Quick viewing(Text Mode)

Regulation of the 24-Hydroxylase Gene Promoter by 1,25-Dihydroxyvitamin D3 Adn Chemothereapeutics Drugs

Regulation of the 24-Hydroxylase Gene Promoter by 1,25-Dihydroxyvitamin D3 Adn Chemothereapeutics Drugs

Regulation of the 24-Hydroxylase Promoter by 1,25-

Dihydroxyvitamin D3 and Chemotherapeutics Drugs.

JOSEPH TAN CHENG TA

BHlthSc, Honours Sc

Thesis submitted for the degree of Doctor of Philosophy

Department of Paediatrics

The University of Adelaide

(Faculty of Health Sciences)

October 2005 Declaration

This work contains no material which has been accepted for the award of any other degree or diploma in any university or other tertiary institution and, to the best of my knowledge and belief, contains no material previously published or written by another person, except where due reference has been made in the text.

I give consent to this copy of my thesis when deposited in the university library, being available for loan and photocopying.

…………………………………….. ………………..

Joseph TanCheng Ta Dated Acknowledgement

Firstly, I would like to take this opportunity to express my sincere thanks to my principal supervisor, Associate Professor Charles Hii for his continuous supervision, support, encouragement and friendship which has made my PhD study an enjoyable experience.

I would also like to thank my co-supervisor Professor Antonio Ferrante for his guidance, patience, encouragement, friendship and support.

My special thanks to Dr Prem Dwivedi, with whom I worked closely, for his valuable supervision, great discussions and long lasting friendship. Also, special thanks to

Professor Brian May, Professor Howard Morris and Dr Peter O’Loughlin for their guidance and for allowing me to use their laboratories to undertake some aspects of my study. Many thanks to Barbara Nutchey, Dr Yong Lee, Dr Sophia Gao and Dr Paul

Anderson for their endless help, great discussion and friendship. A special thanks to all the PhD and honours students in the Department of Immunopathology for their great friendship and jokes that we shared. Also, many thanks to Judy Ferrante, Trish Harvey,

Lily Zedda, Bernadette Miller, Jess Franks, Reneé Mackenzie and other members of the

Department of Immunology, WCH, for their help and in making the laboratory a wonderful place to work in.

Last but not least, I would like to express my gratitude and thanks to God for his support, wisdom and strength, my wonderful family: my parents, my brothers and sisters for their continuous support and encouragement, to my dearest wife Ya-Hui and daughter Catrina, for their endless love, support, encouragement and for making my PhD years a meaningful and worthwhile experience.

ii Summary

Chemotherapy in childhood cancer patients is associated with reduced bone density that can result in osteoporotic fracture in survivors. A significant proportion of paediatric patients experience a reduction in plasma 25-hydroxyvitamin D3 [25(OH)D3] and 1,25- dihydroxyvitamin D3 [1,25(OH)2D3] levels during treatment, the basis of which is unknown. A balance between the bioactivation and degradation of 1,25(OH)2D3 is responsible for maintaining homoeostatic levels of 1,25(OH)2D3 at the correct set-point.

Whereas the , CYP27B1 (25-hydroxyvitamin D3 1α- hydroxylase), catalyses the of the precursor 25(OH)D3 to generate

1,25(OH)2D3, catabolic inactivation and cleavage of 1,25(OH)2D3 is achieved by the mitochondrial cytochrome P450 enzyme, 25-hydroxyvitamin D3 24-hydroxylase

(CYP24), which is highly expressed in bone and kidney cells. Since many of the signalling pathways which regulate the expression of CYP24 are also activated by chemotherapeutic drugs, we hypothesised that the drugs could cause the degradation of plasma 25(OH)D3 and 1,25(OH)2D3 by increasing CYP24 expression, the principal means of facilitating the bio-inactivation and degradation of plasma 25(OH)D3 and

1,25(OH)2D3. Using the kidney cell-lines, COS-1 and HEK293T cells, we now report that chemotherapeutic drugs, represented by daunorubicin hydrochloride (an anthracycline antibiotics), etoposide and vincristine sulphate (vinca and related compounds) and cisplatin (an alkylating agent), were able to enhance CYP24 promoter activity in kidney cell lines transfected with a CYP24 promoter-luciferase construct, either by themselves or in the presence of 1,25(OH)2D3. Dose-response studies with

iii daunorubicin hydrochloride and etoposide, two of the strongest inducers of CYP24 promoter under our experimental conditions, demonstrate that these drugs acted in a concentration-dependent manner. In addition to stimulating promoter activity on their own, the drugs also amplified the induction of the CYP24 promoter by

1,25(OH)2D3. Synergistic increases were generally observed when the cells were treated simultaneously with 1,25(OH)2D3 and a drug. The two kidney cell lines generally responded in a similar manner when challenged with the drugs, either in the presence or absence of 1,25(OH)2D3. Interestingly, the hydroxylated derivative of daunorubicin hydrochloride, hydrochloride which is also a commonly used chemotherapeutic drug, had no effect of promoter activity. Further studies with daunorubicin hydrochloride demonstrated that the effects of the drug per se were not mediated by oxidative stress and the vitamin D receptor was not required for daunorubicin hydrochloride per se to stimulate CYP24 promoter activity. However, daunorubicin hydrochloride caused a modest increase in the expression of the vitamin D receptor and this could contribute to its synergistic activity with 1,25(OH)2D3. In the presence of etoposide, there was also a tendency for the kidney cells to express higher levels of the vitamin D receptor. A key role for the extracellular signal-regulated kinase (ERK)1, ERK2 and ERK5 mitogen-activated protein (MAP) kinases was demonstrated for the inductive action of daunorubicin hydrochloride and etoposide, with

CYP24 promoter-specific transcription factors located in the first –298bp being likely targets of the ERK activity. Studies with a dominant negative mutant of MKK4, one of the two immediate upstream activators of the c-jun N-terminal kinase isoforms, demonstrated that this MAP kinase also played a crucial role in inductive actions of the

iv drugs. Consistent with their use in anti-cancer therapy, all of the above drugs killed the human promyelocytic HL60 leukaemic cells at very low concentrations but had no effect on the viability of kidney or liver cells, either at concentrations used in our experiments or at higher levels. Our data provide novel biochemical evidence that some of the commonly used chemotherapeutic drugs could cause an increase in the transcriptional activation of the promoter, most likely via the MAP kinases activating the transcription factors which bind to the CYP24 promoter. Such an effect could contribute to the reduction in plasma 25(OH)D3 and 1,25(OH)2D3 in some of the patients undergoing .

v Abreviations

1,25(OH)2D3/ 1,25D 1,25-dihydroxyvitamin D3

24,25(OH)2D2 24,25-dihydroxyvitamin D3

25(OH)D3 25-hydroxy vitamin D3

AP1 activator protein 1

BDNF brain-derived neurotrophic factor

CaT1 calcium transport protein

COS-1 African monkey kidney fibroblast cells

Cx43 connexin 43; a gap junction protein

CYP24 24-hydroxylase enzyme

CYP27b1 1α-hydroxylase enzyme

DN dominant negative

EBS Ets-1

ECaC2 epithelial calcium channel protein

EGF epithelial growth factor

ERK extracellular signal-regulated kinase

ERK5 Big MAP kinase 1

EtOH ethanol

Ets E26

FGF-2 fibroblast growth factor-2

HEK293T transform human embryonic kidney cells

HL60 human promyelocytic leukaemic

HPV human papillomaviruses

vi IL interleukin

JNK c-Jun NH2-terminal kinases

LY294002 PI3K inhibitor

MAPK mitogen-activated protein kinase

MEF2 myocyte enhancer factor 2

NCX1 Na+/Ca2+ exchanger pump

NF-Y transcription factor nuclear factor Y

NGF nerve growth factor

OPG osteoprotegerin gene

PDK phosphoinositide-dependent kinase

PI3K phosphatidylinositol 3-kinase

PKA protein kinase A

PKC protein Kinase C

PMA phorbol 12-myristate 13-acetate

PMCA plasma membrane calcium ATPase

PTH parathyroid hormone

RANKL NFkappa-B ligand

Ras small GTP-binding and hydrolyzing

ROS

RXR retinoic X receptor

Sap1 Ets-domain transcription factor

SGK serum- and -inducible kinase

Sp1 specific protein 1

vii TGF-β transforming growth factor-beta

Th T helper

Tpl tumor progression

TRPV transient receptor potential vanilloid

VDR vitamin D receptor

VDRE vitamin D responsive element

VEGF vascular endothelial growth factor

VP-16 etoposide

VSE vitamin D stimulatory element

viii CONTENTS

DECLARATION ...... ii ACKNOWLEDGEMENTS ...... iii THESIS SUMMARY...... iv ABREVIATIONS ...... vi

Chapter 1...... 1

Introduction ...... 1

1.1 General Introduction...... 2 1.2 Overview of the actions of vitamin D and its ...... 4 1.3 Biological Functions of vitamin D ...... 7 1.3.1 Vitamin D and bone...... 7 1.3.2 Vitamin D and Calcium Absorption ...... 14 1.3.3 Vitamin D and Cancer ...... 16 1.3.4 Vitamin D and the Immune system ...... 18 1.4 Vitamin D receptor...... 22 1.4.1 Nuclear VDR...... 22 1.4.2 Plasma membrane VDR ...... 23 1.4.3 Regulation of VDR expression...... 24 1.4.4 Non-classical VDRs ...... 25 1.4.4.1 Annexin II...... 25 1.4.4.2 1,25D3-MARRS ...... 26 1.5 Non-genomic action of vitamin D...... 28 1.6 Vitamin D and intracellular signaling ...... 29 1.6.1 Mitogen-activated protein kinases...... 29 1.6.1.1 ERK1/ERK2 ...... 30 1.6.1.3 JNK...... 34 1.6.2 Phosphathdylinositol 3-kinase...... 36 1.6.3 Protein Kinase C...... 41 1.6.4 Ras ...... 43 1.7 involved in vitamin D Metabolism ...... 45

1 1.7.1 Hepatic 25-hydroxylase enzyme...... 45 1.7.2 Kidney 1α-hydroxylase enzyme...... 46 1.7.2.1 Role of 1α-hydroxylase...... 46 1.7.2.2 Regulation of the 1α-hydroxylase ...... 50 1.7.3 24-hydroxylase enzymes (CYP24) ...... 51 1.7.3.1 Regulation of the CYP24 promoter by vitamin D...... 53 1.7.3.2 Signaling pathways that regulate the CYP24 promoter in response to vitamin D...... 56 1.8 Regulation of CYP24 by compounds other than vitaminD...... 59 1.9 Chemotherapy and plasma vitamin D ...... 61 1.10 Chemotherapeutic drugs...... 63 1.10.1 Categories of chemotherapeutic drugs...... 64 1.10.2 Principal targets of Chemotherapeutic drugs: the cell cycle and ...... 65 1.10.3 Actions of anticancer drugs ...... 66 1.10.3.1 Daunorubicin hydrochloride...... 67 1.10.3.2 Etoposide ...... 69 1.10.3.3 Vincristine Sulfate ...... 70 1.10.3.4 Cisplatin...... 71 1.11 Activation of cell signaling molecules by chemotherapeutic drugs...... 72 1.12 Hypotheses ...... 78 1.13 Aims ...... 79 Chapter 2...... 80

Materials and Methods...... 80

2.1 Materials...... 81 2.1.1 General chemicals, drugs and reagents...... 81 2.1.2 Reagents for Western blots and Kinase assays...... 81 2.1.3 Chemotherapeutic drugs ...... 82 2.1.4 Reagents for lysis, washing and assay buffer, ...... 82 2.1.5 Reagents for transfection ...... 83 2.1.6 Reagents for Luciferase Assay ...... 83 2.1.7 Antibodies...... 83 2.1.8 Serum, Culture media and buffers ...... 83 2.1.9 Flasks and plates for culture ...... 84 2.1.20 Cell lines...... 84 2.1.21 Plasmids/Vectors ...... 84 2.1.22 Buffers and Solutions ...... 88 2.1.23 Reagents for RNA extraction and real time PCR ...... 89

2 2.2 Methods...... 90 2.2.1 Cell Culture ...... 90 2.2.1.1 Culture of HEK293T and COS-1 cells...... 90 2.2.1.2 Cell Culture for kinase assay and for western blotting...... 90 2.2.1.3 Culture of HL 60 cells...... 91 2.2.2 Cell Viability Assay using trypan blue...... 91 2.2.3 Plasmid DNA Preparation and Transfection...... 92 2.2.3.1 Midi-preparation using midi-plasmid DNA preparation kit (Qiagen) ...... 92 2.2.3.2 Large scale preparation by cesium-chloride (CsCl) density gradient procedure...... 92 2.2.3.4 Dual Luciferase assay ...... 94 2.2.4 Preparation of cell lysates...... 94 2.2.5 Lowry’s Protein determination ...... 95 2.2.6 Kinase Assay and Western Blotting ...... 95 2.2.6.1 Purification of GST-jun (1-79) ...... 95 2.2.6.2 JNK kinase assay ...... 96 2.2.6.3 ERK1/ERK2/ERK5 kinase assays...... 97 2.2.6.4 Western Blotting...... 97 2.2.6.5 Western Blot recycling ...... 99 2.2.7 Quantification of mRNA expression using real-time PCR...... 99 2.2.7.1 RNA Extraction ...... 99 2.2.7.2 Generation of cDNA...... 99 2.2.7.3 Real-time PCR...... 100 2.2.8 Statistics...... 100 Chapter 3...... 101 Effects of chemotherapeutic drugs on the induction of the

CYP24 promoter in kidney cells...... 101

3.1 Introduction ...... 102 3.2 Methods for investigating CYP24 promoter induction...... 103 3.3 Results ...... 104 3.3.1 Effects of chemotherapeutic drugs on the activity of the -1.4kb CYP24 promoter in COS-1 and HEK293T kidney cells...... 104 3.3.2 Effects of chemotherapeutic drugs on the activity of the -298 bp CYP24 promoter in COS-1 kidney cells...... 108 3.3.3 Daunorubicin hydrochloride stimulated the –1.4kb CYP24 promoter activity in a dose- dependent manner in COS-1 kidney cells...... 112

3 3.3.4 Dose-dependent effect of daunorubicin hydrochloride on the -298bp CYP24 promoter in COS-1 cells and HEK293T cells...... 114 3.3.5 Etoposide stimulated the -1.4kb CYP24 promoter activity in a dose-dependent manner in COS-1 kidney cells...... 117 3.3.6 Doxorubicin hydrochloride did not stimulate the activity of the -298bp CYP24 promoter or enhance CYP24 induction by 1,25(OH)2D3 in COS-1 cells...... 119 3.4 Summary ...... 121 Chapter 4 Mechanism I...... 124 Mechanism of action of daunorubicin hydrochloride and etoposide on CYP24 promoter activity: Role of reactive oxygen species (ROS), VDR, VDRE and Sp1...... 124

4.1 ROS are not required for CYP24 promoter induction by daunorubicin hydrochloride and etoposide in COS-1 cells...... 125 4.2 Over expression of VDR did not enhance CYP24 promoter induction by daunorubicin hydrochloride in COS-1 and HEK293T cells...... 127 4.3 VDRE and GC box are not involved in CYP24 promoter induction by daunorubicin hydrochloride in COS-1 cells...... 131 4.4 Daunorubicin hydrochloride did not alter the activity of the VDR in COS-1 cells...... 133 4.5 Daunorubicin hydrochloride and etoposide increased the expression of VDR in COS-1 and HEK293T cells...... 135 4.6 Summary ...... 138 Chapter 5 Mechanism II ...... 139 Role of MAP Kinases in 1,25(OH)2D3 mediated transactivation regulation of the CYP24 promoter in COS-1 cells...... 139

5.1 Introduction ...... 140 5.1.1 Vectors used: ...... 142 5.2 Results:...... 143 5.2.1 1,25(OH)2D3 stimulated phosphorylation of ERK1/ERK2 in COS-1 cells...... 143 5.2.2 1,25(OH)2D3 stimulated the activity of ERK5 activity in COS-1 cells ...... 145 5.2.3 Dominant negative ERK1 and MEK5 inhibited 1,25(OH)2D3 mediated induction of the CYP24 promoter...... 148

4 5.2.4 Dominant negative Ras inhibited the stimulation of ERK1/ERK2 activity by 1,25(OH)2D3 in COS-1 cells...... 151 5.2.5 Dominant negative Ras inhibited the stimulation of ERK5 activity by 1,25(OH)2D3 in COS- 1 cells. 152 5.2.6 Dominant negative Ras inhibited CYP24 promoter induction by 1,25(OH)2D3 ...... 155 5.2.7 ERK2 interacted with RXRα independently of 1,25(OH)2D3...... 155 5.2.8 Phosphorylation of RXRα was dependent on the interaction between activated ERK1/ERK2 and RXRα...... 158 5.2.9 RXR phosphorylation played an important role in CYP24 induction by 1,25(OH)2D3. ... 160 5.2.10 Interaction between ERK5 and Ets-1 was 1,25(OH)2D3-dependent...... 162 5.2.11 1,25(OH)2D3 stimulated the phosphorylation of Ets-1 protein...... 164 5.2.12 Phosphorylation of Ets-1 requires the activation of HA-ERK5 by MEK5 and Ras...... 166 5.2.13 CYP24 promoter induction by 1,25(OH)2D3 required the Ets binding site (EBS) and the ERK5 module...... 169 5.3 Summary ...... 171 Chapter 6 Mechanism III...... 173 Role of MAP kinases, PI3K and PKC in mediating CYP24 promoter induction by daunorubicin hydrochloride and etoposide...... 173

6.1 Introduction ...... 174 6.2 Results ...... 174 6.2.1 Dominant negative ERK1 and MEK5 prevent the induction of CYP24 promoter activity by daunorubicin hydrochloride in COS-1 cells...... 174 6.2.2 Daunorubicin hydrochloride stimulated the activities of ERK1/ERK2 and ERK5 in COS-1 cells. 175 6.2.3 Dominant negative MKK4 inhibited daunorubicin hydrochloride induction of CYP24 promoter activity in COS-1 cells...... 177 6.2.4 Daunorubicin hydrochloride stimulated JNK activity in COS-1 cells...... 180 6.2.5 Dominant negative PKCζ inhibited daunorubicin hydrochloride induction of CYP24 promoter activity in COS-1 cells...... 183 6.2.6 Inhibitor of PI3K (LY294002) suppressed daunorubicin hydrochloride induction of CYP24 promoter activity in COS-1 cells...... 186 6.2.7 Inhibition of daunorubicin hydrochloride-mediated transcriptional activation of the CYP24 promoter activity by dominant negative ERK1 and MEK5 in HEK293T cells...... 188

5 6.2.8 Daunorubicin hydrochloride stimulated ERK1/ERK2 and ERK5 activities in HEK293T cells. 191 6.2.9 Effect of dominant negative MKK4 on daunorubicin hydrochloride-and 1,25(OH)2D3- mediated induction of CYP24 promoter activity in HEK293T cells...... 194 6.2.10 Daunorubicin hydrochloride stimulated JNK activity in HEK293T cells...... 196 6.2.11 PKCζ was not required for the induction of the CYP24 promoter by daunorubicin hydrochloride and 1,25(OH)2D3 in HEK293T cells...... 196 6.2.12 Role of MAP Kinases and PKCζ in the induction of CYP24 promoter by etoposide in COS- 1 cells. 198 6.2.13 Dominant negative ERK1 and not MEK5 inhibited etoposide mediated induction of the CYP24 promoter activity in COS-1 cells...... 200 6.2.14 Etoposide stimulated ERK1/ERK2 activity in COS-1 cells...... 202 6.2.15 Dominant negative MKK4 but not PKCζ inhibited etoposide-mediated induction of the CYP24 promoter activity in COS-1 cells...... 202 6.2.16 Etoposide stimulated JNK activity in COS-1 cells...... 205 6.3 Summary ...... 208 Chapter 7...... 209 Effects of chemotherapeutic drugs on endogenous CYP24 mRNA expression in kidney cells ...... 209

7.1 Introduction ...... 210 7.2 Methods...... 210 7.3 Results ...... 211 7.3.1 Daunorubicin hydrochloride and etoposide inhibited the expression of endogenous CYP24 by 1,25(OH)2D3...... 211 7.3.2 Pre-treatment of 1,25(OH)2D3 prevented etoposide from suppressing the induction of CYP24 mRNA expression by 1,25(OH)2D3...... 211 3.3.3 Etoposide stimulated the degradation of CYP24 mRNA...... 214 7.4 Summary ...... 216 Chapter 8...... 217

Discussion ...... 217

8.1 Discussion ...... 218 8.1.1 Anti-cancer drugs and CYP24 promoter activity...... 218 8.1.2 Cytotoxic effects of the drugs in HL60 cells but not in kidney cells...... 220 8.1.3 Lack of a role for the VDR and VDREs in the actions of the drugs...... 221

6 8.1.4 Daunorubicin and etoposide increased the expression of the VDR...... 222 8.1.5 Reactive oxygen species and the promoter inducing effect of daunorubicin and etoposide223 8.1.6 Non-genomic actions of 1,25(OH)2D3 involving intracellular signalling molecules ...... 224 8.1.7 Anti-cancer drugs and intracellular signalling molecules in CYP24 promoter activation.. 227 8.1.8 Suppression of CYP24 mRNA induction by daunorubicin and etoposide ...... 230 8.1.9 Clinical implications of this work and future directions...... 231 8.2 Summary ...... 234 References...... 237

7

Chapter 1

Introduction

1 1.1 General Introduction

The biologically active form of vitamin D, 1,25-dihydroxyvitamin D3 [1,25(OH)2D3], also known as calcitriol, is a secosteroid hormone that plays important roles in a range of physiological processes, including calcium and phosphate homeostasis and bone remodelling (Holick et al., 2003), cellular proliferation and differentiation (Brown et al.,

1999) and immune regulation (Griffin et al., 2003). Thus, it is crucial that the circulating concentration of 1,25(OH)2D3 is maintained at an appropriate level.

Children undergoing chemotherapy has reduced bone mineral mass and density, and are likely to get fractures and skeletal defects during and at the end of the treatment period and beyond (Kaste, 2004). Recent studies have found that these children have reduced plasma levels of 25-hydroxy vitamin D3 (25(OH)D3) and 1,25(OH)2D3 and that the levels of vitamin D metabolites remain low even after the completion of chemotherapy

(Halton et al., 1996; Atkinson et al., 1998; Arikoski et al., 1999a, 1999b).

The homeostatic levels of 25(OH)D3 and 1,25(OH)2D3 are regulated by a series of enzymes which include the 25-hydroxylase, 1α-hydroxylase enzyme (CYP27b1) and 24- hydroxylase enzyme (CYP24). The 25 hydroxylase and the 1α-hydroxylase are involved in the biosynthesis of 25(OH)D3 and the 1,25(OH)2D3, respectively (Jones et al., 1998).

On the other hand, the 24-hydroxylase plays a role in the biodegradation of these hormones (Jones et al., 1998). There are many factors which could potentially lead to the low serum levels of vitamin D seen in these cancer patients. However, based on the

2 evidence that CYP24 biodegrades both 25(OH)D3 and 1,25(OH)2D3, it is possible that

the reduction in plasma 25(OH)D3 and 1,25(OH)2D3 seen in these children is the result

of increased CYP24 action/expression. Currently, it is not known whether

chemotherapeutic drugs alter CYP24 expression or action and this forms the focus of this

thesis.

The CYP24 promoter is highly sensitive to 1,25(OH)2D3 stimulation. The action of

1,25(OH)2D3 is initiated through the binding of the hormone to a vitamin D receptor

(VDR) located either in the plasma membrane or cytoplasm (Norman et al., 2001; Erben et al., 2002). The binding of 1,25(OH)2D3 to VDR in turn activates both genomic and non-genomic responses to induce the transcriptional activation of the CYP24 promoter

(Norman et al., 2001). The genomic response involves the binding of the VDR/retinoic X receptor (RXR) complex to the CYP24 promoter (Dwivedi et al., 2000). The non- genomic response is due to the activation of signaling molecules such as the extracellular signal-regulated kinase (ERK)1 and ERK2, Big MAP kinase 1 (ERK5), c-Jun NH2- terminal kinases (JNK), Protein Kinase C (PKC) and phosphatidylinositol 3-kinase

(PI3K) which have been recently shown to regulate the CYP24 promoter through the phosphorylation of a variety of transcription factors (Dwivedi et al., 2000; Pearson et al.,

2001; Dwivedi et al., 2002; Nutchey et al., 2005). Interestingly, chemotherapeutic drugs

can also stimulate the activities of these signaling molecules (Nowak, 2002; Boldt et al.,

2002; Hershberger et al., 2002; Hayakawa et al., 2003; Mas et al., 2003). Thus, it is

possible that these drugs, through their ability to stimulate the activities of the above

signaling molecules, could increase CYP24 expression.

3 This chapter will review the literature on the physiological significance of 1,25(OH)2D3, how 1,25(OH)2D3 acts and how 1,25(OH)2D3 homeostasis is achieved through a series of enzymes that synthesize and degrade it. This chapter will also provide a discussion on the molecular mechanisms that have been implicated in regulating the transcriptional activation of the CYP24 promoter via both genomic and non-genomic routes, the later involving a series of intracellular signaling molecules. An overview of chemotherapeutic drugs and their modes of action will also be presented, leading to the formulation of my hypothesis and aims of my studies.

1.2 Overview of the actions of vitamin D and its metabolism

1,25(OH)2D3 is crucial for maintaining proper physiological function in human. It acts

via the vitamin D receptor (VDR). A deficiency in 1,25(OH)2D3 can result in

osteomalacia in adults and rickets among children (Reginato and Coquia, 2003).

1,25(OH)2D3 has been shown to play an important role in maintaining calcium and

phosphate homeostasis, cell differentiation and proliferation, bone homeostasis,

regulation of immune cells, and many other important functions (DeLuca, 2004;

Cantorna et al., 2004, Christakos et al., 2003), each of which will be discussed in detail below.

