity by TiPARP, but only a ~40% reduction by ARTD1 or

ARTD12. These results suggested that TiPARP exhibited selective repression of AHR

transactivation compared to other ARTD family members.

Page 115: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

100

Figure 22. CYP1A1-luc reporter gene activity was preferentially inhibited by TiPARP.

HuH7 cells were co-overexpressed with 200 ng or 1500 ng of pcDNA-TiPARP, pCMV-ARTD1

or pCMV-ARTD12 and pCYP1A1-luc and treated with TCDD. Reporter gene activity

significantly lower (P < 0.05) than empty vector transfected cells was denoted with an asterisk.

Reporter gene activity significantly (P < 0.05) greater than TiPARP (1500 ng) transfected cells

was denoted with a pound sign. All luciferase data were presented as means ± S.E.M from three

independent experiments and statistical significance analyzed by One-way ANOVA and Tukey’s

multiple comparisons test. Western blots confirming the presence of increased ARTD1 and

ARTD12 protein expression following transient transfection are presented to the right of bar

graph.

Page 116: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

101

8.2 Ectopic AHR but not ARNT prevented TiPARP-mediated inhibition of AHR transactivation

The AHRR negatively regulates AHR transactivation by competing with AHR for ARNT

dimerization and binding to the AHRE (Mimura, et al., 1999), but also through ARNT-

independent mechanisms (Evans, et al., 2008; Karchner, et al., 2009). To evaluate the role of

ARNT in TiPARP-mediated repression of AHR transactivation we assessed the ability of

TiPARP to repress AHR transactivation in the presence of increasing amounts of AHR or

ARNT. Transient co-transfection of AHR but not ARNT rescued the TiPARP-dependent

repression of CYP1A1-regulated reporter gene induction (Figure 23A and B).

Page 117: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

102

Figure 23. Overexpression of AHR but not ARNT prevented TiPARP-mediated inhibition

of AHR transactivation.

HuH7 cells were co-transfected with pCYP1A1-luc, TiPARP (200 ng) and (A) AHR (250-750

ng) or (B) ARNT (250-750 ng). All transfected cells were treated with 10 nM TCDD for 24 h

and reporter gene activity was determined. Results were shown as means ± S.E.M from three

independent experiments and significance analyzed by One-way ANOVA and Tukey’s multiple

comparisons test. Reporter gene activity of cells co-overexpressing TiPARP significantly lower

than (P < 0.05) transfection- and treatment-matched control cells were denoted with an asterisk.

Western blots confirming the presence of increased AHR and ARNT protein expression

following transient transfection were presented to the right of respective bar graphs.

Page 118: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

103

8.3 Identification of a putative AHR repressor domain of TiPARP To determine the mechanism of TiPARP-dependent repression of AHR, we created a series of

TiPARP truncations and site-directed point mutants, and determined their ability to repress

AHR-dependent reporter gene activity in HuH7 cells. The expression of each truncated product

was confirmed by in vitro translation (Figure 24). N-terminal truncation constructs of TiPARP

expressing amino acids 33-657 to 218-657 repressed CYP1A1-regulated reporter gene activity

similar to full length TiPARP, with the exception of 200-657, which exhibited greater repressor

activity (Figure 25A). The 225-657 truncation construct exhibited a significantly reduced ability

to repress reporter gene activity, whereas overexpression of 245-657 to 445-657 truncations did

not repress AHR-regulated reporter gene activity. A similar pattern of repression was observed

with the CYP1B1-luc reporter (Figure 25B). We created GFP-tagged versions of all N-terminal

truncations to verify expression in transfected cells by Western blot with anti-GFP antibody. We

first tested the ability of the GFP-tagged N-terminal truncations to repress CYP1A1- and

CYP1B1-luc. The GFP-tagged N-terminal truncations displayed a similar pattern of repression of

both reporter genes to their untagged counterparts (Figure 26A and B). Western blot analysis of

the GFP-TiPARP N-terminal truncations showed that the lack of repression was not due to

reduced protein expression of the different GFP-TiPARP truncations since the protein expression

of all N-terminal truncations was greater than full-length GFP-TiPARP despite transfecting and

loading equal amounts (Figure 27).

Because loss of AHR repressive ability was lost between residues 218 and 245, we were then

interested in whether this region of TiPARP was a putative AHR repressor domain that contained

specific residues pertinent to its ability to repress AHR transactivation. We performed an alanine

scan study whereby we mutated every two consecutive residues to alanines between 218 and 241

and determined the ability of each double alanine point mutant to repress TCDD induction of

CYP1A1- and CYP1B1-luc reporter gene activity. Endogenous glycines were not mutated. All

double point mutants significantly repressed CYP1A1- and CYP1B1-luc activity to varying

degrees except for the Y232A; H233A mutant which failed to repress induction of both reporter

genes (Figure 28A and B). These findings indicated that Y232 and H233 were important for the

ability of TiPARP to repress AHR transactivation. Point mutations between residues 220 and 235

significantly reduced the AHR repressive activity of TiPARP indicating these residues may

contribute to repression of AHR as well.

Page 119: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

104

Figure 24. In vitro expression of 35S-labelled N-terminal TiPARP truncations.

N-terminal pcDNA-TiPARP constructs were in vitro translated with 35S-methionine and detected

by SDS-PAGE and autoradiography.

Page 120: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

105

Page 121: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

106

Figure 25. Overexpression of N-terminal TiPARP truncations on AHR-regulated reporter

gene activity.

Reporter gene activity of CYP1A1-luc (A) and CYP1B1-luc (B). HuH7 cells were transiently co-

transfected with pcDNA-TiPARP N-terminal truncation construct and pCYP1A1-luc or

pCYP1B1-luc reporter genes. Cells were treated with DMSO or TCDD for 16 h and reporter

gene activity was determined. Values were normalized to 100% TCDD treatment. Data are

represented as means ± S.E.M. of three independent experiments. Statistical significance was

analyzed by One-way ANOVA and Tukey’s multiple comparisons test. Reporter gene activity

significantly different than (P < 0.05) TCDD-treated empty vector was denoted with an asterisk.

Reporter gene activity significantly different than (P < 0.05) TCDD-treated full-length TiPARP

(1-657) was denoted with a pound sign.

Page 122: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

107

Page 123: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

108

Figure 26. Expression of N-terminal GFP-tagged TiPARP truncations on AHR-regulated

reporter gene activity.

Reporter gene activity of CYP1A1-luc (A) and CYP1B1-luc (B). HuH7 cells were transiently co-

transfected with pEGFP-TiPARP N-terminal truncation construct and pCYP1A1-luc or

pCYP1B1-luc reporter genes. Cells were treated with DMSO or TCDD for 16 h and reporter

gene activity was determined. Values were normalized to 100% TCDD treatment. Data are

represented as means ± S.E.M. of three independent experiments. Statistical significance was

analyzed by One-way ANOVA and Tukey’s multiple comparisons test. Reporter gene activity

significantly different than (P < 0.05) TCDD-treated empty vector was denoted with an asterisk.

Reporter gene activity significantly different than (P < 0.05) TCDD-treated full-length TiPARP

(1-657) was denoted with a pound sign.

Page 124: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

109

Figure 27. Overexpression of GFP-tagged TiPARP N-terminal truncations.

HuH7 cells were overexpressed with pEGFP-TiPARP N-terminal truncation constructs for 48 h

and proteins were detected by Western blot. GFP-TiPARP fusions were detected with anti-GFP

antibody. β-actin was used as a loading control.

Page 125: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

110

Page 126: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

111

Figure 28. Effects of TiPARP double alanine point mutants on AHR-regulated reporter

gene activity.

Reporter gene activity of CYP1A1-luc (A) and CYP1B1-luc (B). HuH7 cells were transiently co-

transfected with pcDNA-TiPARP double alanine point mutant and pCYP1A1-luc or pCYP1B1-

luc. Cells were treated with DMSO or TCDD for 16 h and reporter gene activity was determined.

Values were normalized to 100% TCDD treatment. Data were represented as means ± S.E.M. of

three independent experiments. Statistical significance was analyzed by One-way ANOVA and

Tukey’s multiple comparisons test. Reporter gene activity significantly different than (P < 0.05)

TCDD-treated empty vector was denoted with an asterisk. Reporter gene activity significantly

different than (P < 0.05) TCDD-treated wildtype TiPARP was denoted with a pound sign.

Page 127: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

112

8.4 Repression of AHR by TiPARP required zinc finger function Because the zinc-finger domain (residues 237-264) overlapped with the putative repressor

domain (residues 218-245), we tested the importance of the zinc finger domain in TiPARP-

mediated repression by creating single point mutants of the cysteine residues of CCCH-type

zinc-finger motif. We verified the expression of all zinc finger mutants by in vitro translation

with 35S-methionine (Figure 29A). All three TiPARP cysteine point mutants failed to inhibit

TCDD-induced CYP1A1-luc and CYP1B1-luc reporter gene activity (Figure 29B and C). We

created GFP-tagged versions of all three zinc finger point mutants to test expression in

transiently transfected HuH7 cells. We compared the ability of the GFP-tagged zinc finger point

mutants to untagged. All GFP-tagged point mutants failed to repress induction of CYP1A1- and

CYP1B1-luc, which was similar to the untagged mutants (Figure 29D and E). Expression of

GFP-tagged zinc finger point mutants was greater than wildtype GFP-TiPARP despite equal

loading and transfection of the same amount. These observations demonstrated that the lack of

repression of the zinc finger point mutants was not due to lack of expression (Figure 29F). Zinc

finger function was required for TiPARP-mediated repression of AHR transactivation.

Page 128: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

113

Page 129: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

114

Figure 29. Repression of AHR transactivation required zinc finger function.

(A) In vitro translated 35S-labelled TiPARP zinc finger point mutants. pcDNA-TiPARP zinc

finger point mutant constructs were in vitro translated with 35S-methionine and proteins were

detected by SDS-PAGE and autoradiography. (B-C) CYP1A1-luc and CYP1B1-luc reporter

gene activity was not repressed by TiPARP zinc finger point mutants. HuH7 cells were

transiently co-transfected with pcDNA-TiPARP zinc finger point mutant constructs and

pCYP1A1-luc or pCYP1B1-luc. Cells were treated with DMSO or TCDD for 16 h and reporter

gene activity was determined. (D-E) CYP1A1-luc and CYP1B1-luc reporter gene activity was

not repressed by GFP-tagged TiPARP zinc finger point mutants. HuH7 cells were transiently co-

transfected with pEGFP-TiPARP zinc finger point mutant constructs and pCYP1A1-luc or

pCYP1B1-luc. Cells were treated with DMSO or TCDD for 16 h and reporter gene activity was

determined. All luciferase data were normalized to TCDD treatment, which was set to 100%.

Data were represented as means ± S.E.M. of three independent experiments. Statistical

significance was analyzed by One-way ANOVA and Tukey’s multiple comparisons test.

Reporter gene activity significantly different than (P < 0.05) TCDD-treated empty vector was

denoted with an asterisk. Reporter gene activity significantly different than (P < 0.05) TCDD-

treated wildtype TiPARP was denoted with a pound sign. (F) Western blot of GFP-tagged

TiPARP zinc finger mutants. pEGFP-TiPARP zinc finger point mutant constructs were

overexpressed in HuH7 cells for 48 h. Proteins were detected by Western blot using anti-GFP

antibody. Representative Western blot of three independent replicates. β-actin was used as a

loading control.

Page 130: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

115

8.5 Repression of AHR transactivation by TiPARP required catalytic domain and function

To evaluate the importance of TiPARP catalytic domain in mediating the repression of AHR

transactivation, we created C-terminal truncation and catalytic point mutants, and evaluated them

in our transient transfection reporter gene assay. We in vitro translated both C-terminal

truncations of catalytic mutants constructs with 35S-methionine to confirm expression (Figures

30A and 31A). Deletion of the TiPARP C-terminus containing the catalytic domain (residues

449-657) abolished the ability of TiPARP to repress AHR-dependent reporter gene activity

(Figure 30B). The ability of TiPARP to repress CYP1A1- and CYP1B1-mediated reporter gene

activity was prevented with the H532A and Y564A point mutants, but not I631A (Figure 31B),

which retained catalytic activity (Figure 10B). Similar results were observed with GFP-tagged

TiPARP mutants and immunoblotting confirmed that the lack of repression was not due to lack

of protein expression (Figures 30C-D and 31C-D). Together these data demonstrated TiPARP-

mediated repression of AHR transactivation required catalytic activities.

Page 131: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

116

Page 132: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

117

Figure 30. Repression of AHR transactivation required C-terminal catalytic domain.

(A) In vitro translated 35S-labelled TiPARP C-terminal truncations. pcDNA-TiPARP C-terminal

truncation constructs were in vitro translated with 35S-methionine and proteins were detected by

SDS-PAGE and autoradiography. (B-C) CYP1A1-luc and CYP1B1-luc reporter gene activity

was not repressed by TiPARP C-terminal truncations. HuH7 cells were transiently co-transfected

with pcDNA-TiPARP C-terminal truncation constructs and pCYP1A1-luc or pCYP1B1-luc.

Cells were treated with DMSO or TCDD for 16 h and reporter gene activity was determined. (D-

E) CYP1A1-luc and CYP1B1-luc reporter gene activity was not repressed by GFP-tagged

TiPARP C-terminal truncations. HuH7 cells were transiently co-transfected with pEGFP-

TiPARP C-terminal truncation constructs and pCYP1A1-luc or pCYP1B1-luc. Cells were treated

with DMSO or TCDD for 16 h and reporter gene activity was determined. All luciferase data

were normalized to 100% TCDD treatment. Data are represented as means ± S.E.M. of three

independent experiments. Statistical significance was analyzed by One-way ANOVA and

Tukey’s multiple comparisons test. Reporter gene activity significantly different than (P < 0.05)

TCDD-treated empty vector was denoted with an asterisk. (F) Western blot of GFP-tagged

TiPARP C-terminal truncations. pEGFP-TiPARP C-terminal truncation constructs were

overexpressed in HuH7 cells for 48 h. Proteins were detected by Western blot using anti-GFP

antibody. Representative Western blot of three independent replicates. β-actin was used as a

loading control.

Page 133: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

118

Page 134: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

119

Figure 31. Repression of AHR transactivation required catalytic function.

(A) In vitro translated 35S-labelled TiPARP catalytic point mutants. pcDNA-TiPARP catalytic

point mutant constructs were in vitro translated with 35S-methionine and proteins were detected

by SDS-PAGE and autoradiography. (B-C) CYP1A1-luc and CYP1B1-luc reporter gene activity

was not repressed by TiPARP catalytic point mutants. HuH7 cells were transiently co-transfected

with pcDNA-TiPARP catalytic mutant constructs and pCYP1A1-luc or pCYP1B1-luc. Cells

were treated with DMSO or TCDD for 16 h and reporter gene activity was determined. (D-E)

CYP1A1-luc and CYP1B1-luc reporter gene activity was not repressed by GFP-tagged TiPARP

catalytic point mutants. HuH7 cells were transiently co-transfected with pEGFP-TiPARP

catalytic point mutant constructs and pCYP1A1-luc or pCYP1B1-luc. Cells were treated with

DMSO or TCDD for 16 h and reporter gene activity was determined. All luciferase data were

normalized to 100% TCDD treatment. Data are represented as means ± S.E.M. of three

independent experiments. Statistical significance was analyzed by One-way ANOVA and

Tukey’s multiple comparisons test. Reporter gene activity significantly different than (P < 0.05)

TCDD-treated empty vector was denoted with an asterisk. (F) Western blot of GFP-tagged

TiPARP catalytic point mutants. pEGFP-TiPARP catalytic point mutant constructs were

overexpressed in HuH7 cells for 48 h. Proteins were detected by Western blot using anti-GFP

antibody. Representative Western blot of three independent replicates. β-actin was used as a

loading control.

Page 135: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

120

8.6 TiPARP interaction with AHR To examine direct interactions between AHR and TiPARP we performed co-localization and co-

immunoprecipitation experiments. We first determined the subcellular localization of TiPARP.

Endogenous TiPARP localized as small nuclear foci in HuH7 cells (Figure 32A, panel i).

Overexpressed GFP-TiPARP exhibited focal nuclear expression patterns similar to endogenous

TiPARP (Figure 32A, panel ii). GFP-tagged zinc finger mutant C243A displayed cytosol foci,

suggesting the zinc finger domain is important for nuclear localization (Figure 32A, panel iii).

The GFP-tagged catalytic mutant H532A displayed a diffuse nuclear expression pattern,

suggesting catalytic function is important for foci organization (Figure 32A, panel iv). GFP-

tagged mouse TiPARP expressed a similar nuclear focal pattern to human (Figure 32A, panel

v), while GFP-tagged chicken TiPARP diffusely expressed in the cytosol (Figure 32A, panel

vi).

We then tested whether AHR co-localized with GFP-TiPARP in transfected HuH7 cells treated

with TCDD 1.5 h. In co-transfected cells, a fraction of the nuclear AHR was enriched in nuclear

foci containing GFP-TiPARP (Figure 32B). To verify interaction, experiments were repeated in

COS7 cells to score co-localization (Figure 32C). COS7 were selected for scoring because they

contain low levels of endogenous AHR (Xu, et al., 2007) and cells overexpressing AHR could be

easily distinguished from untransfected cells. GFP-TiPARP and AHR co-localization was scored

based on co-transfection and AHR focal staining. COS7 cells transfected with only AHR and

ARNT and stained for AHR were scored for focal staining. We observed AHR focal staining in

75.6% of GFP-TiPARP and AHR and ARNT co-transfected cells, whereas only 5.5% of cells

expressing AHR and ARNT demonstrated focal staining (Table 8).

Table 8. Frequency of co-localization of GFP-TiPARP and AHR in COS7 cells

Transfection Total number of cells AHR focal staining No AHR focal staining

GFP-TiPARP + AHR/ARNT

113

90 (75.6%)

23 (24.4%)

AHR/ARNT

238

13 (5.5%)

225 (94.5%)

Page 136: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

121

Co-immunoprecipitation experiments were then performed using TREx FLP-IN 293 cells

transfected with GFP-TiPARP, AHR and ARNT and treated with DMSO or TCDD for 1.5 h to

support interaction. Similar to co-localization experiments GFP-TiPARP and AHR co-

immunoprecipitated (Figure 33A). Deletion analysis revealed that AHR interacted with residues

275-448 of TiPARP (Figure 33B). Reciprocal co-immunoprecipitation experiments

demonstrated that both the N-terminus (residue 1-425) and C-terminal transactivation domain

(residue 464-848) of AHR co-immunoprecipitated with GFP-TiPARP (Figure 33C and D).

We then tested the ability of the catalytically inactive and non-repressive point mutant GFP-

TiPARP H532A to interact with AHR. COS7 cells co-overexpressed with GFP-TiPARP H532A

and AHR and ARNT demonstrated co-localization within the nucleus (Figure 33E). Co-

immunoprecipitation experiments demonstrated interaction between GFP-TiPARP H532A and

AHR in the presence and absence of TCDD treatment (Figure 33F). Similarly, GFP-TiPARP

C243A, a mutant that did not repress AHR-dependent reporter gene activity, co-

immunoprecipitated with AHR in the presence and absence of TCDD (Figure 33F).

Interestingly, the repressive N-terminal truncation, GFP-TiPARP 200-657, which co-

immunoprecipitated with AHR (Figure 33B) localized within the cytosol and co-localized with

cytosolic AHR (Figure 33G). These observations suggested that TiPARP interaction with AHR

is not dependent on subcellular localization or the repressive function of TiPARP.

Page 137: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

122

Figure 32. TiPARP interacts with AHR in the nucleus.

(A) Nuclear localization of endogenous TiPARP using anti-PARP7 (84664) (panel i),

overexpressed GFP-TiPARP (panel ii), zinc finger mutant GFP-TiPARP-C243A (panel iii),

catalytic mutant GFP-TiPARP-H532A (panel iv), GFP-mouse TiPARP (GFP-mTiPARP, panel

v) and GFP-chicken TiPARP (GFP-chTiPARP, panel vi) in HuH7 cells. Images were acquired

using an Olympus Fluoview 1000 confocal microscope. (B-C) Co-localization of overexpressed

GFP-TiPARP and AHR in HuH7 (B) and COS7 cells (C). For co-localization with AHR cells

were treated with 10 nM TCDD for 1.5 h and fixed, immunostained for AHR, counterstained

with DAPI and mounted using Vectashield. Images were acquired using an Imager.Z1

epifluorescence microscope and Axiovision software (Zeiss) following deconvolution.

Page 138: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

123

Page 139: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

124

Figure 33. TiPARP truncation and point mutant interaction with AHR.

(A) Reciprocal co-immunoprecipitation experiments of overexpressed GFP-TiPARP and AHR.

TREx-FLP-IN 293 cells were transfected and treated with TCDD (T) or DMSO (D) for 1.5 h.

(B) Co-immunoprecipitation experiments of AHR and GFP-TiPARP truncations in TREx-FLP-

IN 293 cells (upper panel). Co-immunoprecipitation experiments with AHR and GFP in TREx-

FLP-IN 293 cells (lower panel). AHR (H-211) antibody was used for immunoprecipitation and

GFP (JL-8) was used to probe. (C-D) Co-immunoprecipitation experiments of GFP-TiPARP and

AHR truncations in TREx-FLP-IN 293 cells. GFP (632460) antibody was used for

immunoprecipitation and AHR (N-19 or H-211) antibodies were used to probe. (E) Co-

localization of overexpressed GFP-tagged TiPARP catalytic mutant H532A (GFP-H532A) and

AHR in COS7 cells. (F) Co-immunoprecipitation experiments of GFP-TiPARP catalytic

(H532A) and zinc finger (C243A) point mutants with AHR in TREx-FLP-IN 293 cells. (G) Co-

localization of overexpressed GFP-tagged TiPARP N-terminal truncation 200-657 (GFP-200-

657) with AHR. For all co-localization experiments transfected cells were treated with 10 nM

TCDD for 1.5 h and fixed, immunostained for AHR, counterstained with DAPI and mounted

using Vectashield. Images were acquired using an Imager.Z1 epifluorescence microscope and

Axiovision software (Zeiss).

Page 140: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

125

8.7 Identification of nuclear localization signal of N-terminus of TiPARP.

Our co-localization observations suggested the presence of a nuclear localization signal (NLS)

within the N-terminus of TiPARP between residues 1 and 200. We then examined the expression

patterns of the N-terminal and C-terminal GFP-TiPARP truncations. We transiently transfected

GFP-tagged TiPARP truncations into HuH7 cells and observed the subcellular localization. We

first overexpressed GFP-tagged N-terminal truncation 33-657 in HuH7 cells and observed

expression patterns similar to full-length GFP-TiPARP (Figure 34A panels i and ii). GFP-

TiPARP 53-657 expressed in the cytosol as foci within the perinuclear region (Figure 34A panel

iii), whereas GFP-TiPARP 103-657 expressed within the perinuclear region with less defined

foci (Figure 34A panel iv). Both GFP-TiPARP 235-657 and GFP-TiPARP 445-657 truncations

expressed in less defined pattern in the perinuclear region (Figure 34A panels v and vi). We

then overexpressed the C-terminal truncations in HuH7 cells and both GFP-TiPARP 1-448 and

1-234 expressed diffusely and exclusively in the nucleus (Figure 34A panels vii and viii). These

data suggested the presence of a nuclear localization signal between residues 33 and 53 of

TiPARP. Between residues 33 and 53 we identified a short sequence consisting of basic amino

acids (KKKDQKR), which is consistent with a putative NLS sequence (Marfori, et al., 2011).

We created a triple mutant of the three tandem lysines (K41A; K42A; K43A) as well as a single

mutant of the first lysine (K41A) and examined their cellular localization in HuH7 cells. Both the

triple and single NLS mutants expressed in the cytosol suggesting that these residues are part of a

functional NLS sequence (Figure 34A panels ix and x).

We next tested the ability of TiPARP NLS mutants to repress AHR transactivation. Using

reporter gene assays we determined both the triple and single NLS mutants retained the ability to

repress AHR-regulated reporter gene activity similar to wildtype (Figure 34B and C). TiPARP

53-657 also retained the ability to repress reporter gene activity similar to wildtype, which was

expected based on our previous findings that TiPARP truncations 33-657 through 225-657

repressed reporter gene activity. Despite the expression of the NLS mutants and N-terminal

truncations, 53-657 through 225-657, in the cytosol, AHR transactivation was still repressed.

These findings suggested that nuclear localization is not required for TiPARP-mediated

repression of AHR transactivation.

Page 141: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

126

Page 142: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

127

Figure 34. Identification of putative nuclear localization signal (NLS) within TiPARP N-

terminus.

(A) Localization of EGFP-TiPARP N-terminal and C-terminal truncations and nuclear

localization single point mutants in HuH7 cells. Panel i: Full-length GFP-TiPARP (1-657), N-

terminal truncations- panel ii: GFP-TiPARP 33-657, panel iii: GFP-TiPARP 53-657, panel iv:

GFP-TiPARP 103-657, panel v: GFP-TiPARP 235-657, panel vi: GFP-TiPARP 445-657, C-

terminal truncations- panel vii: GFP-TiPARP 1-448, panel viii: GFP-TiPARP 1-234, NLS point

mutants- panel ix: GFP-TiPARP K41A; K42A; K43A, panel x: GFP-TiPARP K41A. All GFP

fusions were overexpressed for 24 h and then fixed in 4% paraformaldehyde. Images were

acquired using an Olympus Fluoview 1000 confocal microscope. (B-C) TiPARP 53-657

truncation and NLS mutants repressed CYP1A1-luc reporter gene activity. HuH7 cells were

transiently co-transfected with pcDNA-TiPARP truncation or point mutant expression constructs

(B) or pEGFP-TiPARP truncation or point mutant expression constructs (C). Cells were treated

with DMSO or TCDD for 16 h and reporter gene activity was determined. All luciferase data

were normalized to 100% TCDD treatment. Data are represented as means ± S.E.M. of three

independent experiments. Statistical significance was analyzed by one-way ANOVA and

Tukey’s multiple comparisons test. Reporter gene activity significantly different than (P < 0.05)

TCDD-treated empty vector was denoted with an asterisk.

Page 143: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

128

8.8 TiPARP is a labile factor that negatively regulates AHR transactivation

TCDD-induced proteolytic degradation of AHR has been proposed to be under the control of a

labile factor, referred to as an AHR degradation promoting factor (ADPF) that when inhibited

increases the expression of AHR target genes, such as CYP1A1 (Ma, 2007; Ma and Baldwin,

2000). In support of these findings, pre-treatment with cycloheximide (CHX) resulted in

enhanced TCDD-dependent CYP1A1 mRNA levels (Ma, 2007; Ma and Baldwin, 2000).

Because we observed that TiPARP was a negative regulator of AHR, we were interested to

determine if TiPARP was a labile factor regulating AHR protein levels. To this end, we tested

the stability of TiPARP mRNA by treating T47D cells with actinomycin D and isolating RNA at

the time indicated (Figure 35A). We observed a 50% loss in TiPARP mRNA after 1 h

actinomycin D treatment (Figure 35A), which was in agreement with a previous report (Ma and

Baldwin, 2002). Due to the lack of a suitable antibody to detect endogenous TiPARP protein

levels, we examined the stability of TiPARP by overexpressing GFP-tagged TiPARP in HuH7

and treating with the 26S proteasome inhibitor, MG-132. After 6 h MG-132 treatment we

observed increased GFP-TiPARP expression compared to transiently transfected but untreated

cells (Figure 35B).

These data suggested that TiPARP was a candidate for the labile regulator of AHR protein levels

and responsible for the increased expression of AHR target genes after pre-treatment with CHX

(Ma, 2007; Ma and Baldwin, 2000). To test this notion, we pre-treated wildtype, Tiparp-/- and

Tiparp+/- MEFs for 1 h with CHX prior to treatment with 6 h TCDD. Because we were interested

in determining the effect of CHX on TCDD-dependent target gene expression, we also

determined the fold increase of TCDD+CHX over TCDD alone for each genotype (Figure 35C

and D). CYP1A1 mRNA levels increased approximately 480- and 300-fold following

TCDD+CHX compared to TCDD alone in wildtype and Tiparp+/- MEFs, respectively. CYP1B1

and AHRR mRNA levels were also significantly increased following TCDD+CHX compared to

TCDD alone, but the fold increases were an order of magnitude lower than those observed for

CYP1A1 (Figure 35C). For Tiparp-/- cells, treatment with CHX+TCDD resulted in significant

but modest increases of approximately 2-fold for each of AHR target gene examined compared

to TCDD alone (Figure 35C). A similar pattern of CYP1A1 and CYP1B1 mRNA

Page 144: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

129

superinduction after TCDD treatment was observed after MG-132 pre-treatment, although the

magnitude of induction was lower than that observed with CHX (Figure 35D). AHRR mRNA

was not significantly induced when co-treated with MG-132 for all MEF lines (Figure 35D).