The physiological levels of 1,25(OH)2D3 are regulated by a series of enzymes which belong to the class of mitochondrial and microsomal cytochrome P450-dependent hydroxylases that participate in the synthetic pathways of the adrenal and gonadal steroid

4 hormones (Henry, 2001). Figure 1.1 shows a cartoon of 1,25(OH)2D3 metabolism. The biologically inactive precursor, vitamin D3 is mainly synthesized from 7- dehydrocholesterol in the basal epidermal layer of the skin through exposure to UV radiation (290nm-315 nm) (Holick, 2004). Vitamin D3 is then transported to the liver by vitamin D binding protein (DBP) where it is hydroxylated by vitamin D3-25-hydroxylase at carbon atom 25 to form 25(OH)D3 (Jones et al., 1998). The 25(OH)D3 is transported to the kidney where it is further hydroxylated by 25-hydroxyvitamin D3 1α-hydroxylase

(CYP27B1 or 1α(OH)ase) at carbon atom 1 to form the biologically active

1,25(OH)2D3. 1,25(OH)2D3 is inactivated or degraded in the kidney by 25- dihydroxyvitamin D3 24-hydroxylase (CYP24) (Jones et al., 1998). Unlike the 25- hydroxylation in the liver which appears to be loosely regulated, both the 1α- hydroxylation and C23/C24 hydroxylation are tightly regulated and appears to play an important role in regulating ambient levels of serum and cellular 1,25(OH)2D3 (Omdahl et al., 2002). Although the CYP27B1 and CYP24 are found in almost all the cell types in the body, the kidney is the major site for the biosynthesis and degradation of circulating

1,25(OH)2D3 (Holick, 1997; Jones et al., 1998). These topics will be discussed in greater detail in the following sections.

5 Liver

UV light

25-hydroxylase Skin 7, 8-dehydrocholesterol Pre-Vitamin D3 Vitamin D3

25(OH)D3

Gastrointestinal Diet CYP27B1

Track OH 1,25(OH)2D3

Kidney CYP24 HO OH

24,25(OH)2D3 1,24,25(OH)3D3 Circulatory 1,25(OH)2D3 Calcitronic acid

Excreted

Figure 1.1 Summary of the biosynthesis and biodegradation of 1,25(OH)2D3

Vitamin D3 is obtained from dietary sources or produced in the skin by the photolytic

cleavage of 7-dehydrocholesterol followed by thermal isomerization. Vitamin D3 is

transported to the liver via the serum vitamin D binding protein, where it is converted to

25-hydroxyvitamin D3, the major circulating metabolite of vitamin D3. The final

activation step, 1α-hydroxylation is carried out by 1α-hydroxylase (CYP27B1) and

occurs primarily, but not exclusively, in the kidney, forming 1,25(OH)2D3, the

hormonally active form of the vitamin. Catabolic inactivation is carried out by 24-

hydroxylase (CYP24), which catalyzes a series of oxidation steps, resulting in side chain

cleavage.

6 1.3 Biological Functions of vitamin D

Traditionally, 1,25(OH)2D3 was considered to play an important role in regulating

calcium homeostasis and bone mineralization. However, more recent studies show that

1,25(OH)2D3 also regulate cell proliferation and differentiation. Additionally,

1,25(OH)2D3 has been demonstrated to regulate the immune system, central nervous

system, reproduction and plays a protective role against cancer (Christakos et al., 2003;

Henry and Norman, 1984).

1.3.1 Vitamin D and bone

1,25(OH)2D3 plays a crucial role in bone development. Bone development is dependent

on the properly regulated differentiation, function and interaction of various types of

bone cells, which include chondrocytes, osteoblasts and osteoclasts cells (Stains and

Civitelli, 2005; Provot and Schipani, 2005). Lack of vitamin D causes rickets in children and osteomalacia in adults.

Vitamin D deficiency in infants can result from a number of factors. These range from an abnormality in vitamin D metabolism, dietary deficiency, metabolic errors to end-organ resistance due to mutations in vitamin D receptor (VDR) (Berry et al., 2002; Pettifor,

2004; Malloy et al., 2004). These defects can result in rickets and osteomalacia. The symptoms include bone pain, failure to thrive, and deformity of the thorax, resulting from severe deformities of the long bone and spine (Wharton and Bishop, 2003). Bone lesions

7 mainly result from inadequate mineralization and deposition of calcium to the growth

plate. This allows the cartilage to continue proliferating, resulting in irregular formation

of under mineralized bone trabeculae in the region directly underlying the growth plate

cartilage. Abnormality in mineralization also occurs in the diaphysis, affecting both the

cortical and trabecular bone. This mineralization defect causes the widening of osteoid

seams, that is, an increase in the amount of un-mineralized bone matrix located between

the layer of osteoblasts and the fully mineralized bone (St-Arnaud and Glorieux, 1997).

Vitamin D deficiency in adults occurs mainly in the sick, the elderly and Middle Eastern

people who lack exposure to sunlight (Lyman, 2005; Richardson, 2005; Flicker et al.,

2003; Venning, 2005; Brock et al., 2004). Low levels of circulating 25(OH)D3 are

normally detected during winter time, especially in people living in Europe (Lamberg-

Allardt et al., 2001; Carnevale et al., 2001). However in Australia, with the heightened awareness of ozone depletion and increasing incidence of skin cancer; there is a general fear among Australians towards exposure to sunlight. With the use of sun creams and clothing to block out UV radiation, there is thus an increase in the number of people, especially people with darker skin, with insufficient levels of 25(OH)D3 (Nowson et al.,

2004)

The abnormality in bone mineralization in humans such as seen in patients with osteomalacia and rickets can be cured by treating the patients with vitamin D metabolites

(Prince and Glendenning, 2004; Harrell et al., 1985). It was observed that the healing of bone lesions only occurs after normalization of blood concentrations of calcium and

8 phosphate. Furthermore, experiments using vitamin D-deficient animals and VDR-null mice showed that normal cartilage and bone mineralization can be achieved after circulating levels of calcium, phosphate and lactose were normalized by intravenous infusion or dietry supplement (Weinstein et al., 1984; Underwood and DeLuca, 1984;

Marie et al., 1982; Holtrop et al., 1986; Amling et al., 1999; Li et al., 1998). Based on the above observations, it was concluded that vitamin D per se is not essential for bone mineralisation and that the main mechanism by which vitamin D restores bone mineralization is due to the restoration of circulating concentrations of calcium and phosphate (i.e. by maintaining intestinal calcium adbsorption). Other observations that support this hypothesis include studies on hereditary vitamin D rickets patients, who carry mutations in the VDR. These patients do not respond to vitamin D treatment.

However, treatment with intravenous calcium and oral phosphate supplementation to normalize their blood concentrations of calcium and phosphate cured the patients of rickets and osteomalacia (Bliziotes et al., 1988; Weisman et al., 1987; Hochberg et al.,

1992; al-Aqeel et al., 1993; Balsan et al., 1986). Thus, it can be concluded that vitamin D plays an important role in bone development by maintaining calcium and phosphate homeostasis.

Nevertheless, various in vitro and in vivo studies have shown that 1,25(OH)2D3 has some roles in bone formation and maintenance. A study by Dardenne et al., (2004) demonstrated that replacement therapy such as feeding with a diet containing calcium, phosphate and lactose to restore normocalcemia in CYP27b1-null mice was not completely effective at normalizing bone growth in a similar manner to 1,25(OH)2D3

9 replacement therapy. This finding suggests that beside maintaining calcium and

phosphate levels, 1,25(OH)2D3 has a direct role in bone growth. It has also been shown

that 1,25(OH)2D3 regulates the proliferation and differentiation of immature osteoblast

(osteoprogenitors cells) into mature osteoblast cells (Siddhanti and Quarles, 1994).

The maintenance of bone homeostasis is a balance between continuous bone formation by osteoblasts and bone resorption by osteoclasts. 1,25(OH)2D3 can also act directly on osteoblasts and osteoclasts to regulate this process. The evidence for the interaction between 1,25(OH)2D3 and bone cells comes from the identification of VDR in osteoblast and osteoclast precursor cells (Mee et al., 1996; van Leeuwen et al., 2001). Furthermore, it was shown that VDR in osteoblast was essential for 1,25(OH)2D3 mediated effect.

Studies by Takeda et al., (1999) demonstrated that osteoblasts from VDR knockout mice were unable to stimulate the differentiation of osteoclasts.

Beside the presence of VDR, bone cells have been shown to possess enzymes that can synthesize and inactivate 1,25(OH)2D3 (Anderson et al., 2005). Thus, CYP24, CYP27b1

and CYP25 have all been found in osteoblast (Staal et al., 1997; Ichikawa et al., 1995;

Panda et al., 2001), whereas, only CYP24 is found in osteoclast (Menaa et al., 2000). The

presence of these enzymes suggests the importance of maintaining 1,25(OH)2D3 levels

within bone cells and that these cell-types may directly regulate plasma 1,25(OH)2D3

levels.

10 1,25(OH)2D3 plays a distinct role at regulating osteoblast in different maturation stages of osteoblasts (Owen et al., 1991). It was shown that 1,25(OH)2D3 could stimulate osteoprogenitors cells (immature osteoblast) to differentiate into mature osteoblasts that are capable of producing matrix protein for bone tissue mineralization

(Aubin and Heersche, 1997). Stimulation of osteoblast differentiation by 1,25(OH)2D3 was found to involve increased expression of bone phenotypic such as bone sialoprotein, collagen type I, osteoprotegerin and transcriptional enhancement of osteocalcin genes in humans and rats and osteopontin gene in mice. (Chen et al., 1996;

Harrison et al., 1989; Demay et al., 1990; Markose et al., 1990; Kerner et al., 1989; Noda et al., 1990).

In contrast, recent in vitro studies that suggest that 1,25(OH)2D3 attenuates the differentiation of osteoprogenotors cells into mature osteoblast cells. A study using mouse osteoblast and rat osteosarcoma cells showed that 1,25(OH)2D3 attenuates expression of Runx2, a protein which is essential for osteoblast differentiation in vitro and skeletal development during embryogenesis (Drissi et al., 2002). This finding was further supported by another study using osteoblasts from VDR knockout mice, which demonstrated the importance of VDR in suppressing osteoblast differentiation (Sooy et al., 2005). Drissi et al., (2002) suggested that the suppression of Runx2 expression plays an important role in moderating 1,25(OH)2D3 induced up-regulation of osteocalcin gene expression. This allows 1,25(OH)2D3 to control the maturation of osteoblast cells in concert with other hormones.

11 Beside having a role in bone formation, 1,25(OH)2D3 plays an important role in inducing

osteoclastic bone resorption by stimulating osteoclast formation and osteoclast activity.

Studies have demonstrated that 1,25(OH)2D3 regulates osteoclasts in an indirect manner

(Suda et al., 2003). 1,25(OH)2D3 acts on pre-osteoblast and stromal cells to increase

production and expression of membrane-bound receptor for activation of NFkappa-B

ligand (RANKL) which binds to RANK, a cell-bound receptor on the osteoclast (see

Figure 1.2) (Anderson et al., 1997; Lacey et al., 1998; Dusso et al., 2005). Through cell- to-cell interaction, binding of RANKL to RANK promotes osteoclast differentiation, activation, survival and adherence to bone surface (Kong Y et al., 1999; Suda et al.,

2003). The up-regulation of RANKL mRNA expression by 1,25(OH)2D3 in human osteoblast cells has been shown to involve the interaction of VDR/retinoic X receptor

(RXR) complex to the vitamin D responsive element (VDRE) on the human RANKL gene promoter (Kitazawa et al., 2003). In mouse osteoblasts cells, 1,25(OH)2D3 has also been shown to inhibit the expression of the osteoprotegerin gene (OPG), a decoy receptor for RANKL that inhibits the stimulatory effect of RANKL (Kondo et al., 2004).

In addition to the direct effect which 1,25(OH)2D3 has on the osteoblast and osteoclast, a range of complex interactions between different vitamin D metabolites, endocrine hormones and various local regulatory systems can influence or alter the effect of

1,25(OH)2D3 on bone cells (Yamanaka et al., 2003; Fratzl-Zelman et al., 2003; Gurlek et

al., 2002 ). Thus it can be seen that 1,25(OH)2D3 has major effects either directly or

indirectly on bone formation and resorption

12

Figure 1.2 1,25(OH)2D3 regulates osteoclastogenesis by reciprocal regulation of

receptor activator of NF-κB (RANK) ligand (RANKL) and osteoprotegerin (OPG).

1,25(OH)2D3-VDR increases the expression of RANKL on the surfaces of the osteoblast. RANKL interaction with its receptor, RANK, promotes maturation of osteoclast progenitor cells to mature osteoclasts, the bone-resorbing cells. 1,25(OH)2D3-

VDR also represses the expression of OPG, a decoy receptor that binds RANKL and prevents RANK-mediated osteoclastogenesis (Dusso et al., 2005).

13 1.3.2 Vitamin D and Calcium Absorption

Calcium is crucial for maintaining cellular metabolic processes (Bianchi et al., 2004;

Olson, 1985) and neuromuscular function, amongst other functions (Gommans et al.,

2002). One of the major biological functions of 1,25(OH)2D3 is to maintain calcium and

phosphate homeostasis. (DeLuca, 2004). The regulation of serum calcium levels by

1,25(OH)2D3 involves increasing the efficiency of calcium and phosphate absorption in the intestine and the reabsorption of calcium from the distal renal tubules (DeLuca,

2004). Earlier studies demonstrated that the hormone enhance the diffusion of calcium in the gastrointestinal mucosa by increasing the fluidity of the microvillus membrane

(Brasitus et al.,1986; Lau et al., 1986). However, recent studies have shown that VDR knockout mice have very high 1,25(OH)2D3 levels without increasing calcium absorption suggesting that the increase in calcium absorption is not related to increase membrane fluidity. Figure 1.3 shows a model for calcium transport across intestinal mucosa. 1,25(OH)2D3 can also enhance the translocation of calcium from both intestinal and kidney cells by stimulating the expression of the cytosolic calcium-binding proteins, calbindins-D9K and D28K (Walters et al 1999; Cao et al., 2002; Hoenderop et al., 2002).

Calbindin is specifically induced by 1,25(OH)2D3 and is thought to be one of the major proteins responsible for altering the flux of calcium across the gastrointestinal mucosa and kidney (Bronner 2003; Zheng et al., 2004). Recently, it was shown that calcium channels, such as the calcium transport protein CaT1 and CaT2, are vitamin D receptor

(VDR)-dependent and their expression is strongly stimulated by 1,25(OH)2D3 (Van

Cromphaut et al., 2001; Okano et al., 2004). CaT1 and CaT2 play an important role in

14 3Na+ NCX1 Ca 2+

Paracellular pathway 1,25(OH) D Transcellular pathway VDR 2 3

ECaC2 ATP

Calbindins-D9K ECaC1 PMCA1b ADP + - - +

Figure 1.3 Molecular model for intestinal calcium absorption.

Calcium absorption from the lumen of intestinal track occurs mainly through the

paracellular pathway or via the transcellular pathway. The rate of passive diffusion of

calcium via the paracellular pathway is dependent on luminal calcium concentration and

the integrity of intracellular tight junction. The transcellular pathway involves the

epithelial calcium entry through specialised epithelial calcium channels (ECaC) such as

ECaC2 and to a much lesser extent ECaC1. The rate of absorption of calcium via these

channels are highly dependent on intracellular free calcium concentration, number of

channel and the constant buffering of intracellular calcium by intracellular calcium

binding proteins such as Calbindin-D9K. 1,25(OH)2D3 can increase the number of both calcium channel and calcium binding proteins to increase the rate of calcium absorption.

The extrusion of calcium into the plasma requires ATP dependent calcium pumps such as plasma membrane calcium ATPase (PMCA1b) and sodium calcium exchanger (NCX)

(Bouillon R.,2003).

15 active transport of calcium into the intestinal and kidney cells, respectively (Bronner,

2003). In vivo studies show considerably reduced expression of CaT1 in the duodenum of

VDR knockout mice. These VDR knockout mice display all the typical characteristics of vitamin D resistance: high circulating levels of 1,25-(OH)2D3, coincident with hypocalcemia, hypophosphatemia, secondary hyperparathyroidism, failure to thrive, rickets and a progressive alopecia. (Okano et al., 2004; Bouillon et al., 2003).

1.3.3 Vitamin D and Cancer

A number of epidemiological studies have demonstrated a protective role for 25(OH)D3 and 1,25(OH)2D3 in preventing the initiation and progression of cancer (Martinez and

Willett, 1998; Grant, 2004; Vanchieri, 2004). These studies found an inverse association between vitamin D intake and subsequent development of cancer. Prostate, breast and colon cancer cells have been found to have reduced 1α-hydroxylase (CYP27B1) activity

(hence lower ability to synthesise 1,25(OH)2D3 ) (Chen et al., 2003; Ma et al., 2004) and have high levels of 24-hydroxylase (CYP24) (greater capability to degrade 1,25(OH)2D3

) (Zhao et al., 1999) and thus have low levels of 1,25(OH)2D3 within the tumor cells.

Serum levels of 25(OH)D3 and 1,25(OH)2D3 have been used as a biomarker for colon cancer development (Guyton et al 2003). Lower levels of serum 1,25(OH)2D3 have also been noted in patients with acute lymphocytic leukemia (ALL) (Fukasawa et al., 2001;

Arikoski et al., 1999). Prostate and breast cancer cells that express high levels of CYP24 are also resistant to 1,25(OH)2D3 treatment (Miller et al., 1995; Miettinen et al., 2004) presumably due to the rapid inactivation of 1,25(OH)2D3. Low levels of CYP27B1

16 mRNA and over-expression of CYP24 have been found in poorly-differentiated colon

cancers (Bareis et al., 2001). Studies have shown that the 20q13.2-q13.3

region, where CYP24 is located, was amplified in a variety of cancers, such as ovarian

cancer, prostate cancer and gastric adenocarcinomas (Tanner et a., 2000; Wolter et al.,

2002; Weiss et al., 2003) Because of these observation, CYP24 has been proposed to be an oncogene (Albertson et al., 2000). Overexpression of CYP24 can abrogate the

1,25(OH)2D3-mediated growth control (Ly et al.,1999; Miller et al., 1995). Increasing the exposure to solar radiation (UV-B) or frequent diet of fish and soy can increase circulatory levels of 25(OH)D3 that has a protective role against cancer (Grant, 2002).

How does 1,25(OH)2D3 suppress cancer growth? In vitro experiments have shown that

1,25(OH)2D3 can function as an antiproliferative, prodifferentiating, and proapoptotic

agent in cancer cells. This anticancer property of 1,25(OH)2D3 is due to the ability of

1,25(OH)2D3 to alter cell cycle control proteins, growth factor signaling,

metalloproteinase activity, angiogenesis, and expression of apoptosis inhibitors (Liu et

al.,1996; Kobayashi et al., 1993; James et al., 1996; Majewski et al., 1993; Haugen et al.,

1996; Kim et al., 1992). Studies have also found that 1,25(OH)2D3 is able to block the growth-promoting pathways initiated by insulin-like growth factor I (IGF-I) and oestrogen-receptor (ER) pathway (Lowe et al., 2003)

It has been found that 1,25(OH)2D3 is effective at inhibiting the growth of certain cancer

types e.g prostate, breast, lung, colon, pancreas, bone, myeloid leukemia and melanoma

(Krishnan et al., 2003). This is due to the fact that prostate, breast, colon cancer and

17 leukemic cells express VDR and thus can respond to 1,25(OH)2D3 treatment. Well-

differentiated breast and colon cancer express higher levels of VDR mRNA than poorly

differentiated cancer cells and this may be used as an indicator for treatment and in

predicting the clinical outcome of the cancer patients (Bareis et al., 2002). However, the

potential use of 1,25(OH)2D3 in the treatment and prevention of cancer is hampered by

the tendency of 1,25(OH)2D3 to cause hypercalcaemia. To overcome this, numerous

synthetic analogues of 1,25(OH)2D3 have been synthesized (e.g 1,25 dihydro-16-ene-23-

yne-vitamin D3, CB1093 and EB1089, a dimethyl side-chain analogue) (Akhter et al.,

2001; Zhou et al., 1990; Lokeshwar et al., 1999). Such analogues have been reported to

exert the same anti-proliferative and pro-apoptotic effects as 1,25(OH)2D3 without

causing hypercalcaemia (Hansen et al., 2001). Other studies have identified the optimal

treatment duration and dosage of 1,25(OH)2D3 in animals models and human trials, and

investigated whether 1,25(OH)2D3 can be used in combination with other

chemotherapeutic agents, cytokines, growth factors, specific inhibitors of CYP24 and

steroid hormones to produce a greater anti-tumour effect. (Ravid et al., 1999; Ly et al.,

1999; Koren et al., 2000; Beer et al., 2001; Evans TR et al., 2002; Ordonez-Moran et al.,

2005). Thus, it appears promising that 1,25(OH)2D3 and its analogues may one day be

used as an effective form of anti-cancer therapy.

1.3.4 Vitamin D and the Immune system

The ability of 1,25(OH)2D3 to regulate immune cells was originally suggested with the identification of VDR in promyelocytes (Kizaki et al., 1991). VDR has subsequently

18 been detected in almost all immune cell-types with significant levels found in T-

lymphocytes, precursor monocytes and peripheral macrophages. CD8+ T-lymphocytes

and immature cells of the thymus have been shown to have the highest amount of VDR

(Veldman et al., 2000). Immune cells such as neutrophils, monocytes and macrophages

were found to have the same 1α-hydroxylase and 24-hydroxylase enzyme as kidney cells,

enzymes that produce and metabolize 1,25(OH)2D3, respectively (Cohen and Gray,

1984). Both enzymes found in these immune cells can be regulated by 1,25(OH)2D3 or cytokines such as interferon (IFN)-γ and tumour necrosis factor (TNF)-α (Reichel et al.,

1987; Pryke et al., 1990). Regulation of the 1α-hydroxylase enzyme by various

inflammatory cytokines has an important role in regulating the levels of 1,25(OH)2D3,

therefore acting as a negative feedback loop in regulating (Overbergh et al.,

2000). A defect in the enzymes that synthesize 1,25(OH)2D3 in the immune cells could

result in autoimmune disease. Studies have found that, macrophages of non-obese

diabetic (NOD) mice have low levels of 1α-hydroxylase enzyme (Overbergh et al.,

2000). This finding suggests that 1,25(OH)2D3 plays a crucial role in suppressing the

immune system to prevent an autoimmune reaction. Indeed, 1,25(OH)2D3 has been

found to prevent autoimmune disease and extending graft survival by regulating the T

cell-mediated immunity (Adorini, 2003; Zhang et al., 2004; Gysemans et al., 2005). The

suppression of T cell-mediated inflammatory response is achieved by redirecting the

immune system towards the T helper (Th)2 pattern. 1,25(OH)2D3 can inhibit the

production of interleukin (IL)-2 and the actions of IL-2 and IFN-γ, resulting in the

suppression of Th1 responses (Jordan et al., 1989; Daniel et al., 2005). In addition,

1,25(OH)2D3 can stimulate IL-4, IL-10 and transforming growth factor-beta (TGF-β)

19 production which constitutes the Th2 response (Hayes et al., 2003; Daniel et al., 2005).

IL-10 is also known to interfere with the induction of Th1 response (Jung et al., 2004). In other studies, 1,25(OH)2D3 was found to suppress the expression of cytokines such as

IL-1β, IL-6, IL-12 and TNF-α in stimulated monocytes/macrophages and inhibit IL-2 and interferon- γ in activated lymphocytes (Cohen et al., 2001; Muller et al., 1992;

Muller et al., 2003 ). 1,25(OH)2D3 can also inhibit the maturation of dendritic cells which play an important role as antigen presenting cells (Griffin et al., 2000).

1,25(OH)2D3 can inhibit adherence of neutrophils to endothelial cells by decreasing the expression of ICAM-1 and ELAM-1 on endothelial cells (Chen, 1995). Studies using murine peritoneal mast cells have demonstrated that 1,25(OH)2D3 can inhibit both the release of histamine from activated mast cells and proliferation of the mast cells (Toyota et al., 1996). In general, studies have shown that 1,25(OH)2D3 acts as an immuno- suppressant. However, this suppression of the immune system by 1,25(OH)2D3 seems to be specific towards the suppression of autoimmune response as tests done on animals have demonstrated that 1,25(OH)2D3 prevents autoimmune disease and prolongs graft survival without interfering with the ability of the animal to act defensively against opportunistic infections (Cantorna et al., 1998). In fact, a number of studies have suggested that, 1,25(OH)2D3 plays an important role in combating infections. Human studies have found that children with rickets or 1,25(OH)2D3 deficiency are susceptible to recurrent infection. These studies have implicated an action of 1,25(OH)2D3 on phagocytes as opposed to lymphocytes. Neutrophils and monocytes of these patients have a deficiency in chemotaxis, reduced motility and impaired phagocytosis (Etzioni et al.,

1989; Lorente et al., 1976). In vitro studies have found that 1,25(OH)2D3 can increase

20 chemotaxis activity, superoxide production and killing activity in human monocytes,

macrophages and neutrophils (Venezio et al., 1988; Deutsch et al., 1995; Levy et al.,

1990; Tokuda et al., 2000). Thus, it appears that while 1,25(OH)2D3 can inhibit the

inflammatory response by inhibiting the interaction between neutrophils and endothelial

cells (Chen, 1995), 1,25(OH)2D3 actually promotes neutrophil anti-microbial activity as

discussed above.

Overall, the above studies show that 1,25(OH)2D3 is important for the proper functioning of the immune system and has clinical potential to suppress or prevent organ rejection in recipients and for treating autoimmune diseases such as encephalomyelitis, rheumatoid arthritis, type I diabetes and inflammatory bowel disease (Hullett et al., 1998; Cantorna et al., 1998; DeLuca and Cantorna, 2001; Hayes et al., 2003).

21 1.4 Vitamin D receptor

The actions of the 1,25(OH)2D3 are mediated by the vitamin D receptor (VDR). VDR knockout mice exhibit similar characteristics to 1,25(OH)2D3-deficient mice such as, hypocalcemia, hyperparathyroidism, rickets, osteomalacia, impaired muscular and motor functions. The of the VDR knockout mice can be prevented with a high calcium, phosphate and lactose diet (Amling et al., 1999; Burne et al., 2005). In bone cells, the levels of 1,25(OH)2D3-mediated induction of the CYP24 enzyme activity has been shown to correspond to the levels of VDR expression (Staal et al., 1997).

1.4.1 Nuclear VDR

The VDR nuclear protein belongs to a large superfamily of nuclear receptors, which includes steroid, retinoid and thyroid hormone receptors. As with other members of the superfamily, the VDR has an N-terminal DNA binding domain that contains two zinc finger DNA binding motifs. The C-terminal portion of the VDR contains a ligand binding domain (LBD) that associates with 1,25(OH)2D3 (Jurutka et al., 2001). The binding of

1,25(OH)2D3 to VDR initiates a conformational change in the VDR that enhances its interaction with one of the several isoforms of retinoid X receptor (RXR) to form a heterodimer that can bind with high affinity to vitamin D-responsive elements (VDREs) on the promoter of 1,25(OH)2D3-responsive genes (Thompson et al., 1998). The VDREs are located in the up-stream promoter region of genes regulated by 1,25(OH)2D3

(Ohyama et al., 1996).

22 Like other steroid hormone receptors, VDR is phosphorylated (Hsieh et al., 1991;

Hilliard et al., 1994; Jurutka et al., 1996). Low levels of phosphorylation are seen in the absence of 1,25(OH)2D3 whereas, 1,25(OH)2D3 treatment can increase phosphorylation of VDR. The functional role of phosphorylation remains unknown. However, it has been suggested that it might be important for nuclear localization, hormone binding and transcriptional activation or repression of the receptor (Whitfield et al., 1995; Jurutka et al., 2002).

1.4.2 Plasma membrane VDR

Beside genomic actions via the nuclear VDR and transcriptional activation, 1,25(OH)2D3 can initiate a non-genomic response, a more rapid response which occurs within minutes.

The non-genomic response includes the opening of chloride channels, activation of protein kinase C (PKC) and mitogen-activated protein (MAP) kinases (Norman et al.,

2001; Fleet, 2004). It has been suggested that the non-genomic response is initiated through binding of 1,25(OH)2D3 to a putative plasma membrane VDR that is distinct from the classical nuclear VDR (Fleet, 1999; Fleet, 2004). Norman et al., (2001) demonstrated that 6-s-trans bowl-shaped 1,25(OH)2D3 binds nuclear VDR whereas planar 6-s-cis ligand shaped 1,25(OH)2D3 binds to plasma membrane VDR. Another study by Nemere et al., (1994) demonstrated that the membrane VDR has different ligand binding affinity and molecular weight from nuclear VDR. The selective binding of

1,25(OH)2D3 and the different binding affinity of plasma membrane VDR suggest the presence of a different type of membrane VDR compared to the classical nuclear VDR.

23 However, it has been suggested that the plasma membrane VDR is derive from nuclear

VDR and that the altered selectivity and ligand specificity are due to interaction of the nuclear VDR with unknown cytoplasmic or plasma membrane protein co-factors (Erben et al., 2002). Erben et al., (2002) showed that both the genomic and non-genomic response to 1,25(OH)2D3 were abrogated in osteoblasts of mice expressing VDR with an intact hormone binding domain, but lacking the first zinc finger necessary for DNA binding. This finding demonstrates that the plasma membrane VDR is derived from the nuclear VDR. Consistent with the data by Erben et al., (2002), Capiati et al., (2002) showed that 1,25(OH)2D3 induce the translocation of VDR from the nuclear to the plasma membrane. This process of relocation involves the transport of nuclear VDR by microtubules and kinases.

1.4.3 Regulation of VDR expression

Up-regulation of VDR expression in the presence of 1,25(OH)2D3 has been shown in

various systems, including rat kidney and intestine, pig kidney, human leukemic cells and

mouse fibroblast cells (Healy et al., 2005; Costa et al., 1985; Mahonen et al., 1990).

Beside 1,25(OH)2D3, forskolin, PTH, genistein and 17beta-estradiol can also up-regulate

VDR expression (Song, 1996; Armbrecht et al., 1998; Lechner and Cross, 2003). Studies

by Barletta et al., (2003) have shown that up-regulation of VDR protein expression by

1,25(OH)2D3 in LLCPK-1 porcine kidney cells can be enhance by phorbol 12-myristate

13-acetate (PMA). They also showed that over expression of PKCα increased the

induction of the –1,500/+60 hVDR promoter construct which was further stimulated by

24 PMA (Barletta et al., 2003). This suggests that PKC plays a role in regulating of VDR expression. Other studies that support transcriptional regulation of VDR by PKC activation include a recent study by Qi X et al., (2002) who identified a functional AP-1 site in the mouse VDR promoter.

However, there are also reports that show that the increase in VDR protein levels by

1,25(OH)2D3 does not involve an increase in transcriptional activation (an increase in

VDR mRNA) but are due to increased stability of hormone occupied VDR (Wiese et al.,

1992; Arbour et al., 1993).

1.4.4 Non-classical VDRs

1.4.4.1 Annexin II

Recent studies have shown that 1,25(OH)2D3 can initiate rapid non-genomic actions by binding to annexin II membrane receptor (Baran et al., 2000). Using an antibody against annexin II, it was shown that annexin II played an important role in mediating the effect of 1,25(OH)2D3 on the activation of MAP kinases by the hormone (Johansen et al.,

2003). However, Mizwicki et al., (2004) reported that annexin II did not specifically bind

1,25(OH)2D3 and thus was not consider a putative receptor for 1,25(OH)2D3. Mizwicki et al. (2004) also suggested that the inhibition of 1,25(OH)2D3-mediated rapid non- genomic effects after addition of anti-annexin II antibody were likely to have been due to a blockade of the ability of annexin II to form complexes with itself and other proteins,

25 and not the inability to specifically bind 1,25(OH)2D3 in the presence of the antibody.

Thus, this issue remains to be resolved.

1.4.4.2 1,25D3-MARRS

1,25(OH)2D3-Membrane associated, rapid response, steroid-binding (1,25D3-MARRS) protein, a 66 kDa protein, was first purified from basolateral membranes of chick intestinal epithelial cells and latter identified as a candidate receptor for 1,25(OH)2D3

(Nemere et al., 1994). 1,25D3-MARRS protein was shown to bind 1,25(OH)2D3 more readily than 25(OH)D3 or 24,25(OH)2D3 (Nemere et al., 1994). At present, the membrane receptor can be detected or block by using a 1,25D3-MARRS antibody

(Ab099), an antibody that recognize the N-terminus of the MARRS protein (Nemere et al, 2004; Nemere et al., 1998). The membrane receptor has been found in intestinal epithelial cells, ameloblasts, odontoblasts, osteoblast and osteoclasts of chicks, mice, rat, human embryos and fetuses and has been shown to play an important role in; phosphate and calcium adsorption, cartilage formation and bone mineralization (Rohe et al., 2005;

Nemere et al., 2004; Berdal et al., 2003; Mesbah et al., 2002; Zhao and Nemere, 2002).

Activation of 1,25D3-MARRS protein by 1,25(OH)2D3 has been shown to stimulate

PKC. A study by Boyan et al, 2003 showed that 1,25(OH)2D3 can stimulate PKC with in minutes in the prehypertrophic and hypertrophic zone of costochondral cartilage of VDR knockout mice, suggesting the presence of a new responsive 1,25(OH)2D3 receptor other than the classical VDR. Furthermore, Boyan et al, 2003 showed that anti-1,25D3-

26 MARRS antibody can attenuate the stimulatory effect of 1,25(OH)2D3 on PKC ativity.

Similar studies perform on isolated intestinal epithelial cells of chicken using ribozyme

MARRS6, previously shown to actively cleave 1,25D3-MARRS sequence, also demonstrated the role played by 1,25D3-MARRS in 1,25(OH)2D3 medated PKC activation (Nemere et al, 2004). At presence, there is no literature on the activation of

MAPK signalling pathway by 1,25D3-MARRS.

1,25D3-MARRS has been shown to play a role in 1,25(OH)2D3 mediated phosphate uptake as the used of MARRS6 and anti-1,25D3-MARRS antibody inhibited

1,25(OH)2D3 mediated phosphate uptake. Similar studies have also shown that PKC is involved in regulating phosphate uptake. PMA has been shown to stimulate phosphate uptake while PKC inhibitor has been shown to inhibit 1,25(OH)2D3 mediated phosphate uptake (Sterling and Nemere, 2005). These suggest that the activation of PKC by 1,25D3-

MARRS protein plays an important role in regulating 1,25(OH)2D3 mediated phosphate uptake.

Unlike the classical VDR promoter, analysis of the 1,25D3-MARRS promoter region in mouse, rat and human failed to find any conventional VDRE sequence within the 1000 bp region immediately upstream of the transcriptional start site. Consistant with this,

1,25(OH)2D3 treatment did not increase mRNA or protein expression of 1,25D3-

MARRS. However, the hormone treatment triggers a rapid nuclear translocation of

1,25D3-MARRS. Interestingly, TGFβ1 was found to stimulate mRNA and protein expression of 1,25D3-MARRS (Rohe B et al., 2005).

27 The evidence for new classes of VDRs is overwhelming. Currently, it is important to characterise the functionality of these receptors. If each of these VDRs has unique functions, there would be potentials for the design of specific inhibitors to target these

VDRs. For example, if 1,25D3-MARRS protein has a unique role in Ca2+ transport, specific inhibitors of 1,25D3-MARRS might diminish the side effect of hypercalcemia when 1,25(OH)2D3 is used to treat cancer.

1.5 Non-genomic action of vitamin D

In addition to the classical manner in which the nuclear VDR, in partnership with RXR, mediate the genomic actions of 1,25(OH)2D3, it is now well accepted that 1,25(OH)2D3 can also exert non-genomic actions as alluded to in the above sections. The stimulation of the non-genomic response by 1,25(OH)D3 plays an important role in regulating cell differentiation, growth, apoptosis and many other cell functions (McGuire et al., 2001;

Morales et al., 2004; Fleet, 2004; Wang et al., 2005). The precise mechanism of how

1,25(OH)D3 mediates the non-genomic response is currently of great interest. This hormone has been shown to activate different cell signaling pathways in a range of cell- types. Examples of the signaling molecules activated by 1,25(OH)D3 in different cells types are: 1) protein kinase C (PKC) in keratinocytes (Bikle et al., 2001), bone cells

(Boyan and Schwartz, 2004), muscle cells (Buitrago et al., 2003), HL60 cells

(Marcinkowska et al., 1997) and macrophages (Higashimoto et al., 1995), 2) MAP kinases in bone cells (Boyan et al., 2002; Chae et al., 2002), muscle cells (Buitrago et al.,

2003), duodenal mucosae (Balogh et al., 2000) and HL60 cells (Marcinkowska et al.,

28 1997) and 3) protein kinase A (PKA) in intestinal cells (Phadnis and Nemere, 2003),

bone cells (Boyan et al., 2002), (Santillan and Boland, 1998) and

pancreatic islet cells (Bourlon et al., 1997). Using pharmacological inhibitors, a number

of studies have suggested that 1,25(OH)2D3 also stimulates the activity of

phosphatidylinositol 3-kinase (PI3K) in bone cells (Vertino et al., 2005), keratinocytes

(Johansen et al., 2003), muscle cells (Buitrago et al., 2002), vascular smooth muscle

(Rebsamen et al., 2002), macrophages (Hmama et al., 2004), T-lymphocytes (Belkowski

et al., 1999), promonocytic U937 cell line (Hmama et al., 1999) and HL60 cells

(Marcinkowska et al., 1998). It seems that different signaling molecules play different

functional roles in different cell-types. Thus, it is difficult to predict whether the same

signaling molecule is involved in mediating the actions of 1,25(OH)2D3 in a given cell-

type (Nutchey et al., 2005).

1.6 Vitamin D and intracellular signaling

1.6.1 Mitogen-activated protein kinases

MAP kinases play an important role in regulating cell growth, differentiation, immune

responses, stress responses and apoptosis (Cobb and Goldsmith, 1995). At least four

distinct MAP kinases pathways are found in mammalian cells (Figure 1.4). These are the

extracellular signal-regulated kinase (ERK)1/ERK2, c-Jun NH2-terminal kinases (JNK), p38 MAP kinase and Big MAP kinase 1 (BMK1) or (ERK5). MAP kinases can be activated by different membrane receptors such as receptor tyrosine kinases, cytokine

29 receptors, antigen receptors, G protein-coupled receptors, integrins and stressors such as radiation, hyperosmotic shock and anisomycin (Pearson et al., 2001; Chang and Karin,

2001).

1.6.1.1 ERK1/ERK2

ERK1/ERK2 have been shown to play an important role in 1,25(OH)2D3 mediated non- genomic action. ERK1/ERK2 was also found to regulate cellular responses to

1,25(OH)2D3 in different cell types. Examples include intestinal permeability (Gentili et al., 2004), HL60 cell differentiation (Marcinkowska, 2001), regulation of Ca2+ levels in skeletal muscle cell, cell contractility, proliferation and differentiation (Morelli et al.,

2001), and stimulation of CYP24 induction in kidney cells (Dwivedi et al., 2002).

Receptor-mediated stimulation of ERK1/ERK2 activity requires the activation of the small GTP binding protein, ras, PKC or PI3K. These in turn act on raf-1, a serine/threonine kinase which phosphorylates and activates the dual specificity MAP kinase/ERK kinase (MEK)1 and MEK2 (Figure 1.4) (Pearson et al., 2001). The MEKs phosphorylate the threonine and tyrosine residues on the TEY sequence in the activation motif of ERK1/ERK2 (Pearson et al., 2001). Activated ERK1/ERK2 then translocate to the nucleus and directly phosphorylate the transcriptional regulatory proteins. The amino acid sequence for ERK1/ERK2 are 90% identical (Boulton et al.,

1990). Since ERK1/ERK2 can regulate cytoplasmic as well as nuclear events, substrates for ERK1/ERK2 include cytoplasmic proteins such as the cytosolic phospholipase A2

30 (Pearson et al., 2001) and transcription factors such as TCF/ElK-1, NF-IL6 (Pearson et al., 2001) and 9-cis retinoid X receptor alpha (RXR alpha) (Dwivedi et al., 2002).

31

Figure 1.4 Mitogen-activated protein kinase (MAPK) signal transduction

pathway. Abbreviations: ASK, apoptosis signal-regulating kinase; MLK, mixed lineage

kinases, MEK, MAP kinase/ERK kinase; MKK; MAP kinase kinase; MAP, mitogen-

activated protein; c-Jun N-terminal protein kinases (JNK); Cx43, connexion 43;

MAPKAP kinase, MAP kinase-activated protein kinase; PRAK, p38-regulated and - activated kinase; SGK, serum- and glucorcoticoid-stimulated kinase; COT, cancer osaka thyroid; Elk-1, E twenty six (ets)-like transcription factor-1; CREB, cAMP response element binding protein, MEF, myocyte enhancing factor; ATF2, activated transcription factor (Hayashi and Lee, 2004)

32 1.6.1.2 BMK1/ERK5

Extracellular signal-regulated kinase 5 (ERK5) is also known as Big MAP kinase 1

(BMK1) because of its large protein size (~ 90 kDa) compared to other MAP kinases

Like other MAP kinases, ERK5 plays an important role in regulating cell survival, cell proliferation and differentiation and cell function (Hayashi and Lee, 2004). ERK5 plays a critical role in regulating embryo development. In vivo studies showed that ERK5 knockout mice die between embryonic days 9.5 and 11.5 due to defects in angiogenesis and cardiovascular development (Regan et al., 2002; Sohn et al., 2002). ERK5 activity has been found to be up-regulated in prostate and breast cancer cells (Weldon et al.,

2002; Mehta et al., 2003). ERK5 has also been shown to regulate CYP24 promoter induction by 1,25(OH)2D3 (Dwivedi et al., 2002). Similar to other MAP kinases, the

ERK5 pathway is activated by a wide variety of mitogens and stress stimuli. Agents that activate ERK5 include serum, epithelial growth factor (EGF), brain-derived neurotrophic factor (BDNF), nerve growth factor (NGF), vascular endothelial growth factor (VEGF), fibroblast growth factor-2 (FGF-2) and phorbol ester (Hayashi and Lee, 2004). The stress stimuli that activate ERK5 include hyperosmotic shock, oxidative stress, laminar flow shear stress and UV irradiation (Hayashi and Lee, 2004). This signaling molecule is exclusively activated by MEK5, which can be phosphorylated by two direct upstream kinases, MEKK2 and MEKK3 (Chao et al., 1999; Sun et al., 2001) (Figure 1.4). These findings were supported by studies using ERK5, MEKK3 and MEKK5 knockout mice.

Consistent with biochemical studies on the ERK5 pathway, the phenotype of ERK5-

33 deficient embryos resembles that of knockout embryos of MEKK3 and MEK5 (Yang et

al., 2000; Wang et al., 2005).

A number of molecules have been identified as substrates of the ERK5 pathway. These

include the MEF2 (myocyte enhancer factor 2) family of transcriptional factors (MEF2A,

MEF2C and MEF2D), Sap1a (Ets-domain transcription factor), SGK (serum- and glucocorticoid-inducible kinase), Cx43 (connexin 43; a gap junction protein), and Bad (a pro-apoptotic member of Bcl-2 family) (Hayashi and Lee, 2004). In 1,25(OH)2D3- stimulated kidney cells, ERK5 was reported to phosphorylate the transcription factor Ets-

1 (E26-1).

1.6.1.3 JNK

The c-Jun N-terminal protein kinases (JNKs) are stress-activated protein kinases

(SAPKs) since they are activated in response to a variety of environmental stress such as

radiation, heat shock, osmotic imbalance, DNA damage and inflammatory cytokines

(Kyriakis and Avruch, 2001). In addition to environmental stressors, JNKs are also

activated in some cell types by agonists such as growth factors, ligands which act via

heterotrimeric G-protein-coupled receptors, phorbol esters and 1,25(OH)2D3 (Ichijo,

1999; Cuenda, 2000; Nutchey et al., 2005). JNK is also activated in T lymphocytes by a

number of ligands (Singh and Zhang, 2004).

Activation of JNK has been associated with cell apoptosis. Studies have shown that

persistent activation of JNK either through over-expression of upstream activators or

34 following exposure to external stimuli can result in apoptosis. Furthermore, inhibition of

JNK signaling by over-expression of the dominant negative mutant of MKK4 (an

upstream regulator of JNK, see Figure 1.4) was shown to enhance cell survival

(Bogoyevitch and Court, 2004). However, there are also studies showing that under

certain circumstances, JNK activation can enhance cell survival (Roulston et al., 1998;

Andreka et al., 2001). Besides regulating cell survival, JNK also has a role in regulating

cell function. JNK was shown to play a role in 1,25(OH)2D3 mediated cell differentiation

in keratinocytes and HL60 cells (Johansen et al., 2003; Wang et al., 2003). A recent

study from our laboratory shows that JNK plays an important role in 1,25(OH)2D3

induction of the CYP24 promoter and its synergistic action with PMA in HEK293T cells

(Nutchey et al., 2005). In T lymphocytes, JNK1 and JNK2 have been proposed to play

essential roles in TH1 and TH2 differentiation (Dong et al., 2002).

At least 10 JNK isoforms have been identified. These include four isoforms of JNK1

(α1, α2, β1, β2) and JNK2 (α1, α2, β1, β2) and two isoforms of JNK3 (α1, α2). These

originate from of three mammalian genes, Jnk1, Jnk2, Jnk3 (Kyriakis and Avruch, 2001). Each of the three JNK genes is expressed on different

(Manning et al., 2002). Jnk1 and Jnk2 are ubiquitously expressed, whereas Jnk3 expression is restricted predominantly to the brain and testes (Kyriakis and Avruch,

2001). The JNK protein is preferentially phosphorylated at Tyr and Thr by MKK4 and

MKK7, respectively (Lawler et al., 1998). MKK4 and MKK7 are in turn activated by

MAP kinase/ERK kinase kinases (MEKKs) which include MEKK1, MEKK2, MEKK3

and MEKK4 (Figure 1.4) (Kyriakis and Avruch, 2001). However, some of these MEKKs

are not specific for the JNK pathway. For example, MEKK2 and MEKK3 are also up-

35 stream activators of p38 and ERK5 (Deacon and Blank, 1997; Deacon and Blank, 1999).

Other activators of MEK4 and MKK7 include tumor progression locus (Tpl)-2 (Salmeron et al., 1996), apoptosis signal-regulating kinase1 (ASK1) (Matsukawa et al., 2004) and mixed lineage kinases (MLK) family (Hirai et al., 1997). Like the MEKKs, these activators can also activate p38 and ERK1/ERK2. PI3K and PKC have also been shown to play a role in regulating JNK signaling. Stimulation of PI3K by 1,25(OH)2D3 was shown to activate JNK in human keratinocytes (Johansen et al., 2003). A study by

Brandlin et al., (2002) has demonstrated that over-expression of PKCε can stimulate JNK activity whereas PKCµ plays an inhibitory role in JNK activation.

A number of molecules have been identified as substrates of the JNK pathway. They include the transcription factor c-Jun, ATF2 (that heterodimerises with c-Jun and stimulates expression of the c-jun gene) (van Dam et al., 1995) and Elk-1 (which is involved in the induction of the c-fos gene, whose product forms the AP-1 heterodimer with c-Jun) (Cavigelli et al., 1995; Shaulian and Karin, 2001). In addition to the transcription factors, other substrates of JNK include Bcl-x(L), an anti-apoptotic protein which is found in the mitochondria (Ito et al., 2001).

1.6.2 Phosphathdylinositol 3-kinase

Phosphathidylinositol 3-kinase (PI3K) plays an important role in regulating a variety of cellular responses, including organization of the cytoskeleton, cell motility, cell

36 differentiation and growth, transformation, cell survival, and insulin action (Fry, 1994;

Toker and Cantley, 1997; Toker and Newton, 2000).

PI3K is a family of lipid kinases that catalyze the phosphorylation of the inositol ring of

phosphoinositides (PI) at the 3′ position. The sugar moiety can be phosphorylated at

different positions to produce different types of phosphatidylinositol (PtdIns)

(Vanhaesebroeck and Alessi, 2000). To date, nine members of the PI3K family,

belonging to 3 classes of PI3K, have been isolated from mammalian cells

(Vanhaesebroeck and Alessi, 2000). Three different lipid products can be generated by

the different PI3K members. They consist of the singly phosphorylated form PtdIns-3-P,

the doubly phosphorylated forms, PtdIns-3, 4-P2 and finally the triply phosphorylated

form PtdIns-3, 4, 5-P3 (Rameh and Cantley, 1999).

Three classes of PI3K are Class I, Class II and Class III PI3K. The Class I PI3K consists

of two subunits, a catalytic subunit and a regulatory adapter subunit. There are four

isoforms of the subunit, p110α, p110β, p110γ and p110δ and three mammalian genes

encode the adapter subunits, p85α, p85β and p85γ (Fruman et al., 1998; Vanhaesebroeck

and Alessi, 2000). Class I PI3K is composed of Class 1a and Class 1b kinases

(Vanhaesebroeck and Alessi, 2000). Whereas Class 1a PI3K appears to be expressed

ubiquitously, Class 1b appears to be restricted to leukocytes. Class II PI3K have

characteristic C-terminal C2 homology domains whereas, Class III PI3K are homologous

to Saccharomyces cervisiae VPS34 (Vacuolar protein sorting mutant) (Schu et al., 1993).

Currently, little is known about Class II and III PI3K.

37

The activation of Class 1a PI3K involves the interaction of a p85 adapter subunit with an

activated receptor tyrosine kinase which in turn recruits the p110 catalytic subunit of

PI3K to the plasma membrane where it can phosphorylate its main PtdIns-4, 5-

P2 (Cantrell, 2001). Besides receptor tyrosine kinase, other non-receptor tyrosine kinases such as Janus kinases in leucocytes can activate PI3K. Binding of p85 to signaling scaffold from by three adapter proteins Shc, Grb2 and Gab2 can also recruit the p110 catalytic subunit to the plasma membrane (Cantrell, 2001). In contrast to Class1a PI3K,

Class 1b kinase is activated by guanine-nucleotide-binding proteins (G proteins)

(Cantrell, 2001). For example, PI3Kγ, the sole member of Class 1b PI3K is activated solely by G-protein coupled receptors via Gβγ. PI3Kγ is composed of a p110γ catalytic subunit and a 101-kDa regulatory subunit (Vanhaesebroeck and Alessi, 2000). Class 1b

PI3K plays an essential role in regulating neutrophil and dendritic cell migration (Del

Prete et al., 2004; Puri et al., 2005).

PI3K has been shown to play a role in the activation of MAP kinase signaling pathways

(Baier et al., 1999; Pandey et al., 1999). A study by (Pandey et al., 1999) using Chinese

hamster ovary cells show that PI3K inhibitors, wortmannin and LY294002, can inhibit

the activation of ras, c-Raf-1, MEK and ERK1/ERK2 by vanadyl sulfate and insulin.

Another study by Baier et al. (1999), using similar PI3K inhibitors, show that retinoic

acid could induce PI3Kγ expression resulting in increased ERK2 activation in the

myelomonocytic cell-line U937. Similar results were also observed in a study using

human keratinocytes which also show that both wortmannin and LY294002 can

completely block the 1,25(OH)2D3 induced phosphorylation of ERK1/ERK2 and JNK1

38 (Johansen et al., 2003). These findings demonstrate that the PI3K is required for the activation of ras/c-Raf1/MEK/ERK or JNK signaling cascade induced by vanadyl sulfate, insulin, retinoic acid and 1,25(OH)2D3 (see Figure 1.5).

Several studies have demonstrated that the activation of certain PKC isozymes is

regulated by the PI3K pathway (Toker et al., 1994; Chou et al., 1998). The inhibition of

PI3K activity by wortmannin and LY294002 which results in inhibition of PKCζ in

HL60 cells, further demonstrate that PKCζ is a downstream target of PI3K (Liu et al.,

1998). This regulation is facilitated by the phosphorylation of PKCζ by phosphoinositide-

dependent kinase (PDK)-1 which is activated by PtdIns-3, 4-P2 and PtdIns-3, 4, 5-P3

(Vanhaesebroeck and Alessi, 2000).

39

Insulin

1,25(OH)2D3

Receptor

Nucleus

Gene Expression

Figure 1.5 Insulin and 1,25(OH)2D3 stimulates gene expression by activating

PI3K/Ras and down stream MEK/ERK1/ERK2 and JNK1 (Modified figure of

Johansen et al., 2003).

40 1,25(OH)2D3 can activate PI3K in different cell types (Hmama et al., 1999; Rebsamen et al., 2002). Studies have shown that PI3K plays a crucial role in 1,25(OH)2D3-induced differentiation of both myeloid cells and HL-60 cells (Marcinkowska et al., 1998;

Hmama et al., 1999). 1,25(OH)2D3 can also activate PI3K to promote vascular smooth muscle cell growth and calcification (Rebsamen et al., 2002). Beside this 1,25(OH)2D3 can also stimulate PI3K in human keratinocytes to promote cell differentiation (Johansen et al., 2003). Currently there are no reports regarding the role PI3K has on the regulation of CYP24 by 1,25(OH)2D3.