These data suggested TiPARP was an important factor regulating the CHX- and MG-132-

dependent increases in ligand-activated AHR transactivation.

Page 145: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

130

Page 146: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

131

Figure 35. TiPARP is a labile negative regulator of AHR.

(A) TiPARP mRNA is rapidly degraded after actinomycin D treatment. T47D cells were treated

with 1 μg/ml actinomycin D for the times indicated and TiPARP mRNA expression levels

determined by qPCR. (B) GFP-TiPARP overexpression is increased after proteasome inhibition.

HuH7 cells were transfected with pEGFP-TiPARP, pcDNA (vector) and pEGFP for 24 h and

treated with 25 μM MG-132 for 6 h and GFP-TiPARP protein levels were determined by

Western blot using anti-GFP antibody. The data presented were from a representative Western

blot from three independent experiments. (C) AHR target gene induction in MEF lines pre-

treated with 10 μg/ml cycloheximide (CHX). MEF cell lines were pre-treated with 10 μg/ml

CHX for 1 h then treated with TCDD for 6 h and gene expression was determined. Data were

normalized to wildtype DMSO. Fold changes between TCDD alone and CHX+TCDD (CHX+T)

were provided for each gene. Gene expression results were shown as means ± S.E.M for three

independent experiments and significance analyzed by One-way ANOVA and Tukey’s multiple

comparisons test. Gene expression levels significantly different (P < 0.05) than TCDD alone

within each genotype were denoted with an asterisk. (D) MEFs lines were pre-treated with 25

μM MG-132 (MG) for 30 min then treated with TCDD for 4 h and gene expression was

determined. Data were normalized to wildtype DMSO. Fold changes between TCDD alone and

MG-132+TCDD (MG+T) were provided for each gene. Gene expression results were shown as

means ± S.E.M for three independent experiments and significance analyzed by One-way

ANOVA and Tukey’s multiple comparisons test. Gene expression levels significantly different

(P < 0.05) than TCDD alone within each genotype were denoted with an asterisk.

Page 147: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

132

9 Comparison of TiPARP- and AHRR-mediated negative regulation of AHR transactivation

Much of our understanding of the negative regulation of AHR comes from studies of the aryl

hydrocarbon receptor repressor (AHRR), which is a part of a negative feedback loop regulating

AHR transactivation (Evans, et al., 2008; Karchner, et al., 2009; Mimura, et al., 1999). Similar

to TiPARP, AHRR expression is induced by ligand-activated AHR (Mimura, et al., 1999). The

AHRR has been proposed to negatively regulate AHR transactivation by competing with AHR

for ARNT heterodimerization and AHRE binding (Mimura, et al., 1999). However, we have

observed in both Tiparp-/- MEFs and Tiparp-/- mouse liver tissue the TCDD-mediated induction

of AHRR mRNA was increased. This observation contradicts the proposed mechanism of

AHRR-mediated repression of AHR transactivation since greater expression of the AHRR in

Tiparp-/- mice should result in greater repression of AHR target transcription rather than greater

induction. To better understand the mechanism(s) of AHR repression we compared TiPARP- and

AHRR-mediated repression of AHR transactivation.

Page 148: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

133

9.1 Kinetics of AHRR induction by TCDD Because both TiPARP and AHRR gene expression are induced by TCDD we first compared the

temporal changes in TiPARP and AHRR mRNA following TCDD treatment in MCF7 breast

carcinoma cells. TCDD-dependent increases in TiPARP mRNA levels were rapid and peaked

after 1.5 h similar to T47D cells, whereas AHRR mRNA an increased over 24 h (Figure 36A).

We detected AHRR protein after 24 h TCDD treatment (Figure 36B); due to lack of suitable

antibody for TiPARP we were unable to detect TiPARP protein expression following TCDD

treatment. We next examined AHRR recruitment to CYP1A1 and CYP1B1 enhancer regions.

Because we detected AHRR protein after 24 h TCDD treatment we treated cells up to 24 h and

then performed ChIP assays. We observed recruitment of AHRR to the CYP1A1 and CYP1B1

regulatory regions only after 24 h TCDD treatment (Figure 36C). AHR/ARNT recruitment

peaked after 45 min TCDD and declined following 6 h, similar to our previous reports (Figure

36D) (Ahmed, et al., 2009; Pansoy, et al. 2010). We previously observed recruitment of

overexpressed GFP-TiPARP to the CYP1A1 enhancer region following TCDD treatment (Figure

18C), indicating that both TiPARP and AHRR are present at AHR regulatory regions in response

to TCDD.

Page 149: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

134

Page 150: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

135

Figure 36. AHRR is induced by TCDD and recruited to AHR-regulated genes.

(A) Temporal induction of AHRR mRNA and TiPARP mRNA in MCF7 cells. MCF7 cells were

treated with TCDD for the times indicated. (B) AHRR protein expression following 24 h TCDD

or DMSO treatment. (C) Recruitment of AHRR to the CYP1A1 and CYP1B1 regulatory regions.

MCF7 cells were treated with TCDD for the times indicated and AHRR recruitment was

determined by ChIP. (D) Recruitment of AHR and ARNT to the CYP1A1 and CYP1B1

regulatory regions. MCF7 cells were treated with TCDD for the times indicated and AHR and

ARNT recruitment was determined.

Page 151: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

136

9.2 AHRR knockdown did not affect AHR transactivation or protein levels

We have demonstrated that RNAi-mediated knockdown of TiPARP resulted in increased TCDD-

induced CYP1A1 and CYP1B1 mRNA expression and we were then interested in the effects of

AHRR knockdown on CYP1A1 and CYP1B1 mRNA levels. We transfected MCF7 cells with

three siAHRR strands for 24 h and treated cells with DMSO or TCDD for 24 h and determined

AHRR mRNA knockdown. All three siAHRR strands reduced AHRR mRNA levels by 38%-

48% following DMSO treatment (Figure 37A). AHRR mRNA levels were also reduced

following TCDD treatment (Figure 37B). Because AHRR protein was not detected in non-

induced (DMSO) MCF7 cells (Figure 36B), we induced AHRR expression by treating

transfected cells with TCDD 24 h post-transfection to determine AHRR protein knockdown. We

observed substantial knockdown of TCDD-induced AHRR protein by all three siAHRR

sequences tested (Figure 37B). We chose to use siAHRR2 and siAHRR3 strands for all

subsequent experiments because they showed greatest AHRR knockdown. We then examined

the effect of AHRR knockdown on AHR target gene expression. AHRR knockdown did not

significantly affect the TCDD-mediated induction of CYP1A1, CYP1B1 or TiPARP mRNA

levels (Figure 37C). We repeated TiPARP knockdown experiments in MCF7 cells to compare

the effects of TiPARP knockdown to AHRR knockdown. Similar to our previous data (Figure

15C) TiPARP knockdown increased TCDD-induced CYP1A1, CYP1B1 and AHRR mRNA

levels (Figure 37D), indicating the mechanism of AHRR-mediated repression of AHR

transactivation may differ from that of TiPARP.

We were then interested in whether AHRR knockdown affected AHR protein levels.

Knockdown of AHRR did not increase constitutive AHR protein levels (DMSO) nor was

TCDD-induced AHR proteasomal degradation reduced, which is in contrast to the effects of

TiPARP knockdown on AHR protein levels (Figure 38A). These results were, however, in

agreement with previous reports demonstrating AHRR overexpression did not alter AHR protein

levels (Evans, et al., 2008). Since ARNT has been previously proposed to play an important role

in AHRR-mediated repression of AHR (Mimura, et al., 1999) we tested ARNT protein levels

following AHRR knockdown. Basal ARNT protein levels following TCDD treatment were not

affected by AHRR knockdown (Figure 38A). TiPARP knockdown increased constitutive AHR

Page 152: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

137

and ARNT levels and decreased TCDD-induced degradation of both proteins, similar to that

observed in T47D cells (Figure 38B). Both basal and TCDD-induced AHRR protein levels were

not affected by TiPARP knockdown despite elevated TCDD-induced AHRR mRNA levels

(Figure 38B). A potential explanation for the discrepancy between AHRR mRNA and protein

levels could be that the increase in AHRR mRNA levels may not be sufficient to observe

increases in AHRR protein.

Page 153: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

138

Page 154: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

139

Figure 37. Effects of AHRR knockdown and TiPARP knockdown on AHR transactivation.

(A) AHRR mRNA and (B) protein knockdown in MCF7 cells. MCF7 cells were transfected with

non-targeting (NT) or three siAHRR strands, following 24 h knockdown cells were treated with

DMSO (D) or TCDD (T) for 24 h and Western blots were performed. (C) Effects of AHRR

knockdown on AHR target gene induction. AHRR was knocked down in MCF7 cells, treated

with DMSO or TCDD for 24 h and AHR target mRNA expression determined. (D) TiPARP

mRNA knockdown in MCF7 cells (E) Effect of TiPARP knockdown on AHR target gene

induction. TiPARP was knocked down in MCF7 cells for 48 h and cells were treated with

DMSO or TCDD for 24 h. Gene expression data were presented as means ± S.E.M. of three

independent replicates. Statistical significance was determined by One-way ANOVA and

Tukey’s multiple comparisons test. mRNA significantly (P < 0.05) different than treatment-

matched non-targeting (NT) was denoted with an asterisk.

Page 155: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

140

Figure 38. AHRR, AHR and ARNT protein levels following AHRR (A) and TiPARP (B)

knockdown.

AHRR and TiPARP were knockdown in MCF7 cells. Cells were transfected with non-targeting

(NT), siAHRR2 and siAHRR3 or siTiPARP12 and siTiPARP13. Following knockdown cells

were treated with DMSO (D) or TCDD (T) for 24 h and protein levels determined by Western

blot.

Page 156: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

141

9.3 AHRR overexpression repressed AHR-regulated reporter gene activity

We next complemented the AHRR knockdown studies with overexpression studies. Based on a

previous report identifying two isoforms of the AHRR we tested both full-length AHRR

(AHRR719) and the predominantly expressed active form of the AHRR, which lacks exon 8

(AHRRΔ8) (Karchner, et al., 2009). We overexpressed increasing amounts of both forms of

AHRR into HuH7 cells and measured the ability of each form to repress TCDD-induced

CYP1A1-luc activity. Similar to a previous report, overexpression of increasing amounts of

AHRR719 did not affect CYP1A1-luc reporter gene activity, whereas AHRRΔ8 potently inhibited

activity (Figure 39A and B) (Karchner, et al., 2009). We next tested the ability of AHR and

ARNT overexpression to rescue AHRRΔ8 repression. Increasing amounts of AHR but not

ARNT prevented the AHRRΔ8-mediated repression of CYP1A1-luc reporter gene activity as

described previously (Karchner, et al., 2009). These results were also similar to rescue

experiments with TiPARP where AHR but not ARNT rescued TCDD-dependent induction of

CYP1A1-luc reporter gene activity (Figure 39C and D).

Page 157: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

142

Page 158: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

143

Figure 39. Overexpression of AHRRΔ8 repressed AHR-regulated reporter gene activity.

HuH7 cells were co-transfected with pCYP1A1-luc and pcDNA-AHRR719 (A) or pcDNA-

AHRRΔ8 (B) at the amounts indicated. Cells were treated with TCDD for 16 h and reporter gene

activity was determined. Transfection efficiency was normalized to β-glactosidase activity. Data

are presented as means ± S.E.M. of three independent replicates. Co-overexpression of AHR (C)

but not ARNT (D) rescued CYP1A1-luc activity. HuH7 cells were co-transfected with 5 ng

pcDNA-AHRRΔ8 construct and 100-500 ng pRc-CMV-AHR or pcDNA4-ARNT. Transfected

cells were treated with DMSO or TCDD for 16 h and CYP1A1-luc reporter gene activity was

determined. Transfection efficiency was normalized to β-glactosidase activity. Data are

presented as means ± S.E.M. of three independent replicates. Reporter gene activity significantly

different than AHRRΔ8 alone was denoted with an asterisk.

Page 159: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

144

9.4 Expression patterns of the AHRR

To examine the cellular localization patterns of the AHRR we created both N- and C-terminal

GFP-tagged expression constructs of AHRRΔ8 and AHRR719. We overexpressed all constructs

in HuH7, COS7 and MCF7 cells. Unexpectedly, both GFP-tagged AHRRΔ8 and AHRR719

expressed in the cytosol in HuH7 and COS7 cells (Figure 40). The location (N- or C-terminally

linked) of the GFP tag on either AHRRΔ8 or AHRR719 did not affect their expression or

localization pattern. We then repeated overexpression of GFP-tagged AHRR constructs in MCF7

cells, GFP-tagged AHRRΔ8 localized to the nucleus, whereas the GFP-tagged AHRR719

localized diffusely to both the nucleus and cytosol (Figure 40). GFP-tagged TiPARP, when

overexpressed in MCF7 cells, appeared as nuclear foci, similar to that observed in HuH7 and

COS7 cells (Figure 32A). Despite the differences in expression patterns of the AHRR in the

three cell lines tested the AHRR significantly repressed AHR-regulated reporter gene activity in

all three cell lines (Karchner, et al., 2009).

Page 160: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

145

Figure 40. Expression of N-terminal and C-terminal GFP-tagged AHRR in MCF7, HuH7

and COS7 cells.

MCF7, HuH7 and COS7 cells were transfected with pEGFP-C2-AHRRΔ8, pEGFP-C2-

AHRR719, pAcGFP-N1-AHRRΔ8, pAcGFP-N1-AHRR719 or pEGFP-C2-TiPARP. All GFP

fusions were overexpressed for 24 h then fixed with 4% paraformaldehyde. Images were

acquired using an Olympus Fluoview 1000 confocal microscope.

Page 161: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

146

Chapter 5: Discussion 10 TiPARP is a mono-ADP-ribosyltransferase Class 2 ARTDs are predicted to be mono-ADP-ribosyltransferases based on the sequence and

structural features of their catalytic domains (Otto, et al., 2005). The presence of a glutamate

within the catalytic HYE triad motif, which is essential for poly(ADP-ribose) chain elongation is

absent from class 2 ARTD members and restricts their enzymatic activities to mono-ADP-

ribosylation (Marsischky, et al., 1995; Rolli, et al., 1997; Ruf, et al., 1998). In agreement with

this notion, the class 2 ARTDs, ARTD7, 8, 10, 12 and 15 have been shown to exhibit mono-

ADP-ribosyltransferase activity (Aguiar, et al., 2005; Di Paola, et al., 2012; Kleine, et al., 2008;

Leung, et al., 2011). TiPARP (ARTD14) has previously demonstrated to be a catalytically active

enzyme; however, whether this catalytic activity was mono-ADP-ribosylation was not

determined (Ma, et al., 2001). We demonstrated that TiPARP is a mono-ADP-ribosyltransferase

that ADP-ribosylates itself and core histones (section 6). Our results indicate the catalytic

activities of TiPARP are dependent on the catalytic histidine and tyrosine of its HYI motif,

because mutation of these residues abolished auto- and hetero-ADP-ribosylation activities. Auto-

ADP-ribosylation of TiPARP is not affected by the presence of activated DNA or RNA, which is

similar to previous reports for ARTD7 and ARTD10 (Aguiar, et al., 2005; Kleine, et al., 2008).

These results suggest TiPARP is a nucleic acid-independent mono-ADP-ribosyltransferase.

We used two assays to examine the auto-ribosylation activity of TiPARP, the 32P-NAD+ assay

and the 35S protein shift assay. The 32P-NAD+ assay uses purified TiPARP incubated with 32P-

labelled NAD+ is an established methodology to detect catalytic activity of ARTDs by indicating

substrate incorporation and shift in mobility by SDS-PAGE (Shah, et al., 2011). The 35S protein

shift assay, uses in vitro translated 35S-labelled TiPARP to detect changes in mobility of 35S-

TiPARP following incubation with or without NAD+ by SDS-PAGE. This small shift in mobility

(or smearing) that we observed for 35S-TiPARP when incubated with β-NAD+ could be

indicative of short oligo-ADP-ribose polymers, dimers or trimers, or mono-ADP-ribose on

multiple residues of TiPARP (Lindahl, et al., 1995; Satoh, et al., 1994). The actual polymer sizes

of auto-ADP-ribosylated TiPARP can be analyzed by sequencing gel (Alvarez-Gonzalez and

Jacobson, 1987; Tanaka, et al., 1978).

Page 162: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

147

For all variants and species of TiPARP as well as conditions tested (with nucleic acids) results

from both 32P-NAD+ and 35S protein shift assays were in agreement except for the catalytic

isoleucine mutant (I631A) and chicken TiPARP. Results from the 32P-NAD+ assay demonstrated

I631A mutant to be catalytically active while the 35S protein shift assay did not show a mobility

shift when incubated with NAD+. This discrepancy may be due to sensitivity differences between

the two assays with the 32P-NAD+ assay being more sensitive than the 35S protein shift assay. A

mutation of the catalytic isoleucine of ARTD10 to a glutamate greatly reduced auto-ADP-

ribosylation activity but did not abolish it (Kleine, et al., 2008). The 35S protein shift assay may

not be sensitive enough to detect a reduction in the level of TiPARP catalytic activity. This

hypothesis may also explain the discrepancy between the two assays with respect to chicken

TiPARP. We observed weaker auto-ADP-ribosylation activity of chicken TiPARP than

mammalian TiPARPs by the 32P-NAD+ assay but no apparent activity in the 35S protein shift

assay, suggesting that sensitivity may be limited in this assay.

We used the 35S protein shift assay to determine whether TiPARP contains an auto-ribosylation

domain. We observed auto-ADP-ribosylation activity was greatly reduced between residues 275

and 328 of TiPARP suggesting the presence of an auto-ribosylation domain. The BRCT (BRCA1

C-terminus-like) domain of ARTD1 is generally referred to as the auto-ribosylation (or auto-

modification) domain (Loeffler, et al., 2011). Careful dissection of the ARTD1 auto-ribosylation

domain revealed three specific residues that are modified (one aspartate and two glutamates)

(Tao, et al., 2009). Other reports have described lysines within the auto-ribosylation and DNA

binding domain to be target residues for modification (Altmeyer, et al., 2009). A recent study

identified 12 auto-ribosylation sites of ARTD1, namely glutamates, aspartates and lysines, 8 of

these auto-ribosylation sites were found outside of the so-called auto-ribosylation domain

(Chapman, et al., 2013). Based on these findings auto-ribosylation of TiPARP may not be

restricted to a specific domain but is rather dispersed throughout the enzyme on multiple

residues. It has been suggested that there are preferred sites of auto-ribosylation of ARTD1 and

when these preferred or primary sites are unavailable secondary sites become prevalent (Tao, et

al., 2009). Results from our 35S mobility shift assay may suggest TiPARP may contain multiple

primary and secondary sites of auto-ribosylation throughout the protein. Currently it is not clear

which residues of ARTD family members serve as modification sites. Glutamate and aspartate

residues were reported to be auto-ribosylation sites on ARTD10 (Kleine, et al., 2008; Rosenthal,

Page 163: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

148

et al., 2013). Other residues that have been reported to be modified by eukaryotic MARTs

include arginine, cysteine, phosphoserine and asparagine and together with glutamate and

aspartate represent potential auto-ribosylated sites on TiPARP (Corda and Di Girolamo, 2003;

Okazaki and Moss, 1996; Ord and Stocken, 1977; Seman, et al., 2004; Smith and Stocken,

1975). Our laboratory is currently determining specific sites of auto-ADP-ribosylation of

TiPARP by mass spectrometry. We have identified 6 distinct peptides that are mono-ADP-

ribosylated. These findings confirm the mono-ADP-ribosyltransferase activity of TiPARP, but

we have not mapped or identified the modified amino acid residues. We have initiated electron

dissociation transfer mass spectrometry to specifically identify the ribosylated residues.

We have demonstrated that TiPARP possesses hetero-ADP-ribosylation activity and ADP-

ribosylates core histones. ARTD1, 2, 3 and 10 have also been described so far to ADP-ribosylate

histones (Kleine, et al., 2008; Messner, et al., 2010; Rulten, et al., 2011). Histone ribosylation

links ADP-ribosylation to other modifications such as acetylation, methylation and

phosphorylation, which constitute an epigenetic code for histones (Messner and Hottiger, 2011).

The contribution of histone ribosylation to the histone code and epigenetic regulation is not well

known. Moreover, ribosylated histones only constitute a small fraction (< 1%) of total histones

(Boulikas, 1989). Recently, macrodomain-containing proteins such as ARTD8, ARTD9,

MacroD1, MacroD2 and C6orf130 have been reported to interact with ADP-ribosylated target

proteins and represent potential ‘readers’ of ADP-ribosylation (Jankevicius, et al., 2013;

Rosenthal, et al., 2013; Sharifi, et al., 2013). MacroD1, MacroD2 and C6orf130 have been

identified as mono-ADP-ribosylhydrolases that bind mono-ADP-ribose and mediate its removal

from target proteins (Rosenthal, et al., 2013). Potential ‘readers’ and mono-ADP-

ribosylhydrolases of ADP-ribosylated histones by TiPARP have not been determined, but would

be an interesting future direction as it will help expand our understanding of the function of

histone ribosylation.

11 TiPARP is a novel negative regulator of AHR transactivation

We determined that TiPARP is a novel negative regulator of AHR transactivation. Our studies

using human cell lines and MEFs isolated from wildtype and Tiparp-deficient mice reveal that

TiPARP modulates ligand-induced proteolytic degradation of AHR and functions as a repressor

Page 164: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

149

of AHR transactivation. In this model, TCDD-activated AHR induces TiPARP expression,

TiPARP is recruited to AHR target genes either through direct DNA binding or tethering through

AHR, increased levels of TiPARP enhance the proteolytic degradation of AHR, by either direct

ribosylation or through the ribosylation of unidentified intermediary factor(s) to repress AHR

target gene transcriptional activation (Figure 41).

Figure 41. Proposed model of repression of AHR transactivation by TiPARP.

The AHR mediates the TCDD-induced expression of TiPARP (1). TiPARP is recruited to the

AHR/ARNT-bound AHRE of targets to repress transcription (CYP1A1 in this example) (2A).

TiPARP represses its own transcription and the regulation of other AHR target genes by binding

to AHR/ARNT-bound AHRE or potentially through direct DNA binding, creating a negative

feedback loop (2B) in which AHR dissociates from AHRE, translocates from the nucleus and is

proteolytically degraded by the 26S proteasome (26S) (3). BTF; basal transcription factors, Co-

act; co-activators.

Page 165: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

150

11.1 Potential co-repression by ADP-ribosylation of AHR The ability of TiPARP to repress AHR transactivation requires its catalytic domain and the direct

mono-ADP-ribosylation of AHR by TiPARP is an obvious mechanism to consider. Previous

studies have reported that ARTDs act as co-regulators (co-activators or co-repressors) for a

number of different transcriptional regulators (see section 2.5.4 for further information). In some

cases, enzymatic activity is required while in others it is not. ADP-ribosylation of Oct-1 and Sp1

electrostatically repels them from DNA resulting in altered transcript expression profiles (Gagne,

et al., 2008; Nie, et al., 1998; Smith, 2001). ADP-ribosylation of p53 and NF-κB prevents their

association with other nuclear factors, which also affected gene expression (Kanai, et al., 2007;

Zerfaoui, et al., 2010). It is possible that mono-ADP-ribosylation of AHR by TiPARP prevents

AHR binding to AHREs or its association with co-regulatory proteins. AHR that is not

complexed with ARNT is susceptible to degradation which could explain the changes in AHR

protein levels we observe when TiPARP expression altered (Davarinos and Pollenz, 1999; Heid,

et al., 2000; Kazlauskas, et al., 2000; Shetty, et al., 2003). The impact of mono-ADP-

ribosylation on transcription factor activity is not well understood and it is unclear whether

mono-ADP-ribosylation can alter transcription factor function by the same mechanism or extent

as poly(ADP-ribose). However it should be noted that mono-ADP-ribosylation of a single amino

acid residue by bacterial toxins, such as diphtheria and cholera toxins, can drastically alter target

protein function leading to severe physiological consequences (Collier, 2001).

The potential that ADP-ribosylation represents a novel posttranslational modification of AHR is

extremely exciting. AHR is subject to posttranslational modifications including SUMOylation,

ubiquitination and phosphorylation (Chen and Tukey, 1996; Pongratz, et al., 1991; Xing, et al.,

2012). Phosphorylation of specific serine and tyrosine residues of AHR is important for DNA-

binding and maximal transactivation, demonstrating modification of AHR affects its function

(Chen and Tukey, 1996; Park, et al., 2000; Pongratz, et al., 1991). Therefore ADP-ribosylation

of AHR could alter many processes involved in the AHR transactivation such as ARNT

dimerization, DNA-binding, association with co-regulatory or basal transcription factors, nuclear

import/export and transactivation. Because the structure and function of the modular domains of

AHR are well understood, mapping of potential mono-ADP-ribosylation sites by TiPARP will be

very informative in determining the effects ADP-ribosylation has on its function. We are

Page 166: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

151

currently optimizing purification of AHR protein to determine if it is hetero-ribosylated by

TiPARP and if so identify ADP-ribosylation using in vitro assays and mass spectrometry.

11.2 Potential targeting of histones and other co-regulatory factors

Core histones, linker histone H1 and chromatin-associated proteins are all targets of ADP-

ribosylation by ARTD1 (Krishnakumar and Kraus, 2010b). ADP-ribosylation (mono- or poly-)

of H1, H2A and H2B by ARTD1 may play a role in the regulation of chromatin structure,

although the general extent of histone modification to transcriptional regulation remains to be

clarified (D'Amours, et al., 1999). ARTD1 can displace the H1 by ADP-ribosylation and block

binding of H1 to promoter chromatin creating a structure permissive to transcription (Ju, et al.,

2004; Kim, et al., 2004; Krishnakumar, et al., 2008). The hetero-ADP-ribosylation of core

histones by TiPARP is a potential mechanism by which TiPARP modulates chromatin structure

at active AHR regulatory regions to repress transcription. TiPARP could also target co-

regulatory proteins with histone-modifying activities that are known to interact with AHR/ARNT

(see Table 3) altering their intrinsic histone-modifying or chromatin-remodelling activities by

ADP-ribosylation. Many of the known AHR co-activators have histone acetyltransferase (HAT)

activity such as NCoA1, 2, 3 and p300 (Hankinson, 2005; Harper, et al., 2006). Histone ADP-

ribosylation by TiPARP could reduce co-activator activity by competing with or blocking

histone acetylation and thus reducing transactivation. Poly-ADP-ribosylation of histone lysine

demethylase KDM5B has been shown to block binding of KDM5B to chromatin and inhibit its

demethylase activity supporting the hypothesis that hetero-ADP-ribosylation by TiPARP may

reduce AHR co-activator function (Krishnakumar and Kraus, 2010a). TiPARP could also

modulate other chromatin-associated proteins to alter chromatin to architecture more favourable

for repression. ARTD1 also hetero-ribosylates DEK, a chromatin-associated protein and

promotes its release from chromatin suggesting TiPARP may potentially modulate other

chromatin-associated proteins besides co-activators to alter chromatin to architecture in favour of

repression (Gamble and Fisher, 2007; Kappes, et al., 2008). Proteomics will be very useful to

identify interacting proteins with TiPARP and provide insights into the mechanism of TiPARP-

mediated repression.

Page 167: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

152

The auto-ribosylation activity of TiPARP may also be important for its repressive function.