1.6.3 Protein Kinase C

Protein Kinase C (PKC) plays an important role in modulating a wide variety of cellular processes such as cell growth, differentiation, secretion, apoptosis and tumor development (Gschwendt, 1999). The activity of PKC is stimulated by a diverse range of agonists which act via G-protein coupled receptors (e.g: bombesin, vascular endothelial growth factor, and formylated peptides), receptor tyrosine kinases (e.g receptors for epidermal growth factor and platelet derived growth factor) and other agonists such as fatty acids (Nishizuka, 1995; Hii et al., 1995; Clerk et al., 2005). With the exception of the fatty acids, receptor-mediated activation of PKC results from the hydrolysis of membrane phosphatidylinositol (4,5) bisphosphate by phospholipase C.

This results in the production and accumulation of diacylglycerol (DAG), and inositol trisphosphate (IP3). While DAG activates PKC (see below), IP3 mobilizes stored Ca2+

(Nishizuka, 1995)

41

The PKC family is subdivided into three groups: conventional, novel and atypical isozymes, based on their requirements for Ca2+ and diacylglycerol (DAG) for activation

(Parker and Murray-Rust 2004). The conventional PKCs (PKC α, βI, βII, γ) are

dependent on calcium and are activated by DAG or PMA. The novel PKC isozymes

(PKC ε, η, δ, θ) are also activated by DAG or PMA but are calcium-independent. The

atypical PKC isozymes (PKC ζ, ι, λ) are calcium-independent and do not respond to

DAG or PMA (Gschwendt, 1999; Parker and Murray-Rust 2004). The difference in

response to DAG and requirement for Ca2+ between the three PKC subfamilies is due to

differences in their molecular structures. The structure of PKC isozymes consists of

conserved domains (C1 –C4) and in between these conserved domains are variable

regions (V1 to V5A). The C1 regions of the conventional and novel PKCs contain two

cysteine-rich sequences (Liu and Heckman, 1998). Ono et al, (1989) showed that

phospholipid and diglyceride/phorbol ester interact with the cysteine-rich sequences of

the conventional PKCs and novel PKCs. Deletion of the C1 region containing the two

cysteine-rich sequences completely abolishes phorbol ester binding (Burns and Bell,

1991). Furthermore, the lack of one of these cysteine-rich sequences in atypical PKC

renders it non-responsive to DAG or PMA (Ono et al., 1989). These findings suggest that

the two cysteine-rich sequences are DAG/phorbol ester binding domains that are

important for PKC activation. Unlike the conventional PKC subfamily, the novel PKCs

lack the C2 region that mediates Ca2+-binding (Ono et al., 1989).

42 1,25(OH)2D3 has been shown to activate PKC in a variety of cell types (Wali et al.,

2003; Shimizu et al., 2002). Although the precise mechanism through which this is

achieved is not known. Activation of PKC by 1,25(OH)2D3 was shown to play a part in

regulating the growth of chondrocytes (Schwartz et al., 2002), reducing insulin-induced

glucose uptake in rat adipocytes (Huang et al., 2002), promoting keratinocytes

differentiation (Bollinger Bollag and Bollag, 2001), increasing calcium influx into

skeletal muscle cells (Capiati et al., 2001) and reducing proliferation, but increasing

apoptosis and differentiation in human colon cancer cell lines (Chen et al., 1999). PKC

has also been shown to play an important role in 1,25(OH)2D3 induction of the CYP24

promoter and its synergistic action with PMA in HEK293T and COS-7 cells (Barletta et

al., 2004; Nutchey et al., 2005). Using dominant negative PKC isozyme mutants,

Nutchey et al., (2005) demonstrated that PKCα, in particular, is crucial for the synergistic action of the CYP24 promoter by 1,25(OH)2D3 and PMA.

1.6.4 Ras

The Ras signaling pathway is an essential signal transduction cascade that controls cell

survival, growth, differentiation and transformation. Three human Ras genes encode the

four isoforms of ras, H-ras, N-ras, K-ras4A and K-ras4B that are the founding members

of a large superfamily of Ras-related proteins (Repasky et al., 2004; Wennerberg et al.,

2005). Ras proteins are small GTP-binding and hydrolyzing proteins (GTPases) that are

targeted to the plasma membrane and exist in two distinct structural and functional

confirmations (Reuther and Der, 2000). Ras is activated when it is bound to GTP and

43 inactivated when bound to GDP. The level of signaling response induced by Ras is

dependent on the effector molecules that bind Ras, such as the MAP kinase kinase kinase,

Raf-1, Ral guanine nucleotide dissociation stimulator, PI 3K, phospholipase C, Ras

inhibitor, RIN1, and polarity protein AF6/Canoe (Ory and Morrison, 2004). The best- characterized Ras effectors are the three Raf serine/threonine kinases, A-Raf, B-Raf and

Raf-1 which stimulate MEK1/MEK2, the immediate upstream regulators of ERK1/ERK2

(Alberola-Ila and Hernandez-Hoyos, 2003; Vojtek and Der, 1998). The process of

ERK1/ERK2 activation by Ras involves the binding of Raf-1 to activated Ras (Ras -

GTP). This process then brings Raf to the plasma membrane where it becomes activated.

The activated Raf then phosphorylates and activates MEK1/MEK2. Various stimuli such

as hormones, cytokines and various drugs have been shown to activate Ras. Studies in

our laboratory have shown that Ras plays a role in the activation of ERK1/ERK2 and

ERK5 by 1,25(OH)2D3 in kidney cells (Dwivedi et al., 2002). Furthermore, using

dominant negative mutant of Ras, we showed that Ras plays a role in regulating the

ability of transcriptional factor, Ets-1, to transactivate the CYP24 promoter (Dwivedi et

al., 2002). Other studies also showed that Ras is involved in the activation of Raf-1 by

1,25(OH)2D3 (Solomon et al., 1999; Buitrago et al., 2003). It was suggested that the activation of Ras and downstream Raf-1 by 1,25(OH)2D3 was due to the hormone inhibiting the Ras -GTP hydrolysis in skeletal muscle cells (Buitrago et al., 2003). The activation of the transcription factor, activator protein 1 (AP1), in keratinocytes by

1,25(OH)2D3 was shown to be dependent on Ras signaling pathway (Johansen et al.,

2003).

44 1.7 Enzymes involved in vitamin D Metabolism

The biosynthesis and biodegradation of 1,25(OH)2D3 have been introduced above

(Figure 1.1), and these processes will be discussed in greater detail in the following sections. The discussion will include a description of the enzymes which are involved in the bio-activation and inactivation of 1,25(OH)2D3, their tissue distribution and how these enzymes are regulated.

1.7.1 Hepatic 25-hydroxylase enzyme.

The main circulatory form of vitamin D, 25-hydroxyvitamin D3 [25(OH)D3] is synthesized by the 25-hydroxylase. Until now, it is still unclear which CYP enzymes catalyse the 25 hydroxylation reaction. Recent studies have identified five CYP enzymes from various species that are capable of vitamin D 25-hydroxylation. These enzymes include the CYP2D11, CYP2D25, CYP2J2/3, CYP3A4 and CYP2R1. However, only the latter three seem to match the enzymatic properties and distribution pattern of the physiologically relevant human liver microsomal enzyme (Prosser and Jones, 2004).

Although the main site for 25-hydroxylation of vitamin D3 occurs in the hepatocytes in the liver, 25-hydroxylase activity has also been found in other tissues such as bone, intestine, kidney and skin (Ichikawa et al., 1995; Hosseinpour et al., 2000; Gascon-Barre et al., 2001). Furthermore, although both hepatocytes and nonparenchymal cells of the rat liver can efficiently take up vitamin D3, only the hepatocytes can carry out the 25- hydroxylation of vitamin D3 (Dueland, 1981).

45

Two types of hepatic 25-hydroxylase enzyme have been identified, the microsomal and

the mitochondrial forms (Dahlback and Wikvall, 1987). Since the microsomal 25-

hydroxylase enzyme has a higher affinity for 25(OH)D3, hence lower Km value than the

mitochondrial enzyme, it is thought that the microsomal 25-hydroxylase is responsible

for maintaining physiological levels of 25(OH)D3 whereas, the mitochondrial enzyme is

used only when very high levels of circulating 25(OH)D3 are required (Holick et al.,

1994; Omdahl et al., 2002). The microsomal 25-hydroxylase uses vitamin D3,

1α(OH)D3, vitamin D2 and 1α(OH)D2 as substrates (Wikvall, 2001). Between

1α(OH)D3 and vitamin D3, the former is a much better substrate for 25-hydroxylase

enzymes than vitamin D3 (Dahlback and Wikvall, 1988). The mitochondrial 25-

hydroxylase has a broad substrate specific and is apparently identical with CYP27, an

obligatory enzyme in bile acid biosynthesis (Andersson et al., 1989). The 25-hydroxylase

is less regulated compared to 1α-hydroxylase and 24-hydroxylase enzyme.

1.7.2 Kidney 1α-hydroxylase enzyme

1.7.2.1 Role of 1α-hydroxylase.

The 1α-hydroxylase is a member of the cytochrome P450 family and is a heme- containing protein designated as CYP27B1 by the Cytochrome P450 gene superfamily

Nomenclature Committee. 1α-hydroxylase plays an important role in the synthesis of

1,25(OH)2D3. 1α-hydroxylase knockout mice has no 1,25(OH)2D3 in the circulation.

46 These mice develop hypocalcemia, secondary hyperparathyroidism, retarded growth,

skeletal abnormalities, and are defective in the reproductive system and immune function

(Panda et al., 2001; Dardenne et al.,2003). Examination of the bone structure reveals alterations to non-collagenous matrix protein expression, reduced numbers of osteoclasts and the presence of rachitic bone lesions. The defects in these 1α-hydroxylase knockout

mice are similar to patients with the genetic disorder, vitamin D-dependent rickets type 1

(Panda et al., 2001).

1α-hydroxylase can catalyze both the C-1 hydroxylation of 25(OH)D3 and

24,25(OH)2D3. 24,25(OH)2D3 is the preferred substrate for 1α-hydroxylase when

compared to 25(OH)D3. However, because of the higher concentration of 25(OH)D3 in

normal physiological conditions (approximately 10 times higher than 24,25(OH)2D3),

more 1,25(OH)2D3 is synthesized than 1,24,25(OH)3D3 (Omdahl et al., 2002).

Subcellular analysis of enzyme activity indicates that the 1α-hydroxylase enzyme is

located in the mitochondria rather than in the microsomal fraction (Henry, 1997). The

biochemical reduction catalysed by 1α-hydroxylase is dependent on NADPH as a co-

factor, as well as NADPH-ferrodoxin reductase and ferrodoxin (Zehnder and

Hewison, 1999).

The process of hydroxylating 25(OH)D3 to 1,25(OH)2D3 occurs mainly in the kidney

which is the main organ for maintaining circulatory levels of 1,25(OH)2D3. 1α-

hydroxylase enzymes are express throughout the kidney such as the proximal tubules,

distal tubules, collecting ducts, cortical and medullary renal tissue (Figure 1.6). It is

47 thought that the production of 1,25(OH)2D3 in the proximal tubules function to support circulating levels of 1,25(OH)2D3, whereas more distal parts of the nephron fulfill an autocrine or paracrine function, i.e. to regulate calcium reabsorption via the calcium- sensing receptor (Hewison et al., 2000).

Extrarenal 1α -hydroxylase is found in a wide variety of tissue. Immunohistochemical staining for 1α-hydroxylase reveals the presence of this enzyme in skin, bone cells, macrophages, colon, ganglia of mesenteric plexus, adrenal gland, pancreas, purkinje cells of the brain, prostate and placental tissue. It is suggested that this enzyme functions in an autocrine or paracrine fashion and may provide 1,25(OH)2D3 to localized area for cell growth and differentiation (Zehnder et al., (2001).

48

Bowman’s Distal Glomerulus capsule tubule Proximal tubule Distal tubule

Proximal tubule

Corte x

Medulla Descending Loop of Limb of loop Henle of Henle

Collecting

duct Ascending Limb of loop of Henle

To renal pelvis

Figure 1.6 Diagram of Nephron

The 1α-hydroxylase enzymes are express in the proximal tubules, distal tubules,

collecting ducts of the cortical and medullary renal tissue.

49 1.7.2.2 Regulation of the 1α-hydroxylase

The renal 1α-hydroxylase is tightly regulated by a wide range of substances such as circulating parathyroid hormone (PTH), insulin-like growth factor I, calcium, phosphorus and 1,25(OH)2D3 (Portale and Miller, 2000; Brenza and DeLuca, 2000; Zoidis et al.,

2002; Zhang et al., 2002). Two of the main regulators of 1α-hydroxylase are PTH and

1,25(OH)2D3. Studies done on avian and mammalian species showed that parathyroidectomy diminishes, whereas PTH administration increases, circulating levels of 1,25(OH)2D3 (Henry, 1997). Studies by Kong et al.,(1999) identified the presence of

PTH responsive regions in the promoter of the 1α-hydroxylase gene and further demonstrated that PTH could stimulate the 1α-hydroxylase promoter activity in a kidney cell line.

In vivo studies using VDR knockout mice showed that mice lacking the VDR develop abnormally high levels of 1-hydroxylase expression and serum concentrations of

1,25(OH)2D3, indicating that 1,25(OH)2D3 inhibits 1α-hydroxylase activity (Takeyama et al., 1997; Murayama et al., 1999). Besides this, 1,25(OH)2D3 was also shown to suppress the 1α-hydroxylase promoter activity in opossum kidney cells (Armbrecht et al.,

2003). The above studies show that PTH has a stimulatory effect on 1α-hydroxylase whereas the effect of 1,25(OH)2D3 is inhibitory. In the physiological system both PTH and 1,25(OH)2D3 work side by side to regulate calcium homeostasis. Low calcium and

1,25(OH)2D3 stimulate the release of PTH which in turn stimulates the production of 1α-

50 hydroxylase. As sufficient 1,25(OH)2D3 is produced, 1,25(OH)2D3 act as a negative

feedback to inhibit the expression of 1α-hydroxylase.

1.7.3 24-hydroxylase enzymes (CYP24)

While the above two enzymes are involved in the bioactivation of vitamin D, CYP24

catalyses the bioinactivation of vitamin D metabolites. CYP24 is mainly found in the

kidney. The major site for renal 24-hydroxylase activities is found in the proximal

convoluted tubule (Iida et al., 1993). However, CYP24 is also found in nearly all cells, or

cells with VDR such as in the immune cells, bone, intestines, skin, brain, muscle and

reproductive organs (Henry and Norman, 1984). The widespread distribution of CYP24

suggests that 24-hydroxylase enzyme play a crucial role in regulating circulating and

local concentration of 25(OH)D3 and 1,25(OH)2D3 by metabolizing them to inactive

forms. CYP24 plays an important role in preventing hypervitaminosis. There are also

findings indicating that the by-product of CYP24, 24,25(OH)2D3 is essential for bone fracture healing (Seo et al., 1997)

In vitro studies on Escherichia coli expressed human and rat CYP24 have demonstrated that human CYP24 was able to catalyze both C-23 and C-24 hydroxylation whereas rat

CYP24 can only catalyze C-24 hydroxylation (Sakaki et al., 2000). Human CYP24 not only catalyzes the 24R-hydroxylation of 25(OH)D3 and 1,25(OH)2D3, but also the subsequent oxidation of the hydroxyl group to a ketone group, (hydroxylation at the C-23 or C-24 position and cleavage of the C23-C24 bond). Studies by Taniguchi et al., (2001)

51 have demonstrated that 1,25(OH)2D3 is the preferred substrate for CYP24. However, a

recent study by Masuda et al., (2004) showed that both 1,25(OH)2D3 and 25(OH)D3 are metabolized at a similar rate by the CYP24 enzyme.

CYP24 knockout mice were found to have high circulating levels of 1,25(OH)2D3

(Masuda et al., 2004 and 2005). These mice also have impaired intramembranous bone formation. Since CYP24 converts 25(OH)D3 to 24,25(OH)2D3, it was suggested that the deficiency in 24,25(OH)2D3 may have contributed to the defect in bone growth and repair

(St-Arnaud, 1999). However, later experiments to rescue the bone phenotype, by cross breeding CYP24 knockout mice with VDR knockout mice revealed that the impaired bone formation in CYP24 knockout mice was not the result of 24,25(OH)2D3 deficiency but because of high levels of 1,25(OH)2D3 (St-Arnaud et al., 2000), acting via as yet uncharacterized mechanisms.

In vivo studies using transgenic rats that constitutively express CYP24 showed a significant reduction in circulating 25(OH)D3 and 24,25(OH)2D3 levels in these rats.

These transgenic animals also developed albuminuria and hyperlipidemia shortly after

weaning (Kasuga et al., 2002; Hosogane et al., 2003). It was suggested that the reduced

24,25(OH)2D3 levels were due to an excessive amount of albumin being excreted from

the glomeruli because of nephritis, therefore causing a decrease in the reuptake of

25(OH)D3 from the proximal tubule of the kidney. The reason for how constitutive

CYP24 expression induces nephritis in rats remains to be elucidated (Hosogane et al.,

2003).

52 1.7.3.1 Regulation of the CYP24 promoter by vitamin D

Of all the compounds that have been found to stimulate CYP24 transcription,

1,25(OH)2D3 is the most studied. The promoter of the CYP24 gene is known as the most transcriptionally responsive vitamin D-inducible gene identified to date. Sequencing of the human CYP24 promoter has mapped out two functional vitamin D response elements

(VDREs) located about 100 bp apart within the –300-bp promoter region (see Figure 1.7)

(Chen and DeLuca 1995; Carlberg, and Polly 1998) whereas, the rat CYP24 promoter contains three VDREs within its promoter region (Zierold et al., 1995). Studies in our laboratory suggest that the proximal VDRE (VDRE-1) is crucial for 1,25(OH)2D3 activation of the CYP24 promoter whereas distal VDRE (VDRE-2) is recruited if further activation is required. For example, in the presence of high concentrations of

1,25(OH)2D3, both VDREs are utilized in a synergistic manner to ensure high levels of

CYP24 expression (Kerry et al., 1996). Activation of the CYP24 promoter by

1,25(OH)2D3 requires the binding of 1,25(OH)2D3 to the VDR, as shown in VDR knock out mice (Takeyama et al., 1997). VDR interacts with RXR to form VDR-RXR heterodimer that then binds to the VDRE in the CYP24 promoter (Zou et al., 1997). In the absence of 1,25(OH)2D3, the heterodimer VDR/RXR binds to one or both VDREs to recruit a corepressor and lower basal expression (Dwivedi et al., 1998; Polly et al., 2000).

Upon 1,25(OH)2D3 binding to the VDR, the corepressor is replaced by a coactivator complex.

53 -2 9 8 G Illus F i C g -2 , u CG CG 58

t a r r n a e CA CA ted a d

1 C C

C . C C VD 7 C C C

G G RE- r A CTGA CTGA e the sequ A 2 D T A A i

a C C s -2 C C g i t 4 r e 4 a

.

m e

G G n m C C c -1 G G e

a 71 T T

G G t a i T T n c VS C C d

G G r

E G G r e e T T p C C l r a -1 A A e t C C 63 i s C C v e G G e n

l t

o a c t i a v -1 t i e 50 CG CG o

o n CCC CCC f s

VD t o

h T T f C C R e

t A A E

h p C C - 1 e T T r

C C o t A A r x a C C -1 i n CT CT m 36 s c a r l i

p r t e -1 i g T T o 2 8 C C i n o CA CA a

n EB l T T

i o C C S n C C f i

T T t t CT CT -1 i h a 1 e t 9 i

o w

n -11 i ACACC

l s d 4 i

t t e

y ,

p C V e G GC D

CCC

p R C E CC Y 1 P CG , -11

2 V 0 4 S - E W -6 , GCTC

T 2 V CCA ( D 2 A 9 R TTGG AT 8 E b

2 p , -5

C ) E

1 p B

r S o +1 , m

p o u t +7 t e a L 4 r t u i . cife v

e ras

e

54 Besides this, recent studies in our laboratory on VDREs of the CYP24 promoter have

identified an E26 (Ets)-1 binding site (EBS) (-128/-119) located downstream of VDRE-1

(-136/-150) that is important for maximal transcriptional activation (see Figure 1.7)

(Dwivedi et al., 2000). Over expressed Ets-1 enhanced 1,25(OH)2D3 induction of

CYP24 expression by 34% whereas, over expression of a mutant Ets-1 (Ets-1 T38A)

resulted in a substantially lowered (42% reduction) CYP24 promoter induction. This

level of reduction is comparable to that observed when the Ets-1 binding site was mutated

(Dwivedi et al., 2002). In addition to the EBS, further analysis by our laboratory of the

CYP24 promoter sequence using a computer program MatInspector (V2.2)

(http://transfac.gbf.de/), has identified a putative GC box (-114/-101) and a putative

CCAAT box (-62/-51) located downstream from the EBS which plays an important role in CYP24 promoter induction (see Figure 1.7) (Dwivedi, Tan, Ferrante, Morris, May, and

Hii, manuscript in preparation). Mutation to GC box reduced basal and 1,25(OH)2D3 induction of the CYP24 promoter by approximately 50% in HEK293T and UMR106 bone cells. Mutation to CCAAT box also reduced the basal and 1,25(OH)2D3 induction of the CYP24 promoter in HEK293T cells by 66% and 75%, respectively (Dwivedi, Tan,

Ferrante, Morris, May, and Hii, manuscript in preparation). Mutation to both GC and

CCAAT box further reduced basal promoter activity and 1,25(OH)2D3 induction of the

CYP24 promoter, (75% and 90% reduction, respectively). In these studies, the transcription factor, Sp1, was shown to interact with the GC box and over expression of

Sp1 increased CYP24 promoter induction, which was then abrogated when the GC box was mutated. Transcription factor nuclear factor Y (NF-Y) was also shown to interact with CCAAT box while over expression of NF-Y increases CYP24 promoter activity that

55 was abrogated when CCAAT box was mutated (Dwivedi, Tan, Ferrante, Morris, May, and Hii, manuscript in preparation). Dwivedi, Tan, Ferrante, Morris, May, and Hii

(manuscript in preparation) using mammalian two-hybrid system showed that Sp1 interacts with Ets-1 efficiently and may play a role in transcriptional cooperation. Studies in our laboratory by Nutchey et al., (2005) identified a vitamin D stimulatory element

(VSE), a sequence of 5’-TGTCGGTCA-3’ which is located about 30bp upstream of

VDRE-1 at –171/-163 in the CYP24 proximal promoter (see Figure 1.7). Mutation to this

VSE site reduced 1,25(OH)2D3 induction of the CYP24 promoter by 60% and completely abolished the synergistic response to PMA and 1,25(OH)2D3 (Nutchey et al.,

2005). This finding shows that the VSE site is not only crucial for 1,25(OH)2D3 induction of the CYP24 promoter but is also required for the synergistic response to PMA and 1,25(OH)2D3.

1.7.3.2 Signaling pathways that regulate the CYP24 promoter in response to vitamin D.

As described earlier (see section 1.7.3.1), the CYP24 promoter contains the VDREs,

EBS, GC box, CCAAT box and VSE site. The stimulation of the CYP24 promoter by

1,25(OH)2D3 involves both genomic and the non-genomic action. The genomic action involves direct interaction of 1,25(OH)2D3 with nuclear VDR which in turn acts on the

VDRE to initiate transcriptional activation of the CYP24 promoter. The non-genomic action of 1,25(OH)2D3 involves the binding of 1,25(OH)2D3 to a putative membrane

VDR which results in the activation of various signaling molecules such as MAP kinases,

PKC and PI3K. These signaling molecules in turn phosphorylate a number of

56 transcription factors (e.g RXRα, Ets-1, Sp1 and NF-Y proteins) which play a crucial role in the CYP24 promoter induction (Dwivedi et al., 2002; Gao et al., 2004). 1,25(OH)2D3 is known to activate different signaling molecules in different cell types (see section 1.5).

Studies from our laboratory and those of others have shown that the MAP Kinases play an important role in the induction of the CYP24 promoter (Dwivedi et al., 2000; Dwivedi et al., 2002; Nutchey et al., 2005). 1,25(OH)2D3 has been shown to activate

ERK1/ERK2 and ERK5 and these MAP kinases play an important role in the induction of the CYP24 promoter in COS-1 cells (Dwivedi et al., 2002; Pearson et al., 2001).

Further more, the dominant negative Ras mutant has been shown to inhibit 1,25(OH)2D3 induction of the CYP24 promoter in COS-1 cells (Dwivedi et al., 2000; Dwivedi et al.,

2002) as well as the activation of down stream ERK1/ERK2 and ERK5 (Dwivedi et al.,

2002). However, the signaling molecules needed to induce the CYP24 promoter can vary between different cell types. For example, a study by Nutchey et al. (2005) showed that in HEK293T cells, the dominant negative ERK1 mutant did not inhibit the CYP24 promoter induction by 1,25(OH)2D3, whereas the dominant negative MKK4 (an upstream activator of JNK) mutant caused a significant reduction in the ability of

1,25(OH)2D3 to stimulate CYP24 promoter activity. Consistent with this, Nutchey et al.

(2005) demonstrated that 1,25(OH)2D3 could stimulate JNK activity and not

ERK1/ERK2 in HEK293T cells. These results seem to differ from the results obtained by

Dwivedi et al. (2002) who showed that in COS-1 cells, 1,25(OH)2D3 could stimulate

ERK1/ERK2 activity and that these kinases were required for 1,25(OH)2D3 induction of the CYP24 promoter. Dwivedi et al. (2002) also found that JNK did not play any role in the CYP24 promoter induction by 1,25(OH)2D3 in COS-1 cells. Differences in the

57 requirement for the type of different cell signaling molecules for CYP24 promoter induction can also be observed in different bone cell lines. Thus, overexpression of Raf-1

(an upstream activator of ERK) led to the stimulation of CYP24 mRNA induction by

1,25(OH)2D3 in MG-63 bone cells whereas overexpression of Raf-1 in MC3T3-E1 bone cells resulted in suppression of CYP24 mRNA (Narayanan et al., 2004). Furthermore,

CYP24 mRNA expression was enhanced when MC3T3-E1 cells were treated with a

MEK inhibitor (U0126) (Narayanan et al., 2004). These findings suggest that ERK can have different effects on CYP24 expression depending on the cell-type involved

(stimulation in MG-63 cells or inhibitory in MC3T3-E1). Narayanan et al. (2004) suggested that this differences in the regulation of the CYP24 promoter induction by

ERK between the different cell-types is dependent on the type of RXR expressed in the cells. There are three different isoforms of RXR: RXRα, RXRβ and RXR γ. MC3T3-E1 cells have low levels of RXRβ and RXRγ whereas MG-63 cells have low levels of

RXRα. These workers suggested that phosphorylation of RXRα by ERK resulted in the

inability of RXRα to heterodimerise to VDR to form VDR/RXR heterodimer and

therefore prevent the transcriptional activation of the CYP24 (Narayanan et al., 2004).