Auto-ribosylation accounts for approximately 90% of ARTD1 total catalytic activity indicating

the enzyme itself is the primary target (D'Amours, et al., 1999). Auto-ribosylated ARTD1

provides a binding platform for DNA repair enzymes and scaffolding proteins (Gagne, et al.,

2008; Hegde, et al., 2008). Therefore it is possible that auto-ribosylated TiPARP recruits co-

repressor proteins including known AHR co-repressors SMRT, SHP or AHRR to promote

transcriptionally repressive complex assembly at the AHREs of target genes (Harper, et al.,

2006; Mimura, et al., 1999; Nguyen, et al., 1999). Alternatively, it is equally possible that auto-

ribosylated TiPARP at AHREs could also prevent the association of AHR/ARNT or basal

transcription factors with co-activators. Identification of potential auto-ribosylation sites on

TiPARP will help clarify their importance in mediating the repressive function of TiPARP.

11.3 TiPARP is a candidate AHR degradation promoting factor (ADPF)

Our proposed model of TiPARP-mediated repression (Figure 41) is reminiscent of the ADPF

hypothesis which proposes the induction of a labile factor that is sensitive to proteasomal and

protein synthesis inhibition and promotes the proteasomal degradation of AHR (Ma and

Baldwin, 2000; Ma and Baldwin, 2002; Ma, et al., 2000). We show by use of transcriptional and

proteasomal inhibitors that TiPARP mRNA and protein is rapidly degraded indicating TiPARP is

labile. Protein synthesis inhibition of Tiparp-/- MEFs did not result in superinduction of AHR

target genes when cells were treated with TCDD which suggested newly generated TiPARP

represses AHR transactivation. We also observed TiPARP knockdown or knockout increases

AHR protein levels and transactivation mimicking the effects of protein synthesis and

proteasomal inhibition (Ma and Baldwin, 2000). Overall, our data support TiPARP as a

candidate labile ADPF.

The ubiquitin-proteasome pathway plays an important role in TCDD-mediated degradation of

AHR (Ma and Baldwin, 2000; Pollenz, 2007). ARTD1 plays a role in the reactive oxygen-

induced activation of the proteasome (Ullrich, et al., 1999). During oxidative stress auto-

ribosylated ARTD1 interacts with and activates nuclear proteasome pathways leading to the

rapid up-regulation of 20S proteasome activity to remove damaged histones more efficiently

(Ullrich, et al., 1999). Inhibition of ARTD5 and ARTD6 activity prevents the degradation of

Page 168: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

153

axin through the ubiquitin-proteasome pathway and this increased half-life of axin promotes the

β-catenin degradation (Mayer-Kuckuk, et al., 1999). These findings demonstrate there is

interplay between the nuclear proteasome proteins and poly-ADP-ribosylation. It is therefore

tempting to hypothesize TiPARP activates components of the proteasome to enhance the TCDD-

mediated AHR degradation. It is also equally possible that TiPARP directly ADP-ribosylates

AHR, targeting it for degradation. Iduna is a poly(ADP-ribose)-dependent E3 ubiquitin ligase

that binds and ubiquitinates both poly-ADP-ribosylated and poly(ADP-ribose)-binding proteins

labelling these proteins for ubiquitin proteasomal degradation and provides an example of how

ADP-ribosylation can be used to target proteins for degradation (Kang, et al., 2011).

11.3.1 TiPARP and the nuclear proteasome

TiPARP localizes to the nucleus and exhibits a focal expression pattern, which can be described

as nuclear bodies (von Mikecz, 2006). It has been established that proteasomes occur in the

nucleus and are associated with nuclear bodies including PML bodies, nucleoplasmic speckles

and focal clusters throughout the nucleoplasm representing proteolytic centres in the nucleus

(Chen, et al., 2002; Fabunmi, et al., 2001; Rockel, et al., 2005). TiPARP expression is stabilized

by proteasomal inhibition resulting in its accumulation in the nucleus (data not shown) and

suggests that TiPARP is subject to proteolytic degradation via the 26S proteasome. All TiPARP

catalytic mutants and cytosolic N-terminal truncations and point mutants (both zinc finger and

NLS mutants) are more highly expressed compared with wildtype indicating that components

within the nucleus and the catalytic activity of TiPARP are responsible for its instability.

Because TiPARP affects AHR protein levels and AHR interacts with TiPARP within the nuclear

foci it is reasonable to hypothesize that TiPARP associates with nuclear bodies containing

proteasomes to target AHR for proteasomal degradation. The catalytic mutant TiPARP-H532A

focal expression pattern was greatly altered but still interacted with AHR. This suggests that

catalytic activity of TiPARP is required for association with proteolytic centres in the nucleus to

degrade AHR, repressing its transactivation. The effects of overexpression of TiPARP catalytic

mutants on AHR protein levels will help clarify whether catalytic activity is required for the

proteolytic degradation of AHR. Together, our findings indicate TiPARP is associated with the

nuclear proteasome and this association is dependent upon its catalytic activity. Currently, we are

Page 169: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

154

determining which nuclear body types are associated with TiPARP to provide evidence that

TiPARP is associated with the nuclear proteasome.

11.3.2 Connecting nuclear localization with repression of AHR transactivation

We identified a functional NLS of TiPARP within its N-terminus between residues 41-47

(Figure 42). The zinc finger domain is also important for nuclear localization of TiPARP. Zinc

finger motifs have been identified as the NLS of several zinc finger proteins (Kuwahara, et al.,

2000; Matheny, et al., 1994; Racca, et al., 2011; Spittau, et al., 2007). Mutation or removal of

either the NLS or zinc finger results in the cytosolic localization of TiPARP. The purpose of

having two different NLSs is currently not known. One significant difference between the NLS

and zinc finger is the NLS mutants still repress AHR transactivation whereas zinc finger mutants

do not. These findings suggest repression is not dependent on nuclear localization but rather zinc

finger function. Zinc finger mutants still interact with AHR presumably in the cytosol indicating

the zinc finger is important for repression and potential proteasomal degradation of AHR. Zinc

finger domains of proteins are known to play a role in protein-protein interactions and thus the

zinc finger domain of TiPARP could play a role in the association with the cytosolic proteasome

rather than with AHR (Kang, et al., 2004; Zhang, et al., 2011). Determining interacting partners

by proteomics and AHR protein levels following overexpression of zinc finger TiPARP mutants

can be used to test this hypothesis.

All C-terminal truncations of TiPARP are nuclear but do not repress AHR transactivation. In

constrast, N-terminal truncations 53-657 to 225-657 are cytosolic and still repess AHR activity.

TiPARP truncation 200-657 which is cytosolic and functional and interacts with AHR in the

cytosol. Because AHR exists in the nucleus and cytosol this observation is not completely

unexpected. It is possible the TiPARP 200-657 truncation interacts with the cytosolic proteasome

to degrade AHR and repress transactivation, representing an alternative pathway to nuclear AHR

degradation. Determining AHR protein levels in cells overexpressed with TiPARP 200-657 will

clarify this alternative pathway. The association with AHR and repression of transactivation by

TiPARP 200-657 may also be independent of proteasomal degradation and instead TiPARP 200-

657 prevents the nuclear translocation of AHR. As far as we know all the truncations and

mutants we created do not exist endogenously in cells and their activity, or lack thereof, is

observed experimentally using overexpression but may not occur biologically. During apoptosis

Page 170: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

155

ARTD1 is cleaved into 24 and 89 kDa fragments and each fragment is functional (Chaitanya, et

al., 2010). The 24 kDa fragment containing two of the zinc fingers localizes to the nucleus and

binds DNA irreversible and the 89 kDa contains the auto-modification and catalytic domains and

retains its catalytic activity but localizes to the cytosol (D'Amours, et al., 2001; Soldani, et al.,

2001). Therefore it is also not completely unexpected that truncated cytosolic TiPARP variants

repress AHR and that potential endogenous TiPARP truncations may exist and exhibit a similar

ability to repress AHR.

Figure 42. Updated schematic of the structure of TiPARP.

NLS; nuclear localization signal, Zn; CCCH-type zinc finger domain, WWE; tryptophan-

tryptophan-glutamate.

11.4 Identification of an AHR repressive domain The use of N-terminal truncations identified a putative AHR repressive domain between residues

218 and 245 (Figure 42). Alanine scan results narrowed this region to between residues 220 and

235. Alanine scanning also identified tyrosine 232 (Y232) and histidine 233 (H233) to be

important for repression of AHR transactivation. The exact function and contribution of Y232

and H233 to repression is not clear and we can only speculate their roles. Both tyrosine and

histidine residues are phosphorylated which can alter protein function and conformation (Miller,

2012; Steeg, et al., 2003). For example, AHR phosphorylation at tyrosine residues is important

for its AHRE-binding activity (Minsavage, et al., 2003; Park, et al., 2000). Therefore it is

possible that phosphorylation of Y232 and/or H233 is important for TiPARP-mediated

repression of AHR transactivation. Tyrosines are commonly sulfated which strengthens protein-

protein interactions, Y232 may also be important for stabilizing interactions with intermediary

Page 171: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

156

proteins involved in the repression of AHR (Seibert and Sakmar, 2008). It is unlikely that Y232

or H233 are important for interaction with AHR since the TiPARP 275-657 truncation interacted

with AHR. Given the proximity of H233 to the zinc finger domain (four residues away) and the

fact that histidine is commonly involved in coordinating the zinc ion of zinc fingers, H233 could

be involved with zinc finger function and its loss resembles a zinc finger mutation (Alberts, et

al., 1998).

12 Comparing TiPARP- with AHRR-mediated transrepression

Much of our understanding of AHR repression comes from studies of AHRR, which is also part

of negative feedback loop regulating AHR function (Evans, et al., 2008; Karchner, et al., 2009;

Mimura, et al., 1999). Despite its discovery more than ten years ago and several reports from

independent laboratories, the precise mechanism of AHRR-mediated transrepression of AHR is

not well understood. Conflicting reports between independent laboratories has been a major

contributing factor for our lack of understanding of the AHRR. We compared the AHR

repressive functions of TiPARP to the AHRR to better understand the mechanisms of negative

regulation of AHR transactivation.

12.1 Gene expression and recruitment As expected, both TiPARP and the AHRR gene expression are induced by TCDD treatment. The

AHRR expression gradually increased over 24 h, whereas TiPARP induction peaked after 1.5 h

and declined thereafter. In line with gene expression kinetics, we detected endogenous AHRR

protein and its recruitment to endogenous AHRE-containing regulatory regions after 24 h TCDD

treatment. This recruitment can be supported by previous reports that used in vitro methods to

determine AHRR interaction with AHRE-containing sequences (Evans, et al., 2008; Mimura, et

al., 1999). Recruitment of the AHRR to AHR target regulatory regions coincides with reduced

AHR/ARNT recruitment, suggesting that AHRR displaces AHR/ARNT binding which is similar

to a previous study that describes the AHRR recruitment to an AHRE-containing regulatory

region of AHRR correlated with AHR/ARNT recruitment loss (Haarmann-Stemmann, et al.,

2007). Similarly, we observed TCDD-dependent recruitment of GFP-TiPARP and AHRR to the

CYP1A1 regulatory region suggesting that both TiPARP and the AHRR can influence the

AHR/ARNT complexes at AHREs. The mechanism of how the AHRR and TiPARP alter

Page 172: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

157

AHR/ARNT complexes at AHREs may differ. TiPARP alters AHR protein levels and the

enhanced AHR proteolytic degradation which could lead to reduced AHR recruitment to its

target genes. The AHRR does not affect the proteolytic degradation of AHR therefore the

reduced AHR/ARNT recruitment we observed after 6 h and 24 h is not likely due AHR

degradation (Evans, et al., 2008).

The direct AHRE-binding of the AHRR is not involved in the mechanism of repression of the

AHRR. This is supported by studies using an AHRR DNA-binding mutant that localizes to the

nucleus and still substantially represses AHR transactivation (Evans, et al., 2008). Direct AHRE-

binding is also not essential for TiPARP-mediated repression as TiPARP NLS mutants and N-

terminal truncations from residue 53 to residue 225 still repress AHR-regulated reporter gene

activity despite localizing to the cytosol. Both TiPARP and the AHRR can directly interact with

AHR; therefore, it is possible that TiPARP and the AHRR directly interact with AHRE-bound as

well as unbound AHR to repress transactivation.

12.2 Overexpression of ARNT and co-regulators Overexpression of AHR but not ARNT significantly reduces repression by AHRR or TiPARP,

suggesting that AHR rather than ARNT is the primary target of both repressors (Evans, et al.,

2008; Karchner, et al., 2009). The lack of involvement of ARNT does not support the original

model AHRR-mediated repression, which proposed the AHRR competes with AHR for ARNT

heterodimerization and AHRE binding (Mimura, et al., 1999). Additionally, overexpression of

many known AHR co-activators did not rescue the inducibility of AHR-mediated reporter gene

activity indicating competition with co-activators is not involved in the AHRR mechanism of

repression. Histone deacetylases 4 and 5 (HDAC4 and HDAC5) are co-repressors that are

associated with the AHRR perhaps recruitment of co-repressors is a mechanism by which the

AHRR can negatively repress AHR transactivation (Oshima, et al., 2009; Oshima, et al., 2007).

12.3 Subcellular localization Previous reports have described that the AHRR is a nuclear protein with an active NLS located

within its N-terminus (Evans, et al., 2008; Kanno, et al., 2007). However we observe GFP-

tagged AHRR constructs to localize to the nucleus in a cell context-dependent manner. Only in

MCF7 cells did we observe the nuclear localization of the AHRR, AHRR localized to the cytosol

Page 173: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

158

in HuH7 and COS7 cells. TiPARP localized to the nucleus in all three cell lines. The reason for

the different localization patterns of the AHRR among the cell lines is currently not known.

ARNT expression has been proposed to influence the localization of the AHRR (Kanno, et al.,

2007). The expression of an AHRR NLS mutant in COS7 cells was shifted from the cytosol to

the nucleus when ARNT was co-overexpressed suggesting ARNT expression can influence

AHRR localization (Kanno, et al., 2007). We observed that AHRR localizes to the cytosol in

both COS7 cells which have low levels of ARNT and in HuH7 cells which express ARNT, so it

does not appear that ARNT expression influences the localization of the AHRR. However, we

did not co-overexpress ARNT with the AHRR in COS7 and HuH7 cells therefore it is possible

that overexpressed ARNT influences subscellular localization of the AHRR. Despite the

difference in localization of the AHRR, the AHRR is a potent repressor of AHR transactivation

in all three cell lines (Evans, et al., 2008; Karchner, et al., 2009). This observation is similar to

that observed with the TiPARP NLS mutants and N-terminal truncations from residue 53 up to

residue 225 that repress AHR transactivation but localize to the cytosol. The AHRR has

previously been reported to not alter nuclear translocation of AHR by comparing the nuclear

translocation of AHR when AHRR was co-overexpressed although these reports did not examine

AHR and AHRR co-localization. Although the subcellular localization of endogenous AHRR

has not yet been determined, our recruitment findings suggest that a proportion of endogenous

AHRR is nuclear.

12.4 Knockdown and knockout models Our knowledge on the AHRR function is largely based on overexpression models, and when

overexpressed the AHRR is an effective repressor of AHR-regulated reporter gene activity.

However, findings from AHRR overexpression studies are not fully supported by knockout

models (Hosoya, et al., 2008; Tigges, et al., 2013). Previous studies that have reported that cell

lines that have low AHR ligand-dependent inducibility of CYP1A1 have high basal levels of

AHRR and knockdown of the AHRR increases basal CYP1A1 expression and AHR ligand

inducibility (Haarmann-Stemmann, et al., 2007; Oshima, et al., 2007). We observed the

constitutive levels of CYP1A1 mRNA in MCF7 cells were not significantly affected by AHRR

knockdown. Our results are supported by a recent report that describes a non-causal relationship

between AHRR expression and CYP1A1 induction and activity human skin fibroblasts (Tigges,

et al., 2013). Reports using fibroblasts isolated from Ahrr-deficient mice demonstrated

Page 174: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

159

impressive CYP1A1 induction by AHR ligands which suggest that the knockdown levels we

achieved in MCF7 cells were not sufficient to significantly increase inducibility of AHR target

genes (Hosoya, et al., 2008; Tigges, et al., 2013). Alternatively, given that the increased

CYP1A1 expression levels by AHR ligand treatment was limited to the heart, spleen and skin of

Ahrr-deficient mice our choice of cell type (MCF7 breast carcinoma) may have limited results.

TiPARP knockdown in all cell types that we examined which includes two human breast

carcinoma, hepatoma and gastric carcinoma lines and the Tiparp-deficient MEFs displayed

significantly greater AHR target inducibility which suggests TiPARP to be a general negative

regulator of AHR transactivation and the AHRR to be a tissue/cell type-specific repressor.

Interestingly, Ahrr-deficient MEFs demonstrated significantly greater constitutive AHR mRNA

levels compared to wildtype MEFs (Tigges, et al., 2013). In contrast, Tiparp-deficient MEFs do

not display significantly different basal AHR mRNA levels and altered transcriptional expression

of the AHR represents a potential difference between the modes of repression of TiPARP and

AHRR.

12.5 Potential overlap of TiPARP and the AHRR repressive pathways

A comparison of what is known about the AHR repressive functions of TiPARP and the AHRR

reveals similarities but also some notable differences. Similarities between TiPARP- and AHRR-

mediated inhibition of AHR transactivation include: (1) both genes are induced by TCDD and

regulated by AHR; (2) both interact with AHR; and (3) increased ARNT expression fails to

prevent AHRR- or TiPARP-dependent repression of AHR transactivation (Evans, et al., 2008;

Karchner, et al., 2009; Lo, et al., 2011; Mimura, et al., 1999). In contrast, AHRR repression of

AHR action does not involve increased AHR degradation, direct binding to AHREs, or inhibition

of the AHR nuclear translocation (Evans, et al., 2008; Hahn, et al., 2009). These similarities and

differences suggest there could be pathway overlap and interplay between TiPARP and the

AHRR.

The kinetic profiles of TCDD-regulated expression of TiPARP and AHRR suggest that TiPARP

induction could be an immediate mechanism to repress AHR transactivation, while the AHRR is

a more long-term mechanism of repression. We observe the expression and recruitment of

TiPARP after 45 min of TCDD treatment, whereas the expression and recruitment of the AHRR

Page 175: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

160

occurs after 24 h TCDD treatment. The rapid induction and recruitment of TiPARP can be a

primary response to repress AHR and the more gradual induction of AHRR expression and

recruitment to AHR target genes may represent a secondary mechanism of repression. TiPARP

knockdown and knockout increased TCDD-induced expression of AHRR mRNA and other AHR

target genes. Initially this might seem counterintuitive since greater expression of AHRR should

decrease AHR target expression rather than enhance it; however, this could indicate that AHRR

requires TiPARP for optimal repression. Because in the absence of TiPARP the increase of

AHRR expression alone is not sufficient to repress AHR target induction this suggests TiPARP

and the AHRR function in tandem or co-operatively. Both TiPARP and AHRR can directly

interact with AHR and overexpression of AHR can restore induction of AHR targets when co-

overexpressed with either TiPARP or AHRR suggesting that AHR is a common target of both

repressors (Karchner, et al., 2009). Our finding that TCDD induction of AHR targets in Tiparp-

deficient MEFs with protein synthesis inhibition was approximately 2-fold greater than TCDD

treatment alone indicates that TiPARP induction is not solely responsible for the observed

repression and that inhibition of AHRR synthesis might also contribute to the increases. Double

knockout of TiPARP and AHRR will be very helpful in determining potential co-operatively

between TiPARP and the AHRR to repress AHR transactivation.

13 TiPARP is an in vivo negative regulator of AHR transactivation

Hepatic tissue from Tiparp-deficient mice show significantly greater TCDD-induced AHR target

gene expression compared with tissue isolated from wildtype mice, supporting that TiPARP is an

in vivo negative regulator of AHR transactivation. This AHR target induction profile was not

observed in hepatic tissue from 3MC-treated Ahrr-deficient mice (Hosoya, et al., 2008).

Increased CYP1A1 induction of Ahrr-deficient mice was only observed in spleen, heart and skin

tissues, which indicates that AHRR is a tissue-specific repressor (Hosoya, et al., 2008).

Preliminary results indicate that kidney and spleen tissue from 6 h TCDD-treated Tiparp-

deficient mice show a similar AHR target induction profile to hepatic tissue suggesting TiPARP

to potentially be a more wide-spread repressor than the AHRR (data not shown). The tissue

distribution and increased induction CYP1A1 of Tiparp-deficient mice is similar to CA-AHR

transgenic mice which display heightened CYP1A1 mRNA in many tissues including liver,

kidney and spleen (Andersson, et al., 2002; Brunnberg, et al., 2006; Moennikes, et al., 2004).

Page 176: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

161

However it should be noted that beyond CYP1A1, the gene induction profile of hepatic tissue

from CA-AHR mice is very different from hepatic tissue isolated from TCDD-treated wildtype

mice and most of the TCDD-responsive genes, including TiPARP, were not detected in CA-

AHR mouse livers (Dere, et al., 2006; Kurachi, et al., 2002; Moennikes, et al., 2004; Zeytun, et

al., 2002). TCDD-treated Tiparp-deficient mice demonstrate significantly greater induction of

many TCDD-responsive genes indicating that uninhibited AHR transactivation by the lack of

repression of AHR is different from continuous AHR activation as observed in the CA-AHR

mouse model.

13.1 Role of TiPARP in TCDD-mediated toxicity

13.1.1 Purported effects of TiPARP on the NAD+ pool and SIRT activity

Our knowledge of a role of TiPARP in TCDD-mediated toxicity has been limited to studies that

use chicken (Gallus gallus) TiPARP to examine the effects of TiPARP on the suppression of

gluconeogenesis that is associated with wasting syndrome (Diani-Moore, et al., 2010; Diani-

Moore, et al., 2013). Chicken TiPARP was identified as a mediator of the TCDD-induced

suppression of gluconeogenesis that consumes hepatic NAD+, lowering SIRT activity, repressing

gluconeogenic genes G6Pase and PEPCK, and reducing glucose output (Diani-Moore, et al.,

2010). TiPARP induction was also purported to be involved in AHR-mediated sensitivity to

NASH by depleting NAD+ levels which lowers SIRT3 activity leading decreased SOD2 activity

and less superoxide scavenging (He, et al., 2013). Conclusions from both of these studies were

largely based observations that TCDD treatment increases TiPARP gene expression and

decreases NAD+ levels suggesting that NAD+ levels control gene expression (Diani-Moore, et

al., 2010; He, et al., 2013). We considered whether lower NAD+ levels rather than TiPARP

directly could be responsible for the repression of AHR transactivation; however, addition of

nicotinamide, a NAD+ precursor also represses induction of AHR targets indicating that NAD+

levels and TiPARP-mediated repression of AHR targets are independent (Diani-Moore, et al.,

2010).

TiPARP was found to contribute to TCDD-mediated suppression of gluconeogenesis because

TCDD treatment induces TiPARP mRNA and decreases gluconeogenic gene expression and

subsequently reduces glucose output (Diani-Moore, et al., 2010). TiPARP overexpression

reduces PEPCK mRNA and glucose output, mimicking the effects of TCDD, leading the authors

Page 177: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

162

to conclude that TiPARP mediates TCDD-induced suppression of gluconeogenesis (Diani-

Moore, et al., 2010). If TiPARP is a mediator of TCDD toxicity, Tiparp-deficient mice should be

more resistant to TCDD-induced gluconeogenesis suppression and wasting; however,

preliminary studies suggest that Tiparp-deficient mice are more sensitive to TCDD suggesting

TiPARP plays a protective rather than a mediator role in TCDD-mediated toxicity (data not

shown). Toxicity studies with Tiparp-deficient mice are currently ongoing.

Chicken TiPARP is reported to be a poly-ADP-ribosylating ARTD that is able to deplete the

cellular NAD+ pool (Diani-Moore, et al., 2010; Diani-Moore, et al., 2013). We have

demonstrated using catalytic assays that chicken TiPARP is a mono-ADP-ribosyltransferase,

which agrees with the structural prediction of its catalytic activity (Otto, et al., 2005). The choice

of assay to demonstrate catalytic activity is most likely the reason for this discrepancy. We use

purified GST-tagged chicken TiPARP incubated with 32P-NAD+, the generally accepted method

to examine catalytic activity of a specific ARTD whereas Diani-Moore et al. used an indirect

method by immunoprecipitating poly(ADP-ribose) from whole cell extracts transfected with

TiPARP or treated with TCDD as an indirect means to study TiPARP catalytic activity (Diani-

Moore, et al., 2010; Diani-Moore, et al., 2013). The anti-poly(ADP-ribose) antibodies are able to

detect six or more ADP-ribose units and cannot detect mono-ADP-ribose accurately, therefore it

is unclear why an anti-poly(ADP-ribose) antibody is able to detect catalytic activity directly from

TiPARP (Kleine, et al., 2008). It is still unclear how TiPARP, which is a MART, is capable of

significantly depleting the NAD+ pool. ARTD1 accounts for 85-90% of cellular ADP-ribose

synthesis suggesting that TiPARP is not the primary consumer of cellular NAD+ (Shieh, et al.,

1998). The reported depletion of cellular NAD+ could be due to stress caused by transfection of

TiPARP which leads to the activation of ARTD1 and increase in poly(ADP-ribose) activity and

thus decreases cellular NAD+ levels. Additionally, we observe chicken TiPARP is less

catalytically active than mammalian TiPARPs making TiPARP less likely of a candidate in

chicken cells to substantially deplete the NAD+ pool.

Another difference between chicken TiPARP and mammalian TiPARPs is when transfected into

mammalian cells chicken TiPARP did not repress AHR transactivation or form nuclear foci. The

lack of activity could be due to cell context, since we transfected chicken TiPARP into

immortalized human cell lines. These findings suggest that TiPARP exhibits species- or model-

specific differences. Chicken TiPARP is also less catalytically active than mammalian TiPARPs

Page 178: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

163

illustrating another species difference between mammalian and chicken TiPARP. Since chickens

are more sensitive to TCDD than mice (LD50 <25 μg/kg bw for chickens versus 182 μg/kg bw

for B6 strain mice), it is possible that the lower TiPARP activity could be a potential contributing

factor to the higher TCDD sensitivity reported in chickens (Chapman and Schiller, 1985; Greig,

et al., 1973).

14 TiPARP-mediated regulation of other transcription factors

The ability of TiPARP to regulate transcription is most likely not limited to AHR. ARTD1 and

ARTD2 regulate many transcriptional factors acting as co-activators and co-repressors in

different pathways (section 2.5.4) (Krishnakumar and Kraus, 2010b; Szanto, et al., 2012).

Despite its name, the AHRR has also shown activity toward transcription factors other than AHR

including hypoxia inducible factor 1 alpha (HIF-1α) (Karchner, et al., 2009). TiPARP is

regulated by other transcription factors including estrogen receptor alpha (ERα) and

glucocorticoid receptor supporting the possibility that TiPARP could be involved in these

pathways (Kininis, et al., 2007; Reddy, et al., 2009). Ongoing studies in our laboratory reveal

that TiPARP interacts with ERα and negatively regulates ERα-mediated transactivation,

whereas TiPARP co-activates liver X receptor (LXR)-mediated transactivation (data not shown).

All together our results indicate that TiPARP is a multifunctional transcription factor co-

regulator.

15 Limitations and future considerations

15.1 TiPARP antibody Our studies of TiPARP function were highly dependent on transient transfection to overexpress

or knockdown TiPARP. We were unable to verify TiPARP protein overexpression, knockdown

of endogenous TiPARP or TCDD-induced endogenous TiPARP due to the lack of suitable

commercially available antibody able to detect TiPARP. To verify knockdown we resorted to

overexpressing GFP-tagged TiPARP with siTiPARP sequences and using an anti-GFP antibody

to detect GFP-TiPARP protein levels. Similarly, we overexpressed GFP-TiPARP to perform

ChIP assays and test GFP-TiPARP protein stability. Plasmid-derived overexpressed protein is

not identical to endogenous protein, which is why a high quality anti-TiPARP antibody is critical

Page 179: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

164

for studying the endogenous TiPARP protein behaviour. An anti-TiPARP antibody would also

be extremely useful to detect TCDD-induced TiPARP expression, recruitment to AHR target

regulatory regions and co-immunoprecipitation with endogenous AHR. Due to the lack of a

suitable antibody we were also unable to confirm complete knockout in the MEFs isolated from

Tiparp-deficient mice at the protein level, leaving open the possibility Tiparp-deficient mice to

express a truncated version of TiPARP. To date the only evidence we have of endogenous

TiPARP protein are confocal images of putative TiPARP staining in the nuclei of HuH7 cells by

a commercially available anti-TiPARP antibody that in our hands only works for

immunofluorescence. The specificity of the antibody was tested with overexpressing GFP-

TiPARP and staining for TiPARP (data not shown). We are currently purifying TiPARP protein

for generation of our own antibody. Once generated, we intend to use the antibody to detect and

confirm TiPARP protein expression in all our models. We also plan to use the antibody to

perform ChIP-sequencing to determine TiPARP recruitment and potential overlapping

recruitment with AHR and ARNT. A proper anti-TiPARP antibody will certainly be an

important tool for our laboratory.

15.2 ARTD catalytic activity assays The in vitro ADP-ribosylation assays, which involve incubation of purified ARTD protein and 32P-NAD+ are relatively simple and is a well-accepted method to demonstrate ARTD activity.

Although catalytic activity assays provide insights into auto- and hetero-ADP-ribosylation of

ARTDs, they do not identify specific sites/residues of ADP-ribosylation. Furthermore, although

ARTD catalytic assays do illustrate extensive poly-ADP-ribosylation with the reduced mobility

of modified protein by SDS-PAGE they do not provide information about polymer length.

Because the molecular weight of a single ADP-ribose moiety is quite small (540 Da) mono-

ADP-ribosylated and short oligo-ADP-ribosylated proteins are visually indistinguishable (Laing,

et al., 2011). Sequencing gels of modified proteins can be used to illustrate the lengths of ADP-

ribose chains. We used catalytic activity assays to demonstrate TiPARP has auto- and hetero-

ADP-ribosylation activities and TiPARP is most likely a mono-ADP-ribosyltransferase.

However, we did not quantify the amount of ADP-ribose units TiPARP is able to synthesize

leaving open the possibility that TiPARP can synthesis short oligo-ADP-ribose polymers that

display no visual slowing of migration by SDS-PAGE. Because catalytic activity assays are non-

quantitative, small differences in the level of activity may not be visually observed. For instance,

Page 180: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

165

the TiPARP-I631A and –I631E mutants are catalytically active; however, it is not clear if the

level of catalytic activity of either mutant differs from that of wildtype. Our GST-purified

TiPARP and mutant protein preparations contain breakdown products making quantification of

full-length enzyme difficult. To compare the activities of wildtype TiPARP and these isoleucine

mutants requires highly pure protein to ensure equal quantities of wildtype and mutants are used

for each assay. We are currently optimizing our protein purification protocol for TiPARP to

reduce breakdown products and increase purity.

Our catalytic assays with TiPARP also did not identify specific sites or residues on TiPARP that

were auto-ADP-ribosylated or sites of core histones that were hetero-ADP-ribosylated by

TiPARP. Determining the sites or residues of modification will demonstrate whether TiPARP

has a preference for a particular amino acid. We are currently using mass spectrometry to

determine sites of auto-ADP-ribosylation on TiPARP, by using mass spectrometry we are able to

identify specific residues of TiPARP that are auto-ADP-ribosylated and confirm that these sites

are in fact mono-ADP-ribosylated. Once we purify large enough quantities of AHR protein we

will determine whether it is hetero-ADP-ribosylated by TiPARP and use mass spectrometry to

identify specific sites of AHR that are mono-ADP-ribosylated.

In vitro catalytic assays may be artificial and may not accurately reflect modification in a cellular

context or in vivo. Direct evidence of intracellular protein mono-ADP-ribosylation is quite

limited due to the lack of antibody to detect mono-ADP-ribosylated proteins (Feijs, et al., 2013).

The challenge of providing direct evidence for intracellular mono-ADP-ribosylation is

compounded by the lack of specific mono-ARTD inhibitors (Feijs, et al., 2013). Although

TiPARP displays auto- and hetero-ADP-ribosylation in vitro our understanding of whether this

occurs in cell or in vivo is limited.

16 Overall significance of research The objective of this research was to understand TiPARP and its function in AHR

transactivation. We provided evidence that TiPARP is a nuclear MART that mono-ADP-

ribosylates itself and core histones. We identified TiPARP as a novel AHR ligand-induced

negative regulator of AHR transactivation in vitro and in vivo. We also identified its NLS

sequence providing further evidence for its nuclear localization. TiPARP specifically targets

AHR and promotes its proteolytic degradation and represents a candidate for the elusive ADPF.

Page 181: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

166

The catalytic and zinc finger functions are essential for TiPARP to repress AHR transactivation.

Comparison with the AHRR reveals some notable similarities and differences suggesting

TiPARP and the AHRR use partially overlapping but also distinct mechanisms to repress AHR

activity. Overall we have established a new mechanism of negative feedback regulation of AHR

transactivation and identified a nuclear mono-ADP-ribosyltransferase.

Page 182: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

167

References Abbott, B. D., Birnbaum, L. S. and Perdew, G. H. (1995). Developmental expression of two

members of a new class of transcription factors: I. Expression of aryl hydrocarbon receptor in the C57BL/6N mouse embryo. Dev Dyn 204; 133-43.

Abdellatif, M. (2012). Sirtuins and pyridine nucleotides. Circulation research 111; 642-56.

Adachi, J., Mori, Y., Matsui, S., Takigami, H., Fujino, J., Kitagawa, H., Miller, C. A., 3rd, Kato, T., Saeki, K. and Matsuda, T. (2001). Indirubin and indigo are potent aryl hydrocarbon receptor ligands present in human urine. J Biol Chem 276; 31475-8.

Adamietz, P. and Rudolph, A. (1984). ADP-ribosylation of nuclear proteins in vivo. Identification of histone H2B as a major acceptor for mono- and poly(ADP-ribose) in dimethyl sulfate-treated hepatoma AH 7974 cells. J Biol Chem 259; 6841-6.

Aguiar, R. C., Takeyama, K., He, C., Kreinbrink, K. and Shipp, M. A. (2005). B-aggressive lymphoma family proteins have unique domains that modulate transcription and exhibit poly(ADP-ribose) polymerase activity. J Biol Chem 280; 33756-65.

Aguiar, R. C., Yakushijin, Y., Kharbanda, S., Salgia, R., Fletcher, J. A. and Shipp, M. A. (2000). BAL is a novel risk-related gene in diffuse large B-cell lymphomas that enhances cellular migration. Blood 96; 4328-34.

Ahmed, S., Valen, E., Sandelin, A. and Matthews, J. (2009). Dioxin increases the interaction between aryl hydrocarbon receptor and estrogen receptor alpha at human promoters. Toxicol Sci 111; 254-66.

Akiyama, T., Takasawa, S., Nata, K., Kobayashi, S., Abe, M., Shervani, N. J., Ikeda, T., Nakagawa, K., Unno, M., Matsuno, S. and Okamoto, H. (2001). Activation of Reg gene, a gene for insulin-producing beta -cell regeneration: poly(ADP-ribose) polymerase binds Reg promoter and regulates the transcription by autopoly(ADP-ribosyl)ation. Proc Natl Acad Sci U S A 98; 48-53.

Alberts, I. L., Nadassy, K. and Wodak, S. J. (1998). Analysis of zinc binding sites in protein crystal structures. Protein science : a publication of the Protein Society 7; 1700-16.

Alkhatib, H. M., Chen, D. F., Cherney, B., Bhatia, K., Notario, V., Giri, C., Stein, G., Slattery, E., Roeder, R. G. and Smulson, M. E. (1987). Cloning and expression of cDNA for human poly(ADP-ribose) polymerase. Proc Natl Acad Sci U S A 84; 1224-8.

Altmeyer, M., Barthel, M., Eberhard, M., Rehrauer, H., Hardt, W. D. and Hottiger, M. O. (2010). Absence of poly(ADP-ribose) polymerase 1 delays the onset of Salmonella enterica serovar Typhimurium-induced gut inflammation. Infection and immunity 78; 3420-31.

Altmeyer, M., Messner, S., Hassa, P. O., Fey, M. and Hottiger, M. O. (2009). Molecular mechanism of poly(ADP-ribosyl)ation by PARP1 and identification of lysine residues as ADP-ribose acceptor sites. Nucleic Acids Res 37; 3723-38.

Alvarez-Gonzalez, R. and Althaus, F. R. (1989). Poly(ADP-ribose) catabolism in mammalian cells exposed to DNA-damaging agents. Mutat Res 218; 67-74.

Alvarez-Gonzalez, R. and Jacobson, M. K. (1987). Characterization of polymers of adenosine diphosphate ribose generated in vitro and in vivo. Biochemistry 26; 3218-24.

Page 183: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

168

Ambrose, H. E., Papadopoulou, V., Beswick, R. W. and Wagner, S. D. (2007). Poly-(ADP-ribose) polymerase-1 (Parp-1) binds in a sequence-specific manner at the Bcl-6 locus and contributes to the regulation of Bcl-6 transcription. Oncogene 26; 6244-52.

Ame, J. C., Fouquerel, E., Gauthier, L. R., Biard, D., Boussin, F. D., Dantzer, F., de Murcia, G. and Schreiber, V. (2009). Radiation-induced mitotic catastrophe in PARG-deficient cells. J Cell Sci 122; 1990-2002.

Ame, J. C., Rolli, V., Schreiber, V., Niedergang, C., Apiou, F., Decker, P., Muller, S., Hoger, T., Menissier-de Murcia, J. and de Murcia, G. (1999). PARP-2, A novel mammalian DNA damage-dependent poly(ADP-ribose) polymerase. J Biol Chem 274; 17860-8.

Ame, J. C., Spenlehauer, C. and de Murcia, G. (2004). The PARP superfamily. Bioessays 26; 882-93.

Amiri, K. I., Ha, H. C., Smulson, M. E. and Richmond, A. (2006). Differential regulation of CXC ligand 1 transcription in melanoma cell lines by poly(ADP-ribose) polymerase-1. Oncogene 25; 7714-22.

Anderson, M. G., Scoggin, K. E., Simbulan-Rosenthal, C. M. and Steadman, J. A. (2000). Identification of poly(ADP-ribose) polymerase as a transcriptional coactivator of the human T-cell leukemia virus type 1 Tax protein. Journal of virology 74; 2169-77.

Andersson, P., McGuire, J., Rubio, C., Gradin, K., Whitelaw, M. L., Pettersson, S., Hanberg, A. and Poellinger, L. (2002). A constitutively active dioxin/aryl hydrocarbon receptor induces stomach tumors. Proc Natl Acad Sci U S A 99; 9990-5.

Andersson, P., Ridderstad, A., McGuire, J., Pettersson, S., Poellinger, L. and Hanberg, A. (2003). A constitutively active aryl hydrocarbon receptor causes loss of peritoneal B1 cells. Biochem Biophys Res Commun 302; 336-41.

Antonsson, C., Arulampalam, V., Whitelaw, M. L., Pettersson, S. and Poellinger, L. (1995a). Constitutive function of the basic helix-loop-helix/PAS factor Arnt. Regulation of target promoters via the E box motif. J Biol Chem 270; 13968-72.

Antonsson, C., Whitelaw, M. L., McGuire, J., Gustafsson, J. A. and Poellinger, L. (1995b). Distinct roles of the molecular chaperone hsp90 in modulating dioxin receptor function via the basic helix-loop-helix and PAS domains. Mol Cell Biol 15; 756-65.

Aravind, L. (2001). The WWE domain: a common interaction module in protein ubiquitination and ADP ribosylation. Trends Biochem Sci 26; 273-5.

Atasheva, S., Akhrymuk, M., Frolova, E. I. and Frolov, I. (2012). New PARP gene with an anti-alphavirus function. Journal of virology 86; 8147-60.

Augustin, A., Spenlehauer, C., Dumond, H., Menissier-De Murcia, J., Piel, M., Schmit, A. C., Apiou, F., Vonesch, J. L., Kock, M., Bornens, M. and De Murcia, G. (2003). PARP-3 localizes preferentially to the daughter centriole and interferes with the G1/S cell cycle progression. J Cell Sci 116; 1551-62.

Baba, T., Mimura, J., Gradin, K., Kuroiwa, A., Watanabe, T., Matsuda, Y., Inazawa, J., Sogawa, K. and Fujii-Kuriyama, Y. (2001). Structure and expression of the Ah receptor repressor gene. J Biol Chem 276; 33101-10.

Page 184: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

169

Bai, C. and Tolias, P. P. (1996). Cleavage of RNA hairpins mediated by a developmentally regulated CCCH zinc finger protein. Mol Cell Biol 16; 6661-7.

Bai, P., Canto, C., Brunyanszki, A., Huber, A., Szanto, M., Cen, Y., Yamamoto, H., Houten, S. M., Kiss, B., Oudart, H., Gergely, P., Menissier-de Murcia, J., Schreiber, V., Sauve, A. A. and Auwerx, J. (2011). PARP-2 regulates SIRT1 expression and whole-body energy expenditure. Cell metabolism 13; 450-60.

Bai, P., Houten, S. M., Huber, A., Schreiber, V., Watanabe, M., Kiss, B., de Murcia, G., Auwerx, J. and Menissier-de Murcia, J. (2007). Poly(ADP-ribose) polymerase-2 [corrected] controls adipocyte differentiation and adipose tissue function through the regulation of the activity of the retinoid X receptor/peroxisome proliferator-activated receptor-gamma [corrected] heterodimer. J Biol Chem 282; 37738-46.

Barbarulo, A., Iansante, V., Chaidos, A., Naresh, K., Rahemtulla, A., Franzoso, G., Karadimitris, A., Haskard, D. O., Papa, S. and Bubici, C. (2012). Poly(ADP-ribose) polymerase family member 14 (PARP14) is a novel effector of the JNK2-dependent pro-survival signal in multiple myeloma. Oncogene

Bell, C. E. and Eisenberg, D. (1996). Crystal structure of diphtheria toxin bound to nicotinamide adenine dinucleotide. Biochemistry 35; 1137-49.

Benjamin, R. C. and Gill, D. M. (1980). Poly(ADP-ribose) synthesis in vitro programmed by damaged DNA. A comparison of DNA molecules containing different types of strand breaks. J Biol Chem 255; 10502-8.

Berger, F., Ramirez-Hernandez, M. H. and Ziegler, M. (2004). The new life of a centenarian: signalling functions of NAD(P). Trends Biochem Sci 29; 111-8.

Berger, N. A. (1985). Poly(ADP-ribose) in the cellular response to DNA damage. Radiation research 101; 4-15.

Bernshausen, T., Jux, B., Esser, C., Abel, J. and Fritsche, E. (2006). Tissue distribution and function of the Aryl hydrocarbon receptor repressor (AhRR) in C57BL/6 and Aryl hydrocarbon receptor deficient mice. Archives of toxicology 80; 206-11.

Bick, M. J., Carroll, J. W., Gao, G., Goff, S. P., Rice, C. M. and MacDonald, M. R. (2003). Expression of the zinc-finger antiviral protein inhibits alphavirus replication. Journal of virology 77; 11555-62.

Birnbaum, L. S. (1994). The mechanism of dioxin toxicity: relationship to risk assessment. Environ Health Perspect 102 Suppl 9; 157-67.

Birnbaum, L. S. and Tuomisto, J. (2000). Non-carcinogenic effects of TCDD in animals. Food additives and contaminants 17; 275-88.

Bittinger, M. A., Nguyen, L. P. and Bradfield, C. A. (2003). Aspartate aminotransferase generates proagonists of the aryl hydrocarbon receptor. Mol Pharmacol 64; 550-6.

Bjeldanes, L. F., Kim, J. Y., Grose, K. R., Bartholomew, J. C. and Bradfield, C. A. (1991). Aromatic hydrocarbon responsiveness-receptor agonists generated from indole-3-carbinol in vitro and in vivo: comparisons with 2,3,7,8-tetrachlorodibenzo-p-dioxin. Proc Natl Acad Sci U S A 88; 9543-7.

Page 185: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

170

Blackshear, P. J. (2002). Tristetraprolin and other CCCH tandem zinc-finger proteins in the regulation of mRNA turnover. Biochem Soc Trans 30; 945-52.

Blander, G. and Guarente, L. (2004). The Sir2 family of protein deacetylases. Annu Rev Biochem 73; 417-35.

Boehler, C., Gauthier, L. R., Mortusewicz, O., Biard, D. S., Saliou, J. M., Bresson, A., Sanglier-Cianferani, S., Smith, S., Schreiber, V., Boussin, F. and Dantzer, F. (2011). Poly(ADP-ribose) polymerase 3 (PARP3), a newcomer in cellular response to DNA damage and mitotic progression. Proc Natl Acad Sci U S A 108; 2783-8.

Bonicalzi, M. E., Haince, J. F., Droit, A. and Poirier, G. G. (2005). Regulation of poly(ADP-ribose) metabolism by poly(ADP-ribose) glycohydrolase: where and when? Cellular and molecular life sciences : CMLS 62; 739-50.

Boulikas, T. (1989). DNA strand breaks alter histone ADP-ribosylation. Proc Natl Acad Sci U S A 86; 3499-503.

Boutros, P. C., Bielefeld, K. A., Pohjanvirta, R. and Harper, P. A. (2009). Dioxin-dependent and dioxin-independent gene batteries: comparison of liver and kidney in AHR-null mice. Toxicological sciences : an official journal of the Society of Toxicology 112; 245-56.

Boverhof, D. R., Burgoon, L. D., Tashiro, C., Sharratt, B., Chittim, B., Harkema, J. R., Mendrick, D. L. and Zacharewski, T. R. (2006). Comparative toxicogenomic analysis of the hepatotoxic effects of TCDD in Sprague Dawley rats and C57BL/6 mice. Toxicol Sci 94; 398-416.

Broadbent, T. A. and Broadbent, H. S. (1998). 1 - 1. The chemistry and pharmacology of indole-3-carbinol (indole-3-methanol) and 3-(methoxymethyl)indole. [Part I]. Curr Med Chem 5; 337-52.

Brochu, G., Shah, G. M. and Poirier, G. G. (1994). Purification of poly(ADP-ribose) glycohydrolase and detection of its isoforms by a zymogram following one- or two-dimensional electrophoresis. Analytical biochemistry 218; 265-72.

Brunnberg, S., Andersson, P., Lindstam, M., Paulson, I., Poellinger, L. and Hanberg, A. (2006). The constitutively active Ah receptor (CA-Ahr) mouse as a potential model for dioxin exposure--effects in vital organs. Toxicology 224; 191-201.

Brunnberg, S., Andersson, P., Poellinger, L. and Hanberg, A. (2011). The constitutively active Ah receptor (CA-AhR) mouse as a model for dioxin exposure - effects in reproductive organs. Chemosphere 85; 1701-6.

Bunger, M. K., Glover, E., Moran, S. M., Walisser, J. A., Lahvis, G. P., Hsu, E. L. and Bradfield, C. A. (2008). Abnormal liver development and resistance to 2,3,7,8-tetrachlorodibenzo-p-dioxin toxicity in mice carrying a mutation in the DNA-binding domain of the aryl hydrocarbon receptor. Toxicological sciences : an official journal of the Society of Toxicology 106; 83-92.

Bunger, M. K., Moran, S. M., Glover, E., Thomae, T. L., Lahvis, G. P., Lin, B. C. and Bradfield, C. A. (2003). Resistance to 2,3,7,8-tetrachlorodibenzo-p-dioxin toxicity and abnormal liver development in mice carrying a mutation in the nuclear localization sequence of the aryl hydrocarbon receptor. J Biol Chem 278; 17767-74.

Page 186: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

171

Burbach, K. M., Poland, A. and Bradfield, C. A. (1992). Cloning of the Ah-receptor cDNA reveals a distinctive ligand-activated transcription factor. Proc Natl Acad Sci U S A 89; 8185-9.

Burd, C. G. and Dreyfuss, G. (1994). Conserved structures and diversity of functions of RNA-binding proteins. Science 265; 615-21.

Burkle, A. and Virag, L. (2013). Poly(ADP-ribose): PARadigms and PARadoxes. Molecular aspects of medicine

Butler, A. J. and Ordahl, C. P. (1999). Poly(ADP-ribose) polymerase binds with transcription enhancer factor 1 to MCAT1 elements to regulate muscle-specific transcription. Mol Cell Biol 19; 296-306.

Camicia, R., Bachmann, S. B., Winkler, H. C., Beer, M., Tinguely, M., Haralambieva, E. and Hassa, P. O. (2013). BAL1/ARTD9 represses the anti-proliferative and pro-apoptotic IFNgamma-STAT1-IRF1-p53 axis in diffuse large B-cell lymphoma. J Cell Sci 126; 1969-80.

Carver, L. A., Jackiw, V. and Bradfield, C. A. (1994). The 90-kDa heat shock protein is essential for Ah receptor signaling in a yeast expression system. J Biol Chem 269; 30109-12.

Casper, R. F., Quesne, M., Rogers, I. M., Shirota, T., Jolivet, A., Milgrom, E. and Savouret, J. F. (1999). Resveratrol has antagonist activity on the aryl hydrocarbon receptor: implications for prevention of dioxin toxicity. Mol Pharmacol 56; 784-90.

Cervellera, M. N. and Sala, A. (2000). Poly(ADP-ribose) polymerase is a B-MYB coactivator. J Biol Chem 275; 10692-6.

Chaitanya, G. V., Steven, A. J. and Babu, P. P. (2010). PARP-1 cleavage fragments: signatures of cell-death proteases in neurodegeneration. Cell communication and signaling : CCS 8; 31.

Chalmers, A., Johnston, P., Woodcock, M., Joiner, M. and Marples, B. (2004). PARP-1, PARP-2, and the cellular response to low doses of ionizing radiation. Int J Radiat Oncol Biol Phys 58; 410-9.

Chambon, P., Weill, J. D. and Mandel, P. (1963). Nicotinamide mononucleotide activation of new DNA-dependent polyadenylic acid synthesizing nuclear enzyme. Biochem Biophys Res Commun 11; 39-43.

Chapman-Smith, A., Lutwyche, J. K. and Whitelaw, M. L. (2004). Contribution of the Per/Arnt/Sim (PAS) domains to DNA binding by the basic helix-loop-helix PAS transcriptional regulators. J Biol Chem 279; 5353-62.

Chapman, D. E. and Schiller, C. M. (1985). Dose-related effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in C57BL/6J and DBA/2J mice. Toxicol Appl Pharmacol 78; 147-57.

Chapman, J. D., Gagne, J. P., Poirier, G. G. and Goodlett, D. R. (2013). Mapping PARP-1 Auto-ADP-ribosylation Sites by Liquid Chromatography-Tandem Mass Spectrometry. Journal of proteome research

Chen, I., Safe, S. and Bjeldanes, L. (1996). Indole-3-carbinol and diindolylmethane as aryl hydrocarbon (Ah) receptor agonists and antagonists in T47D human breast cancer cells. Biochem Pharmacol 51; 1069-76.

Page 187: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

172

Chen, M., Rockel, T., Steinweger, G., Hemmerich, P., Risch, J. and von Mikecz, A. (2002). Subcellular recruitment of fibrillarin to nucleoplasmic proteasomes: implications for processing of a nucleolar autoantigen. Molecular biology of the cell 13; 3576-87.

Chen, Y. H. and Tukey, R. H. (1996). Protein kinase C modulates regulation of the CYP1A1 gene by the aryl hydrocarbon receptor. J Biol Chem 271; 26261-6.

Chi, N. W. and Lodish, H. F. (2000). Tankyrase is a golgi-associated mitogen-activated protein kinase substrate that interacts with IRAP in GLUT4 vesicles. J Biol Chem 275; 38437-44.

Chiang, Y. J., Hsiao, S. J., Yver, D., Cushman, S. W., Tessarollo, L., Smith, S. and Hodes, R. J. (2008). Tankyrase 1 and tankyrase 2 are essential but redundant for mouse embryonic development. PloS one 3; e2639.

Cho, S. H., Ahn, A. K., Bhargava, P., Lee, C. H., Eischen, C. M., McGuinness, O. and Boothby, M. (2011). Glycolytic rate and lymphomagenesis depend on PARP14, an ADP ribosyltransferase of the B aggressive lymphoma (BAL) family. Proc Natl Acad Sci U S A 108; 15972-7.

Ciolino, H. P., Daschner, P. J. and Yeh, G. C. (1999). Dietary flavonols quercetin and kaempferol are ligands of the aryl hydrocarbon receptor that affect CYP1A1 transcription differentially. Biochem J 340 ( Pt 3); 715-22.

Cohen-Armon, M., Visochek, L., Rozensal, D., Kalal, A., Geistrikh, I., Klein, R., Bendetz-Nezer, S., Yao, Z. and Seger, R. (2007). DNA-independent PARP-1 activation by phosphorylated ERK2 increases Elk1 activity: a link to histone acetylation. Mol Cell 25; 297-308.

Collart, C., Remacle, J. E., Barabino, S., van Grunsven, L. A., Nelles, L., Schellens, A., Van de Putte, T., Pype, S., Huylebroeck, D. and Verschueren, K. (2005). Smicl is a novel Smad interacting protein and cleavage and polyadenylation specificity factor associated protein. Genes Cells 10; 897-906.

Collier, R. J. (2001). Understanding the mode of action of diphtheria toxin: a perspective on progress during the 20th century. Toxicon : official journal of the International Society on Toxinology 39; 1793-803.

Cook, B. D., Dynek, J. N., Chang, W., Shostak, G. and Smith, S. (2002). Role for the related poly(ADP-Ribose) polymerases tankyrase 1 and 2 at human telomeres. Mol Cell Biol 22; 332-42.

Corda, D. and Di Girolamo, M. (2002). Mono-ADP-ribosylation: a tool for modulating immune response and cell signaling. Science's STKE : signal transduction knowledge environment 2002; pe53.

Corda, D. and Di Girolamo, M. (2003). Functional aspects of protein mono-ADP-ribosylation. Embo J 22; 1953-8.

Cortes, U., Tong, W. M., Coyle, D. L., Meyer-Ficca, M. L., Meyer, R. G., Petrilli, V., Herceg, Z., Jacobson, E. L., Jacobson, M. K. and Wang, Z. Q. (2004). Depletion of the 110-kilodalton isoform of poly(ADP-ribose) glycohydrolase increases sensitivity to genotoxic and endotoxic stress in mice. Mol Cell Biol 24; 7163-78.

Page 188: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

173

Crews, S. T. (1998). Control of cell lineage-specific development and transcription by bHLH-PAS proteins. Genes Dev 12; 607-20.

D'Amours, D., Desnoyers, S., D'Silva, I. and Poirier, G. G. (1999). Poly(ADP-ribosyl)ation reactions in the regulation of nuclear functions. Biochem J 342 ( Pt 2); 249-68.

D'Amours, D., Sallmann, F. R., Dixit, V. M. and Poirier, G. G. (2001). Gain-of-function of poly(ADP-ribose) polymerase-1 upon cleavage by apoptotic proteases: implications for apoptosis. J Cell Sci 114; 3771-8.

Dalton, T. P., Puga, A. and Shertzer, H. G. (2002). Induction of cellular oxidative stress by aryl hydrocarbon receptor activation. Chem Biol Interact 141; 77-95.

Dantzer, F., de La Rubia, G., Menissier-De Murcia, J., Hostomsky, Z., de Murcia, G. and Schreiber, V. (2000). Base excision repair is impaired in mammalian cells lacking Poly(ADP-ribose) polymerase-1. Biochemistry 39; 7559-69.

Dantzer, F., Mark, M., Quenet, D., Scherthan, H., Huber, A., Liebe, B., Monaco, L., Chicheportiche, A., Sassone-Corsi, P., de Murcia, G. and Menissier-de Murcia, J. (2006). Poly(ADP-ribose) polymerase-2 contributes to the fidelity of male meiosis I and spermiogenesis. Proc Natl Acad Sci U S A 103; 14854-9.

Davarinos, N. A. and Pollenz, R. S. (1999). Aryl hydrocarbon receptor imported into the nucleus following ligand binding is rapidly degraded via the cytosplasmic proteasome following nuclear export. J Biol Chem 274; 28708-15.