Besides the MAP kinases, PKC has also been shown to play an important role in the

induction of the CYP24 promoter by 1,25(OH)2D3 (Partridge et al., 1998; Armbrecht et

al., 2001). PMA a direct activator of PKC, was shown to enhance the 1,25(OH)2D3

induction of CYP24 mRNA expression in rat intestinal epithelial cells (Koyama et al.,

1994; Armbrecht et al., 2001). In the presence of 1,25(OH)2D3, PMA was shown to

synergistically increase CYP24 promoter activity in kidney cells (Barletta et al., 2003;

58 Chen et al., 1993). In addition, a recent study by Nutchey et al. (2005) show that PKC is also required for the synergistic increase in CYP24 promoter activity caused by PMA and

1,25(OH)2D3 in HEK293T cells. These workers also investigated the roles played by the different isoforms of PKC (PKCα, PKCβ1, PKCδ, and PKCε) on CYP24 induction by

PMA and 1,25(OH)2D3. Results from these studies demonstrate that PKCα plays an important role in the synergistic induction of the CYP24 promoter by PMA. In contrast,

PKCβ1 was found to be required for 1,25(OH)2D3 to stimulate CYP24 promoter activity

(Nutchey et al., 2005). Thus, there is overwhelming evidence that show that PKC, ERK and JNK are involve in stimulating CYP24 induction and the signaling molecules that are required depend on the cell-type being studied.

1.8 Regulation of CYP24 by compounds other than vitaminD.

Besides 1,25(OH)2D3, the CYP24 promoter has been shown to be regulated by a variety of substances including parathyroid hormone (PTH), retinoid X receptor (RXR) ligand

(LG100268), genistein, forskolin and PMA (Armbrecht et al., 2003; Allegretto et al.,

1995; Kallay et al., 2002; Christakos et al., 1996; Barletta et al., 2003). PTH was shown to decrease both CYP24 mRNA and enzyme activity in kidney cells by reducing the stability of the CYP24 mRNA (Zierold et al., 2003). The effect of PTH on CYP24 in bone cells is different from kidney cells. PTH by itself has no effect on the CYP24 levels in bone cells but acts synergistically with 1,25(OH)2D3 to increase the CYP24 mRNA and protein (Armbrecht et al., 1998). The synergistic increase in CYP24 mRNA caused by 1,25(OH)2D3 and PTH was shown to be mediated through the cAMP signaling

59 pathway (Armbrecht et al., 1998). Consistent with the role of cAMP, a study using HL60

cells has demonstrated that the activation of cAMP by forskolin resulted in an increase in the CYP24 enzyme activity (Smith et al., 1999). In contrast, forskolin was shown to inhibit CYP24 enzyme activity in a dose dependent manner in mouse renal tubule

(Mandla and Tenenhouse, 1992). The basis for this difference in action for forskolin is not known.

Genistein, a plant derived isoflavone has been shown to reduce CYP24 mRNA in tumor

cells by down regulating CYP24 promoter induction (Farhan et al., 2003). Currently, the

mechanism for the reduction in CYP24 promoter induction by genistein is unknown.

The RXR ligand (LG100268) was shown to increase CYP24 promoter activity by binding

to RXR which in turn act through the VDRE within the CYP24 promoter to stimulate its

induction (Zou et al., 1997).

There has been quite a number of studies that investigated the action of PMA on CYP24

induction, either in the presence or absence of 1,25(OH)2D3. PMA was shown to

enhance 1,25(OH)2D3 induction of CYP24 promoter activity in COS-7 and LLCPK-1

cells by acting through PKC and MAP kinase to increase VDR expression and

phosphorylation (Barletta et al., 2004). However, there are also studies that suggest that

the increase in CYP24 induction by PMA is independent of VDR expression but through

a direct action on the CYP24 promoter (Koyama et al., 1994). Similarly, PMA was also

shown to increase CYP24 enzyme activity and in the presence of 1,25(OH)2D3,

synergistically increase CYP24 promoter activity in kidney cells. PKC, ERK and JNK

60 have all been shown to play a role in up-regulating the expression of CYP24 (Mandla et

al., 1990; Nutchey et al., 2005)

1.9 Chemotherapy and plasma vitamin D

Cancer is the most common cause of disease-related death in children in industrialized

societies (Parker et. al., 1997). With the improvement in the of anticancer drugs,

surgical techniques, radiation therapy and the development of intricate imaging

technologies for earlier detection and monitoring of the disease, the survival rate of

childhood cancer patients has approached 60-70% (Green, 2002; Parisi et al., 1999,

Meadows et al., 1986 and Dreyer et al., 2002) and is projected to reach 85% by the year

2010 (Harras, 1996). With the improved survival rate, therapy-associated complications have become more evident and thus have become a major management challenge. These patients can develop skeletal defects, reduced bone mineral mass and density, osteoporosis and fractures during and at the end of the treatment period. Many of these skeletal defects, such as scoliosis, craniofacial dysplasia, and limb-length discrepancy, are readily apparent. However, other side effects such as osteoporosis and osteonecrosis are silent until they reach advanced stages where they may compromise quality and duration of survival and be more difficult to ameliorate (Kaste, 2004). The basis for the reduction in bone mass and density are likely to be multi-factorial. Cancer treatments such as corticosteroids, antineoplastic drugs and radiotherapy have been reported to reduce bone mass and density (Abrams et al., 2004 and van Leeuwen et al., 2000)

Besides this, endocrinopathies (e.g growth hormone deficiency and hypogonadism),

nutritional deficits and relatively sedentary lifestyle during and even after therapy can

61 interfere with bone-mineral accretion and skeletal development (Tillmann et al., 2002,

Arikoski et al., 1999, Warner et al., 1998). Another factor could involve a disturbance in the metabolism of 1,25(OH)2D3 which is an essential regulator of calcium homeostasis

(Glorieux, 1990).

Studies by Arikoski and colleagues have demonstrated that children undergoing chemotherapy for leukemia as well as solid tumours had reduced plasma levels of

25(OH)D3 and 1,25(OH)2D3 and that the vitamin D metabolites remain low even after the completion of chemotherapy (Arikoski et al., 1999a, 1999b). Halton and colleagues reported that more than 70% of paediatric patients with acute lymphoblastic leukemia showed reduced levels of plasma 1,25(OH)2D3 following chemotherapy (Halton et al.,

1996; Atkinson et al., 1998). Although most of the reports for the reduction of serum levels of 1,25(OH)2D3 were found in children, there is some evidence that chemotherapy can also reduce plasma 1,25(OH)2D3 in adults patients with gynaecological malignancies

(Gao et al., 1993). Given the importance of 1,25(OH)2D3 in the maintenance of calcium levels and bone density, chemotherapy-induced disturbance in 1,25(OH)2D3 metabolism could be a major contributing factor to the reduction in bone mass during and after treatment.

The mechanism for the reduction in 1,25(OH)2D3 is not known. Halton et al., (1996) speculated that the reduction in 1,25(OH)2D3 was due to receptor-mediated binding of

1,25(OH)2D3 to leukemic cells. Other possible mechanisms suggested by Halton and colleagues included the down-regulation of 1α-hydroxylase which is required for the

62 biosynthesis of 1,25(OH)2D3 (Halton et al., 1996) or up-regulation of metabolism to inactive products. Arikoski et al., (1999b) suggested that the reduction in 1,25(OH)2D3 was due to factors such as malnutrition and diminished sunlight exposure, vitamin D malabsorption due to intestinal mucositis and deficiency in 25(OH)2D3. This is because

Arikoski et al., (1999a and b) found that the levels of plasma 25(OH)D3 and

1,25(OH)2D3 were reduced in patients with leukemia as well as with solid tumor even at completion of their chemotherapy treatment. Other proposed mechanisms for the reduction in serum 1,25(OH)2D3 during chemotherapy include impairment of mitochondrial function in the proximal tubules of the kidneys by chemotherapeutic drugs such as cisplatin (Gao et al., 1993). However, none of these possibilities have been proven and identifying the causes remains a major challenge.

1.10 Chemotherapeutic drugs

The realization that a large number of children who are undergoing or have completed chemotherapy have reduced levels of 25(OH)D3 and 1,25(OH)2D3 metabolites, suggests to us that chemotherapeutic drugs may have a direct effect on the metabolism of vitamin

D. The following sections briefly review the major classes of drugs and examples, their effects and associated mechanisms of action, including the signaling molecules which they have been reported to stimulate.

63 1.10.1 Categories of chemotherapeutic drugs

Alkylating agents, include mustargen-nitrogen mustard, cyclophosphamide (cytoxan), melphalan (alkeran), chlorambucil (leukeran), cis-platinum ("non-classical" alkylating agent), carbo-platinum ("non-classical" alkylating agent), carmustine (BCNU), thiotepa and busulfan (myleran).

Vinca alkaloids and related substances, include vincristine, vinblastine and VP-16.

Anthracycline antibiotics, include doxorubicin (adriamycin), actinomycin D, daunorubicin (daunomycin), bleomycin, idarubicin and mitoxantrone.

Glucocorticoids, include prednisone/prednisolone, and triamcinolone

(vetalog).

Inhibitors of protein/DNA/RNA synthesis, include Methotrexate, 6-thioguanine, 5- fluorouracil (5-FU), cytosine arabinoside (ara-C, cytosar), L-asparaginase (Elspar), dacarbazine (DTIC), hydroxyurea (hydrea) and procarbazine (matulane).

Miscellaneous agents, includes paclitaxel.

64 1.10.2 Principal targets of Chemotherapeutic drugs: the cell cycle and

apoptosis.

The ultimate aims of anti-cancer therapy are to disrupt cancer cell cycling and kill the

cancer cells. Cancer cells have the advantage over normal cells of disregulated growth

due, predominantly, to their resistance to the normal regulation of apoptosis

(programmed cell death in which the cell commits suicide by a highly regulated

biochemical process). Consequently, anti-cancer drugs are designed to target the cell

cycle and to increase the sensitivity of cancer cells to death inducers and apoptotic

effectors. The cell cycle can be divided into: Mitosis (M) phase (cell division phase,

duration 0.5-1 hr), growth 1 (G1) postmitotic phase (duration depending on G0. A phase

for RNA and protein synthesis in preparation for S phase), growth zero (G0) phase

(variable duration time and is temporary or permanent if the cell has reached the end of

its cycling capability), DNA synthesis (S) phase (duration 10-20 hrs, DNA replication occurs during this phase), growth 2 (G2) premitotic phase (RNA and protein synthesis).

The anti-cancer drugs can have actions either on a specific phase (e.g vincristine which has a specificity on the M phase but can work in S or G1 phase at high concentrations) or be non-specific (e.g. daunorubicin and doxorubicin, but most effective in S phase).

Apoptosis comes from the Greek words apo = from and ptosis = falling. This is a key

process in the body which responds to cell damage, infection, stress or DNA damage.

Apoptosis is required for homeostasis, development and immune cell regulation. It can

occur, for instance, when a cell is damaged beyond repair, or infected with a virus. The

65 "decision" for apoptosis can come from the cell itself, from its surrounding tissue or from

a cell that is part of the immune system such as the monocyte/macrophage-derived cytokine, tumour necrosis factor α, and CD8+ cytototic T lymphocytes. If the ability of the cell to undergo apoptosis is impaired (e.g. by mutation), or if the initiation of apoptosis is blocked (by a virus), the damaged cell can continue to divide without restrictions and develops into cancer. For example, as part of the hijacking of the cell's

genetic system carried out by human papillomaviruses (HPV), the E6 gene is expressed and this protein causes the degradation of p53 protein, a vital protein which regulates the

decision for apoptosis. This interference of the apoptotic capability of cells is believed to be a major cause for the development of cervical cancer in individuals with persistent oncogenic HPV infection. Owing to its crucial role in cancer development and progression, many of the above listed anti-cancer drugs have also been developed due to their ability to not only block the cell cycle but also to cause apoptosis or sensitize cancer cells to apoptosis-inducing agents, some of which will be described in greater detail below.

1.10.3 Actions of anticancer drugs

For this discussion, we will focus on four commonly used and better characterized anti-

cancer drugs which we have selected from the above-listed classes of drugs. These four

drugs are also the subject of investigation in this thesis.

66 1.10.3.1 Daunorubicin hydrochloride

Daunorubicin hydrochloride is a member of the anthracycline family of antibiotics that is

produced from a strain of Streptomyces coeruleorubidus. Its molecular structure is shown

in Figure 1.8 A. It is one of the major anti-tumour agents which are widely used in the

treatment of acute myeloid leukemias and acute lymphocytic leukemia (Tallman et al.,

2005). The drug can significantly increase the rate of complete remission compared to

others such as vincristine, prednisone, and L-asparaginase alone.

Daunorubicin hydrochloride is rapidly taken up by different tissues. Typically, it has been

found that the highest concentrations of daunorubicin hydrochloride are found in the

spleen, kidney, liver, lung and heart. Daunorubicin hydrochloride and its hydroxylated

derivative, doxorubicin hydrochloride, are both anti-mitotic and cytotoxic by virtue of

their ability to cause DNA damage by intercalating between base pairs and to generate

free radicals which can cause DNA damage (Minotti et al., 2004; Gervasoni et al., 2004).

The drugs also inhibit topoisomerase II activity by stabilizing the DNA topoisomerase II

complex (Minotti et al., 2004). Topoisomerase II is a ubiquitous enzyme which is

essential for the survival of all eukaryotic organisms and plays critical roles in virtually

every aspect of DNA metabolism. The enzyme unknots and untangles DNA by passing

an intact helix through a transient double-stranded break that it generates in a separate

helix. Owing to its central role in DNA replication, topoisomerase II is the target for some of the most active and widely prescribed anticancer drugs currently utilized for the treatment of human cancers (Wang, 2002). The list of potential toxic effects includes

67 A: Daunorubicin Hydrochloride

B: Etoposide

C: Vincristine Sulfate

D: Cisplatin

Figure 1.8 Molecular structures of 4 selected chemotherapeutics drugs

68 persistent and severe myelosuppression, and cardiac toxicity which can lead to heart

failure, fetal abnormality, secondary leukemias and tissue necrosis.

Daunorubicin hydrochloride can activate a number of cell signaling molecules, including

MAP Kinases, PI3K and PKC which regulate cell survival and apoptosis (Mas et al.,

2003). Indeed, numerous reports have demonstrated that this drug can trigger apoptosis of

cancer cells in vitro (Masquelier et al., 2004; Gervasoni et al., 2004; Dimanche-Boitrel et al., 2005). This will be discussed further below.

1.10.3.2 Etoposide

Etoposide (also known as VP-16) is a semisynthetic derivative of podophyllotoxin. The

molecular structure of etoposide is shown in Figure 1.8 B. Etoposide is used to treat a

variety of malignancies and shows considerable activity in treating leukemias,

neuroblastomas and many other malignancies such as refractory testicular tumors and

small cell lung cancer (Frappaz et al., 1992; Hijiya et al., 2004; Tsuchiya et al., 2005).

The cytotoxicity of etoposide is due, in part, to its ability to inhibit topoisomerase II,

resulting in the inhibition of DNA synthesis. Etoposide can also activate apoptosis

signaling via activation of the MAP kinases, PKC and p53 signalling molecules (Shimada

et al., 2003; DeVries et al., 2002). Toxic side effects of etoposide include

myelosuppression, hypotension, fetal malformation, gastrointestinal toxicities and

alopecia. Toxicity studies in non-cancer cells such as hepatocytes, seem to show that non-

cancer cells are less sensitive to etoposide than cancer cells such as HL-60 and liver

69 cancer cells. For example, Zhuo et al. (2004) showed that 40 µM of etoposide was not

cytotoxic to normal human hepatocytes. Furthermore, a study using normal rat

hepatocytes showed that etoposide at concentrations below 176 µM did not induce

apoptosis (Gomez-Lechon et al., 2002). The concentration of etoposide used in these

studies were higher than those needed to kill cancer cells such as human hepatocellular

carcinoma cell line (less than 17 µM of etoposide) (Miao et al., 2003) and HL60 (30µM

of etoposide) (Bjorling-Poulsen and Issinger, 2003). Thus, like most anti-cancer drugs,

etoposide shows selectivity towards cancer cells.

1.10.3.3 Vincristine Sulfate

Vincristine Sulfate is derived from a common flowering plant, the periwinkle plant. The

molecular structure of vincristine sulfate is shown in Figure 1.8 C. The mechanism of

anti-cancer action of vincristine sulfate is via the inhibition of microtubule formation in

the mitotic spindle. In addition, vincristine sulfate can promote apoptosis by regulating

different signaling molecules such as ras, Raf, MAP Kinases, PKA and PKC (Wang et

al., 1999; Stone and Chambers, 2000). Furthermore, the PI3K inhibitor, LY294002, has been demonstrated to synergistically enhance the cytotoxicity of vincristine sulfate towards human malignant glioma cell lines, suggesting that the PI3K pathway promotes resistance towards vincristine sulfate (Shingu et al., 2003). Indeed, some types of cancers have developed resistance to apoptosis by upregulating their expression of Acute

Transforming (Akt)/protein kinase B (PKB) which is an effector of PI3K (Siddik, 2003;

Cenni et al., 2004).

70 Vincristine sulfate is normally used together with other chemotherapeutic drugs to treat acute leukemia, Hodgkin’s disease, non-Hodgkin’s malignant lymphomas, rhabdomyosarcoma, neuroblastoma and Wilms’ tumor (Zoubek et al., 1994; Zinzani et al., 2002; Bisogno et al., 2005; Mori et al., 2005). Toxic side effects of vincristine sulfate include acute uric acid nephropathy, acute shortness of breath and severe bronchospasm, impairment of fertility, hair loss, hypertension or hypotension, sensory impairment, neuritic pain and motor difficulties.

1.10.3.4 Cisplatin

Cisplatin is a heavy metal complex containing a central atom of platinum surrounded by two chloride atoms and two ammonia molecules in the cis position (see Figure 1.8 D for molecular structure). The anti-cancer effect of cisplatin is due to the drug reacting with the DNA, causing DNA damage. Cisplatin reacts with DNA to produce mostly intra- strand crosslinks between adjacent guanine base pairs or adjacent guanine and adenine base pairs (Nowosielska and Marinus 2005). This formation of intra-strand crosslinks between DNA blocks the action of DNA polymerases, resulting in the inhibition of DNA synthesis (Pinto and Lippard, 1985). Cisplatin can also promote the formation of crosslinks between DNA and proteins, thus shielding the cisplatin-modified DNA from being repaired. The damaged DNA triggers cell-cycle arrest and apoptosis (Jordan and

Carmo-Fonseca, 2000). Cisplatin can also activate signaling molecules such as the MAP kinases and PKC to initiate apoptosis (Wang et al., 2004; Sedletska et al., 2005; Iioka et al.,2005).

71

Cisplatin has been used together with other approved chemotherapeutic drugs to treat a variety of cancers, including metastatic testicular tumor, metastatic ovarian tumor and advanced bladder cancer (Einhorn, 2002; Rosenberg et al., 2005). Toxic side effects of cisplatin include low levels of white blood cell counts which can lead to increased risk of infection, low levels of red blood cell (anemia), low platelet levels which can result in increased risk of bleeding, nausea and vomiting, kidney damage, hearing loss and infertility.

1.11 Activation of cell signaling molecules by chemotherapeutic drugs.

A number of studies have shown that chemotherapeutic drugs such as those mentioned above can activate PKC and MAP Kinases, ERK1/ERK2 and JNK (Nowak, 2002; Boldt et al., 2002; Hershberger et al., 2002; Hayakawa et al., 2003; Mas et al., 2003). For example, Nowak (2002) found that 50 µM of cisplatin caused the activation PKCα in renal proximal tubular cells. A detectable increase in kinase was observed at 30 min and kinase activation peaked at around 1 h. Cisplatin also stimulates the activity of

ERK1/ERK2 MAP kinases (Nowak, 2002). Increases in ERK1/ERK2 activity could be observed at 30 min after the addition of the drug and kinase activity peaked at around

6hrs and this was maintained for up to 24hrs (Nowak, 2002). The effect of cisplatin on

ERK1/ERK2 activity was shown to be PKC-dependent since the PKC inhibitor, Go 6976, inhibited the stimulation of ERK1/ERK2 activity by cisplatin.

72 Etoposide, as mentioned above, has also been reported to stimulate the activity of

ERK1/ERK2. Thus, Tang et al. (2002) reported that etoposide (100µM) caused the activation of ERK1/ERK2 in mouse embryonic fibroblasts. This effect was evident within 1 h of treatment and was maintain for at least 8 h. Studies by Boldt et al. (2002)

showed that the chemotherapeutic drugs, taxol and etoposide, could stimulate the

activities of ERK1/ERK2, JNK and/or p38 MAP kinases in three human cancer cell lines:

human squamous carcinoma (A431), human cervix carcinoma (HeLa) and human breast

epithelial carcinoma (MCF7). At 68 µM, etoposide was found to stimulate the activity of

ERK1/ERK2 in A431, HeLa and MCF7 cells at 3 h, 15 min and 3 h, respectively (Boldt et al., 2002). Increases in JNK activity following treatment with etoposide was observed at 3, 1 and 3 h in A431, HeLa and MCF7, respectively. These authors also investigated the effect of taxol on the activities of the MAP kinases. In A431 and HeLa cells, the actions of 50 nM taxol was found to be somewhat similar to those of etoposide, with the exception that in HeLa and MCF7 cells, stimulation of ERK1/ERK2 activity was biphasic. The second phase of activation occurred between 12-18 h after treatment. JNK activation in MCF7 cells the presence of taxol was biphasic, with the second phase commencing at 9-12 h. Etoposide also stimulated the activation of p38 in these cell lines, assessed by measuring MAPKAP-2 activity. Kinetics studies in HeLa and MCF7 cells demonstrate that etoposide increased p38 activity at 1 h after the commencement of treatment. In A431 cells, the increase occurred at 6 h (Boldt et al., 2002). In contrast, taxol did not have much of an effect on p38 activity in the three cell lines. Despite these increases, it was found that suppression of the drug-induced activation of the MAP kinases did not have a consistent effect of cancer cell survival or death. These results

73 prompted Boldt et al. (2002) to propose that the original view of ERK promoting cell

survival, and JNK and p38 furthering apoptosis is too simplistic, and in this form, cannot

be exploited as a general tool to enhance the effectiveness of chemotherapy. There is also

a functional dependency on which cancer cell-type and drug are being investigated.

Daunorubicin hydrochloride has been reported to stimulate the activities of ERK1/ERK2,

the MAP kinase kinase kinase, raf-1, and PI3K in U937 cells (Mas et al., 2003). Increases

in ERK1/ERK2 and raf-1 activities were detectable within 10-30 min of the addition of

the drug. Daunorubicin hydrochloride also stimulated the hydrolysis of

phosphatidylcholine (PC) by a PC-specific phospholipase C and the activation of PKCζ in these cells (Mas et al., 2003). The stimulatory effect of the drug on ERK1/ERK2 activity required PKCζ and PI3K since a kinase-dead mutant of PKCζ blocked the

activation of ERK1/ERK2 and cytoxicity. Wortmannin, an inhibitor of PI3K, also

blocked ERK1/ERK2 activation by daunorubicin hydrochloride (Mas et al., 2003). A role

for reactive oxygen species (ROS) in the activation of the MAP kinase and PI3K by

daunorubicin hydrochloride was also demonstrated. Thus, Mas et al. (2003) not only

showed that daunorubicin hydrochloride stimulated H2O2 generation but activation of

ERK1/ERK2 and PI3K was blocked by the antioxidant, N-acetyl cysteine. This is

consistent with ROS being generated by and involved in the actions of chemotherapeutic

agents (Sinha and Mimnaugh, 1990). Studies in human malignant testicular germ cells by

Schweyer et al. (2004a) have also shown that cisplatin can stimulate the production of

ROS which in turn activates Raf-1 and its downstream MEK and ERK signaling

pathway.

74 Currently, there is much interest in the roles MAP kinases play in regulating cell survival and apoptosis. The ERK signaling pathway is normally regarded as having a protective role during stress condition (Rosseland et al., 2005) and it promotes cell survival (Boldt et al., 2002). It is known that some tumour cell-types such as melanoma cells can constitutively exhibit high levels of ERK activity (Smalley, 2003). Consequently, there has been much interest in the use of selective inhibitors of the MAP kinase in conjunction with chemotherapeutic drugs to increase the efficacy of the anti-cancer drugs (Kohno and

Pouyssegur, 2003). Consistent with this view, the inhibition of MEK1/MEK2 or

ERK1/ERK2 has been shown to increase the sensitivity of tumour cells to chemotherapeutic drugs. At the same time, over expression of ERK1/ERK2 was shown to increase the survival of tumour cells when they are exposed to chemotherapeutic drugs

(Brantley-Finley et al., 2003; Mas et al., 2003). Studies by Brantly-Finley et al. (2003) showed that inhibition of the ERK1/ERK2 pathway by U0126 in KB-3 carcinoma cells strongly enhanced the cytotoxicity of vinblastine and doxorubicin but had no effect on etoposide-treated cells. Similarly, studies by Mas et al. (2003) showed that the MEK inhibitor, PD98059, increased the sensitivity of U937 cells to killing by daunorubicin hydrochloride. These studies support the view that ERK1/ERK2 plays an important role in mediating cell survival and promoting resistance against chemotherapeutic drugs.

In contrast, there are also studies which show that ERK1/ERK2 plays a role in initiating apoptosis of tumour cells and enhancing the cytotoxicity of chemotherapeutic drugs (Yeh et al., 2002; Tang et al., 2002; Schweyer et al., 2004b). Studies by Schweyer et al.