David, K. K., Andrabi, S. A., Dawson, T. M. and Dawson, V. L. (2009). Parthanatos, a messenger of death. Frontiers in bioscience : a journal and virtual library 14; 1116-28.

Davidovic, L., Vodenicharov, M., Affar, E. B. and Poirier, G. G. (2001). Importance of poly(ADP-ribose) glycohydrolase in the control of poly(ADP-ribose) metabolism. Experimental cell research 268; 7-13.

De Lucia, F., Mennella, M. R., Quesada, P. and Farina, B. (1996). Poly(ADPribosyl)ation system in transcriptionally active rat testis chromatin fractions. J Cell Biochem 63; 334-41.

de Murcia, G. and Menissier de Murcia, J. (1994). Poly(ADP-ribose) polymerase: a molecular nick-sensor. Trends Biochem Sci 19; 172-6.

de Murcia, G., Schreiber, V., Molinete, M., Saulier, B., Poch, O., Masson, M., Niedergang, C. and Menissier de Murcia, J. (1994). Structure and function of poly(ADP-ribose) polymerase. Molecular and cellular biochemistry 138; 15-24.

de Waard, P. W., Peijnenburg, A. A., Baykus, H., Aarts, J. M., Hoogenboom, R. L., van Schooten, F. J. and de Kok, T. M. (2008). A human intervention study with foods containing natural Ah-receptor agonists does not significantly show AhR-mediated effects as measured in blood cells and urine. Chem Biol Interact 176; 19-29.

Delescluse, C., Lemaire, G., de Sousa, G. and Rahmani, R. (2000). Is CYP1A1 induction always related to AHR signaling pathway? Toxicology 153; 73-82.

Denison, M. S., Fisher, J. M. and Whitlock, J. P., Jr. (1988). The DNA recognition site for the dioxin-Ah receptor complex. Nucleotide sequence and functional analysis. J Biol Chem 263; 17221-4.

Page 189: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

174

Denison, M. S. and Heath-Pagliuso, S. (1998). The Ah receptor: a regulator of the biochemical and toxicological actions of structurally diverse chemicals. Bull Environ Contam Toxicol 61; 557-68.

Denison, M. S. and Nagy, S. R. (2003). Activation of the aryl hydrocarbon receptor by structurally diverse exogenous and endogenous chemicals. Annu Rev Pharmacol Toxicol 43; 309-34.

Denison, M. S., Soshilov, A. A., He, G., DeGroot, D. E. and Zhao, B. (2011). Exactly the same but different: promiscuity and diversity in the molecular mechanisms of action of the aryl hydrocarbon (dioxin) receptor. Toxicological sciences : an official journal of the Society of Toxicology 124; 1-22.

Denu, J. M. (2005). The Sir 2 family of protein deacetylases. Curr Opin Chem Biol 9; 431-40.

Dere, E., Boverhof, D. R., Burgoon, L. D. and Zacharewski, T. R. (2006). In vivo-in vitro toxicogenomic comparison of TCDD-elicited gene expression in Hepa1c1c7 mouse hepatoma cells and C57BL/6 hepatic tissue. BMC genomics 7; 80.

Dere, E., Lo, R., Celius, T., Matthews, J. and Zacharewski, T. R. (2011). Integration of genome-wide computation DRE search, AhR ChIP-chip and gene expression analyses of TCDD-elicited responses in the mouse liver. BMC genomics 12; 365.

Desmarais, Y., Menard, L., Lagueux, J. and Poirier, G. G. (1991). Enzymological properties of poly(ADP-ribose)polymerase: characterization of automodification sites and NADase activity. Biochim Biophys Acta 1078; 179-86.

Desnoyers, S., Shah, G. M., Brochu, G., Hoflack, J. C., Verreault, A. and Poirier, G. G. (1995). Biochemical properties and function of poly(ADP-ribose) glycohydrolase. Biochimie 77; 433-8.

DeVito, M. J., Ma, X., Babish, J. G., Menache, M. and Birnbaum, L. S. (1994). Dose-response relationships in mice following subchronic exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin: CYP1A1, CYP1A2, estrogen receptor, and protein tyrosine phosphorylation. Toxicol Appl Pharmacol 124; 82-90.

Di Girolamo, M., Dani, N., Stilla, A. and Corda, D. (2005). Physiological relevance of the endogenous mono(ADP-ribosyl)ation of cellular proteins. Febs J 272; 4565-75.

Di Paola, S., Micaroni, M., Di Tullio, G., Buccione, R. and Di Girolamo, M. (2012). PARP16/ARTD15 is a novel endoplasmic-reticulum-associated mono-ADP-ribosyltransferase that interacts with, and modifies karyopherin-ss1. PloS one 7; e37352.

Diani-Moore, S., Ram, P., Li, X., Mondal, P., Youn, D. Y., Sauve, A. A. and Rifkind, A. B. (2010). Identification of the aryl hydrocarbon receptor target gene TiPARP as a mediator of suppression of hepatic gluconeogenesis by 2,3,7,8-tetrachlorodibenzo-p-dioxin and of nicotinamide as a corrective agent for this effect. J Biol Chem 285; 38801-10.

Diani-Moore, S., Zhang, S., Ram, P. and Rifkind, A. B. (2013). Aryl hydrocarbon receptor activation by dioxin targets phosphoenolpyruvate carboxykinase (PEPCK) for ADP-ribosylation via TCDD-inducible poly(ADP-ribose) polymerase (TiPARP). J Biol Chem

Dolwick, K. M., Swanson, H. I. and Bradfield, C. A. (1993). In vitro analysis of Ah receptor domains involved in ligand-activated DNA recognition. Proc Natl Acad Sci U S A 90; 8566-70.

Page 190: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

175

Doly, J. and Mandel, P. (1967). [Demonstration of the biosynthesis in vivo of a compound polymer, polyadenosine diphosphoribose in the nucleus of the liver of chickens]. Comptes rendus hebdomadaires des seances de l'Academie des sciences. Serie D: Sciences naturelles 264; 2687-90.

Dong, L., Ma, Q. and Whitlock, J. P., Jr. (1997). Down-regulation of major histocompatibility complex Q1b gene expression by 2,3,7,8-tetrachlorodibenzo-p-dioxin. J Biol Chem 272; 29614-9.

Durkacz, B. W., Omidiji, O., Gray, D. A. and Shall, S. (1980). (ADP-ribose)n participates in DNA excision repair. Nature 283; 593-6.

Dynek, J. N. and Smith, S. (2004). Resolution of sister telomere association is required for progression through mitosis. Science 304; 97-100.

El-Khamisy, S. F., Masutani, M., Suzuki, H. and Caldecott, K. W. (2003). A requirement for PARP-1 for the assembly or stability of XRCC1 nuclear foci at sites of oxidative DNA damage. Nucleic Acids Res 31; 5526-33.

Elliott, G. and Rechsteiner, M. (1975). Pyridine nucleotide metabolism in mitotic cells. Journal of cellular physiology 86 Suppl 2; 641-51.

Erdelyi, K., Bai, P., Kovacs, I., Szabo, E., Mocsar, G., Kakuk, A., Szabo, C., Gergely, P. and Virag, L. (2009). Dual role of poly(ADP-ribose) glycohydrolase in the regulation of cell death in oxidatively stressed A549 cells. FASEB journal : official publication of the Federation of American Societies for Experimental Biology 23; 3553-63.

Evans, B. R., Karchner, S. I., Allan, L. L., Pollenz, R. S., Tanguay, R. L., Jenny, M. J., Sherr, D. H. and Hahn, M. E. (2008). Repression of aryl hydrocarbon receptor (AHR) signaling by AHR repressor: role of DNA binding and competition for AHR nuclear translocator. Mol Pharmacol 73; 387-98.

Evans, B. R., Karchner, S. I., Franks, D. G. and Hahn, M. E. (2005). Duplicate aryl hydrocarbon receptor repressor genes (ahrr1 and ahrr2) in the zebrafish Danio rerio: structure, function, evolution, and AHR-dependent regulation in vivo. Arch Biochem Biophys 441; 151-67.

Fabunmi, R. P., Wigley, W. C., Thomas, P. J. and DeMartino, G. N. (2001). Interferon gamma regulates accumulation of the proteasome activator PA28 and immunoproteasomes at nuclear PML bodies. J Cell Sci 114; 29-36.

Fan, F. and Rozman, K. K. (1994). Relationship between acute toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) and disturbance of intermediary metabolism in the Long-Evans rat. Archives of toxicology 69; 73-8.

Feijs, K. L., Verheugd, P. and Luscher, B. (2013). Expanding functions of intracellular, resident mono-ADP-ribosylation in cell physiology. Febs J

Fernandez-Salguero, P., Pineau, T., Hilbert, D. M., McPhail, T., Lee, S. S., Kimura, S., Nebert, D. W., Rudikoff, S., Ward, J. M. and Gonzalez, F. J. (1995). Immune system impairment and hepatic fibrosis in mice lacking the dioxin-binding Ah receptor. Science 268; 722-6.

Fernandez-Salguero, P. M., Ward, J. M., Sundberg, J. P. and Gonzalez, F. J. (1997). Lesions of aryl-hydrocarbon receptor-deficient mice. Vet Pathol 34; 605-14.

Page 191: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

176

Fisher, A. E., Hochegger, H., Takeda, S. and Caldecott, K. W. (2007). Poly(ADP-ribose) polymerase 1 accelerates single-strand break repair in concert with poly(ADP-ribose) glycohydrolase. Mol Cell Biol 27; 5597-605.

Fletcher, N., Wahlstrom, D., Lundberg, R., Nilsson, C. B., Nilsson, K. C., Stockling, K., Hellmold, H. and Hakansson, H. (2005). 2,3,7,8-Tetrachlorodibenzo-p-dioxin (TCDD) alters the mRNA expression of critical genes associated with cholesterol metabolism, bile acid biosynthesis, and bile transport in rat liver: a microarray study. Toxicol Appl Pharmacol 207; 1-24.

Forgacs, A. L., Dere, E., Angrish, M. M. and Zacharewski, T. R. (2013). Comparative analysis of temporal and dose-dependent TCDD-elicited gene expression in human, mouse, and rat primary hepatocytes. Toxicological sciences : an official journal of the Society of Toxicology 133; 54-66.

Forst, A. H., Karlberg, T., Herzog, N., Thorsell, A. G., Gross, A., Feijs, K. L., Verheugd, P., Kursula, P., Nijmeijer, B., Kremmer, E., Kleine, H., Ladurner, A. G., Schuler, H. and Luscher, B. (2013). Recognition of mono-ADP-ribosylated ARTD10 substrates by ARTD8 macrodomains. Structure 21; 462-75.

Frericks, M., Temchura, V. V., Majora, M., Stutte, S. and Esser, C. (2006). Transcriptional signatures of immune cells in aryl hydrocarbon receptor (AHR)-proficient and AHR-deficient mice. Biol Chem 387; 1219-26.

Fritsche, E., Schafer, C., Calles, C., Bernsmann, T., Bernshausen, T., Wurm, M., Hubenthal, U., Cline, J. E., Hajimiragha, H., Schroeder, P., Klotz, L. O., Rannug, A., Furst, P., Hanenberg, H., Abel, J. and Krutmann, J. (2007). Lightening up the UV response by identification of the arylhydrocarbon receptor as a cytoplasmatic target for ultraviolet B radiation. Proc Natl Acad Sci U S A 104; 8851-6.

Frueh, F. W., Hayashibara, K. C., Brown, P. O. and Whitlock, J. P., Jr. (2001). Use of cDNA microarrays to analyze dioxin-induced changes in human liver gene expression. Toxicol Lett 122; 189-203.

Frye, R. A. (1999). Characterization of five human cDNAs with homology to the yeast SIR2 gene: Sir2-like proteins (sirtuins) metabolize NAD and may have protein ADP-ribosyltransferase activity. Biochem Biophys Res Commun 260; 273-9.

Gagne, J. P., Isabelle, M., Lo, K. S., Bourassa, S., Hendzel, M. J., Dawson, V. L., Dawson, T. M. and Poirier, G. G. (2008). Proteome-wide identification of poly(ADP-ribose) binding proteins and poly(ADP-ribose)-associated protein complexes. Nucleic Acids Res 36; 6959-76.

Gamble, M. J. and Fisher, R. P. (2007). SET and PARP1 remove DEK from chromatin to permit access by the transcription machinery. Nat Struct Mol Biol 14; 548-55.

Gambone, C. J., Hutcheson, J. M., Gabriel, J. L., Beard, R. L., Chandraratna, R. A., Soprano, K. J. and Soprano, D. R. (2002). Unique property of some synthetic retinoids: activation of the aryl hydrocarbon receptor pathway. Mol Pharmacol 61; 334-42.

Gao, G., Guo, X. and Goff, S. P. (2002). Inhibition of retroviral RNA production by ZAP, a CCCH-type zinc finger protein. Science 297; 1703-6.

Page 192: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

177

Garcia-Salcedo, J. A., Gijon, P., Nolan, D. P., Tebabi, P. and Pays, E. (2003). A chromosomal SIR2 homologue with both histone NAD-dependent ADP-ribosyltransferase and deacetylase activities is involved in DNA repair in Trypanosoma brucei. Embo J 22; 5851-62.

Gasiewicz, T. A., Holscher, M. A. and Neal, R. A. (1980). The effect of total parenteral nutrition on the toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin in the rat. Toxicol Appl Pharmacol 54; 469-88.

Gibson, B. A. and Kraus, W. L. (2012). New insights into the molecular and cellular functions of poly(ADP-ribose) and PARPs. Nature reviews. Molecular cell biology 13; 411-24.

Gill, D. M., Pappenheimer, A. M., Jr., Brown, R. and Kurnick, J. T. (1969). Studies on the mode of action of diphtheria toxin. VII. Toxin-stimulated hydrolysis of nicotinamide adenine dinucleotide in mammalian cell extracts. The Journal of experimental medicine 129; 1-21.

Gillner, M., Bergman, J., Cambillau, C., Alexandersson, M., Fernstrom, B. and Gustafsson, J. A. (1993). Interactions of indolo[3,2-b]carbazoles and related polycyclic aromatic hydrocarbons with specific binding sites for 2,3,7,8-tetrachlorodibenzo-p-dioxin in rat liver. Mol Pharmacol 44; 336-45.

Glowacki, G., Braren, R., Firner, K., Nissen, M., Kuhl, M., Reche, P., Bazan, F., Cetkovic-Cvrlje, M., Leiter, E., Haag, F. and Koch-Nolte, F. (2002). The family of toxin-related ecto-ADP-ribosyltransferases in humans and the mouse. Protein science : a publication of the Protein Society 11; 1657-70.

Goenka, S. and Boothby, M. (2006). Selective potentiation of Stat-dependent gene expression by collaborator of Stat6 (CoaSt6), a transcriptional cofactor. Proc Natl Acad Sci U S A 103; 4210-5.

Goenka, S., Cho, S. H. and Boothby, M. (2007). Collaborator of Stat6 (CoaSt6)-associated poly(ADP-ribose) polymerase activity modulates Stat6-dependent gene transcription. J Biol Chem 282; 18732-9.

Gonzalez, F. J. and Fernandez-Salguero, P. (1998). The aryl hydrocarbon receptor: studies using the AHR-null mice. Drug Metab Dispos 26; 1194-8.

Goodwin, P. M., Lewis, P. J., Davies, M. I., Skidmore, C. J. and Shall, S. (1978). The effect of gamma radiation and neocarzinostatin on NAD and ATP levels in mouse leukaemia cells. Biochim Biophys Acta 543; 576-82.

Goryo, K., Suzuki, A., Del Carpio, C. A., Siizaki, K., Kuriyama, E., Mikami, Y., Kinoshita, K., Yasumoto, K., Rannug, A., Miyamoto, A., Fujii-Kuriyama, Y. and Sogawa, K. (2007). Identification of amino acid residues in the Ah receptor involved in ligand binding. Biochem Biophys Res Commun 354; 396-402.

Gradin, K., Toftgard, R., Poellinger, L. and Berghard, A. (1999). Repression of dioxin signal transduction in fibroblasts. Identification Of a putative repressor associated with Arnt. J Biol Chem 274; 13511-8.

Grahnert, A., Friedrich, M., Pfister, M., Haag, F., Koch-Nolte, F. and Hauschildt, S. (2002). Mono-ADP-ribosyltransferases in human monocytes: regulation by lipopolysaccharide. Biochem J 362; 717-23.

Page 193: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

178

Greig, J. B., Jones, G., Butler, W. H. and Barnes, J. M. (1973). Toxic effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin. Food and cosmetics toxicology 11; 585-95.

Grose, K. R. and Bjeldanes, L. F. (1992). Oligomerization of indole-3-carbinol in aqueous acid. Chem Res Toxicol 5; 188-93.

Grubisha, O., Smith, B. C. and Denu, J. M. (2005). Small molecule regulation of Sir2 protein deacetylases. Febs J 272; 4607-16.

Gu, Y. Z., Hogenesch, J. B. and Bradfield, C. A. (2000). The PAS superfamily: sensors of environmental and developmental signals. Annu Rev Pharmacol Toxicol 40; 519-61.

Guo, H. L., Zhang, C., Liu, Q., Li, Q., Lian, G., Wu, D., Li, X., Zhang, W., Shen, Y., Ye, Z., Lin, S. Y. and Lin, S. C. (2012). The Axin/TNKS complex interacts with KIF3A and is required for insulin-stimulated GLUT4 translocation. Cell research 22; 1246-57.

Guo, X., Carroll, J. W., Macdonald, M. R., Goff, S. P. and Gao, G. (2004). The zinc finger antiviral protein directly binds to specific viral mRNAs through the CCCH zinc finger motifs. Journal of virology 78; 12781-7.

Guo, X., Ma, J., Sun, J. and Gao, G. (2007). The zinc-finger antiviral protein recruits the RNA processing exosome to degrade the target mRNA. Proc Natl Acad Sci U S A 104; 151-6.

Haag, F. and Koch-Nolte, F. (1998). Endogenous relatives of ADP-ribosylating bacterial toxins in mice and men: potential regulators of immune cell function. Journal of biological regulators and homeostatic agents 12; 53-62.

Haarmann-Stemmann, T., Bothe, H., Kohli, A., Sydlik, U., Abel, J. and Fritsche, E. (2007). Analysis of the transcriptional regulation and molecular function of the aryl hydrocarbon receptor repressor in human cell lines. Drug Metab Dispos 35; 2262-9.

Hahn, M. E. (2002). Aryl hydrocarbon receptors: diversity and evolution. Chem Biol Interact 141; 131-60.

Hahn, M. E., Allan, L. L. and Sherr, D. H. (2009). Regulation of constitutive and inducible AHR signaling: complex interactions involving the AHR repressor. Biochem Pharmacol 77; 485-97.

Haigis, M. C., Mostoslavsky, R., Haigis, K. M., Fahie, K., Christodoulou, D. C., Murphy, A. J., Valenzuela, D. M., Yancopoulos, G. D., Karow, M., Blander, G., Wolberger, C., Prolla, T. A., Weindruch, R., Alt, F. W. and Guarente, L. (2006). SIRT4 inhibits glutamate dehydrogenase and opposes the effects of calorie restriction in pancreatic beta cells. Cell 126; 941-54.

Hakme, A., Wong, H. K., Dantzer, F. and Schreiber, V. (2008). The expanding field of poly(ADP-ribosyl)ation reactions. 'Protein Modifications: Beyond the Usual Suspects' Review Series. EMBO Rep 9; 1094-100.

Han, S. and Tainer, J. A. (2002). The ARTT motif and a unified structural understanding of substrate recognition in ADP-ribosylating bacterial toxins and eukaryotic ADP-ribosyltransferases. Int J Med Microbiol 291; 523-9.

Han, W., Li, X. and Fu, X. (2011). The macro domain protein family: structure, functions, and their potential therapeutic implications. Mutat Res 727; 86-103.

Page 194: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

179

Hankinson, O. (1995). The aryl hydrocarbon receptor complex. Annu Rev Pharmacol Toxicol 35; 307-40.

Hankinson, O. (2005). Role of coactivators in transcriptional activation by the aryl hydrocarbon receptor. Arch Biochem Biophys 433; 379-86.

Hao, N., Lee, K. L., Furness, S. G., Bosdotter, C., Poellinger, L. and Whitelaw, M. L. (2012). Xenobiotics and loss of cell adhesion drive distinct transcriptional outcomes by aryl hydrocarbon receptor signaling. Mol Pharmacol 82; 1082-93.

Harper, P. A., Riddick, D. S. and Okey, A. B. (2006). Regulating the regulator: factors that control levels and activity of the aryl hydrocarbon receptor. Biochem Pharmacol 72; 267-79.

Hassa, P. O., Haenni, S. S., Elser, M. and Hottiger, M. O. (2006). Nuclear ADP-ribosylation reactions in mammalian cells: where are we today and where are we going? Microbiol Mol Biol Rev 70; 789-829.

Hassa, P. O. and Hottiger, M. O. (2002). The functional role of poly(ADP-ribose)polymerase 1 as novel coactivator of NF-kappaB in inflammatory disorders. Cellular and molecular life sciences : CMLS 59; 1534-53.

Hassoun, E. A., Wang, H., Abushaban, A. and Stohs, S. J. (2002). Induction of oxidative stress in the tissues of rats after chronic exposure to TCDD, 2,3,4,7,8-pentachlorodibenzofuran, and 3,3',4,4',5-pentachlorobiphenyl. Journal of toxicology and environmental health. Part A 65; 825-42.

Hassoun, Z., Deschenes, M., Lafortune, M., Dufresne, M. P., Perreault, P., Lepanto, L., Gianfelice, D., Bui, B. and Pomier-Layrargues, G. (2001). Relationship between pre-TIPS liver perfusion by the portal vein and the incidence of post-TIPS chronic hepatic encephalopathy. The American journal of gastroenterology 96; 1205-9.

Hayes, K. R., Zastrow, G. M., Nukaya, M., Pande, K., Glover, E., Maufort, J. P., Liss, A. L., Liu, Y., Moran, S. M., Vollrath, A. L. and Bradfield, C. A. (2007). Hepatic transcriptional networks induced by exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin. Chem Res Toxicol 20; 1573-81.

He, F., Tsuda, K., Takahashi, M., Kuwasako, K., Terada, T., Shirouzu, M., Watanabe, S., Kigawa, T., Kobayashi, N., Guntert, P., Yokoyama, S. and Muto, Y. (2012). Structural insight into the interaction of ADP-ribose with the PARP WWE domains. FEBS Lett 586; 3858-64.

He, J., Hu, B., Shi, X., Weidert, E. R., Lu, P., Xu, M., Huang, M., Kelley, E. E. and Xie, W. (2013). Activation of the aryl hydrocarbon receptor sensitizes mice to nonalcoholic steatohepatitis by deactivating mitochondrial sirtuin deacetylase sirt3. Mol Cell Biol 33; 2047-55.

Heeres, J. T. and Hergenrother, P. J. (2007). Poly(ADP-ribose) makes a date with death. Curr Opin Chem Biol 11; 644-53.

Hegde, M. L., Hazra, T. K. and Mitra, S. (2008). Early steps in the DNA base excision/single-strand interruption repair pathway in mammalian cells. Cell research 18; 27-47.

Page 195: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

180

Heid, S. E., Pollenz, R. S. and Swanson, H. I. (2000). Role of heat shock protein 90 dissociation in mediating agonist-induced activation of the aryl hydrocarbon receptor. Mol Pharmacol 57; 82-92.

Helferich, W. G. and Denison, M. S. (1991). Ultraviolet photoproducts of tryptophan can act as dioxin agonists. Mol Pharmacol 40; 674-8.

Henry, E. C. and Gasiewicz, T. A. (2003). Agonist but not antagonist ligands induce conformational change in the mouse aryl hydrocarbon receptor as detected by partial proteolysis. Mol Pharmacol 63; 392-400.

Henry, E. C., Welle, S. L. and Gasiewicz, T. A. (2010). TCDD and a putative endogenous AhR ligand, ITE, elicit the same immediate changes in gene expression in mouse lung fibroblasts. Toxicological sciences : an official journal of the Society of Toxicology 114; 90-100.

Hetz, C. (2012). The unfolded protein response: controlling cell fate decisions under ER stress and beyond. Nature reviews. Molecular cell biology 13; 89-102.

Holbourn, K. P., Shone, C. C. and Acharya, K. R. (2006). A family of killer toxins. Exploring the mechanism of ADP-ribosylating toxins. Febs J 273; 4579-93.

Honjo, T., Nishizuka, Y. and Hayaishi, O. (1968). Diphtheria toxin-dependent adenosine diphosphate ribosylation of aminoacyl transferase II and inhibition of protein synthesis. J Biol Chem 243; 3553-5.

Hosoya, T., Harada, N., Mimura, J., Motohashi, H., Takahashi, S., Nakajima, O., Morita, M., Kawauchi, S., Yamamoto, M. and Fujii-Kuriyama, Y. (2008). Inducibility of cytochrome P450 1A1 and chemical carcinogenesis by benzo[a]pyrene in AhR repressor-deficient mice. Biochem Biophys Res Commun 365; 562-7.

Hottiger, M. O., Hassa, P. O., Luscher, B., Schuler, H. and Koch-Nolte, F. (2010). Toward a unified nomenclature for mammalian ADP-ribosyltransferases. Trends Biochem Sci 35; 208-19.

Hsia, M. T. and Kreamer, B. L. (1985). Delayed wasting syndrome and alterations of liver gluconeogenic enzymes in rats exposed to the TCDD congener 3,3', 4,4'-tetrachloroazoxybenzene. Toxicol Lett 25; 247-58.

Hsiao, S. J. and Smith, S. (2008). Tankyrase function at telomeres, spindle poles, and beyond. Biochimie 90; 83-92.

Huang, K., Tidyman, W. E., Le, K. U., Kirsten, E., Kun, E. and Ordahl, C. P. (2004). Analysis of nucleotide sequence-dependent DNA binding of poly(ADP-ribose) polymerase in a purified system. Biochemistry 43; 217-23.

Huang, S. M., Mishina, Y. M., Liu, S., Cheung, A., Stegmeier, F., Michaud, G. A., Charlat, O., Wiellette, E., Zhang, Y., Wiessner, S., Hild, M., Shi, X., Wilson, C. J., Mickanin, C., Myer, V., Fazal, A., Tomlinson, R., Serluca, F., Shao, W., Cheng, H., Shultz, M., Rau, C., Schirle, M., Schlegl, J., Ghidelli, S., Fawell, S., Lu, C., Curtis, D., Kirschner, M. W., Lengauer, C., Finan, P. M., Tallarico, J. A., Bouwmeester, T., Porter, J. A., Bauer, A. and Cong, F. (2009). Tankyrase inhibition stabilizes axin and antagonizes Wnt signalling. Nature 461; 614-20.

Page 196: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

181

Huletsky, A., de Murcia, G., Muller, S., Hengartner, M., Menard, L., Lamarre, D. and Poirier, G. G. (1989). The effect of poly(ADP-ribosyl)ation on native and H1-depleted chromatin. A role of poly(ADP-ribosyl)ation on core nucleosome structure. J Biol Chem 264; 8878-86.

Ikejima, M., Noguchi, S., Yamashita, R., Ogura, T., Sugimura, T., Gill, D. M. and Miwa, M. (1990). The zinc fingers of human poly(ADP-ribose) polymerase are differentially required for the recognition of DNA breaks and nicks and the consequent enzyme activation. Other structures recognize intact DNA. J Biol Chem 265; 21907-13.

Ikuta, T., Eguchi, H., Tachibana, T., Yoneda, Y. and Kawajiri, K. (1998). Nuclear localization and export signals of the human aryl hydrocarbon receptor. J Biol Chem 273; 2895-904.

Ikuta, T., Kobayashi, Y. and Kawajiri, K. (2004). Phosphorylation of nuclear localization signal inhibits the ligand-dependent nuclear import of aryl hydrocarbon receptor. Biochem Biophys Res Commun 317; 545-50.

Imai, S. and Guarente, L. (2010). Ten years of NAD-dependent SIR2 family deacetylases: implications for metabolic diseases. Trends in pharmacological sciences 31; 212-20.

Israel, D. I., Estolano, M. G., Galeazzi, D. R. and Whitlock, J. P., Jr. (1985). Superinduction of cytochrome P1-450 gene transcription by inhibition of protein synthesis in wild type and variant mouse hepatoma cells. J Biol Chem 260; 5648-53.