(2004b) in human malignant germ cells have demonstrated that cisplatin mediates

75 apoptosis by stimulating MEK and its downstream ERK1/ERK2 to activate caspase-3

(executing caspase). Consistent with this observation, Nowak (2002) using MEK and

PKC inhibitors have also shown that PKCα and ERK1/ERK2 play a role in mediating

cisplatin-induced renal cell apoptosis by causing a sustained decrease in mitochondrial

function and by inhibiting Na+/K+-ATPase activity. Tang et al. (2002) reported that

ERK1/ERK2 could play a role in promoting NIH3T3 cell apoptosis. These authors showed that etoposide and adriamycin not only stimulated the activity of ERK1/ERK2 but that the MEK inhibitor, PD98059, significantly reduced the degree of apoptosis caused by etoposide and adriamycin. Furthermore, the expression of constitutively active

ERKs could considerably increase the sensitivity of immortalized NIH3T3 cells to etoposide-induced cell-death (Tang et al., 2002). As mentioned above, Boldt et al.

(2002), using a clonogenic assay to determine the ability of cancer cells to form colonies after taxol treatment, showed that in HeLa cells, ERK1/ERK2 has a pro-apoptotic role whereas in A431 and to a lesser degree in MCF7 cells, both ERK and p38 play a pro- survival role. These results suggest that the role played by the ERKs is dependent on the

cell type in question as well as the chemotherapeutic drug being used. This stresses the

need to investigate cell signalling in each cell-type of interest. What occurs in one cell-

type may not occur in another.

Besides ERK1/ERK2, it has also been demonstrated that chemotherapeutic drugs can

activate JNK and that the JNK signaling pathway may play an important role in

regulating drug-induced cell death. The activation of JNK is normally linked to cell death

under stressful stimuli. Thus, Brantley-Finley et al. (2003) showed that the JNK inhibitor,

76 SP600125, protected KB-3 carcinoma cells against the cytotoxic effects of vinblastine and doxorubicin hydrochloride and almost abrogated etoposide-induced cell death. The notion that JNK plays a destructive role in the mechanism of action of chemotherapeutic drugs is further supported by observations in various other tumour cell-types, including prostate cancer cells and breast cancer cells (Shimada et al., 2003; Kim et al., 2003).

Furthermore, the ability of 1,25(OH)2D3 to enhance the killing of murine squamous cell carcinoma by cisplatin has been reported to be mediated by the JNK pathway

(Hershberger et al., 2002). These data are consistent with JNK playing a role as a pro- apoptotic signalling molecule.

However, there are also some studies which showed that JNK could play a role in promoting cell survival (Nishina et al., 2004). NH2-terminal phosphorylation of c-Jun by

JNK was shown to play a role in DNA repair and promoting cell survival after cisplatin treatment in human cancer cell lines such as glioblastoma, ovarian, breast, and prostate carcinoma cells (Gjerset et al., 1999; Hayakawa et al., 1999; Potapova et al., 2001). A study by Hayakawa et al., (2003) showed that inhibition of JNK increased the sensitivity of tumour cells to cisplatin and etoposide and that JNK played a role in drug resistance by enhancing DNA repair in the human breast cancer BT474 cell line. Recent studies have suggested that the different JNK isotypes may play opposing roles in regulating cell survival and death (Dietrich et al., 2004; Ventura et al., 2004; She et al., 2002). Thus, the above results could be due, at least in part, to which JNK isoform was activated in a particular cell-type by a particular drug.

77 From the above discussion, it is therefore clear that many of the chemotherapeutic drugs are able to stimulate the activities of signalling molecules such as PKC, PI3K and the

ERK1/ERK2, JNK and p38 MAP kinases, to mention a few. The roles these signalling molecules play in the cytotoxic action of the drugs are dependent, to a large degree, on the cell-type and the drug under investigation. Given that the MAP kinases have moved into the limelight of drug discovery, modulation of the MAPK pathways to enhance the efficacy of chemotherapeutic drugs is likely to require some tailoring to suit the specific combination of drug and cancer.

Currently, there are no reports on whether chemotherapeutic drugs can regulate the

CYP24 promoter. However, given the roles PKC and the MAP kinases play in mediating the effect of 1,25(OH)2D3 on the CYP24 promoter, and the ability of chemotherapeutic drugs to stimulate the activities of these signalling molecules, it is possible that the drugs could alter the expression of CYP24 and hence the rate of 1,25(OH)2D3 metabolism.

1.12 Hypotheses

Based on the above information, I hypothesise that chemotherapeutic drugs cause a reduction in the circulating levels of 1,25(OH)2D3 by up-regulating the expression of cytochrome p450C24 (CYP24) in kidney cells. This is due to their ability to stimulate the activities of the MAP kinases. Blocking the activities of these kinases inhibits the ability of the drugs to stimulate the transcriptional activation of the CYP24 promoter.

78 1.13 Aims

The aims of my project are to investigate whether:

chemotherapeutic drugs such as daunorubicin hydrochloride, etoposide, cisplatin and vincristine sulfate are able to stimulate CYP24 promoter activation in kidney cell-lines. the above drugs alter the ability of 1,25(OH)2D3 to stimulate the CYP24 promoter. the chemotherapeutic drugs are able to stimulate the activities of ERK1/ERK2, ERK5 and JNK in the kidney cells.

inhibition of ERK1/ERK2/ERK5 and JNK activities blocks the ability of the chemotherapeutic drugs per se to stimulate CYP24 promoter activity and whether this is modulated by the presence of 1,25(OH)2D3.

79

Chapter 2

Materials and Methods

80 2.1 Materials

2.1.1 General chemicals, drugs and reagents

Trizma base (tris- (Hydroxymethyl) aminomethane), EGTA, EDTA, Triton X-100,

Ponceau S, Nonidet-P40, D-glucose, phorbol 12-myristate 13-acetate (PMA), Folin and

Ciocalteu’s Phenol Reagent, sodium diatrizoate were from Sigma-Aldrich Pty. Ltd,

(Sydney, NSW, Australia). Tween-20, sucrose, NaCl, Na2CO3 were from Asia Pacific

Specialty Chemical Limited, (Seven Hills, NSW, Australia). β-mercaptoethanol, acetic acid, Na/K-tartrate, sodium hydroxide (NaOH), diethyl ether, hexane, acetic acid, and acetone were from AJAX Finechem, (Seven Hills, NSW, Australia). CuSO4 was from

BDH Chemicals, (Port Fairy, VIC, Australia). Dimethyl sulphoxide (DMSO) was from

Merck, Germany. Trypsin-EDTA was from Gibco, (Grand Island, NY, USA). HEPES was from Paton Scientific, (Stepney, SA, Australia). Bovine Serum Albumin was from

Boehringer Manheim, (Mannheim, Germany). Trypan Blue was from Sigma, (St Louis,

MO, USA).

2.1.2 Reagents for Western blots and Kinase assays

Acrylamide (N, N’-Methylbisacrylamide 30%), ammonium persulfate (APS), sodium dodecyl sulfate (SDS), N,N,N’,N’,-Tetramethylethylenediamine (TEMED), Bio-Rad

Minigel Apparatus, and low range molecular markers were from Bio-Rad Laboratories,

(Sydney, NSW, Australia). Chemiluminescence reagent was from Perkin Elmer Life

Sciences, (Boston, MA, USA). Western Blot recycling kit was from Alpha Diagnostic

International, (San Antonio, TX, USA). Nitrocellulose membrane (0.2 mm thickness

Optitran BA-S83, reinforced NC) was from Schleicher & Schuell, (Dassel, Germany). γ- 81 32P ATP (specific activity 4000 Ci/mM) was from Geneworks (Adelaide, SA, Australia).

X-ray film (Curix Ortho HT-G 100 NIF) was obtained from AGFA, (Mortsel, Belgium).

2.1.3 Chemotherapeutic drugs

Daunorubicin hydrochloride, doxorubicin hydrochloride, vincristine sulfate salt, etoposide and cis-platinum ІІ diammine diachloride (cisplatin) were from Sigma, (St

Louis, MO, USA).

2.1.4 Reagents for lysis, washing and assay buffer,

Leupeptin, pepstatin A, benzamidine hydrochloride, DL-dithiothreitol (DTT), PMSF

(phenylmethylsulfonyl fluoride) and Sigma 104 (phosphatase substrate) were from

Sigma-Aldrich Pty. Ltd, (Sydney, NSW, Australia). Aprotinin (bovine lung, 100,000 U/5 ml) was from Calbiochem, (La Jolla, CA, USA).

82 2.1.5 Reagents for transfection

DOTAP Liposomal Transfection Reagent was from Roche, (Indianapolis, IN, USA).

2.1.6 Reagents for Luciferase Assay

Dual luciferase assay kit was from Promega (Madison, WI, USA)

2.1.7 Antibodies

Anti-mouse IgG, (HRP conjugated) was from DAKO Corporation, (Carpinteria, CA,

USA). Anti-rabbit IgG (HRP conjugated) was from Chemicon, (Boronia, VIC, Australia).

Anti-ERK2, anti-active ERK and anti-β actin antibodies were from Santa Cruz

Biotechnology, (Santa Cruz, CA USA). Anti-ERK5 antibody was from Sigma-Aldrich

Pty. Ltd., (Sydney, NSW, Australia). Anti-VDR and anti-rat IgG antibodies were kind gifts from Dr EM Gardiner (Garvan Institute of Medical Research, NSW, Australia).

2.1.8 Serum, Culture media and buffers

Fetal calf serum, RPMI-1640 and L-glutamine (200 mM) was from JRH biosciences,

(Lenexa, KS, USA). Dulbecco’s Minimal Essential Media (DMEM), RPMI-1640 without phenol red and Trypsin EDTA were from Gibco BRL, (Carlsbad, California, USA).

Penicillin/streptomycin (5000 U/ml) was from Commonwealth Serum Laboratories

83 (CSL), (Parkville, Victoria, Australia). Phosphate buffered saline (PBS) was from Cell

Image, (Adelaide, SA, Australia).

2.1.9 Flasks and plates for culture

Culture flasks (75 cm2, 125 cm2) were from Corning, (Corning, NY, USA) and Becton

Dickinson Biosciences, (Franklin Lakes, NJ USA). Tissue culture plates (100x20 mm) were from TPP, (Trasadingen, Switzerland). Multi-well plates (24 wells) were from

Becton Dickinson Labware, (Franklin Lakes, NJ, USA).

2.1.20 Cell lines

Human primary embryonic kidney cell line (HEK-293T) was from J Health, Oxford

University, (Oxford, UK, England). Monkey kidney fibroblast cell line (COS-1) and human promyelocytic leukemic (HL60) cells were obtained from American Type Culture

Collection (ATCC), (Rockville, MD, USA). Isolated primary rat hepatocytes were from

Dr Roland Gregory, Department of Medical Biochemistry (Flinders University, SA,

Australia). They were prepared fresh on the day of use.

2.1.21 Plasmids/Vectors

The rat CYP24 promoter luciferase constructs pCYP24WT(-1.4kb)-Luc) and pCYP24WT(-298bp)-Luc contain the promoter sequence from either -1.4 kb or -298 bp to the transcription start site as well as 74 bp of 5´-untranslated region cloned upstream of

84 the firefly luciferase reporter gene in the pGL3 basic vector, as described previously

(Hahn et al., 1994; Dwivedi et al., 2000) (Figure 2.1 A and B).

The pCYP24mVDRE1&2(-298bp)-Luc (where Luc stands for luciferase) containing mutations to VDRE1 and VDRE2 has been described previously (Dwivedi et al., 2000).

(Figure 2.1 C).

The pCYP24(mEBS)(-298bp)-Luc reporter construct containing a mutation in the EBS has been described previously (Dwivedi et al., 2000) (Figure 2.1 D).

The pCYP24mGC(-298bp)-Luc containing a mutation to the GC box at -114/-101 has been described previously (Gao et al., 2004) (Figure 2.1 E).

85

VDRE-2 VDRE-1 EBS GC-Box Luciferase A +1 -1.4kb pCYP24WT(-1.4kb)-Luc +74

B VDRE-2 VDRE-1 EBS GC-Box Luciferase +1 -298bp pCYP24WT(-298bp)-Luc +74

C VDRE-2 VDRE-1 EBS GC-Box Luciferase +1 pCYP24mVDRE1&2-Luc +74 -298bp

D VDRE-2 VDRE-1 EBS GC-Box Luciferase +1 -298bp pCYP24mEBS-Luc +74

E VDRE-2 VDRE-1 EBS GC-Box Luciferase +1 -298bp pCYP24mGC-Luc +74

Figure 2.1 Diagram showing the CYP24WT(-1.4kb)-Luc, CYP24WT(-298bp)-Luc

and the site-directed mutagenesis for the VDRE1&2, EBS and putative GC site of

the CYP24WT(-298bp)-Luc.

86 The pCYP24WT(-298bp)-Luc, pCYP24WT(-1.4kb)-Luc, pCYP24mVDRE1&2-Luc, pCYP24mEBS-Luc, pRSV-hVDR, pRSV-hRXR S260A, VP16-RXR and pCYP24mGC-

Luc were provided by Dr. P Dwivedi (School of Molecular Sciences, University of

Adelaide, SA, Australia) (Hahn et al., 1994; Dwivedi et al., 2000; Dwivedi et al., 2002).

The promoterless pRL-null Renilla luciferase vector (pRL-Null-Luc) and HSV-thymidine kinase promoter pRL-TK Renilla luciferase vector (pRL-TK-Luc) was purchased from

Promega (Madison, WI, USA).

Dominant negative PKC-ζ, PKCζK281M, was generously provided by Dr. JW Soh

(Columbia University, USA) (Soh et al., 1999).

Dominant negative Ras, Ras17N, and dominant negative ERK1, ERK1K71R were kindly provided by Prof. C. J. Der (University of North Carolina at Chapel Hill, NC, USA)

(Quilliam et al., 1994; Westwick et al., 1994)

Dominant negative MEK5, MEK5A, was kindly provided by Prof. E. Nishida (Kyoto

University, Kyoto, Japan) (Kamakura et al., 1999).

Dominant negative MKK4, MKK4K116R, was provided by Prof. D. Riches (National

Jewish Medical and Research Center, Denver, CO, U.S.A.) (Chan et al., 1999)

GST-Ets-1 was kindly provided by Professor Ismail Kola (Institute of Reproduction and

Development, Monash University, Victoria, Australia) (Dwivedi et al., 2002)

87 His-Ets-1 and GST Ets-1T38A mutant were provided by Dr. P Dwivedi (School of

Molecular Sciences, University of Adelaide, SA, Australia) (Dwivedi et al., 2002).

2.1.22 Buffers and Solutions

The following buffers were used in this thesis:

Lysis buffer – 20 mM HEPES, pH 7.2, NP-40 0.5% (v/v), 100 mM NaCl, 2 mM Sigma

104, 1 mM DTT, 1 mM PMSF, 10 µg/ml aprotinin, 10 µg/ml leupeptin, 10 µg/ml pepstatin A and 10 µg/ml benzamidine hydrochloride.

Washing buffer – 10 mM HEPES, pH 7, 100 mM NaCl, 2 mM DTT, 2 mM Sigma 104,

1 mM PMSF and 10 µg/ml each of aprotinin, leupeptin, pepstatin A and benzamidine.

Assay buffer – 20 mM HEPES, pH 7.2, 20 mM β-glycerophosphate, 10 mM Sigma 104,

10 mM MgCl2, 1 mM DTT, 50 µM orthovanadate, 20 µM ATP.

Laemmli buffer – 20 mM Tris-HCl, pH 6.8, 40% sucrose, 6% SDS and 10 mM β- mercaptoethanol.

Blocking buffer for western blots - 5% skim milk, 25 mM Tris/HCl pH 7.4, 100 mM

NaCl.

Washing buffer for western blots – blocking buffer + 0.1% Tween-20.

GTE buffer - 50 mM glucose, 25 mM Tris pH 8.0, 10 mM EDTA pH 8.0.

88

1x Tris-EDTA buffer (1xTE buffer) – 10 mM Tris-HCl, 1.0 mM EDTA pH 8.0.

Lowry’s Solution - 2% Na2CO3, 1% SDS, 0.4% NaOH, 0.16% Na/K-tartrate.

Solution II for plasmid DNA extraction - 0.2 M NaOH, 1% SDS.

2.1.23 Reagents for RNA extraction and real time PCR

Trizol Reagent and SuperScript III reverse transcriptase was from Invitrogen (Mount

Waverley, Vic, Australia). Random hexamer primers, dNTPs, and PCR primers were purchased from Geneworks (Adelaide, SA, Australia). Primers designed to amplify cDNA for CYP24 (F: 5’-cctgctgccagattctctggaa-3’; R: 5’-atacttcttgtggtactccacca-3’) and

GAPDH (F: 5’-acccagaagactgtggatgg-3’; R: 5’-cagtgagc ttcccgttcag-3’) were based on

Genebank sequences (CYP24 Acc: NM000782; GAPDH Acc: NM002046). The iQ™

SYBR® Green Supermix was purchased from Bio-Rad Laboratories (Regents Park, NSW,

Australia).

89 2.2 Methods

2.2.1 Cell Culture

2.2.1.1 Culture of HEK293T and COS-1 cells

HEK293T and COS-1 cells were maintained in DMEM supplemented with 10% heat- inactivated FCS, 100 U/ml penicillin/streptomycin, 4 mM L-glutamine. They were incubated in a humidified atmosphere of 5% CO2 in air at 37°C.

2.2.1.2 Cell Culture for kinase assay and for western blotting.

HEK293T and COS-1 cells were seeded at 2.5 x 106 cells per 10 cm2 culture plate in 6 ml of DMEM supplemented with 10% heat inactivated FCS, 100 U/ml penicillin/streptomycin, 4 mM L-glutamine. The cells were incubated at 37°C in a humidified atmosphere of 5% CO2 in air overnight before being washed with 4 ml serum- free RPMI-1640. The cells were then incubated with 6 ml of serum-free RPMI-1640 overnight before being treated with chemotherapeutic drugs and/or 1,25(OH)2D3.

90 2.2.1.3 Culture of HL 60 cells

HL-60 cells were maintained in RPMI-1640 supplemented with 10% heat-inactivated

FCS, 100 U/ml penicillin/streptomycin, 4 mM L-glutamine. They were incubated in a

6 humidified atmosphere of 5% CO2 in air at 37°C, and kept at a density of <1x10 cells/ml.

2.2.2 Cell Viability Assay using trypan blue.

HEK293T cells, COS-1 cells, isolated hepatocytes and HL60 cells were seeded at 75,000 cells in 400 µl of DMEM containing 10% FCS in a 24 well plates. The cells were incubated overnight before being treated with one of the following chemotherapeutic drugs: etoposide, cis-platinum, vincristin sulfate or daunorubicin hydrochloride at concentrations indicated in the Results section. After 24 hours, the cells were harvested and 44 µl of Trypan blue (1% w/v) was added. The total number of cells and the trypan blue-positive cells were counted using a haemocytometer under a light microscope. The percentage of dead cells was calculated based on the number of cells that had taken up trypan blue relative to the total number of cells counted.

91 2.2.3 Plasmid DNA Preparation and Transfection

2.2.3.1 Midi-preparation using midi-plasmid DNA preparation kit (Qiagen)

Isolation of up to 100 µg plasmid DNA was performed using the midi-preparation kit according to the manufacture’s instruction.

2.2.3.2 Large scale preparation by cesium-chloride (CsCl) density gradient procedure

High concentrations of plasmid DNA can be obtain using a modified version of

Sambrook et al. (1989). The desired clone culture e.g.20 ml was inoculated into 250 ml of

LB medium with appropriate antibiotics (100 µg/ml Ampicillin) and incubated at 37oC with shaking. The bacterial cells were then pelleted at 4oC in a SS-34 rotor at 6000 rpm for 5 min. Plasmid DNA was extracted using the alkaline lysis procedure: 10 ml of freshly prepared GTE buffer with lysozyme (2 mg/ml) was used to lyse the cells on ice for 20 min; 20 ml of freshly prepared solution II was added and mixed with the cell lysate by inverting the tubes and then 15 ml of NaAc (3M, pH 4.6) was added, mixed and left on ice for 10 min. The supernatant was collected and passed through a funnel containing cloth bandage and the volume recorded in a 50 ml measuring cylinder. Plasmid DNA was precipitated by the addition of 0.6 volume of isopropanol and incubated on ice for 10 min.

The plasmid DNA was pelleted at 9000 rpm for 20 min and the supernatant discarded.

The pelleted plasmid DNA was allowed to air dry and resuspended in 4 ml of 1xTE buffer. To separate plasmid DNA from RNA and genomic DNA, the plasmid was further purified using the CsCl/ethidium bromide density gradient procedure. CsCl (4.2 g) was mixed with the dissolved plasmid DNA by vortexing thoroughly. Ethidium bromide (100

µl of 10 mg/ml) was added as the marker of plasmid DNA. The mixture was then

92 centrifuged in a Beckman TL-100 bench top ultracentrifuge and TLA-100.2 rotor at

80,000 rpm at 20oC overnight. Plasmid DNA with ethidium bromide (the lower most central red band) was havested using syringe and needles. The ethidium bromide was removed by butanol extraction. The plasmid DNA was then precipitated with 2.5 volumes of 100% ethanol, rinsed with 70% ethanol, resuspended in 1xTE buffer, and quantified by spectrophotometry (260 nm) and analysed by agarose gel electrophoresis to confirm concentration and supercoiling.

2.2.3.3 Transient transfection of mammalian cells and treatment with 1,25(OH)2D3 and/or chemotherapeutic drugs.

Transfection of cells was performed with the liposomal transfection reagent, DOTAP, according to Dwivedi et al., (2002). In preparation for transfection, approximately 5x104 cells were seeded into 24 well plates in 400 µl of DMEM containing 10% FCS to achieve

60-70% confluence. Following incubation for 4-5 h to allow the cells to attach, the wells were washed with 250 µl of serum-free RPMI-1640 and the medium replaced with 400 µl of serum-free RPMI-1640. Transfection was performed in triplicate using 200 ng of the

CYP24 promoter luciferase construct and 50 ng of pRL-Null-Renilla luciferase vector

(pRL-Null-Luc) as control for transfection efficiency. Whenever indicated, 200 ng of dominant negative mutants or blank plasmids were also co-transfected. All plasmid DNA was diluted in HEPES buffer (20 mM, pH 7.4) to a final volume of 5 µl and mixed with

1.5 µl of DOTAP diluted to 3.5 µl with HEPES buffer. DNA-DOTAP complex formation was achieved by incubating the mixture for 15-20 min at room temperature. A total of 10

µl DNA-DOTAP complex was transfected into the cells. After an overnight incubation, the media was replace with 400 µl serum-free medium RPMI-1640. The cells were treated with different chemotherapeutic drugs, 1,25(OH)2D3 or vehicle (H2O, ethanol or

93 DMSO). The maximum amount of vehicle was 0.1% (v/v), which did not affect promoter activity. After 24 h, the medium was aspirated and the cells were lysed using 50 µl of 1X passive lysis buffer (Promega) for 20 min at 25oC.

2.2.3.4 Dual Luciferase assay

Dual luciferase activity was measured using a TD-20/20 luminometer. The first luminescence from the firefly luciferase (representing CYP24 promoter activity) was recorded by adding 25 µl of LARII substrate into the 2 µl of cell lysate in an eppendorf tube. The second luminescence for Renilla luciferase (representing reference luminerscence) was quantified by the addition of 25 µl of Stop and Glo substrate to quench the first reaction and simultaneously initiate the Renilla luciferase reaction. Data were recorded as relative luciferase activity, the ratio of the first luminescence over the second luminescence. Results were then presented as fold induction over vehicle (EtOH or DMSO)-treated control cells. Analysis of renilla luciferase activity from experiments in which chemotherapeutic drugs were used revealed that the drugs did not significantly alter this activity.

2.2.4 Preparation of cell lysates

After treatment the plates were placed on ice and the medium was removed using vacuum suction. The cells were harvested using a rubber policeman in 300 µL of lysis buffer. The samples were rocked for 2 h at 4o C, centrifuged (14,000 g x 5 min, 4o C) and the supernatants were collected and stored at –70o C until assayed. An aliquot was taken for protein estimation.

94 2.2.5 Lowry’s Protein determination

Protein content of samples was determined using the Lowry method (Harrington, 1990).

Protein standards were obtained using serial dilutions of BSA (1 mg/ml) with H2O, resulting in 6 standards, 0 µg, 3.125 µg, 6.25 µg, 12.5 µg, 25 µg and 50 µg. Each unknown sample (5 µl) was diluted in 45 µl of H2O, then 150 µl of CuSO4/Lowry’s solution (1:100) was added to all tubes and these were left to stand for 10 min at room temperature. Folin and Ciocalteu’s Phenol Reagent/H2O (1:1), 15 µl, was then added and the tubes were left to stand for 30 min. 180 µl of each sample was transferred to a 96 well plate and the absorbance at 570 nm was determined in a plate reader (Dynatech MR5000,

Guernsey, Channel Islands, UK). The absorbance of the protein standards was plotted on a linear regression nomogram (Instat, GraphPad Software Incorporated, San Diego, CA,

USA) and the protein content of the test samples was calculated.

2.2.6 Kinase Assay and Western Blotting

2.2.6.1 Purification of GST-jun (1-79)

A plasmid encoding the 1-79 amino acid sequence of c-jun was kindly provided by Prof.

C.J. Der (University of North Carolina, Chapel Hill, USA). After transformation, a single colony was inoculated into 100 ml of Luria-Bertani (LB) broth supplemented with 100

µg/ml ampicillin (CSL, Victoria, Australia) and shaken at 200 rpm for 16 h at 37oC in a

G24 Environmental Incubator shaker (New Brunswick Scientific, Edison, USA). After 16 h the entire inoculum was transferred into 400 ml of fresh LB broth supplemented with

100 µg/mL ampicilin and shaken at 200 rpm for 1 h at 37oC. The expression of GST c-jun

95 (1-79) fusion protein was induced by the addition of 1 mM Isopropyl-β-D- thiogactopyranoside (Boehringer Mannheim, Castle Hill, Australia) to the culture. The flasks were shaken for another 3 h at 200 rpm at 37oC. The cells were then collected by centrifugation (10 min x 12,500 g) at 4oC. The pellets were resuspended in 20 ml of ice cold phosphate buffer (22 mM, pH 7.0) and the cells disrupted using a French Pressure

Cell (SLM Instruments). The lysate was cleared by centrifugation (30 min x 48,000 g) at

4oC. The clarified lysate was then added to glutathione-sepharose CL4B (Pharmacia

Biotech, Uppsala, Sweden) and incubated overnight with constant mixing at 4oC. The

GST-c-jun (1-79)-bound glutathione sepharose was then washed with 50 ml of PBS (3 x

5 min, 600 g) and finally resuspended in PBS (1:1, PBS:beads). The GST-c-jun (1-79)- bound glutathione sepharose was stored at –70o C.

2.2.6.2 JNK kinase assay

A solid phase assay (Hii et al., 1998) was employed to determine the activity of JNK.

Equal amounts of lysate protein (0.8 mg to 1.2 mg) were added to 50 µl of GST-c-jun (1-

79) glutathione sepharose and MgCl2 and ATP added to a final concentration of 15 mM and 10 µM, respectively, and the samples rocked for 2 h or overnight at 4o C. The samples were washed once each with 400 µl with lysis buffer, washing buffer and assay buffer. The assay was initiated by adding 30 µl of reaction mixture (assay buffer, containing 33 mM of DTT, 3.8 mM of Sigma 104 and 10 µCi of [γ-32P] ATP) to the tubes, which were then incubated at 30o C for 20 min. The reaction was terminated by the addition of Laemmli buffer (ratio of 1:2 of laemmli buffer: assay mixture) and the samples heated at 100o C for 5 min. Phosphorylated GST c-jun (1-79) was fractionated on a 12% SDS-PAGE gel (BioRad, NSW, Australia) (175 volts for 90 min). An aliquot of

Low Range molecular weight markers (Bio-Rad, Sydney, NSW, Australia) was loaded in

96 a separate lane to provide a range of molecular weight markers. The radioactivity incorporated in the GST-c-jun (1-79) peptide was quantified using a Packard Instant

Imager (Packard, Canberra, Australia) and integrated using the Imager software (version

2.0.5).

2.2.6.3 ERK1/ERK2/ERK5 kinase assays

Equal amounts of lysate protein (0.8 mg to 1.2 mg) were added to 15 µl of protein A

Sepharose (Sigma Chemical Company, St Louis, USA) and the samples rocked for 30 min at 4oC and centrifuged (15,000 g x 10 s, 4oC) to pre-clear the samples. The supernatants were removed and incubated with 15 µl of protein A Sepharose and 3 µg of anti-ERK2 or anti-ERK5 antibody. After rocking for 2 h or overnight at 4oC, the samples were centrifuged (15,000 g x10 s, 4 o C) and the supernatants were discarded. The pellets were washed as described above for JNK assay. The assay was initiated by adding 30 µl of reaction mixture (assay buffer containing, 33 mM of DTT, 3.8 mM of Sigma 104, 0.03 mg/ml MBP and 10 µCi of [γ-32P] ATP). Samples were incubated at 30o C for 20 min.

The reaction was terminated by the addition of Laemmli buffer (15 µl) followed by heating at 100oC for 5 min. Phosphorylated MBP was separated from unincorporated 32P

ATP on a 16% SDS-PAGE gel run at 175 volts for approximately 1.5 h. Phosphorylated

MBP was visualised in a Packard Instant Imager (Packard, Canberra, Australia) and the amount of radiolabel was quantified using the Imager software (version 2.0.5).

2.2.6.4 Western Blotting

Western blotting was used to determine the levels of ERK1/ERK2, ERK5, active ERK and VDR protein. We used a Bio-Rad Mini gel apparatus or CBS Scientific MGV-201

97 mini-gel unit to fractionate ERK5 (10% SDS-PAGE gel), ERK1/ERK2 and VDR (12%

SDS-PAGE gel). Lysates were mixed with Laemmli buffer, boiled for 5 min at 100° C and between 20 to 40 µg of proteins were loaded per lane, depending on the experiment.

Within each experiment, the same amount of lysate proteins were loaded for each lane.

The gel was run at 175 volts for approximately 1 h or until the dye front had migrated out of the gel. The separated proteins were subsequently transferred onto nitrocellulose membrane in a Bio-Rad Mini Trans-Blot cell (Bio-Rad, Sydney, NSW, Australia) at 100

V for 1.5 h. The blots were stained with Ponceau S (0.1% in 5% acetic acid) to visualise the transferred proteins and confirm even transfer of proteins to the membrane. After destaining, the membrane was blocked for 1 h at 37° C or overnight at 4° C in blocking buffer, then incubated with primary antibody (Table 2.1) diluted in Tris-Tween (pH 7.4) for 1 h. The membrane was washed (3 x 10 min) in washing buffer and then incubated for

1 h with horseradish peroxidase-conjugated secondary antibody (Table 2.1), diluted in blocking buffer. The membrane was washed again (3 x 10 min) with washing buffer. The target protein was visualized by enhanced chemiluminescence (Hii et al., 1995). The relative density of each band was determined using Image Quant software, version 3.3

(Molecular Dynamics, Charlottesville, VA, USA).

Table 2.1: Antibodies types and dilution for primary and secondary antibodies.

Primary Antibodies Secondary Antibodies

Anti ERK5 1:10,000 Anti Rabbit 1:2000

Anti ERK2 1: 2000 Anti Goat 1: 2000

Anti-active ERK 1:10,000 Anti Rabbit 1:2000

Anti VDR 1:1000 Anti Rat 1:50000

Anti β-actin 1:10,000 Anti Mouse 1:2000

98 2.2.6.5 Western Blot recycling

Following Western blotting, nitrocellulose membranes were stored at -20° C. When required the blots were stripped using a Western blot recycling kit. The membrane was incubated in stripping solution (1:10) for 10 min, then blocked using the supplied blocking solution (1:20) (2x5 min). The membrane was then re-probed with primary and secondary antibodies as previously described. Stripping and re-probing was possible up to

3 times. The bands were again subjected to densitometric analysis. Densitometry permits the correction of unequal protein loading and the results are expressed as a ratio of a target protein: reference protein.

2.2.7 Quantification of mRNA expression using real-time PCR

2.2.7.1 RNA Extraction

Following treatment, cells were lysed in 1 ml of Trizol Reagent. RNA was than partitioned into the aqueous phase by adding chloroform (0.2 ml), mixed and centrifuged

(15min x 12,000 g). The aqueous phase was than transferred to a clean eppendolf tube.

RNA was precipitated by adding 0.5 ml of isopropanol and 10 µg of RNase-free glycogen. After incubating overnight at -70° C, the sample was centrifuged (15min x

12,000 g). Supernatent was discared and RNA pellet washed with 1 ml of 75% EtOH.

The RNA pellet was than air dry and resuspended in 15 µl of H2O. Total amount of RNA was quantified using a spectrophotometer (by reading the absorption at 260nm).

2.2.7.2 Generation of cDNA cDNA was generated by reverse PCR. Briefly, 1 µg of RNA was mixed with Reaction

Mixture I which consisted of 2 µl of Random Hexamer, 2 µl of 10mM dNTPs and topped

99 up to a final volume of 23 µl with H2O. After incubating for 5 min at 65° C, the mixture was place on ice for 1 min. Mixture II consisting of 8 µl of 5x first strand Buffer, 2 µl of

0.1 M DTT and 2 µl of Superscript III Reverse Transcriptase was than added and the mixture incubated at 50° C for 60 min. Following completion of reaction, the enzyme was than inactivated by heating to 70° C for 15 min.

2.2.7.3 Real-time PCR

The cDNA samples were amplified by mixing 1 µl of cDNA together with 5 µl of iQ

SYBR Green Supermix (2x), 0.5 µl of CYP24 or GAPDH forward primer (5µM), 0.5 µl of reverse primer (5µM) and 3 µl of H2O. The reaction mixture was than placed into a

Rotor-Gene Real-Time PCR thermo cycler and the following programme was run. 15 min at 94° C to melt the cDNA, 35 cycle of 30 sec at 94° C, 30 sec at 60° C, 30 sec at 72° C to amplify cDNA and 4 min at 72° C to deactivate enzyme. To check purity a melt curve was generated by heating for 15 sec at 72° C and 90° C for 5 sec. Relative expression between samples was calculated using the comparative cycle threshold (CT) method

(∆CT) (Livak and Schmittgen, 2001).

2.2.8 Statistics

ANOVA, Tukey-Kramer multiple comparisons test and Dunnett’s multiple comparisons test and Student’s t-test were conducted using Graph Pad InStat V2.02 (Graph Pad

Software, San Diego, CA, USA) and SPSS 11 student version (SPSS, Chicago, IL, USA).

Results were considered statistically significant when p<0.05.

100

Chapter 3

Effects of chemotherapeutic drugs on the induction of the CYP24 promoter in kidney

cells.

101 3.1 Introduction

As discussed in Chapter 1.11, chemotherapeutics drugs have been reported to stimulate the activities of MAP kinases which we have been previously reported to regulate the activity of the CYP24 promoter by 1,25(OH)2D3. This raises the possibility that the drugs could induce CYP24 promoter activity by themselves. The aim of this chapter is to investigate whether chemotherapeutic drugs such as daunorubicin hydrochloride, etoposide, cisplatin, vincristine sulfate, doxorubicin hydrochloride and dexamethasone can stimulate the CYP24 promoter activity in kidney cells. The effect of the drugs were tested in two different CYP24 promoter constructs, the -1.4kb and the -298bp promoter.

Unlike the -1.4kb promoter, the -298bp promoter has been studied extensively and has been very well characterized. However, we chose to investigate the -1.4 kb promoter construct first because it resembles more of the endogenous promoter in length.

COS-1 and HEK293T cells were used in this study. Both kidney cell lines have been previously validated for studying the regulation of CYP24 promoter activity (Dwivedi et al., 2000 and Nutchey et al., 2005). The concentrations of drugs used in this study are in the range of levels found in the serum of patients receiving chemotherapy. List et al.

(2001) found that the peak serum concentration of daunorubicin hydrochloride in patients with acute myeloid leukemia to be around 0.9 µM – 1.1 µM. Kato et al. (2003) found that etoposide levels in Japanese children and adolescents were around 16 µM – 54 µM.

Marina et al. (2002) found that the peak serum concentration of doxorubicin hydrochloride in children with refractory tumours to be around 63 µM – 83 µM. Peng et al. (1997) reported that the peak concentration of cisplatin in pediatric patients was around 13 µM – 20 µM. These concentrations were also used in a study to investigate the

102 effects of the drugs on apoptosis in osteoblasts (10 µM etoposide, 0.1 µM vincristine sulfate and 0.1 µM daunorubicin hydrochloride) (Davies et al., 2003).

Prior to undertaking the promoter induction studies, the effects of the drugs on cell viability were investigated to exclude the possibility that these drugs are cytotoxic to cells which either synthesise or metabolise 1,25(OH)2D3. Using the trypan blue exclusion test to determine viability, the data demonstrate that whereas daunorubicin hydrochloride, etoposide, cisplatin and vincristine sulfate were effective at killing HL60 human promyelocytic leukaemic cells at the above-listed dose ranges, these drugs did not affect the viability of COS-1, HEK293T and primary hepatocyte cells, even at concentrations higher than those found in the serum of patients receiving chemotherapy (data not shown). Thus, it is unlikely that the reduction in serum 25(OH)2D3 and 1,25(OH)2D3 levels seen in children undergoing chemotherapy is the result of the cytotoxic effects of the drugs on hepatocytes and kidney cells which produce the vitamin D metabolites.

3.2 Methods for investigating CYP24 promoter induction

COS-1 and HEK293T kidney cells were transiently transfected with a construct containing the –298bp or the –1.4kb CYP24 promoter sequence fused to the firefly luceferase reporter gene (pCYP24WT (-298bp)-Luc and pCYP24WT (-1.4kb)-Luc). A recombinant vector pRL-Null (0) expressing renilla luciferase was also co-transfected as an internal control for transfection efficiency. Cells were then treated with daunorubicin hydrochloride, etoposide, cisplatin or vincristine sulfate alone or in combination with

1,25(OH)2D3. Data obtain (ratio of firefly:renilla luciferase activity) were expressed as fold induction, the ratio of drug treated cells over ethanol treated cells (control group)

103 3.3 Results

3.3.1 Effects of chemotherapeutic drugs on the activity of the -1.4kb CYP24 promoter in COS-1 and HEK293T kidney cells.

COS-1 and HEK293T cells were transiently cotransfected with a construct containing the

-1.4 kb CYP24 promoter-luciferase gene (pCYP24WT (-1.4 kb-Luc) and the pRL-0 vector expressing Renilla luciferase. Cells were treated with daunorubicin hydrochloride, etoposide, cisplatin or vincristine sulfate alone or in combination with 1,25(OH)2D3. The data obtain were expressed as fold induction, the ratio of drug treated cells over ethanol treated cells (control group).

The data show that daunorubicin hydrochloride by itself increase the activity of the -1.4 kb promoter in COS-1 and HEK293T cells by 2 ± 0.5 and 2.4 ± 0.4 fold, respectively over ethanol treated cells (Figure 3.1a). As expected, 1,25(OH)2D3 stimulated the activity of the CYP24 promoter in both cell lines. The induction by 1,25(OH)2D3 in

HEK293T cells was substantially higher than in COS-1 cells. Treatment with daunorubicin hydrochloride and 1,25(OH)2D3 together caused an additive effect on

CYP24 promoter activity in COS-1 cells but no further increase in CYP24 promoter activity was observed in HEK293T cells under this condition (Figure 3.1a).

The effect of etoposide was investigated next. This drug increased the -1.4 kb CYP24 promoter activity in COS-1 and HEK293T cells by 3 ± 0.6 and 2.8 ± 0.5 fold, respectively (Figure 3.1b). Interestingly, a synergistic increase in CYP24 promoter activity was observed when COS-1 and HEK293T cells were treated with both etoposide and 1,25(OH)2D3, giving approximately 7.4 and 4.3 fold, respectively, higher induction than that caused by 1,25(OH)2D3 (Figure 3.1b).

104

Vincristine sulfate alone increase the -1.4 kb CYP24 promoter activity by 2.9 ± 0.5 fold in COS-1 cells (Figure 3.1c). However, the drug did not affect the activity of the CYP24 promoter in HEK293T cells. A synergistic increase in CYP24 promoter activity was observed when COS-1 and HEK293T cells were treated with both vincristine sulfate and

1,25(OH)2D3, approximately 4.2 and 2 fold respectively higher than induction cause by

1,25(OH)2D3 (Figure 3.1c).

Cisplatin alone did not stimulate the -1.4 kb CYP24 promoter activity in both cell lines

(Figure 3.1d). However, the treatment of the cells with cisplatin and 1,25(OH)2D3 resulted in a synergistic increase in CYP24 promoter activity in COS-1 cells, approximately a 6.3 fold higher than induction cause by 1,25(OH)2D3. No further increase in promoter activity was observed in HEK293T cells (Figure 3.1d).

105 a b

25 ### 80 ) n on) i o i t ct c

u 20 ***

d COS-1 du n I n

COS-1 HEK293T I 60 d l old HEK293T

(F 15 n n (Fo o i t io t c c u 40 u d d n I n I

10 r ## e t oter o m

om *** r # 20

5 pro

* 4 * * 2 P *** CYP24 p CY 0 0 Daunorubicin - + - + Etoposide - + - + 1,2 5(OH)2D3 - - + + 1,25(OH)2D3 - - + +

c d

40 25

) ### ### ) n ction io t u

c d 20 *** n du I 30 COS-1 COS-1 n I old

HEK293T d HEK293T l F ( o F 15 on i t n ( c io

*** t c du 20 u n d I r n

### I

10

r ote e t o om r m p 10 o * pr 24 5 4 * * 2 P CYP

CY 0 0 Vincr istine - + - + Cisplatin - + - + 1,2 5(OH)2D3 - - + + 1,25(OH)2D3 - - + +

106 Figure 3.1 Effects of chemotherapeutic drugs on the -1.4kb CYP24 promoter in

COS-1 and HEK293T cells. Cells were transiently transfected with 200 ng of pCYP24WT (-1.4kb)-Luc and incubated with daunorubicin hydrochloride (2 µΜ), etoposide (100 µΜ), vincristine sulfate (10 µΜ), cisplatin (100 µΜ) and/or 1,25(OH)2D3

(10 nM) for 24 h. Control cells received ethanol (0.1%, v/v). The results are presented as fold induction over ethanol-treated control cells. Data are mean ± SEM of three independent experiments performed in triplicate. Significance of difference between control and drug or 1,25(OH)2D3: *p<0.05, **p<0.01, ***p<0.001; Significance of difference between 1,25(OH)2D3 and 1,25(OH)2D3 + drug: #p<0.05, ##p<0.01,

###p<0.001 (ANOVA and Tukey-Kramer multiple comparisons test). The response observed with HEK293T cells in the presence of 1,25(OH)2D3 either alone or in the presence of a drug was generally greater than that seen with COS-1.

107

3.3.2 Effects of chemotherapeutic drugs on the activity of the -298 bp CYP24 promoter in COS-1 kidney cells.

In this study, we compared the effects of daunorubicin hydrochloride, etoposide, vincristine sulfate and cisplatin on the -298bp and -1.4kb CYP24 promoter in COS-1 cells. Treatment of COS-1 cells with daunorubicin hydrochloride resulted in an increased the activity of the -298 bp CYP24 promoter activity by 2.7 ± 0.5 fold. This result is similar to the results seen in the -1.4 kb promoter. As expected, 1,25(OH)2D3 up- regulated the -298bp CYP24 promoter activity. Unlike the -1.4 kb CYP24 promoter, treatment with both daunorubicin hydrochloride and 1,25(OH)2D3 resulted in a substantial increase in the -298bp CYP24 promoter activity, (a 13.3 ± 1.1 fold higher than ethanol treated cells, and approximately 7 fold higher than the induction caused by

1,25(OH)2D3) (Figure 3.2a).

Likewise to the data obtain from the -1.4 kb promoter, etoposide was also able to increase the -298bp CYP24 promoter induction by 2.1 ± 0.2 fold over ethanol treated cells. When combined with 1,25(OH)2D3, etoposide treatment caused a synergistic increase in -298bp

CYP24 promoter activity (a 26.3 ± 3.9 fold increase over ethanol treated cells, and approximately 2.9 fold higher than CYP24 induction by 1,25(OH)2D3) (Figure 3.2b).

This result is similar to the result obtained in the -1.4 kb promoter.

108

a b

16 35 ### ) ### n o n)

cti 28 ## ctio du u n

12 d 298bp n d I

I

l 298bp

o 1.4KB F

1.4KB old (

n 21 (F io n t c u

ctio

d 8 u n d I r In

e 14 t

o # ** m oter o m o pr

4 ** * r 4

2 7 * 4 p P * * * * CY

CYP2 0 0

Daunorubicin - + - + Etoposide - + - + 1 ,25(OH)2D3 - - + + 1,25(OH)2D3 - - + +

c d

20 24 ###

ion) t on) c i t u 298bp 20 c

d 16 u n 298bp I 1.4KB ###

Ind 1.4KB old

(F 16 old

n (F

12 o i n t ** o c i t c du 12 ###

du n I er In 8 ot er ot 8 om r p om r * * 4 * * 4 P24 p * CYP24 Y C

0 0 Vincristine - + - + Cisplatin - + - + 1,25(OH)2D3 - - + + 1,25(OH)2D3 - - + +

109 Figure 3.2 Comparison of the effects of chemotherapeutics drugs on the -298bp and

-1.4kb CYP24 promoter constructs in COS-1 cells. Cells were transiently transfected with 200 ng of pCYP24WT (-298bp)-Luc or pCYP24WT (-1.4kb)-Luc and incubated with daunorubicin hydrochloride (2 µΜ), etoposide (100 µΜ), vincristine sulfate (10 µΜ), cisplatin (100 µΜ) and/or 1,25(OH)2D3 (10 nM) for 24 h. Control cells received ethanol

(0.1%, v/v). The results are presented as fold induction over control cells. Data are mean

± SEM of three independent experiments performed in triplicate. Significance of difference between control and drug or 1,25(OH)2D3: *p<0.05, **p<0.01, ***p<0.001; significance of difference between 1,25(OH)2D3 and 1,25(OH)2D3 + drug: #p<0.05,

##p<0.01, ###p<0.001 (ANOVA and Tukey-Kramer multiple comparisons test).

110 Vincristine sulphate can also increased the -298bp CYP24 promoter activity in COS-1 cells by 1.9 ± 0.3 fold over ethanol treated cells (Figure 3.2c). However, the addition of vincristine sulphate and 1,25(OH)2D3 did not further increase the -298bp CYP24 promoter activity, which is in contrast to data obtain from the -1.4 kb promoter which showed a synergistic increase in promoter activity.

In contrast to the other drug, cisplatin alone did not affect the activity of either the -298bp or the -1.4 kb CYP24 promoter. However, when incubated together with 1,25(OH)2D3, a synergistic increase in the -298bp promoter activity was observed. This amounted to a 9.2

± 1.5 fold increase over ethanol treated cells which is approximately 4.9 fold higher than

CYP24 induction by 1,25(OH)2D3 alone (Figure 3.2d).

We also tested cisplatin at a lower concentration (10 µM). The results show that cisplatin alone did not stimulate the -298bp CYP24 promoter but in the presence of 1,25(OH)2D3 the drug synergistically increased the CYP24 promoter activity which was 3 fold higher than induced by 1,25(OH)2D3 alone (data not shown). These results demonstrate that while there are subtle differences between the two promoter constructs, there is a general consensus between the two constructs.

111 3.3.3 Daunorubicin hydrochloride stimulated the –1.4kb CYP24 promoter activity in a dose-dependent manner in COS-1 kidney cells.

COS-1 cells were transiently transfected with pCYP24WT (-1.4 kb)-Luc and the pRL-0 vector and treated with 0.2, 0.5, 1 and 2 µΜ of daunorubicin hydrochloride ±

1,25(OH)2D3. Results from these experiments (Figure 3.3) show that daunorubicin hydrochloride per se can increase the –1.4 kb CYP24 promoter activity in a dose dependent manner. Increases in CYP24 promoter activity by 3.6 ± 0.8 fold and 5.3 ± 1.5 fold were observed at 1 µΜ and 2 µΜ of daunorubicin hydrochloride respectively. In the presence of 1,25(OH)2D3, the drug demonstrated an enhancing effect on the promoter activity especially at 1 and 2 µΜ of daunorubicin hydrochloride. However, the effect was not synergistic. In addition, lower doses of daunorubicin hydrochloride (0.2 µΜ and 0.5

µΜ) did not alter the ability of 1,25(OH)2D3 to stimulate the activity of the -1.4 kb

CYP24 promoter (Figure 3.3).

These results demonstrate that daunorubicin hydrochloride can stimulate the -1.4 kb

CYP24 promoter in a dose dependent manner and can enhance the induction by

1,25(OH)2D3 in COS-1 cells.

112

12 #

#

n 1,25(OH)2D3 o i

ct 9 ) du n n o I i

t c * du oter

n

I 6 om r Vehicle * (Fold

* CYP24 p 3

0

00.511.52

Daunorubicin Hydrochloride (µM)

Figure 3.3 Dose-response profile of the effects of daunorubicin hydrochloride on the activity of the –1.4kb CYP24 promoter in COS-1 cells. Cells were transiently transfected with 200 ng of pCYP24WT (-1.4kb)-Luc and incubated with daunorubicin hydrochloride (0.2 µΜ, 0.5 µΜ, 1 µΜ and 2 µΜ) ± 1,25(OH)2D3 (10 nM) for 24 h. Data shown are mean ± SEM of three independent experiments performed in triplicate.

Significance of difference between untreated cells and drug- or 1,25(OH)2D3-treated cells: *p<0.05. Significance of difference between 1,25(OH)2D3 and drug +

1,25(OH)2D3: #p<0.05 (ANOVA and Tukey-Kramer multiple comparisons test).

113 3.3.4 Dose-dependent effect of daunorubicin hydrochloride on the -298bp

CYP24 promoter in COS-1 cells and HEK293T cells.

The dose-response profile of the effect of daunorubicin hydrochloride was also investigated using the -298bp promoter construct. The results in Figure 3.4 show a dose dependent increase in the CYP24 promoter activity in COS-1 cells incubated with daunorubicin hydrochloride. At 1 µΜ of daunorubicin hydrochloride, CYP24 promoter induction was increased by 3.7 ± 0.4 fold when compared to untreated cells. The threshold of stimulation was around 0.5 µΜ. Synergistic increases in CYP24 promoter activity were observed in cells treated with 1,25(OH)2D3 and low concentrations (0.13 and 0.25 µΜ) of daunorubicin hydrochloride. However, the enhancing effect of daunorubicin hydrochloride on the effect of 1,25(OH)2D3 diminished as the concentrations of daunorubicin hydrochloride were increased.

In contrast to data obtain from COS-1 cells, daunorubicin hydrochloride alone did not stimulate the activity of the -298bp promoter in HEK293T cells in the dose range tested.

However, treatment of HEK293T cells with both 1,25(OH)2D3 and daunorubicin hydrochloride produced a synergistic increase in CYP24 promoter activity, which amounted to approximately 2.7, 3.0, 3.7 and 2.7 fold induction at 0.13, 0.25, 0.5 and 1

µΜ daunorubicin hydrochloride, respectively, over the effect of 1,25(OH)2D3 alone.

These data demonstrate that in this cell line, the predominant effect of daunorubicin hydrochloride was observed in the presence of 1,25(OH)2D3, thereby confirming that

114 COS-1 8 ##

n o 6 i # t c

n) ** o i Indu

duct 1,25(OH)2D3 oter 4 n I Vehicle (Fold * 2 CYP24 prom

0 0 0.2 0.4 0.6 0.8 1

Daunorubicin Hydrochloride (µM)

18 HEK293T ##

15 ## n ## 1,25(OH)2D3 o i t ## duc n) 12 n o I i

t c du oter

n 9 I om r

(Fold 6 ***

CYP24 p 3 Vehicle

0 0 0.2 0.4 0.6 0.8 1 Daunorubicin Hydrochloride (µM)

115 Figure 3.4 Dose-response profile of the effect of daunorubicin hydrochloride on the

-298bp CYP24 promoter in the presence or absence of 1,25(OH)2D3 in COS-1 and

HEK293T cells. COS-1 and HEK293T cells were transfected with 200 ng of vector containing pCYP24WT (-298bp)-Luc promoter. Daunorubicin hydrochloride and/or

1,25(OH)2D3 were added and promoter activity was assayed 24 h later as described under Materials and Methods. Results shown are mean + SEM of 3 experiments conducted in triplicate determinations. Significance of difference between control and drug or 1,25(OH)2D3 treatment: *p<0.05, **p<0.01, ***p<0.001; significance of difference between 1,25(OH)2D3 and 1,25(OH)2D3 + drug: #p<0.05, ##p<0.01

(ANOVA followed by Tukey-Kramer multiple comparisons test).