Israel, D. I. and Whitlock, J. P., Jr. (1983). Induction of mRNA specific for cytochrome P1-450 in wild type and variant mouse hepatoma cells. J Biol Chem 258; 10390-4.

Ito, T., Nagai, H., Lin, T. M., Peterson, R. E., Tohyama, C., Kobayashi, T. and Nohara, K. (2006). Organic Chemicals Adsorbed onto Diesel Exhaust Particles Directly Alter the Differentiation of Fetal Thymocytes Through Arylhydrocarbon Receptor but Not Oxidative Stress Responses. J Immunotoxicol 3; 21-30.

Jain, S., Dolwick, K. M., Schmidt, J. V. and Bradfield, C. A. (1994). Potent transactivation domains of the Ah receptor and the Ah receptor nuclear translocator map to their carboxyl termini. J Biol Chem 269; 31518-24.

Jankevicius, G., Hassler, M., Golia, B., Rybin, V., Zacharias, M., Timinszky, G. and Ladurner, A. G. (2013). A family of macrodomain proteins reverses cellular mono-ADP-ribosylation. Nat Struct Mol Biol 20; 508-14.

Jepsen, K. and Rosenfeld, M. G. (2002). Biological roles and mechanistic actions of co-repressor complexes. J Cell Sci 115; 689-98.

Jeuken, A., Keser, B. J., Khan, E., Brouwer, A., Koeman, J. and Denison, M. S. (2003). Activation of the Ah receptor by extracts of dietary herbal supplements, vegetables, and fruits. J Agric Food Chem 51; 5478-87.

Johansson, M. (1999). A human poly(ADP-ribose) polymerase gene family (ADPRTL): cDNA cloning of two novel poly(ADP-ribose) polymerase homologues. Genomics 57; 442-5.

Jonsson, G. G., Menard, L., Jacobson, E. L., Poirier, G. G. and Jacobson, M. K. (1988). Effect of hyperthermia on poly(adenosine diphosphate-ribose) glycohydrolase. Cancer Res 48; 4240-3.

Ju, B. G., Solum, D., Song, E. J., Lee, K. J., Rose, D. W., Glass, C. K. and Rosenfeld, M. G. (2004). Activating the PARP-1 sensor component of the groucho/ TLE1 corepressor

Page 197: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

182

complex mediates a CaMKinase IIdelta-dependent neurogenic gene activation pathway. Cell 119; 815-29.

Juarez-Salinas, H., Levi, V., Jacobson, E. L. and Jacobson, M. K. (1982). Poly(ADP-ribose) has a branched structure in vivo. J Biol Chem 257; 607-9.

Juarez-Salinas, H., Mendoza-Alvarez, H., Levi, V., Jacobson, M. K. and Jacobson, E. L. (1983). Simultaneous determination of linear and branched residues in poly(ADP-ribose). Analytical biochemistry 131; 410-8.

Juszczynski, P., Kutok, J. L., Li, C., Mitra, J., Aguiar, R. C. and Shipp, M. A. (2006). BAL1 and BBAP are regulated by a gamma interferon-responsive bidirectional promoter and are overexpressed in diffuse large B-cell lymphomas with a prominent inflammatory infiltrate. Mol Cell Biol 26; 5348-59.

Jwa, M. and Chang, P. (2012). PARP16 is a tail-anchored endoplasmic reticulum protein required for the PERK- and IRE1alpha-mediated unfolded protein response. Nature cell biology 14; 1223-30.

Kafafi, S. A., Afeefy, H. Y., Said, H. K. and Kafafi, A. G. (1993). Relationship between aryl hydrocarbon receptor binding, induction of aryl hydrocarbon hydroxylase and 7-ethoxyresorufin O-deethylase enzymes, and toxic activities of aromatic xenobiotics in animals. A new model. Chem Res Toxicol 6; 328-34.

Kaminker, P. G., Kim, S. H., Taylor, R. D., Zebarjadian, Y., Funk, W. D., Morin, G. B., Yaswen, P. and Campisi, J. (2001). TANK2, a new TRF1-associated poly(ADP-ribose) polymerase, causes rapid induction of cell death upon overexpression. J Biol Chem 276; 35891-9.

Kanai, M., Hanashiro, K., Kim, S. H., Hanai, S., Boulares, A. H., Miwa, M. and Fukasawa, K. (2007). Inhibition of Crm1-p53 interaction and nuclear export of p53 by poly(ADP-ribosyl)ation. Nature cell biology 9; 1175-83.

Kanai, M., Miwa, M., Kuchino, Y. and Sugimura, T. (1982). Presence of branched portion in poly(adenosine diphosphate ribose) in vivo. J Biol Chem 257; 6217-23.

Kang, H. C., Lee, Y. I., Shin, J. H., Andrabi, S. A., Chi, Z., Gagne, J. P., Lee, Y., Ko, H. S., Lee, B. D., Poirier, G. G., Dawson, V. L. and Dawson, T. M. (2011). Iduna is a poly(ADP-ribose) (PAR)-dependent E3 ubiquitin ligase that regulates DNA damage. Proc Natl Acad Sci U S A 108; 14103-8.

Kang, X., Falick, A. M., Nelson, R. E., Gao, G., Rogers, K., Aphasizhev, R. and Simpson, L. (2004). Disruption of the zinc finger motifs in the Leishmania tarentolae LC-4 (=TbMP63) L-complex editing protein affects the stability of the L-complex. J Biol Chem 279; 3893-9.

Kannan, P., Yu, Y., Wankhade, S. and Tainsky, M. A. (1999). PolyADP-ribose polymerase is a coactivator for AP-2-mediated transcriptional activation. Nucleic Acids Res 27; 866-74.

Kanno, Y., Miyama, Y., Takane, Y., Nakahama, T. and Inouye, Y. (2007). Identification of intracellular localization signals and of mechanisms underlining the nucleocytoplasmic shuttling of human aryl hydrocarbon receptor repressor. Biochem Biophys Res Commun 364; 1026-31.

Page 198: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

183

Kappes, F., Fahrer, J., Khodadoust, M. S., Tabbert, A., Strasser, C., Mor-Vaknin, N., Moreno-Villanueva, M., Burkle, A., Markovitz, D. M. and Ferrando-May, E. (2008). DEK is a poly(ADP-ribose) acceptor in apoptosis and mediates resistance to genotoxic stress. Mol Cell Biol 28; 3245-57.

Karchner, S. I., Franks, D. G., Powell, W. H. and Hahn, M. E. (2002). Regulatory interactions among three members of the vertebrate aryl hydrocarbon receptor family: AHR repressor, AHR1, and AHR2. J Biol Chem 277; 6949-59.

Karchner, S. I., Jenny, M. J., Tarrant, A. M., Evans, B. R., Kang, H. J., Bae, I., Sherr, D. H. and Hahn, M. E. (2009). The active form of human aryl hydrocarbon receptor (AHR) repressor lacks exon 8, and its Pro 185 and Ala 185 variants repress both AHR and hypoxia-inducible factor. Mol Cell Biol 29; 3465-77.

Karlberg, T., Thorsell, A. G., Kallas, A. and Schuler, H. (2012). Crystal structure of human ADP-ribose transferase ARTD15/PARP16 reveals a novel putative regulatory domain. J Biol Chem 287; 24077-81.

Karras, G. I., Kustatscher, G., Buhecha, H. R., Allen, M. D., Pugieux, C., Sait, F., Bycroft, M. and Ladurner, A. G. (2005). The macro domain is an ADP-ribose binding module. Embo J 24; 1911-20.

Katoh, M. and Katoh, M. (2003). Identification and characterization of human TIPARP gene within the CCNL amplicon at human chromosome 3q25.31. Int J Oncol 23; 541-7.

Kawaichi, M., Ueda, K. and Hayaishi, O. (1981). Multiple autopoly(ADP-ribosyl)ation of rat liver poly(ADP-ribose) synthetase. Mode of modification and properties of automodified synthetase. J Biol Chem 256; 9483-9.

Kazlauskas, A., Poellinger, L. and Pongratz, I. (1999). Evidence that the co-chaperone p23 regulates ligand responsiveness of the dioxin (Aryl hydrocarbon) receptor. J Biol Chem 274; 13519-24.

Kazlauskas, A., Poellinger, L. and Pongratz, I. (2000). The immunophilin-like protein XAP2 regulates ubiquitination and subcellular localization of the dioxin receptor. J Biol Chem 275; 41317-24.

Keil, C., Grobe, T. and Oei, S. L. (2006). MNNG-induced cell death is controlled by interactions between PARP-1, poly(ADP-ribose) glycohydrolase, and XRCC1. J Biol Chem 281; 34394-405.

Kelly, S. M., Pabit, S. A., Kitchen, C. M., Guo, P., Marfatia, K. A., Murphy, T. J., Corbett, A. H. and Berland, K. M. (2007). Recognition of polyadenosine RNA by zinc finger proteins. Proc Natl Acad Sci U S A 104; 12306-11.

Kewley, R. J., Whitelaw, M. L. and Chapman-Smith, A. (2004). The mammalian basic helix-loop-helix/PAS family of transcriptional regulators. Int J Biochem Cell Biol 36; 189-204.

Kickhoefer, V. A., Siva, A. C., Kedersha, N. L., Inman, E. M., Ruland, C., Streuli, M. and Rome, L. H. (1999). The 193-kD vault protein, VPARP, is a novel poly(ADP-ribose) polymerase. The Journal of cell biology 146; 917-28.

Kikuchi, Y., Ohsawa, S., Mimura, J., Ema, M., Takasaki, C., Sogawa, K. and Fujii-Kuriyama, Y. (2003). Heterodimers of bHLH-PAS protein fragments derived from AhR, AhRR, and

Page 199: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

184

Arnt prepared by co-expression in Escherichia coli: characterization of their DNA binding activity and preparation of a DNA complex. J Biochem 134; 83-90.

Kim, I. K., Kiefer, J. R., Ho, C. M., Stegeman, R. A., Classen, S., Tainer, J. A. and Ellenberger, T. (2012). Structure of mammalian poly(ADP-ribose) glycohydrolase reveals a flexible tyrosine clasp as a substrate-binding element. Nat Struct Mol Biol 19; 653-6.

Kim, M. Y., Mauro, S., Gevry, N., Lis, J. T. and Kraus, W. L. (2004). NAD+-dependent modulation of chromatin structure and transcription by nucleosome binding properties of PARP-1. Cell 119; 803-14.

Kim, M. Y., Zhang, T. and Kraus, W. L. (2005). Poly(ADP-ribosyl)ation by PARP-1: 'PAR-laying' NAD+ into a nuclear signal. Genes Dev 19; 1951-67.

Kininis, M., Chen, B. S., Diehl, A. G., Isaacs, G. D., Zhang, T., Siepel, A. C., Clark, A. G. and Kraus, W. L. (2007). Genomic analyses of transcription factor binding, histone acetylation, and gene expression reveal mechanistically distinct classes of estrogen-regulated promoters. Mol Cell Biol 27; 5090-104.

Kleine, H., Herrmann, A., Lamark, T., Forst, A. H., Verheugd, P., Luscher-Firzlaff, J., Lippok, B., Feijs, K. L., Herzog, N., Kremmer, E., Johansen, T., Muller-Newen, G. and Luscher, B. (2012). Dynamic subcellular localization of the mono-ADP-ribosyltransferase ARTD10 and interaction with the ubiquitin receptor p62. Cell communication and signaling : CCS 10; 28.

Kleine, H. and Luscher, B. (2009). Learning how to read ADP-ribosylation. Cell 139; 17-9.

Kleine, H., Poreba, E., Lesniewicz, K., Hassa, P. O., Hottiger, M. O., Litchfield, D. W., Shilton, B. H. and Luscher, B. (2008). Substrate-assisted catalysis by PARP10 limits its activity to mono-ADP-ribosylation. Mol Cell 32; 57-69.

Kleman, M. I., Poellinger, L. and Gustafsson, J. A. (1994). Regulation of human dioxin receptor function by indolocarbazoles, receptor ligands of dietary origin. J Biol Chem 269; 5137-44.

Koch-Nolte, F., Kernstock, S., Mueller-Dieckmann, C., Weiss, M. S. and Haag, F. (2008). Mammalian ADP-ribosyltransferases and ADP-ribosylhydrolases. Frontiers in bioscience : a journal and virtual library 13; 6716-29.

Koh, D. W., Lawler, A. M., Poitras, M. F., Sasaki, M., Wattler, S., Nehls, M. C., Stoger, T., Poirier, G. G., Dawson, V. L. and Dawson, T. M. (2004). Failure to degrade poly(ADP-ribose) causes increased sensitivity to cytotoxicity and early embryonic lethality. Proc Natl Acad Sci U S A 101; 17699-704.

Korkalainen, M., Linden, J., Tuomisto, J. and Pohjanvirta, R. (2005). Effect of TCDD on mRNA expression of genes encoding bHLH/PAS proteins in rat hypothalamus. Toxicology 208; 1-11.

Korkalainen, M., Tuomisto, J. and Pohjanvirta, R. (2004). Primary structure and inducibility by 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) of aryl hydrocarbon receptor repressor in a TCDD-sensitive and a TCDD-resistant rat strain. Biochem Biophys Res Commun 315; 123-31.

Kraus, W. L. (2008). Transcriptional control by PARP-1: chromatin modulation, enhancer-binding, coregulation, and insulation. Curr Opin Cell Biol 20; 294-302.

Page 200: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

185

Kraus, W. L. and Lis, J. T. (2003). PARP goes transcription. Cell 113; 677-83.

Krishnakumar, R., Gamble, M. J., Frizzell, K. M., Berrocal, J. G., Kininis, M. and Kraus, W. L. (2008). Reciprocal binding of PARP-1 and histone H1 at promoters specifies transcriptional outcomes. Science 319; 819-21.

Krishnakumar, R. and Kraus, W. L. (2010a). PARP-1 regulates chromatin structure and transcription through a KDM5B-dependent pathway. Mol Cell 39; 736-49.

Krishnakumar, R. and Kraus, W. L. (2010b). The PARP side of the nucleus: molecular actions, physiological outcomes, and clinical targets. Mol Cell 39; 8-24.

Kuimov, A. N., Kuprash, D. V., Petrov, V. N., Vdovichenko, K. K., Scanlan, M. J., Jongeneel, C. V., Lagarkova, M. A. and Nedospasov, S. A. (2001). Cloning and characterization of TNKL, a member of tankyrase gene family. Genes and immunity 2; 52-5.

Kulkarni, P. S., Crespo, J. G. and Afonso, C. A. (2008). Dioxins sources and current remediation technologies--a review. Environment international 34; 139-53.

Kumar, M. B. and Perdew, G. H. (1999). Nuclear receptor coactivator SRC-1 interacts with the Q-rich subdomain of the AhR and modulates its transactivation potential. Gene Expr 8; 273-86.

Kumar, M. B., Ramadoss, P., Reen, R. K., Vanden Heuvel, J. P. and Perdew, G. H. (2001). The Q-rich subdomain of the human Ah receptor transactivation domain is required for dioxin-mediated transcriptional activity. J Biol Chem 276; 42302-10.

Kumar, M. B., Tarpey, R. W. and Perdew, G. H. (1999). Differential recruitment of coactivator RIP140 by Ah and estrogen receptors. Absence of a role for LXXLL motifs. J Biol Chem 274; 22155-64.

Kurachi, M., Hashimoto, S., Obata, A., Nagai, S., Nagahata, T., Inadera, H., Sone, H., Tohyama, C., Kaneko, S., Kobayashi, K. and Matsushima, K. (2002). Identification of 2,3,7,8-tetrachlorodibenzo-p-dioxin-responsive genes in mouse liver by serial analysis of gene expression. Biochem Biophys Res Commun 292; 368-77.

Kurosaki, T., Ushiro, H., Mitsuuchi, Y., Suzuki, S., Matsuda, M., Matsuda, Y., Katunuma, N., Kangawa, K., Matsuo, H., Hirose, T. and et al. (1987). Primary structure of human poly(ADP-ribose) synthetase as deduced from cDNA sequence. J Biol Chem 262; 15990-7.

Kuwahara, J., Watanabe, Y., Kayasuga, T. and Itoh, K. (2000). Zn finger and nuclear localization of transcription factor Sp1. Nucleic acids symposium series 265-6.

Lahvis, G. P., Lindell, S. L., Thomas, R. S., McCuskey, R. S., Murphy, C., Glover, E., Bentz, M., Southard, J. and Bradfield, C. A. (2000). Portosystemic shunting and persistent fetal vascular structures in aryl hydrocarbon receptor-deficient mice. Proc Natl Acad Sci U S A 97; 10442-7.

Laing, S., Unger, M., Koch-Nolte, F. and Haag, F. (2011). ADP-ribosylation of arginine. Amino Acids 41; 257-69.

Langelier, M. F., Planck, J. L., Roy, S. and Pascal, J. M. (2012). Structural basis for DNA damage-dependent poly(ADP-ribosyl)ation by human PARP-1. Science 336; 728-32.

Page 201: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

186

LaPres, J. J., Glover, E., Dunham, E. E., Bunger, M. K. and Bradfield, C. A. (2000). ARA9 modifies agonist signaling through an increase in cytosolic aryl hydrocarbon receptor. J Biol Chem 275; 6153-9.

Lee, J. H., Wada, T., Febbraio, M., He, J., Matsubara, T., Lee, M. J., Gonzalez, F. J. and Xie, W. (2010). A novel role for the dioxin receptor in fatty acid metabolism and hepatic steatosis. Gastroenterology 139; 653-63.

Leppard, J. B., Dong, Z., Mackey, Z. B. and Tomkinson, A. E. (2003). Physical and functional interaction between DNA ligase IIIalpha and poly(ADP-Ribose) polymerase 1 in DNA single-strand break repair. Mol Cell Biol 23; 5919-27.

Leung, A. K., Vyas, S., Rood, J. E., Bhutkar, A., Sharp, P. A. and Chang, P. (2011). Poly(ADP-ribose) regulates stress responses and microRNA activity in the cytoplasm. Mol Cell 42; 489-99.

Levy-Wilson, B. (1981). ADP-ribosylation of trout testis chromosomal proteins: distribution of ADP-ribosylated proteins among DNase I-sensitive and -resistant chromatin domains. Arch Biochem Biophys 208; 528-34.

Liang, J., Wang, J., Azfer, A., Song, W., Tromp, G., Kolattukudy, P. E. and Fu, M. (2008). A novel CCCH-zinc finger protein family regulates proinflammatory activation of macrophages. J Biol Chem 283; 6337-46.

Liaudet, L., Pacher, P., Mabley, J. G., Virag, L., Soriano, F. G., Hasko, G. and Szabo, C. (2002). Activation of poly(ADP-Ribose) polymerase-1 is a central mechanism of lipopolysaccharide-induced acute lung inflammation. American journal of respiratory and critical care medicine 165; 372-7.

Lin, T. M., Ko, K., Moore, R. W., Buchanan, D. L., Cooke, P. S. and Peterson, R. E. (2001). Role of the aryl hydrocarbon receptor in the development of control and 2,3,7,8-tetrachlorodibenzo-p-dioxin-exposed male mice. J Toxicol Environ Health A 64; 327-42.

Lindahl, T., Satoh, M. S., Poirier, G. G. and Klungland, A. (1995). Post-translational modification of poly(ADP-ribose) polymerase induced by DNA strand breaks. Trends Biochem Sci 20; 405-11.

Lindebro, M. C., Poellinger, L. and Whitelaw, M. L. (1995). Protein-protein interaction via PAS domains: role of the PAS domain in positive and negative regulation of the bHLH/PAS dioxin receptor-Arnt transcription factor complex. Embo J 14; 3528-39.

Liszt, G., Ford, E., Kurtev, M. and Guarente, L. (2005). Mouse Sir2 homolog SIRT6 is a nuclear ADP-ribosyltransferase. J Biol Chem 280; 21313-20.

Lo, R., Celius, T., Forgacs, A. L., Dere, E., MacPherson, L., Harper, P., Zacharewski, T. and Matthews, J. (2011). Identification of aryl hydrocarbon receptor binding targets in mouse hepatic tissue treated with 2,3,7,8-tetrachlorodibenzo-p-dioxin. Toxicol Appl Pharmacol 257; 38-47.

Lo, R. and Matthews, J. (2012). High-resolution genome-wide mapping of AHR and ARNT binding sites by ChIP-Seq. Toxicological sciences : an official journal of the Society of Toxicology 130; 349-61.

Loeffler, P. A., Cuneo, M. J., Mueller, G. A., DeRose, E. F., Gabel, S. A. and London, R. E. (2011). Structural studies of the PARP-1 BRCT domain. BMC structural biology 11; 37.

Page 202: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

187

Lonskaya, I., Potaman, V. N., Shlyakhtenko, L. S., Oussatcheva, E. A., Lyubchenko, Y. L. and Soldatenkov, V. A. (2005). Regulation of poly(ADP-ribose) polymerase-1 by DNA structure-specific binding. J Biol Chem 280; 17076-83.

Loseva, O., Jemth, A. S., Bryant, H. E., Schuler, H., Lehtio, L., Karlberg, T. and Helleday, T. (2010). PARP-3 is a mono-ADP-ribosylase that activates PARP-1 in the absence of DNA. J Biol Chem 285; 8054-60.

Lu, H., Cui, W. and Klaassen, C. D. (2011). Nrf2 protects against 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)-induced oxidative injury and steatohepatitis. Toxicol Appl Pharmacol 256; 122-35.

Luo, X. and Kraus, W. L. (2012). On PAR with PARP: cellular stress signaling through poly(ADP-ribose) and PARP-1. Genes Dev 26; 417-32.

Lusska, A., Wu, L. and Whitlock, J. P., Jr. (1992). Superinduction of CYP1A1 transcription by cycloheximide. Role of the DNA binding site for the liganded Ah receptor. J Biol Chem 267; 15146-51.

Ma, Q. (2002). Induction and superinduction of 2,3,7,8-tetrachlorodibenzo-rho-dioxin-inducible poly(ADP-ribose) polymerase: role of the aryl hydrocarbon receptor/aryl hydrocarbon receptor nuclear translocator transcription activation domains and a labile transcription repressor. Arch Biochem Biophys 404; 309-16.

Ma, Q. (2007). Aryl hydrocarbon receptor degradation-promoting factor (ADPF) and the control of the xenobiotic response. Mol Interv 7; 133-7.

Ma, Q. and Baldwin, K. T. (2000). 2,3,7,8-tetrachlorodibenzo-p-dioxin-induced degradation of aryl hydrocarbon receptor (AhR) by the ubiquitin-proteasome pathway. Role of the transcription activaton and DNA binding of AhR. J Biol Chem 275; 8432-8.

Ma, Q. and Baldwin, K. T. (2002). A cycloheximide-sensitive factor regulates TCDD-induced degradation of the aryl hydrocarbon receptor. Chemosphere 46; 1491-500.

Ma, Q., Baldwin, K. T., Renzelli, A. J., McDaniel, A. and Dong, L. (2001). TCDD-inducible poly(ADP-ribose) polymerase: a novel response to 2,3,7,8-tetrachlorodibenzo-p-dioxin. Biochem Biophys Res Commun 289; 499-506.

Ma, Q., Renzelli, A. J., Baldwin, K. T. and Antonini, J. M. (2000). Superinduction of CYP1A1 gene expression. Regulation of 2,3,7, 8-tetrachlorodibenzo-p-dioxin-induced degradation of Ah receptor by cycloheximide. J Biol Chem 275; 12676-83.

Ma, Q. and Whitlock, J. P., Jr. (1997). A novel cytoplasmic protein that interacts with the Ah receptor, contains tetratricopeptide repeat motifs, and augments the transcriptional response to 2,3,7,8-tetrachlorodibenzo-p-dioxin. J Biol Chem 272; 8878-84.

MacDonald, C. J., Ciolino, H. P. and Yeh, G. C. (2001). Dibenzoylmethane modulates aryl hydrocarbon receptor function and expression of cytochromes P50 1A1, 1A2, and 1B1. Cancer Res 61; 3919-24.

MacPherson, L. and Matthews, J. (2010). Inhibition of aryl hydrocarbon receptor-dependent transcription by resveratrol or kaempferol is independent of estrogen receptor alpha expression in human breast cancer cells. Cancer Lett 299; 119-29.

Page 203: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

188

Maeda, Y., Hunter, T. C., Loudy, D. E., Dave, V., Schreiber, V. and Whitsett, J. A. (2006). PARP-2 interacts with TTF-1 and regulates expression of surfactant protein-B. J Biol Chem 281; 9600-6.

Magni, G., Amici, A., Emanuelli, M., Orsomando, G., Raffaelli, N. and Ruggieri, S. (2004). Enzymology of NAD+ homeostasis in man. Cellular and molecular life sciences : CMLS 61; 19-34.

Marfori, M., Mynott, A., Ellis, J. J., Mehdi, A. M., Saunders, N. F., Curmi, P. M., Forwood, J. K., Boden, M. and Kobe, B. (2011). Molecular basis for specificity of nuclear import and prediction of nuclear localization. Biochim Biophys Acta 1813; 1562-77.

Marsischky, G. T., Wilson, B. A. and Collier, R. J. (1995). Role of glutamic acid 988 of human poly-ADP-ribose polymerase in polymer formation. Evidence for active site similarities to the ADP-ribosylating toxins. J Biol Chem 270; 3247-54.

Matheny, C., Day, M. L. and Milbrandt, J. (1994). The nuclear localization signal of NGFI-A is located within the zinc finger DNA binding domain. J Biol Chem 269; 8176-81.

Mathis, G. and Althaus, F. R. (1987). Release of core DNA from nucleosomal core particles following (ADP-ribose)n-modification in vitro. Biochem Biophys Res Commun 143; 1049-54.

Matsui, T., Segall, J., Weil, P. A. and Roeder, R. G. (1980). Multiple factors required for accurate initiation of transcription by purified RNA polymerase II. J Biol Chem 255; 11992-6.

Mayer-Kuckuk, P., Ullrich, O., Ziegler, M., Grune, T. and Schweiger, M. (1999). Functional interaction of poly(ADP-ribose) with the 20S proteasome in vitro. Biochem Biophys Res Commun 259; 576-81.

Mazen, A., Menissier-de Murcia, J., Molinete, M., Simonin, F., Gradwohl, G., Poirier, G. and de Murcia, G. (1989). Poly(ADP-ribose)polymerase: a novel finger protein. Nucleic Acids Res 17; 4689-98.

McGuire, J., Okamoto, K., Whitelaw, M. L., Tanaka, H. and Poellinger, L. (2001). Definition of a dioxin receptor mutant that is a constitutive activator of transcription: delineation of overlapping repression and ligand binding functions within the PAS domain. J Biol Chem 276; 41841-9.

McMillan, B. J. and Bradfield, C. A. (2007). The aryl hydrocarbon receptor is activated by modified low-density lipoprotein. Proc Natl Acad Sci U S A 104; 1412-7.

Mehrotra, P., Riley, J. P., Patel, R., Li, F., Voss, L. and Goenka, S. (2011). PARP-14 functions as a transcriptional switch for Stat6-dependent gene activation. J Biol Chem 286; 1767-76.

Mendoza-Alvarez, H. and Alvarez-Gonzalez, R. (2001). Regulation of p53 sequence-specific DNA-binding by covalent poly(ADP-ribosyl)ation. J Biol Chem 276; 36425-30.

Mendoza-Alvarez, H. and Alvarez-Gonzalez, R. (2004). The 40 kDa carboxy-terminal domain of poly(ADP-ribose) polymerase-1 forms catalytically competent homo- and heterodimers in the absence of DNA. J Mol Biol 336; 105-14.

Page 204: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

189

Menissier de Murcia, J., Ricoul, M., Tartier, L., Niedergang, C., Huber, A., Dantzer, F., Schreiber, V., Ame, J. C., Dierich, A., LeMeur, M., Sabatier, L., Chambon, P. and de Murcia, G. (2003). Functional interaction between PARP-1 and PARP-2 in chromosome stability and embryonic development in mouse. Embo J 22; 2255-63.

Messner, S., Altmeyer, M., Zhao, H., Pozivil, A., Roschitzki, B., Gehrig, P., Rutishauser, D., Huang, D., Caflisch, A. and Hottiger, M. O. (2010). PARP1 ADP-ribosylates lysine residues of the core histone tails. Nucleic Acids Res 38; 6350-62.

Messner, S. and Hottiger, M. O. (2011). Histone ADP-ribosylation in DNA repair, replication and transcription. Trends Cell Biol 21; 534-42.

Meyer-Ficca, M. L., Meyer, R. G., Coyle, D. L., Jacobson, E. L. and Jacobson, M. K. (2004). Human poly(ADP-ribose) glycohydrolase is expressed in alternative splice variants yielding isoforms that localize to different cell compartments. Experimental cell research 297; 521-32.

Meyer, B. K. and Perdew, G. H. (1999). Characterization of the AhR-hsp90-XAP2 core complex and the role of the immunophilin-related protein XAP2 in AhR stabilization. Biochemistry 38; 8907-17.

Meyer, B. K., Pray-Grant, M. G., Vanden Heuvel, J. P. and Perdew, G. H. (1998). Hepatitis B virus X-associated protein 2 is a subunit of the unliganded aryl hydrocarbon receptor core complex and exhibits transcriptional enhancer activity. Mol Cell Biol 18; 978-88.

Mhin, B. J., Lee, J. E. and Choi, W. (2002). Understanding the congener-specific toxicity in polychlorinated dibenzo-p-dioxins: chlorination pattern and molecular quadrupole moment. J Am Chem Soc 124; 144-8.

Michishita, E., Park, J. Y., Burneskis, J. M., Barrett, J. C. and Horikawa, I. (2005). Evolutionarily conserved and nonconserved cellular localizations and functions of human SIRT proteins. Molecular biology of the cell 16; 4623-35.

Miller, W. T. (2012). Tyrosine kinase signaling and the emergence of multicellularity. Biochim Biophys Acta 1823; 1053-7.

Mimura, J., Ema, M., Sogawa, K. and Fujii-Kuriyama, Y. (1999). Identification of a novel mechanism of regulation of Ah (dioxin) receptor function. Genes Dev 13; 20-5.

Mimura, J. and Fujii-Kuriyama, Y. (2003). Functional role of AhR in the expression of toxic effects by TCDD. Biochim Biophys Acta 1619; 263-8.

Mimura, J., Yamashita, K., Nakamura, K., Morita, M., Takagi, T. N., Nakao, K., Ema, M., Sogawa, K., Yasuda, M., Katsuki, M. and Fujii-Kuriyama, Y. (1997). Loss of teratogenic response to 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in mice lacking the Ah (dioxin) receptor. Genes Cells 2; 645-54.

Min, W., Cortes, U., Herceg, Z., Tong, W. M. and Wang, Z. Q. (2010). Deletion of the nuclear isoform of poly(ADP-ribose) glycohydrolase (PARG) reveals its function in DNA repair, genomic stability and tumorigenesis. Carcinogenesis 31; 2058-65.

Minsavage, G. D., Vorojeikina, D. P. and Gasiewicz, T. A. (2003). Mutational analysis of the mouse aryl hydrocarbon receptor tyrosine residues necessary for recognition of dioxin response elements. Arch Biochem Biophys 412; 95-105.

Page 205: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

190

Miwa, M., Ishihara, M., Takishima, S., Takasuka, N., Maeda, M., Yamaizumi, Z., Sugimura, T., Yokoyama, S. and Miyazawa, T. (1981). The branching and linear portions of poly(adenosine diphosphate ribose) have the same alpha(1 leads to 2) ribose-ribose linkage. J Biol Chem 256; 2916-21.

Miyamoto, T., Kakizawa, T. and Hashizume, K. (1999). Inhibition of nuclear receptor signalling by poly(ADP-ribose) polymerase. Mol Cell Biol 19; 2644-9.

Moennikes, O., Loeppen, S., Buchmann, A., Andersson, P., Ittrich, C., Poellinger, L. and Schwarz, M. (2004). A constitutively active dioxin/aryl hydrocarbon receptor promotes hepatocarcinogenesis in mice. Cancer Res 64; 4707-10.

Moss, J., Oppenheimer, N. J., West, R. E., Jr. and Stanley, S. J. (1986). Amino acid specific ADP-ribosylation: substrate specificity of an ADP-ribosylarginine hydrolase from turkey erythrocytes. Biochemistry 25; 5408-14.

Moss, J., Stanley, S. J., Nightingale, M. S., Murtagh, J. J., Jr., Monaco, L., Mishima, K., Chen, H. C., Williamson, K. C. and Tsai, S. C. (1992). Molecular and immunological characterization of ADP-ribosylarginine hydrolases. J Biol Chem 267; 10481-8.

Mostoslavsky, R., Chua, K. F., Lombard, D. B., Pang, W. W., Fischer, M. R., Gellon, L., Liu, P., Mostoslavsky, G., Franco, S., Murphy, M. M., Mills, K. D., Patel, P., Hsu, J. T., Hong, A. L., Ford, E., Cheng, H. L., Kennedy, C., Nunez, N., Bronson, R., Frendewey, D., Auerbach, W., Valenzuela, D., Karow, M., Hottiger, M. O., Hursting, S., Barrett, J. C., Guarente, L., Mulligan, R., Demple, B., Yancopoulos, G. D. and Alt, F. W. (2006). Genomic instability and aging-like phenotype in the absence of mammalian SIRT6. Cell 124; 315-29.

Mueller-Dieckmann, C., Kernstock, S., Lisurek, M., von Kries, J. P., Haag, F., Weiss, M. S. and Koch-Nolte, F. (2006). The structure of human ADP-ribosylhydrolase 3 (ARH3) provides insights into the reversibility of protein ADP-ribosylation. Proc Natl Acad Sci U S A 103; 15026-31.

Muller, S., Moller, P., Bick, M. J., Wurr, S., Becker, S., Gunther, S. and Kummerer, B. M. (2007). Inhibition of filovirus replication by the zinc finger antiviral protein. Journal of virology 81; 2391-400.

Mullins, D. W., Jr., Giri, C. P. and Smulson, M. (1977). Poly(adenosine diphosphate-ribose) polymerase: the distribution of a chromosome-associated enzyme within the chromatin substructure. Biochemistry 16; 506-13.

Nebert, D. W. (1989). The Ah locus: genetic differences in toxicity, cancer, mutation, and birth defects. Crit Rev Toxicol 20; 153-74.

Nebert, D. W., Dalton, T. P., Okey, A. B. and Gonzalez, F. J. (2004). Role of aryl hydrocarbon receptor-mediated induction of the CYP1 enzymes in environmental toxicity and cancer. J Biol Chem 279; 23847-50.

Nebert, D. W., Petersen, D. D. and Fornace, A. J., Jr. (1990). Cellular responses to oxidative stress: the [Ah] gene battery as a paradigm. Environ Health Perspect 88; 13-25.

Nebert, D. W., Roe, A. L., Dieter, M. Z., Solis, W. A., Yang, Y. and Dalton, T. P. (2000). Role of the aromatic hydrocarbon receptor and [Ah] gene battery in the oxidative stress response, cell cycle control, and apoptosis. Biochem Pharmacol 59; 65-85.

Page 206: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

191

Neuvonen, M. and Ahola, T. (2009). Differential activities of cellular and viral macro domain proteins in binding of ADP-ribose metabolites. J Mol Biol 385; 212-25.

Nguyen, T. A., Hoivik, D., Lee, J. E. and Safe, S. (1999). Interactions of nuclear receptor coactivator/corepressor proteins with the aryl hydrocarbon receptor complex. Arch Biochem Biophys 367; 250-7.

Nie, J., Sakamoto, S., Song, D., Qu, Z., Ota, K. and Taniguchi, T. (1998). Interaction of Oct-1 and automodification domain of poly(ADP-ribose) synthetase. FEBS Lett 424; 27-32.

Niere, M., Kernstock, S., Koch-Nolte, F. and Ziegler, M. (2008). Functional localization of two poly(ADP-ribose)-degrading enzymes to the mitochondrial matrix. Mol Cell Biol 28; 814-24.

Niere, M., Mashimo, M., Agledal, L., Dolle, C., Kasamatsu, A., Kato, J., Moss, J. and Ziegler, M. (2012). ADP-ribosylhydrolase 3 (ARH3), not poly(ADP-ribose) glycohydrolase (PARG) isoforms, is responsible for degradation of mitochondrial matrix-associated poly(ADP-ribose). J Biol Chem 287; 16088-102.

Nirodi, C., NagDas, S., Gygi, S. P., Olson, G., Aebersold, R. and Richmond, A. (2001). A role for poly(ADP-ribose) polymerase in the transcriptional regulation of the melanoma growth stimulatory activity (CXCL1) gene expression. J Biol Chem 276; 9366-74.

Nishihashi, H., Kanno, Y., Tomuro, K., Nakahama, T. and Inouye, Y. (2006). Primary structure and organ-specific expression of the rat aryl hydrocarbon receptor repressor gene. Biol Pharm Bull 29; 640-7.

Oei, S. L., Griesenbeck, J., Schweiger, M., Babich, V., Kropotov, A. and Tomilin, N. (1997). Interaction of the transcription factor YY1 with human poly(ADP-ribosyl) transferase. Biochem Biophys Res Commun 240; 108-11.

Ohbayashi, T., Oikawa, K., Iwata, R., Kameta, A., Evine, K., Isobe, T., Matsuda, Y., Mimura, J., Fujii-Kuriyama, Y., Kuroda, M. and Mukai, K. (2001). Dioxin induces a novel nuclear factor, DIF-3, that is implicated in spermatogenesis. FEBS Lett 508; 341-4.

Oikawa, K., Ohbayashi, T., Mimura, J., Fujii-Kuriyama, Y., Teshima, S., Rokutan, K., Mukai, K. and Kuroda, M. (2002). Dioxin stimulates synthesis and secretion of IgE-dependent histamine-releasing factor. Biochem Biophys Res Commun 290; 984-7.

Oka, J., Ueda, K., Hayaishi, O., Komura, H. and Nakanishi, K. (1984). ADP-ribosyl protein lyase. Purification, properties, and identification of the product. J Biol Chem 259; 986-95.

Oka, S., Kato, J. and Moss, J. (2006). Identification and characterization of a mammalian 39-kDa poly(ADP-ribose) glycohydrolase. J Biol Chem 281; 705-13.

Okayama, H., Honda, M. and Hayaishi, O. (1978). Novel enzyme from rat liver that cleaves an ADP-ribosyl histone linkage. Proc Natl Acad Sci U S A 75; 2254-7.

Okazaki, I. J. and Moss, J. (1996). Mono-ADP-ribosylation: a reversible posttranslational modification of proteins. Advances in pharmacology 35; 247-80.

Okazaki, I. J. and Moss, J. (1999). Characterization of glycosylphosphatidylinositiol-anchored, secreted, and intracellular vertebrate mono-ADP-ribosyltransferases. Annu Rev Nutr 19; 485-509.

Page 207: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

192

Okey, A. B. (1990). Enzyme induction in the cytochrome P-450 system. Pharmacol Ther 45; 241-98.

Okey, A. B., Bondy, G. P., Mason, M. E., Nebert, D. W., Forster-Gibson, C. J., Muncan, J. and Dufresne, M. J. (1980). Temperature-dependent cytosol-to-nucleus translocation of the Ah receptor for 2,3,7,8-tetrachlorodibenzo-p-dioxin in continuous cell culture lines. J Biol Chem 255; 11415-22.

Olabisi, O. A., Soto-Nieves, N., Nieves, E., Yang, T. T., Yang, X., Yu, R. Y., Suk, H. Y., Macian, F. and Chow, C. W. (2008). Regulation of transcription factor NFAT by ADP-ribosylation. Mol Cell Biol 28; 2860-71.

Ono, T., Kasamatsu, A., Oka, S. and Moss, J. (2006). The 39-kDa poly(ADP-ribose) glycohydrolase ARH3 hydrolyzes O-acetyl-ADP-ribose, a product of the Sir2 family of acetyl-histone deacetylases. Proc Natl Acad Sci U S A 103; 16687-91.

Ord, M. G. and Stocken, L. A. (1977). Adenosine diphosphate ribosylated histones. Biochem J 161; 583-92.

Oshima, M., Mimura, J., Sekine, H., Okawa, H. and Fujii-Kuriyama, Y. (2009). SUMO modification regulates the transcriptional repressor function of aryl hydrocarbon receptor repressor. J Biol Chem 284; 11017-26.

Oshima, M., Mimura, J., Yamamoto, M. and Fujii-Kuriyama, Y. (2007). Molecular mechanism of transcriptional repression of AhR repressor involving ANKRA2, HDAC4, and HDAC5. Biochem Biophys Res Commun 364; 276-82.

Otto, H., Reche, P. A., Bazan, F., Dittmar, K., Haag, F. and Koch-Nolte, F. (2005). In silico characterization of the family of PARP-like poly(ADP-ribosyl)transferases (pARTs). BMC genomics 6; 139.

Pansoy, A., Ahmed, S., Valen, E., Sandelin, A. and Matthews, J. (2010). 3-methylcholanthrene induces differential recruitment of aryl hydrocarbon receptor to human promoters. Toxicol Sci

Park, S., Henry, E. C. and Gasiewicz, T. A. (2000). Regulation of DNA binding activity of the ligand-activated aryl hydrocarbon receptor by tyrosine phosphorylation. Arch Biochem Biophys 381; 302-12.

Pavri, R., Lewis, B., Kim, T. K., Dilworth, F. J., Erdjument-Bromage, H., Tempst, P., de Murcia, G., Evans, R., Chambon, P. and Reinberg, D. (2005). PARP-1 determines specificity in a retinoid signaling pathway via direct modulation of mediator. Mol Cell 18; 83-96.

Perdew, G. H. (1988). Association of the Ah receptor with the 90-kDa heat shock protein. J Biol Chem 263; 13802-5.

Petrulis, J. R., Hord, N. G. and Perdew, G. H. (2000). Subcellular localization of the aryl hydrocarbon receptor is modulated by the immunophilin homolog hepatitis B virus X-associated protein 2. J Biol Chem 275; 37448-53.

Phelan, D., Winter, G. M., Rogers, W. J., Lam, J. C. and Denison, M. S. (1998). Activation of the Ah receptor signal transduction pathway by bilirubin and biliverdin. Arch Biochem Biophys 357; 155-63.

Page 208: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

193

Pohjanvirta, R., Korkalainen, M., McGuire, J., Simanainen, U., Juvonen, R., Tuomisto, J. T., Unkila, M., Viluksela, M., Bergman, J., Poellinger, L. and Tuomisto, J. (2002). Comparison of acute toxicities of indolo[3,2-b]carbazole (ICZ) and 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in TCDD-sensitive rats. Food Chem Toxicol 40; 1023-32.

Pohjanvirta, R. and Tuomisto, J. (1994). Short-term toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin in laboratory animals: effects, mechanisms, and animal models. Pharmacological reviews 46; 483-549.

Poirier, G. G., de Murcia, G., Jongstra-Bilen, J., Niedergang, C. and Mandel, P. (1982). Poly(ADP-ribosyl)ation of polynucleosomes causes relaxation of chromatin structure. Proc Natl Acad Sci U S A 79; 3423-7.

Poland, A. and Knutson, J. C. (1982). 2,3,7,8-tetrachlorodibenzo-p-dioxin and related halogenated aromatic hydrocarbons: examination of the mechanism of toxicity. Annu Rev Pharmacol Toxicol 22; 517-54.

Pollak, N., Dolle, C. and Ziegler, M. (2007). The power to reduce: pyridine nucleotides--small molecules with a multitude of functions. Biochem J 402; 205-18.

Pollenz, R. S. (2002). The mechanism of AH receptor protein down-regulation (degradation) and its impact on AH receptor-mediated gene regulation. Chem Biol Interact 141; 41-61.

Pollenz, R. S. (2007). Specific blockage of ligand-induced degradation of the Ah receptor by proteasome but not calpain inhibitors in cell culture lines from different species. Biochem Pharmacol 74; 131-43.

Pollenz, R. S. and Buggy, C. (2006). Ligand-dependent and -independent degradation of the human aryl hydrocarbon receptor (hAHR) in cell culture models. Chem Biol Interact 164; 49-59.

Pollenz, R. S., Santostefano, M. J., Klett, E., Richardson, V. M., Necela, B. and Birnbaum, L. S. (1998). Female Sprague-Dawley rats exposed to a single oral dose of 2,3,7,8-tetrachlorodibenzo-p-dioxin exhibit sustained depletion of aryl hydrocarbon receptor protein in liver, spleen, thymus, and lung. Toxicological sciences : an official journal of the Society of Toxicology 42; 117-28.

Pollenz, R. S., Sattler, C. A. and Poland, A. (1994). The aryl hydrocarbon receptor and aryl hydrocarbon receptor nuclear translocator protein show distinct subcellular localizations in Hepa 1c1c7 cells by immunofluorescence microscopy. Mol Pharmacol 45; 428-38.

Pongratz, I., Antonsson, C., Whitelaw, M. L. and Poellinger, L. (1998). Role of the PAS domain in regulation of dimerization and DNA binding specificity of the dioxin receptor. Mol Cell Biol 18; 4079-88.

Pongratz, I., Mason, G. G. and Poellinger, L. (1992). Dual roles of the 90-kDa heat shock protein hsp90 in modulating functional activities of the dioxin receptor. Evidence that the dioxin receptor functionally belongs to a subclass of nuclear receptors which require hsp90 both for ligand binding activity and repression of intrinsic DNA binding activity. J Biol Chem 267; 13728-34.

Page 209: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

194

Pongratz, I., Stromstedt, P. E., Mason, G. G. and Poellinger, L. (1991). Inhibition of the specific DNA binding activity of the dioxin receptor by phosphatase treatment. J Biol Chem 266; 16813-7.

Puga, A., Maier, A. and Medvedovic, M. (2000). The transcriptional signature of dioxin in human hepatoma HepG2 cells. Biochem Pharmacol 60; 1129-42.

Purnell, M. R., Stone, P. R. and Whish, W. J. (1980). ADP-ribosylation of nuclear proteins. Biochem Soc Trans 8; 215-27.

Racca, A. C., Camolotto, S. A., Ridano, M. E., Bocco, J. L., Genti-Raimondi, S. and Panzetta-Dutari, G. M. (2011). Kruppel-like factor 6 expression changes during trophoblast syncytialization and transactivates sshCG and PSG placental genes. PloS one 6; e22438.

Rannug, A., Rannug, U., Rosenkranz, H. S., Winqvist, L., Westerholm, R., Agurell, E. and Grafstrom, A. K. (1987). Certain photooxidized derivatives of tryptophan bind with very high affinity to the Ah receptor and are likely to be endogenous signal substances. J Biol Chem 262; 15422-7.

Realini, C. A. and Althaus, F. R. (1992). Histone shuttling by poly(ADP-ribosylation). J Biol Chem 267; 18858-65.

Rechsteiner, M., Hillyard, D. and Olivera, B. M. (1976). Turnover at nicotinamide adenine dinucleotide in cultures of human cells. Journal of cellular physiology 88; 207-17.

Reddy, T. E., Pauli, F., Sprouse, R. O., Neff, N. F., Newberry, K. M., Garabedian, M. J. and Myers, R. M. (2009). Genomic determination of the glucocorticoid response reveals unexpected mechanisms of gene regulation. Genome research 19; 2163-71.

Reeder, R. H., Ueda, K., Honjo, T., Nishizuka, Y. and Hayaishi, O. (1967). Studies on the polymer of adenosine diphosphate ribose. II. Characterization of the polymer. J Biol Chem 242; 3172-9.

Rich, P. R. (2003). The molecular machinery of Keilin's respiratory chain. Biochem Soc Trans 31; 1095-105.

Rippmann, J. F., Damm, K. and Schnapp, A. (2002). Functional characterization of the poly(ADP-ribose) polymerase activity of tankyrase 1, a potential regulator of telomere length. J Mol Biol 323; 217-24.

Robert, I., Dantzer, F. and Reina-San-Martin, B. (2009). Parp1 facilitates alternative NHEJ, whereas Parp2 suppresses IgH/c-myc translocations during immunoglobulin class switch recombination. The Journal of experimental medicine 206; 1047-56.

Roberts, B. J. and Whitelaw, M. L. (1999). Degradation of the basic helix-loop-helix/Per-ARNT-Sim homology domain dioxin receptor via the ubiquitin/proteasome pathway. J Biol Chem 274; 36351-6.

Rockel, T. D., Stuhlmann, D. and von Mikecz, A. (2005). Proteasomes degrade proteins in focal subdomains of the human cell nucleus. J Cell Sci 118; 5231-42.

Rolli, V., O'Farrell, M., Menissier-de Murcia, J. and de Murcia, G. (1997). Random mutagenesis of the poly(ADP-ribose) polymerase catalytic domain reveals amino acids involved in polymer branching. Biochemistry 36; 12147-54.

Page 210: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

195

Rosenthal, F., Feijs, K. L., Frugier, E., Bonalli, M., Forst, A. H., Imhof, R., Winkler, H. C., Fischer, D., Caflisch, A., Hassa, P. O., Luscher, B. and Hottiger, M. O. (2013). Macrodomain-containing proteins are new mono-ADP-ribosylhydrolases. Nat Struct Mol Biol 20; 502-7.

Rouleau, M., Aubin, R. A. and Poirier, G. G. (2004). Poly(ADP-ribosyl)ated chromatin domains: access granted. J Cell Sci 117; 815-25.

Rouleau, M., McDonald, D., Gagne, P., Ouellet, M. E., Droit, A., Hunter, J. M., Dutertre, S., Prigent, C., Hendzel, M. J. and Poirier, G. G. (2007). PARP-3 associates with polycomb group bodies and with components of the DNA damage repair machinery. J Cell Biochem 100; 385-401.

Rowlands, J. C., McEwan, I. J. and Gustafsson, J. A. (1996). Trans-activation by the human aryl hydrocarbon receptor and aryl hydrocarbon receptor nuclear translocator proteins: direct interactions with basal transcription factors. Mol Pharmacol 50; 538-48.

Ruf, A., Rolli, V., de Murcia, G. and Schulz, G. E. (1998). The mechanism of the elongation and branching reaction of poly(ADP-ribose) polymerase as derived from crystal structures and mutagenesis. J Mol Biol 278; 57-65.

Rulten, S. L., Fisher, A. E., Robert, I., Zuma, M. C., Rouleau, M., Ju, L., Poirier, G., Reina-San-Martin, B. and Caldecott, K. W. (2011). PARP-3 and APLF function together to accelerate nonhomologous end-joining. Mol Cell 41; 33-45.

Safe, S. (1990). Polychlorinated biphenyls (PCBs), dibenzo-p-dioxins (PCDDs), dibenzofurans (PCDFs), and related compounds: environmental and mechanistic considerations which support the development of toxic equivalency factors (TEFs). Crit Rev Toxicol 21; 51-88.

Safe, S. H. (1986). Comparative toxicology and mechanism of action of polychlorinated dibenzo-p-dioxins and dibenzofurans. Annu Rev Pharmacol Toxicol 26; 371-99.

Satoh, M. S., Poirier, G. G. and Lindahl, T. (1994). Dual function for poly(ADP-ribose) synthesis in response to DNA strand breakage. Biochemistry 33; 7099-106.

Savas, U., Bhattacharyya, K. K., Christou, M., Alexander, D. L. and Jefcoate, C. R. (1994). Mouse cytochrome P-450EF, representative of a new 1B subfamily of cytochrome P-450s. Cloning, sequence determination, and tissue expression. J Biol Chem 269; 14905-11.

Sbodio, J. I. and Chi, N. W. (2002). Identification of a tankyrase-binding motif shared by IRAP, TAB182, and human TRF1 but not mouse TRF1. NuMA contains this RXXPDG motif and is a novel tankyrase partner. J Biol Chem 277; 31887-92.

Sbodio, J. I., Lodish, H. F. and Chi, N. W. (2002). Tankyrase-2 oligomerizes with tankyrase-1 and binds to both TRF1 (telomere-repeat-binding factor 1) and IRAP (insulin-responsive aminopeptidase). Biochem J 361; 451-9.

Schaldach, C. M., Riby, J. and Bjeldanes, L. F. (1999). Lipoxin A4: a new class of ligand for the Ah receptor. Biochemistry 38; 7594-600.

Schmidt, J. V., Su, G. H., Reddy, J. K., Simon, M. C. and Bradfield, C. A. (1996). Characterization of a murine Ahr null allele: involvement of the Ah receptor in hepatic growth and development. Proc Natl Acad Sci U S A 93; 6731-6.

Page 211: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

196

Schoggins, J. W., Wilson, S. J., Panis, M., Murphy, M. Y., Jones, C. T., Bieniasz, P. and Rice, C. M. (2011). A diverse range of gene products are effectors of the type I interferon antiviral response. Nature 472; 481-5.

Schreiber, V., Ame, J. C., Dolle, P., Schultz, I., Rinaldi, B., Fraulob, V., Menissier-de Murcia, J. and de Murcia, G. (2002). Poly(ADP-ribose) polymerase-2 (PARP-2) is required for efficient base excision DNA repair in association with PARP-1 and XRCC1. J Biol Chem 277; 23028-36.

Schreiber, V., Dantzer, F., Ame, J. C. and de Murcia, G. (2006). Poly(ADP-ribose): novel functions for an old molecule. Nature reviews. Molecular cell biology 7; 517-28.

Schreiber, V., Hunting, D., Trucco, C., Gowans, B., Grunwald, D., De Murcia, G. and De Murcia, J. M. (1995). A dominant-negative mutant of human poly(ADP-ribose) polymerase affects cell recovery, apoptosis, and sister chromatid exchange following DNA damage. Proc Natl Acad Sci U S A 92; 4753-7.

Schrenk, D. (1998). Impact of dioxin-type induction of drug-metabolizing enzymes on the metabolism of endo- and xenobiotics. Biochem Pharmacol 55; 1155-62.

Seefeld, M. D., Corbett, S. W., Keesey, R. E. and Peterson, R. E. (1984). Characterization of the wasting syndrome in rats treated with 2,3,7,8-tetrachlorodibenzo-p-dioxin. Toxicol Appl Pharmacol 73; 311-22.

Seefeld, M. D. and Peterson, R. E. (1984). Digestible energy and efficiency of feed utilization in rats treated with 2,3,7,8-tetrachlorodibenzo-p-dioxin. Toxicol Appl Pharmacol 74; 214-22.

Seibert, C. and Sakmar, T. P. (2008). Toward a framework for sulfoproteomics: Synthesis and characterization of sulfotyrosine-containing peptides. Biopolymers 90; 459-77.

Seidel, S. D., Winters, G. M., Rogers, W. J., Ziccardi, M. H., Li, V., Keser, B. and Denison, M. S. (2001). Activation of the Ah receptor signaling pathway by prostaglandins. J Biochem Mol Toxicol 15; 187-96.

Seimiya, H. and Smith, S. (2002). The telomeric poly(ADP-ribose) polymerase, tankyrase 1, contains multiple binding sites for telomeric repeat binding factor 1 (TRF1) and a novel acceptor, 182-kDa tankyrase-binding protein (TAB182). J Biol Chem 277; 14116-26.

Seman, M., Adriouch, S., Haag, F. and Koch-Nolte, F. (2004). Ecto-ADP-ribosyltransferases (ARTs): emerging actors in cell communication and signaling. Curr Med Chem 11; 857-72.

Sewall, C. H. and Lucier, G. W. (1995). Receptor-mediated events and the evaluation of the Environmental Protection Agency (EPA) of dioxin risks. Mutat Res 333; 111-22.

Shah, G. M., Kandan-Kulangara, F., Montoni, A., Shah, R. G., Brind'amour, J., Vodenicharov, M. D. and Affar el, B. (2011). Approaches to detect PARP-1 activation in vivo, in situ, and in vitro. Methods in molecular biology 780; 3-34.

Sharifi, R., Morra, R., Denise Appel, C., Tallis, M., Chioza, B., Jankevicius, G., Simpson, M. A., Matic, I., Ozkan, E., Golia, B., Schellenberg, M. J., Weston, R., Williams, J. G., Rossi, M. N., Galehdari, H., Krahn, J., Wan, A., Trembath, R. C., Crosby, A. H., Ahel, D., Hay, R., Ladurner, A. G., Timinszky, G., Williams, R. S. and Ahel, I. (2013). Deficiency of

Page 212: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

197

terminal ADP-ribose protein glycohydrolase TARG1/C6orf130 in neurodegenerative disease. Embo J 32; 1225-37.

Shen, Z., Wells, R. L., Liu, J. and Elkind, M. M. (1993). Identification of a cytochrome P450 gene by reverse transcription--PCR using degenerate primers containing inosine. Proc Natl Acad Sci U S A 90; 11483-7.

Shertzer, H. G., Nebert, D. W., Puga, A., Ary, M., Sonntag, D., Dixon, K., Robinson, L. J., Cianciolo, E. and Dalton, T. P. (1998). Dioxin causes a sustained oxidative stress response in the mouse. Biochem Biophys Res Commun 253; 44-8.

Shetty, P. V., Bhagwat, B. Y. and Chan, W. K. (2003). P23 enhances the formation of the aryl hydrocarbon receptor-DNA complex. Biochem Pharmacol 65; 941-8.

Shieh, W. M., Ame, J. C., Wilson, M. V., Wang, Z. Q., Koh, D. W., Jacobson, M. K. and Jacobson, E. L. (1998). Poly(ADP-ribose) polymerase null mouse cells synthesize ADP-ribose polymers. J Biol Chem 273; 30069-72.

Simbulan-Rosenthal, C. M., Rosenthal, D. S., Luo, R., Samara, R., Espinoza, L. A., Hassa, P. O., Hottiger, M. O. and Smulson, M. E. (2003). PARP-1 binds E2F-1 independently of its DNA binding and catalytic domains, and acts as a novel coactivator of E2F-1-mediated transcription during re-entry of quiescent cells into S phase. Oncogene 22; 8460-71.

Skidmore, C. J., Davies, M. I., Goodwin, P. M., Halldorsson, H., Lewis, P. J., Shall, S. and Zia'ee, A. A. (1979). The involvement of poly(ADP-ribose) polymerase in the degradation of NAD caused by gamma-radiation and N-methyl-N-nitrosourea. European journal of biochemistry / FEBS 101; 135-42.

Slade, D., Dunstan, M. S., Barkauskaite, E., Weston, R., Lafite, P., Dixon, N., Ahel, M., Leys, D. and Ahel, I. (2011). The structure and catalytic mechanism of a poly(ADP-ribose) glycohydrolase. Nature 477; 616-20.

Slattery, E., Dignam, J. D., Matsui, T. and Roeder, R. G. (1983). Purification and analysis of a factor which suppresses nick-induced transcription by RNA polymerase II and its identity with poly(ADP-ribose) polymerase. J Biol Chem 258; 5955-9.

Slezak, B. P., Hatch, G. E., DeVito, M. J., Diliberto, J. J., Slade, R., Crissman, K., Hassoun, E. and Birnbaum, L. S. (2000). Oxidative stress in female B6C3F1 mice following acute and subchronic exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD). Toxicological sciences : an official journal of the Society of Toxicology 54; 390-8.

Smith, B. C. and Denu, J. M. (2006). Sir2 protein deacetylases: evidence for chemical intermediates and functions of a conserved histidine. Biochemistry 45; 272-82.

Smith, J. A. and Stocken, L. A. (1975). Chemical and metabolic properties of adenosine diphosphate ribose derivatives of nuclear proteins. Biochem J 147; 523-9.

Smith, S. (2001). The world according to PARP. Trends Biochem Sci 26; 174-9.

Smith, S. and de Lange, T. (2000). Tankyrase promotes telomere elongation in human cells. Current biology : CB 10; 1299-302.

Smith, S., Giriat, I., Schmitt, A. and de Lange, T. (1998). Tankyrase, a poly(ADP-ribose) polymerase at human telomeres. Science 282; 1484-7.

Page 213: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

198

Sogawa, K., Iwabuchi, K., Abe, H. and Fujii-Kuriyama, Y. (1995). Transcriptional activation domains of the Ah receptor and Ah receptor nuclear translocator. J Cancer Res Clin Oncol 121; 612-20.

Soldani, C., Lazze, M. C., Bottone, M. G., Tognon, G., Biggiogera, M., Pellicciari, C. E. and Scovassi, A. I. (2001). Poly(ADP-ribose) polymerase cleavage during apoptosis: when and where? Experimental cell research 269; 193-201.

Spittau, B., Wang, Z., Boinska, D. and Krieglstein, K. (2007). Functional domains of the TGF-beta-inducible transcription factor Tieg3 and detection of two putative nuclear localization signals within the zinc finger DNA-binding domain. J Cell Biochem 101; 712-22.

Stahl, B. U., Beer, D. G., Weber, L. W. and Rozman, K. (1993). Reduction of hepatic phosphoenolpyruvate carboxykinase (PEPCK) activity by 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) is due to decreased mRNA levels. Toxicology 79; 81-95.

Steeg, P. S., Palmieri, D., Ouatas, T. and Salerno, M. (2003). Histidine kinases and histidine phosphorylated proteins in mammalian cell biology, signal transduction and cancer. Cancer Lett 190; 1-12.

Sugihara, K., Kitamura, S., Yamada, T., Okayama, T., Ohta, S., Yamashita, K., Yasuda, M., Fujii-Kuriyama, Y., Saeki, K., Matsui, S. and Matsuda, T. (2004). Aryl hydrocarbon receptor-mediated induction of microsomal drug-metabolizing enzyme activity by indirubin and indigo. Biochem Biophys Res Commun 318; 571-8.

Sugimura, T., Fujimura, S., Hasegawa, S. and Kawamura, Y. (1967). Polymerization of the adenosine 5'-diphosphate ribose moiety of NAD by rat liver nuclear enzyme. Biochim Biophys Acta 138; 438-41.

Sun, J., Maresso, A. W., Kim, J. J. and Barbieri, J. T. (2004). How bacterial ADP-ribosylating toxins recognize substrates. Nat Struct Mol Biol 11; 868-76.

Sutter, T. R., Tang, Y. M., Hayes, C. L., Wo, Y. Y., Jabs, E. W., Li, X., Yin, H., Cody, C. W. and Greenlee, W. F. (1994). Complete cDNA sequence of a human dioxin-inducible mRNA identifies a new gene subfamily of cytochrome P450 that maps to chromosome 2. J Biol Chem 269; 13092-9.

Svensson, C. and Lundberg, K. (2001). Immune-specific up-regulation of adseverin gene expression by 2,3,7,8-tetrachlorodibenzo-p-dioxin. Mol Pharmacol 60; 135-42.

Swanson, H. I., Chan, W. K. and Bradfield, C. A. (1995). DNA binding specificities and pairing rules of the Ah receptor, ARNT, and SIM proteins. J Biol Chem 270; 26292-302.

Szanto, M., Brunyanszki, A., Kiss, B., Nagy, L., Gergely, P., Virag, L. and Bai, P. (2012). Poly(ADP-ribose) polymerase-2: emerging transcriptional roles of a DNA-repair protein. Cellular and molecular life sciences : CMLS 69; 4079-92.

Takada, T., Iida, K. and Moss, J. (1993). Cloning and site-directed mutagenesis of human ADP-ribosylarginine hydrolase. J Biol Chem 268; 17837-43.

Tanaka, M., Hayashi, K., Sakura, H., Miwa, M., Matsushima, T. and Sugimura, T. (1978). Demonstration of high molecular weight poly (adenosine diphosphate ribose). Nucleic Acids Res 5; 3183-94.

Page 214: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

199

Tanny, J. C., Dowd, G. J., Huang, J., Hilz, H. and Moazed, D. (1999). An enzymatic activity in the yeast Sir2 protein that is essential for gene silencing. Cell 99; 735-45.

Tanuma, S. and Endo, H. (1990). Identification in human erythrocytes of mono(ADP-ribosyl) protein hydrolase that cleaves a mono(ADP-ribosyl) Gi linkage. FEBS Lett 261; 381-4.

Tao, Z., Gao, P. and Liu, H. W. (2009). Identification of the ADP-ribosylation sites in the PARP-1 automodification domain: analysis and implications. J Am Chem Soc 131; 14258-60.

Tigges, J., Weighardt, H., Wolff, S., Gotz, C., Forster, I., Kohne, Z., Huebenthal, U., Merk, H. F., Abel, J., Haarmann-Stemmann, T., Krutmann, J. and Fritsche, E. (2013). Aryl hydrocarbon receptor repressor (AhRR) function revisited: repression of CYP1 activity in human skin fibroblasts is not related to AhRR expression. J Invest Dermatol 133; 87-96.

Tijet, N., Boutros, P. C., Moffat, I. D., Okey, A. B., Tuomisto, J. and Pohjanvirta, R. (2006). Aryl hydrocarbon receptor regulates distinct dioxin-dependent and dioxin-independent gene batteries. Mol Pharmacol 69; 140-53.

Tomita, S., Sinal, C. J., Yim, S. H. and Gonzalez, F. J. (2000). Conditional disruption of the aryl hydrocarbon receptor nuclear translocator (Arnt) gene leads to loss of target gene induction by the aryl hydrocarbon receptor and hypoxia-inducible factor 1alpha. Mol Endocrinol 14; 1674-81.

Tsuchiya, Y., Nakajima, M., Itoh, S., Iwanari, M. and Yokoi, T. (2003). Expression of aryl hydrocarbon receptor repressor in normal human tissues and inducibility by polycyclic aromatic hydrocarbons in human tumor-derived cell lines. Toxicological sciences : an official journal of the Society of Toxicology 72; 253-9.

Tulin, A. and Spradling, A. (2003). Chromatin loosening by poly(ADP)-ribose polymerase (PARP) at Drosophila puff loci. Science 299; 560-2.

Tuncel, H., Tanaka, S., Oka, S., Nakai, S., Fukutomi, R., Okamoto, M., Ota, T., Kaneko, H., Tatsuka, M. and Shimamoto, F. (2012). PARP6, a mono(ADP-ribosyl) transferase and a negative regulator of cell proliferation, is involved in colorectal cancer development. Int J Oncol 41; 2079-86.

Tuomisto, J. (2005). Does mechanistic understanding help in risk assessment--the example of dioxins. Toxicol Appl Pharmacol 207; 2-10.

Tuppurainen, K. and Ruuskanen, J. (2000). Electronic eigenvalue (EEVA): a new QSAR/QSPR descriptor for electronic substituent effects based on molecular orbital energies. A QSAR approach to the Ah receptor binding affinity of polychlorinated biphenyls (PCBs), dibenzo-p-dioxins (PCDDs) and dibenzofurans (PCDFs). Chemosphere 41; 843-8.

Uchida, K., Morita, T., Sato, T., Ogura, T., Yamashita, R., Noguchi, S., Suzuki, H., Nyunoya, H., Miwa, M. and Sugimura, T. (1987). Nucleotide sequence of a full-length cDNA for human fibroblast poly(ADP-ribose) polymerase. Biochem Biophys Res Commun 148; 617-22.

Ullrich, O., Reinheckel, T., Sitte, N., Hass, R., Grune, T. and Davies, K. J. (1999). Poly-ADP ribose polymerase activates nuclear proteasome to degrade oxidatively damaged histones. Proc Natl Acad Sci U S A 96; 6223-8.

Van den Berg, M., Birnbaum, L., Bosveld, A. T., Brunstrom, B., Cook, P., Feeley, M., Giesy, J. P., Hanberg, A., Hasegawa, R., Kennedy, S. W., Kubiak, T., Larsen, J. C., van Leeuwen,

Page 215: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

200

F. X., Liem, A. K., Nolt, C., Peterson, R. E., Poellinger, L., Safe, S., Schrenk, D., Tillitt, D., Tysklind, M., Younes, M., Waern, F. and Zacharewski, T. (1998). Toxic equivalency factors (TEFs) for PCBs, PCDDs, PCDFs for humans and wildlife. Environ Health Perspect 106; 775-92.

Verheugd, P., Forst, A. H., Milke, L., Herzog, N., Feijs, K. L., Kremmer, E., Kleine, H. and Luscher, B. (2013). Regulation of NF-kappaB signalling by the mono-ADP-ribosyltransferase ARTD10. Nature communications 4; 1683.

von Mikecz, A. (2006). The nuclear ubiquitin-proteasome system. J Cell Sci 119; 1977-84.

Wacker, D. A., Ruhl, D. D., Balagamwala, E. H., Hope, K. M., Zhang, T. and Kraus, W. L. (2007). The DNA binding and catalytic domains of poly(ADP-ribose) polymerase 1 cooperate in the regulation of chromatin structure and transcription. Mol Cell Biol 27; 7475-85.

Wahl, M., Lahni, B., Guenther, R., Kuch, B., Yang, L., Straehle, U., Strack, S. and Weiss, C. (2008). A technical mixture of 2,2',4,4'-tetrabromo diphenyl ether (BDE47) and brominated furans triggers aryl hydrocarbon receptor (AhR) mediated gene expression and toxicity. Chemosphere 73; 209-15.

Walisser, J. A., Bunger, M. K., Glover, E., Harstad, E. B. and Bradfield, C. A. (2004). Patent ductus venosus and dioxin resistance in mice harboring a hypomorphic Arnt allele. J Biol Chem 279; 16326-31.

Waller, C. L. and McKinney, J. D. (1995). Three-dimensional quantitative structure-activity relationships of dioxins and dioxin-like compounds: model validation and Ah receptor characterization. Chem Res Toxicol 8; 847-58.

Wattenberg, L. W. and Loub, W. D. (1978). Inhibition of polycyclic aromatic hydrocarbon-induced neoplasia by naturally occurring indoles. Cancer Res 38; 1410-3.

Weber, L. W., Lebofsky, M., Stahl, B. U., Gorski, J. R., Muzi, G. and Rozman, K. (1991). Reduced activities of key enzymes of gluconeogenesis as possible cause of acute toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in rats. Toxicology 66; 133-44.

Wejheden, C., Brunnberg, S., Larsson, S., Lind, P. M., Andersson, G. and Hanberg, A. (2010). Transgenic mice with a constitutively active aryl hydrocarbon receptor display a gender-specific bone phenotype. Toxicological sciences : an official journal of the Society of Toxicology 114; 48-58.

Welsby, I., Hutin, D. and Leo, O. (2012). Complex roles of members of the ADP-ribosyl transferase super family in immune defences: looking beyond PARP1. Biochem Pharmacol 84; 11-20.

Wentworth, J. N., Buzzeo, R. and Pollenz, R. S. (2004). Functional characterization of aryl hydrocarbon receptor (zfAHR2) localization and degradation in zebrafish (Danio rerio). Biochem Pharmacol 67; 1363-72.

White, S. S. and Birnbaum, L. S. (2009). An overview of the effects of dioxins and dioxin-like compounds on vertebrates, as documented in human and ecological epidemiology. Journal of environmental science and health. Part C, Environmental carcinogenesis & ecotoxicology reviews 27; 197-211.

Page 216: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

201

Whitelaw, M. L., Gottlicher, M., Gustafsson, J. A. and Poellinger, L. (1993). Definition of a novel ligand binding domain of a nuclear bHLH receptor: co-localization of ligand and hsp90 binding activities within the regulable inactivation domain of the dioxin receptor. Embo J 12; 4169-79.

Whitelaw, M. L., McGuire, J., Picard, D., Gustafsson, J. A. and Poellinger, L. (1995). Heat shock protein hsp90 regulates dioxin receptor function in vivo. Proc Natl Acad Sci U S A 92; 4437-41.

Whitlock, J. P., Jr. (1999). Induction of cytochrome P4501A1. Annu Rev Pharmacol Toxicol 39; 103-25.

Whitlock, J. P., Jr. and Gelboin, H. V. (1973). Induction of aryl hydrocarbon (benzo(a)pyrene) hydroxylase in liver cell culture by temporary inhibition of protein synthesis. J Biol Chem 248; 6114-21.

Williams, G. T., Lau, K. M., Coote, J. M. and Johnstone, A. P. (1985). NAD metabolism and mitogen stimulation of human lymphocytes. Experimental cell research 160; 419-26.

Wincent, E., Amini, N., Luecke, S., Glatt, H., Bergman, J., Crescenzi, C., Rannug, A. and Rannug, U. (2009). The suggested physiologic aryl hydrocarbon receptor activator and cytochrome P4501 substrate 6-formylindolo[3,2-b]carbazole is present in humans. J Biol Chem 284; 2690-6.

Wincent, E., Bengtsson, J., Mohammadi Bardbori, A., Alsberg, T., Luecke, S., Rannug, U. and Rannug, A. (2012). Inhibition of cytochrome P4501-dependent clearance of the endogenous agonist FICZ as a mechanism for activation of the aryl hydrocarbon receptor. Proc Natl Acad Sci U S A 109; 4479-84.

Wormke, M., Stoner, M., Saville, B. and Safe, S. (2000). Crosstalk between estrogen receptor alpha and the aryl hydrocarbon receptor in breast cancer cells involves unidirectional activation of proteasomes. FEBS Lett 478; 109-12.

Xing, X., Bi, H., Chang, A. K., Zang, M. X., Wang, M., Ao, X., Li, S., Pan, H., Guo, Q. and Wu, H. (2012). SUMOylation of AhR modulates its activity and stability through inhibiting its ubiquitination. Journal of cellular physiology 227; 3812-9.

Xu, M., Li, D., Lu, Y. and Chen, G. Q. (2007). Leukemogenic AML1-ETO fusion protein increases carcinogen-DNA adduct formation with upregulated expression of cytochrome P450-1A1 gene. Experimental hematology 35; 1249-55.

Yamamoto, J., Ihara, K., Nakayama, H., Hikino, S., Satoh, K., Kubo, N., Iida, T., Fujii, Y. and Hara, T. (2004). Characteristic expression of aryl hydrocarbon receptor repressor gene in human tissues: organ-specific distribution and variable induction patterns in mononuclear cells. Life sciences 74; 1039-49.

Yan, Q., Xu, R., Zhu, L., Cheng, X., Wang, Z., Manis, J. and Shipp, M. A. (2013). BAL1 and its partner E3 ligase, BBAP, link Poly(ADP-ribose) activation, ubiquitylation, and double-strand DNA repair independent of ATM, MDC1, and RNF8. Mol Cell Biol 33; 845-57.

Yelamos, J., Monreal, Y., Saenz, L., Aguado, E., Schreiber, V., Mota, R., Fuente, T., Minguela, A., Parrilla, P., de Murcia, G., Almarza, E., Aparicio, P. and Menissier-de Murcia, J. (2006). PARP-2 deficiency affects the survival of CD4+CD8+ double-positive thymocytes. Embo J 25; 4350-60.

Page 217: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

202

Yu, M., Schreek, S., Cerni, C., Schamberger, C., Lesniewicz, K., Poreba, E., Vervoorts, J., Walsemann, G., Grotzinger, J., Kremmer, E., Mehraein, Y., Mertsching, J., Kraft, R., Austen, M., Luscher-Firzlaff, J. and Luscher, B. (2005). PARP-10, a novel Myc-interacting protein with poly(ADP-ribose) polymerase activity, inhibits transformation. Oncogene 24; 1982-93.

Yu, W., Ginjala, V., Pant, V., Chernukhin, I., Whitehead, J., Docquier, F., Farrar, D., Tavoosidana, G., Mukhopadhyay, R., Kanduri, C., Oshimura, M., Feinberg, A. P., Lobanenkov, V., Klenova, E. and Ohlsson, R. (2004). Poly(ADP-ribosyl)ation regulates CTCF-dependent chromatin insulation. Nat Genet 36; 1105-10.

Zahradka, P. and Ebisuzaki, K. (1982). A shuttle mechanism for DNA-protein interactions. The regulation of poly(ADP-ribose) polymerase. European journal of biochemistry / FEBS 127; 579-85.

Zaniolo, K., Desnoyers, S., Leclerc, S. and Guerin, S. L. (2007). Regulation of poly(ADP-ribose) polymerase-1 (PARP-1) gene expression through the post-translational modification of Sp1: a nuclear target protein of PARP-1. BMC molecular biology 8; 96.

Zerfaoui, M., Errami, Y., Naura, A. S., Suzuki, Y., Kim, H., Ju, J., Liu, T., Hans, C. P., Kim, J. G., Abd Elmageed, Z. Y., Koochekpour, S., Catling, A. and Boulares, A. H. (2010). Poly(ADP-ribose) polymerase-1 is a determining factor in Crm1-mediated nuclear export and retention of p65 NF-kappa B upon TLR4 stimulation. Journal of immunology 185; 1894-902.

Zeytun, A., McKallip, R. J., Fisher, M., Camacho, I., Nagarkatti, M. and Nagarkatti, P. S. (2002). Analysis of 2,3,7,8-tetrachlorodibenzo-p-dioxin-induced gene expression profile in vivo using pathway-specific cDNA arrays. Toxicology 178; 241-60.

Zhang, F., Wang, Y., Wang, L., Luo, X., Huang, K., Wang, C., Du, M., Liu, F., Luo, T. and Huang, D. (2013). Poly(ADP-ribose) polymerase 1 is a key regulator of estrogen receptor alpha-dependent gene transcription. J Biol Chem 288; 11348-57.

Zhang, Q., Xiao, H., Chai, S. C., Hoang, Q. Q. and Lu, H. (2011). Hydrophilic residues are crucial for ribosomal protein L11 (RPL11) interaction with zinc finger domain of MDM2 and p53 protein activation. J Biol Chem 286; 38264-74.

Zhang, Z., Hildebrandt, E. F., Simbulan-Rosenthal, C. M. and Anderson, M. G. (2002). Sequence-specific binding of poly(ADP-ribose) polymerase-1 to the human T cell leukemia virus type-I tax responsive element. Virology 296; 107-16.

Zhao, K., Chai, X. and Marmorstein, R. (2003). Structure of the yeast Hst2 protein deacetylase in ternary complex with 2'-O-acetyl ADP ribose and histone peptide. Structure 11; 1403-11.

Zhao, K., Harshaw, R., Chai, X. and Marmorstein, R. (2004). Structural basis for nicotinamide cleavage and ADP-ribose transfer by NAD(+)-dependent Sir2 histone/protein deacetylases. Proc Natl Acad Sci U S A 101; 8563-8.

Zhu, Y., Chen, G., Lv, F., Wang, X., Ji, X., Xu, Y., Sun, J., Wu, L., Zheng, Y. T. and Gao, G. (2011). Zinc-finger antiviral protein inhibits HIV-1 infection by selectively targeting multiply spliced viral mRNAs for degradation. Proc Natl Acad Sci U S A 108; 15834-9.

Zhu, Y. and Gao, G. (2008). ZAP-mediated mRNA degradation. RNA biology 5; 65-7.

Page 218: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

203

Ziegler, M. (2000). New functions of a long-known molecule. Emerging roles of NAD in cellular signaling. European journal of biochemistry / FEBS 267; 1550-64.

Page 219: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

204

Copyright Acknowledgements

OXFORD UNIVERSITY PRESS LICENSE TERMS AND CONDITIONS Jun 11, 2013

This is a License Agreement between Laura M MacPherson ("You") and Oxford University

Press ("Oxford University Press") provided by Copyright Clearance Center ("CCC"). The

license consists of your order details, the terms and conditions provided by Oxford

University Press, and the payment terms and conditions.

All payments must be made in full to CCC. For payment instructions, please see information listed at the bottom of this form. License Number 3165990488601 License date Jun 11, 2013 Licensed content publisher Oxford University Press

Licensed content publication Nucleic Acids Research

Licensed content title

2,3,7,8-Tetrachlorodibenzo-p-dioxin poly(ADP-ribose) polymerase (TiPARP, ARTD14) is a mono-ADP-ribosyltransferase and repressor of aryl hydrocarbon receptor transactivation:

Licensed content author

Laura MacPherson, Laura Tamblyn, Sharanya Rajendra, Fernando Bralha, J. Peter McPherson, Jason Matthews

Licensed content date 02/01/2013

Type of Use Thesis/Dissertation

Institution name

Title of your work Characterization of 2,3,7,8-tetracholordibenzon-p-dioxin-inducible poly(ADP-ribose) polymerase and its role in aryl hydrocarbon receptor transactivation

Publisher of your work n/a

Expected Sep 2013

Page 220: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

205

publication date Permissions cost 0.00 USD Value added tax 0.00 USD Total 0.00 USD Total 0.00 USD Terms and Conditions

STANDARD TERMS AND CONDITIONS FOR REPRODUCTION OF MATERIAL

FROM AN OXFORD UNIVERSITY PRESS JOURNAL

1. Use of the material is restricted to the type of use specified in your order details. 2. This permission covers the use of the material in the English language in the following territory: world. If you have requested additional permission to translate this material, the terms and conditions of this reuse will be set out in clause 12. 3. This permission is limited to the particular use authorized in (1) above and does not allow you to sanction its use elsewhere in any other format other than specified above, nor does it apply to quotations, images, artistic works etc that have been reproduced from other sources which may be part of the material to be used. 4. No alteration, omission or addition is made to the material without our written consent. Permission must be re-cleared with Oxford University Press if/when you decide to reprint. 5. The following credit line appears wherever the material is used: author, title, journal, year, volume, issue number, pagination, by permission of Oxford University Press or the sponsoring society if the journal is a society journal. Where a journal is being published on behalf of a learned society, the details of that society must be included in the credit line. 6. For the reproduction of a full article from an Oxford University Press journal for whatever purpose, the corresponding author of the material concerned should be informed of the proposed use. Contact details for the corresponding authors of all Oxford University Press journal contact can be found alongside either the abstract or full text of the article concerned, accessible from www.oxfordjournals.org Should there be a problem clearing these rights, please contact [email protected] 7. If the credit line or acknowledgement in our publication indicates that any of the figures, images or photos was reproduced, drawn or modified from an earlier source it will be necessary for you to clear this permission with the original publisher as well. If this permission has not been obtained, please note that this material cannot be included in your publication/photocopies. 8. While you may exercise the rights licensed immediately upon issuance of the license at the end of the licensing process for the transaction, provided that you have disclosed complete and accurate details of your proposed use, no license is finally effective unless and until full payment is received from you (either by Oxford University Press or by Copyright Clearance Center (CCC)) as provided in CCC's Billing and Payment terms and conditions. If full payment is not received on a timely basis, then any license preliminarily granted shall be deemed automatically revoked and shall be void as if never granted. Further, in the event that you breach any of these terms and conditions or any of CCC's Billing and Payment terms and conditions, the license is automatically revoked and shall be void as if never granted. Use of materials as described in a revoked license, as well as any use of the materials beyond the scope of an unrevoked license, may constitute copyright infringement

Page 221: 2,3,7,8-TETRACHLORODIBENZO-p-DIOXIN-INDUCIBLE POLY(ADP … › bitstream › 1807 › 65686 › 1 › MacP… · Laura Mariko MacPherson . Doctor of Philosophy . Department of Pharmacology

206

and Oxford University Press reserves the right to take any and all action to protect its copyright in the materials. 9. This license is personal to you and may not be sublicensed, assigned or transferred by you to any other person without Oxford University Press’s written permission. 10. Oxford University Press reserves all rights not specifically granted in the combination of (i) the license details provided by you and accepted in the course of this licensing transaction, (ii) these terms and conditions and (iii) CCC’s Billing and Payment terms and conditions. 11. You hereby indemnify and agree to hold harmless Oxford University Press and CCC, and their respective officers, directors, employs and agents, from and against any and all claims arising out of your use of the licensed material other than as specifically authorized pursuant to this license.

12. Other Terms and Conditions:

v1.4

If you would like to pay for this license now, please remit this license along with your payment made payable to "COPYRIGHT CLEARANCE CENTER" otherwise you will be invoiced within 48 hours of the license date. Payment should be in the form of a check or money order referencing your account number and this invoice number RLNK501040863. Once you receive your invoice for this order, you may pay your invoice by credit card. Please follow instructions provided at that time. Make Payment To: Copyright Clearance Center Dept 001 P.O. Box 843006 Boston, MA 02284-3006 For suggestions or comments regarding this order, contact RightsLink Customer Support: [email protected] or +1-877-622-5543 (toll free in the US) or +1-978-646-2777. Gratis licenses (referencing $0 in the Total field) are free. Please retain this printable license for your reference. No payment is required.