116 daunorubicin hydrochloride could enhance CYP24 promoter induction under our experimental conditions.

3.3.5 Etoposide stimulated the -1.4kb CYP24 promoter activity in a dose- dependent manner in COS-1 kidney cells.

Etoposide had the greatest effect amongst the other three drugs on the induction of the -

1.4 kb promoter (Figure 3.1 and 3.2). Thus, it was of interest to also investigate the relationship between varying dosage of etoposide and CYP24 induction in COS-1 cells using the -1.4 kb promoter. Cells were co-transfected with pCYP24WT (-1.4kb)-Luc and the pRL-0 vector and were then treated with 12.5 and 100 µΜ of etoposide and/or 10 nM of 1,25(OH)2D3. The data in Figure 3.5 demonstrate that 12.5 and 100 µΜ of etoposide increased the transcriptional activity of the CYP24 promoter by 2.7 ± 0.7 and 1.8 ± 0.4 fold respectively over untreated cells. Treatment of the cells with both etoposide and

1,25(OH)2D3 produced a synergistic increase in CYP24 promoter activity. These data show that etoposide per se could stimulate CYP24 promoter induction, whereas in the presence of 1,25(OH)2D3 could enhance CYP24 promoter induction.

117

36 ## ##

30 1,25(OH)2D3 n o i

ct ) du n 24 n o I i

ct er du ot n

I 18

om r old

(F P24 p 12 Y ** C

6 * Vehicle *

0 0 20 40 60 80 100

Etoposide (µM)

Figure 3.5 Effects of different concentrations of etoposide on the activity of the -

1.4kb CYP24 promoter in COS-1 cells in the presence or absence of 1,25(OH)2D3.

Cells were transiently transfected with 200 ng of pCYP24WT (-1.4kb)-Luc and incubated with etoposide (12.5 µΜ and 100 µΜ) and/or 1,25(OH)2D3 (10 nM). The results are presented as fold induction. Data shown are mean ± SEM of three independent experiments performed in triplicate. Significance of difference between control and drug or 1,25(OH)2D3: *p<0.05 and **p<0.001; significance of difference between

1,25(OH)2D3 and 1,25(OH)2D3 + drug: #p<0.05 (ANOVA followed by Tukey-Kramer multiple comparisons test).

118 3.3.6 Doxorubicin hydrochloride did not stimulate the activity of the -298bp

CYP24 promoter or enhance CYP24 induction by 1,25(OH)2D3 in COS-1 cells.

Doxorubicin hydrochloride (adriamycin), the hydroxylated derivative of daunorubicin hydrochloride, is another anthracycline anti-cancer drug that is commonly used to treat various form of cancers including leukemia. Since the above data show that daunorubicin hydrochloride increased both basal and 1,25(OH)2D3 induction of CYP24 promoter in

COS-1 cells, this experiment was conducted to determine whether doxorubicin hydrochloride could increase the transcriptional activation of the CYP24 promoter in

COS-1 kidney cells. Results from this study (Figure 3.6) show that doxorubicin hydrochloride at all concentrations tested did not increase CYP24 promoter activity or enhanced CYP24 induction by 1,25(OH)2D3 in COS-1 cells. Thus, not all drugs are able to stimulate CYP24 promoter activity. In this situation, hydroxylation of the daunorubicin structure resulted in the loss of promoter-inducing activity of the drug.

119

5

4 ion t

c u ) d n n o I i t

c 3 1,25(OH)2D3 du oter * In

om r old p 2 4

(F Vehicle 2

CYP 1

0 00.511.52 Doxorubicin Hydrochloride (µM)

Figure 3.6 Lack of effect of doxorubicin hydrochloride on the -298bp CYP24 promoter in COS-1 cells. Cells were transiently transfected with 200 ng of pCYP24WT

(-298bp)-Luc and incubated with doxorubicin hydrochloride at the indicated concentrations and/or 1,25(OH)2D3 (10 nM). Data are mean ± SEM of three independent experiments, performed in triplicate. Significance of difference between untreated and

1,25(OH)2D3 treated cells: * p<0.05. No significant differences were detected between untreated and drug treated cells or between 1,25(OH)2D3 and 1,25(OH)2D3 + drug

(ANOVA followed by Tukey-Kramer multiple comparisons test).

120

3.4 Summary

The data demonstrate that in COS-1 cells, daunorubicin hydrochloride, etoposide and vincristine sulfate per se could stimulate the -1.4kb promoter. In contrast, cisplatin did not. Similarly, in HEK293T cells daunorubicin hydrochloride and etoposide per se were both able to stimulate the -1.4kb CYP24 promoter. However, both cisplatin and vincristine sulfate did not stimulate the -1.4kb promoter in HEK293T cells.

The data with the -1.4kb promoter construct also showed that in COS-1 cells, daunorubicin hydrochloride was only able to produce an additive effect in the presence of

1,25(OH)2D3. On the other hand, etoposide, vicristine sulfate and cisplatin were able to synergistically up-regulate the -1.4kb CYP24 promoter in the presence of 1,25(OH)2D3.

However in HEK293T cells, unlike daunorubicin hydrochloride and cisplatin, both etoposide and vincristine sulfate were able to synergistically up-regulate the CYP24 promoter induction by 1,25(OH)2D3.

Results from the -298bp CYP24 promoter demonstrate that daunorubicin hydrochloride, etoposide and vincristine sulfate can also stimulate the promoter activity by itself in COS-

1 cells. In contrast, cisplatin was not able to stimulate the -298bp CYP24 promoter. These data are consistent with the data from the -1.4kb promoter.

Furthermore, the data also show that daunorubicin hydrochloride, cisplatin and etoposide in the presence of 1,25(OH)2D3, synergistically up-regulated the -298bp CYP24 promoter activity in COS-1 cells. On the other hand, vincristine sulfate was only able to cause an additive effect on the CYP24 promoter activity in the presence of 1,25(OH)2D3.

121 The effect of daunorubicin hydrochloride per se on both the -298bp and the -1.4kb promoter activity in COS-1 cells was dose dependent. In the presence of 1,25(OH)2D3, no synergistic increase in -1.4kb promoter activity was observed at all different concentrations of daunorubicin hydrochloride. In contrast, synergistic response were only observed at lower concentrations of daunorubicin hydrochloride in the -298bp promoter.

The dose response profile for daunorubicin hydrochloride was some what different between COS-1 and HEK293T cells. Unlike in COS-1 cells, daunorubicin hydrochloride alone did not affect the -298bp promoter activity in HEK293T cells. However, in the presence of 1,25(OH)2D3, a synergistic increase in CYP24 promoter activity was detected.

Although etoposide per se was less able to stimulate the -1.4kb CYP24 promoter in COS-

1 cells over the concentration range tested, etoposide was found to synergistically stimulate the -1.4kb promoter at higher concentrations (12.5 µM and 100 µM) in the presence of 1,25(OH)2D3.

The investigations with doxorubicin hydrochloride (a hydroxylated derivative of daunorubicin hydrochloride) showed that the drug did not stimulate the CYP24 promoter alone or in the presence of 1,25(OH)2D3. Similarly, investigation with dexamethasone also showed that the drug neither stimulate the CYP24 promoter alone or in the presence of 1,25(OH)2D3 (data not shown). These results demonstrate that some chemotherapeutic drugs do not stimulate CYP24 induction.

Overall, these results demonstrate that a number of chemotherapeutic drugs were able to activate the CYP24 promoter and have the potential to further stimulate CYP24 induction in the presence of 1,25(OH)2D3. Furthermore, the effects of the chemotherapeutic drugs

122 on CYP24 promoter varied somewhat between two different kidney cell-lines and different promoter lengths.

123

Chapter 4 Mechanism I

Mechanism of action of daunorubicin hydrochloride and etoposide on CYP24 promoter activity: Role of reactive oxygen species (ROS),

VDR, VDRE and Sp1

124 4.1 ROS are not required for CYP24 promoter induction by daunorubicin hydrochloride and etoposide in COS-1 cells.

As discussed in Chapter 1.10, anticancer drugs such as daunorubicin hydrochloride, etoposide and cisplatin can activate MAP kinases and PKC through the production of reactive oxygen species (ROS). It was therefore possible that these drugs could act via

ROS in the induction of CYP24 promoter. To determine if the stimulation of CYP24 promoter activity by daunorubicin hydrochloride and etoposide involved ROS, COS-1 cells were transiently co-transfected with the -298bp CYP24 promoter-Luc construct and a vector containing renilla luciferase. The cells were then treated with the anti-oxidant, N- acetyl L-cysteine (10 mM), for one hour before being treated with ethanol (EtOH),

1,25(OH)2D3 (10 –8 M), daunorubicin hydrochloride (2 µM) or etoposide (100 µM) for

24 h. This concentration of N-acetyl L-cysteine has previously been reported to be effective in inhibiting the production of oxygen radicals in cells (Ravid et al., 1999;

Gouaze et al., 2001; Kurz et al., 2004; Corna et al., 2004). The data in Figure 4.1 demonstrate that N-acetyl L-cysteine did not reduce CYP24 promoter induction by daunorubicin hydrochloride or etoposide, suggesting that ROS did not play a role in mediating the effects of the drugs on CYP24 promoter activity in the kidney cells.

However, N-acetyl L-cysteine enhanced CYP24 promoter induction by 1,25(OH)2D3, demonstrating that the anti-oxidant was active, Although the precise reason for the increase CYP24 promoter activity in

125

20 * RPMI -1640 16 N-acetyN-acetyl lL-cystein L- Cysteinee y y it it tiv tiv c c n) n) 12 oter a oter a ductio ductio m m d In d In l l 4 pro 4 pro (Fo (Fo 8 P2 P2 Y Y C C

4

0

n e 5D ci tOH 2 i id E 1, ub os or top un E Da

Figure 4.1 Effect of N-acetyl L-cysteine on CYP24 promoter induction by daunorubicin hydrochloride, etoposide and 1,25(OH)2D3 in COS-1 cells. Cells were transiently transfected with 200ng of pCYP24-WT (298bp)-Luc and treated with N-acetyl

L-cysteine (10 mM) for 1hr prior to incubation with daunorubicin hydrochloride (2 µΜ), etoposide (100 µΜ) or 1,25(OH)2D3 (10-8 M) for 24 h. Results are presented as fold induction over EtOH treated cells. Data are representative of three independent experiments performed in triplicates. Significance of difference between RPMI-1640 and

N-acetyl L-cysteine treatment. * p<0.01 (Student’s t-test)

126 the presence of N-acetyl L-cysteine remains unclear, it is possible that, ROS production in the presence of 1,25(OH)2D3 could inhibit CYP24 promoter induction by the hormone and this was relieved by N-acetyl L-cysteine. Consistent with this possibility, there are reports that 1,25(OH)2D3 can deplete cancer cells of reduced glutathione and this has been proposed as a mechanism for the anti-cancer action of 1,25(OH)2D3. 1,25(OH)2D3 has also been shown to suppress the expression of the antioxidant enzyme, Cu 2+/Zn 2+ superoxide dismutase, an enzyme that is responsible for cellular defense against ROS and this results in a built up of ROS within cells (Ravid A et al., 1999; Ravid A & Koren R

2003). Our data demonstrate that treatment of the cells with N-acetyl L-cysteine resulted in a higher induction of the CYP24 promoter activity by 1,25(OH)2D3, suggesting the generation of ROS by the hormone in the kidney cells.

4.2 Over expression of VDR did not enhance CYP24 promoter induction by daunorubicin hydrochloride in COS-1 and HEK293T cells.

In this set of experiments, only daunorubicin hydrochloride was tested. Since ectopic expression of the nuclear VDR greatly enhances the CYP24 promoter induction by

1,25(OH)2D3 (Dwivedi et al., 2002), we investigated whether the VDR was required for

CYP24 promoter induction by daunorubicin hydrochloride. COS-1 cells were co- transfected with a plasmid carrying the nuclear VDR or empty vector, pcDNA3, the -

298bp CYP24 promoter sequence fused to the firefly luceferase reporter gene

(pCYP24WT (298)-Luc) and a vector (pRL-0) expressing Renilla luciferase. The cells were treated with either daunorubicin hydrochloride, 1,25(OH)2D3 or both. The data obtained were expressed as fold induction over control. As reported previously (Dwivedi

127 et al.,2002), CYP24 promoter activation by 1,25(OH)2D3 was greatly increased in VDR- transfected cells compared to pcDNA3-transfected cells (Figure 4.2). In contrast, VDR overexpression inhibited the ability of daunorubicin hydrochloride per se to stimulate the induction of CYP24 promoter.

Similar results were observed in HEK293T cells. Thus cells transfected with the VDR exhibited a greater induction of CYP24 promoter activity than control cells when treated with 1,25(OH)2D3 (Figure 4.3) as was observed in COS-1 cells (Figure 4.2).

Interestingly, VDR overexpression did not affect promoter activity when the cells were treated with both daunorubicin hydrochloride and 1,25(OH)2D3. The synergistic effect of

1,25(OH)2D3 and daunorubicin hydrochloride was still evident in the VDR overexpressing cells.

These results suggest that unlike 1,25(OH)2D3, daunorubicin hydrochloride did not require the nuclear VDR to stimulate the CYP24 promoter. The lack of a further increase in the VDR overexpressing cells treated with 1,25(OH)2D3 and daunorubicin hydrochloride suggest that maximum induction had been reached in these cells.

128

30 **

n o i t

c u

d 20 n on) i I t

er ot

om r P

4 10 (Fold Induc 2 P * Y C

0

VDR (200ng) --++

Daunorubicin (2µM) ++ --

1,25D (10-8 M) -- ++

Figure 4.2 Effect of overexpression of VDR on CYP24 promoter induction by daunorubicin hydrochloride or 1,25(OH)2D3 in COS-1 cells. Cells were co- transfected with 200ng of pCYP24-WT (298bp)-Luc and 200ng of VDR or empty vector pcDNA3 and treated with daunorubicin hydrochloride or 1,25(OH)2D3 for 24 h. The results are presented as fold induction over that seen with EtOH-treated cells. Data are representative of three independent experiments performed in triplicates. Significance of difference between pcDNA3 and VDR transfected cells. * p < 0.05 and ** p < 0.01

(Student’s t-test).

129

50

40 ppCDNcDNA3A n

o i t

c VDR

u d n on) i I t 30 * er

ot om r

P 20 (Fold Induc 24 P Y

C 10

0

EEtOHtOH 1,25(OH)2D3 Daunorubicin Daunorubicin + 1,25(OH)2D3

Figure 4.3 Effect of over expression of VDR on CYP24 induction by daunorubicin and/or 1,25(OH)2D3 in HEK293T cells. Cells were co-transfected with 200 ng of pCYP24-WT (298bp)-Luc and 200 ng of VDR or empty vector and treated with daunorubicin hydrochloride (0.25 µM) and/or 1,25(OH)2D3 (10-8 M) for 24 h. The results are presented as fold induction over the response seen in EtOH-treated cells. Data are representative of three independent experiments performed in triplicates. Significance of difference between pcDNA3 and VDR transfected cells. * p<0.01 (Student’s t-test)

130 4.3 VDRE and GC box are not involved in CYP24 promoter induction by daunorubicin hydrochloride in COS-1 cells.

Again, only the effect of daunorubicin hydrochloride was tested in this set of experiments. The vitamin D responsive element (VDRE) and GC box element (see

Chapter 1.7.3.1) play an important role in the activation of the CYP24 promoter by

1,25(OH)2D3. The transactivation of the CYP24 promoter by 1,25(OH)2D3 requires the binding of both VDR/RXR heterodimers and specific protein 1 (Sp1) transcription factor to the two VDREs and GC box, respectively, within the CYP24 promoter. Studies in our laboratory (Dwivedi et al., 2000); Dwivedi et al., manuscript in preparation) have demonstrated that mutation of the VDREs abolishes CYP24 promoter induction by

1,25(OH)2D3 and mutation of the GC box reduces basal as well as 1,25(OH)2D3- stimulated promoter activity. We therefore investigated whether these promoter regions were required for promoter induction by daunorubicin hydrochloride.

Cells were transiently transfected with wild type CYP24 (pCYP24WT-Luc), mutant

VDRE1 and VDRE2 (pCYP24mVDRE1+2–Luc) or CYP24 promoter with mutated GC box (pCYP24mGC-Luc) and a vector expressing Renilla luciferase. Cells were treated with 1,25(OH)2D3 or daunorubicin hydrochloride for 24 h. As expected, CYP24 promoter induction by 1,25(OH)2D3 was significantly reduced when VDRE1+2 or the

GC box was mutated (Figure 4.4). However, the ability of daunorubicin hydrochloride to stimulate CYP24 promoter activity was not affected by mutation of VDRE1+2 or the GC box. These results demonstrate that unlike 1,25(OH)2D3, daunorubicin hydrochloride did not require the VDRE or GC box to stimulate CYP24 promoter activity.

131

18 pCYP24WT-Luc

pCYP24mVDRE-Luc 15 pCYP24mGC-Luc

on

ti c 12 u d n) In io

t er

ot 9 om d Induc r l 4 P (Fo 2 6 P

Y C

3 **

0 EtOH 1,25(OH)2D3 Daunorubicin

Figure 4.4 Effects of mutant VDRE and mutant GC-box on CYP24 induction by daunorubicin hydrochloride and 1,25(OH)2D3 in COS-1 cells. COS-1 cells were co- transfected with wild type pCYP24WT (-298bp)-Luc, a construct carrying mutant

VDRE1+2 (pCYP24mVDRE-Luc) or a vector carrying mutant GC box (pCYP24mGC-

Luc). Cells were treated with 1,25(OH)2D3 (10-8 M) or daunorubicin hydrochloride (2

µM) for 24 h and luciferase activity determined. Results are presented as fold induction, over EtOH-treated cells. Significance of difference between CYP24WT, CYP24mVDRE and CYP24mGC-box. * p <0.001 (ANOVA and Dunnet multiple comparisons test).

132

4.4 Daunorubicin hydrochloride did not alter the activity of the VDR in COS-1 cells

To confirm that daunorubicin hydrochloride did not have a direct effect on the activity of

VDR itself, COS-1 cells were transiently transfected with a vector which contained the consensus vitamin D response element (VDRE) of the DR3-type (two AGGTCA motifs separated by 3bp) cloned upstream of the TK promoter driving the firefly luciferase, and pRL-0 carrying Renilla luciferase. The data in Figure 4.5 demonstrate that daunorubicin hydrochloride did not affect VDR activity because the drug did not affect reporter expression when added alone or in combination with 1,25(OH)2D3. As expected,

1,25(OH)2D3 stimulated the activity of the VDR. The smaller response seen in the presence of the hormone was due to the absence of accessory elements such as the Ets binding site (please see Chapter 5), GC-Box and the VSE (Nutchey et al, 2005) in this construct. The data imply that daunorubicin hydrochloride does not act via the VDREs and that elements other than the VDREs are required for synergism.

133

3

ion * 2 * duct

In er t

old Induction)

(F 1

DR3 promo

0 Daunorubicin - - + +

1,25(OH)2D3 - + - +

Figure 4.5 Effect of daunorubicin hydrochloride on the activity of VDR in COS-1 cells. COS-1 cells were transiently transfected with a vector which contained the consensus vitamin D response element (VDRE) of the DR3-type and pRL-0. Cells were treated with 1,25(OH)2D3 (10-8 M) ± daunorubicin hydrochloride (2 µM) for 24 h and luciferase activity determined. Results are presented as fold induction, over EtOH-treated cells. Significance of difference between control and 1,25(OH)2D3-, or 1,25(OH)2D3 ± daunorubicin hydrochloride-treated cells p<0.01 (ANOVA followed by Dunnet’s multiple comparisons test).

134 4.5 Daunorubicin hydrochloride and etoposide increased the expression of VDR in COS-1 and HEK293T cells.

The data in Figure 4.3 demonstrate that the ability of the cells to respond to 1,25(OH)2D3 was enhanced when the level of the VDR was increased. It is therefore possible that the drugs could enhance CYP24 promoter induction by 1,25(OH)2D3 by increasing the level of VDR in the kidney cells. We therefore investigated whether daunorubicin hydrochloride and etoposide could increase the expression of VDR in these cells. COS-1 cells were incubated with daunorubicin hydrochloride or etoposide for 24 h, lysed and the expression of VDR was determined by Western blotting using an anti-VDR antibody. The data in Figure 4.6 demonstrate that COS-1 cells treated with daunorubicin hydrochloride or etoposide had increased VDR expression. Equal protein loading was also demonstrated by reprobing with anti β-actin antibody (Figure 4.6).

Similar results were obtained in HEK293T cells treated with daunorubicin hydrochloride or etoposide, where both drugs caused an increase in VDR expression compared to ethanol-treated cells (Figure 4.7). These results could explain, at least in part, the enhancing effects of daunorubicin hydrochloride and etoposide on CYP24 promoter induction by 1,25(OH)2D3.

135

EtOH Daunorubicin Etoposide

Figure 4.6 Effects of daunorubicin hydrochloride and etoposide on vitamin D receptor (VDR) expression in COS-1 cells. COS-1 cells were treated with ethanol

(EtOH), daunorubicin hydrochloride (2 µΜ) and etoposide (100 µM) for 24 h. The cells were lysed and samples were Western blotted using an anti-VDR antibody (upper panel).

The blots were stripped and reprobed with anti-β-actin antibody to assess loading (lower panel). The data above are representative of 3 experiments with similar results.

136

EtOH Daunorubicin Etoposide

Figure 4.7 Effects of daunorubicin hydrochloride and etoposide on vitamin D receptor (VDR) expression in HEK293T cells. HEK293T cells were treated with ethanol (EtOH), daunorubicin hydrochloride (0.25 µΜ) and etoposide (100 µM) for 24 h.

The cells were lysed and samples were Western blotted using an anti-VDR antibody

(upper panel). The blots were stripped and reprobed with anti-β-actin antibody to assess loading (lower panel). The data above are representative of 3 experiments with similar results.

137 4.6 Summary

Overall, the above results demonstrate that the drugs tested did not act via ROS,

VDRE1/VDRE2 or the GC Box to stimulate the CYP24 promoter activity. However, an increase in VDR expression caused by daunorubicin hydrochloride and etoposide could mediate some of the enhancing effects of the drugs on promoter induction by

1,25(OH)2D3.

138

Chapter 5 Mechanism II

Role of MAP Kinases in 1,25(OH)2D3 mediated transactivation regulation of the CYP24 promoter in

COS-1 cells

139 5.1 Introduction

The data in Chapter 3 show that a selection of chemotherapeutic drugs could stimulate

CYP24 promoter activity and in the presence of 1,25(OH)2D3 cause a synergistic up- regulation of the CYP24 promoter activity. To understand how these chemotherapeutic drugs up-regulated the CYP24 promoter, it is important to first understand how

1,25(OH)2D3 regulates the CYP24 promoter.

The classical mode for the induction of CYP24 by 1,25(OH)2D3 involves the binding of

1,25(OH)2D3 to VDR-RXR heterodimer, which in turn binds to vitamin D response elements (VDRE) to mediate the activation of the CYP24 promoter. The rat CYP24 promoter has two VDRE (Hahn et al 1994), VDRE1 at –150/-136 and VDRE2 at –258/-

244. The two VDREs can synergise with each other in response to 1,25(OH)2D3 (Kerry et al., 1996). Although VDRE1 has a lower binding affinity than VDRE2 towards the

VDR-RXR complex, it is able to mediate 1,25(OH)2D3 induction of the CYP24 promoter to a greater extent than VDRE2. Computer analysis of the promoter region proximal to

VDRE1 led to the identification of an Ets-binding site (EBS) (-128/-119) located downstream of VDRE1. Mutation to the EBS resulted in a significant reduction in CYP24 promoter activity, indicating that the EBS plays an important role in the transcriptional activation of the CYP24 promoter especially at lower concentrations of 1,25(OH)2D3

(Dwivedi et al., 2000). Studies by Dwivedi et al. (2000) also showed that over expression of Ets-1 could enhance 1,25(OH)2D3 induction of a truncated CYP24 (-186bp) promoter containing just VDRE1 and the adjacent EBS. In our studies conducted in collaboration with Dwivedi et al. (2002), we showed that transfection of kidney cells with the Ets-1

140 mutant, Ets-1 T38A, or mutant EBS reduced 1,25(OH)2D3-mediated induction of the

CYP24 promoter by similar extents. Mutation to both VDRE1 and VDRE2 reduced promoter activity to near basal levels. These data show that, unlike the EBS, VDRE1 and

VDRE2 are crucial for the transcriptional activation of the CYP24 promoter. Over expression of exogenous RXR did not further increase CYP24 promoter activity whereas mutation of RXR (RXRS260A) significantly reduced promoter activity. These data suggest that, although RXR plays an important role in the induction of the CYP24 promoter, unlike Ets-1, RXR is not limiting in these cells.

1,25(OH)2D3 has been shown to activate the mitogen-activated protein (MAP) kinases and protein kinase C (PKC) (Chapter 1; Lissoos et al., 1993; Beno et al., 1995; Schwartz et al., 2002). However, the roles these MAP kinases play in the activation of the CYP24 promoter by 1,25(OH)2D3 are poorly understood. Furthermore, the targets of the MAP kinases within the context of CYP24 promoter activation by 1,25(OH)2D3 are not known. In addition to ERK1/ERK2, recent studies have demonstrated that the recently discovered MAP kinase, ERK5, is expressed in many cell-types and many ligands which stimulate ERK1/ERK2 also stimulate ERK5 activity. However, this has not been reported for 1,25(OH)2D3. These issues will be addressed in this Chapter in collaboration with Dr

Prem Dwivedi and Prof Brian May (School of Molecular Bioscience, Adelaide

University). These data have been presented by Dwivedi et al. (2002).

141 5.1.1 Vectors used:

A number of vectors are described in this Chapter. (See also Dwivedi et al. (2002) for further description) These include:

pCYP24WT (-298bp)-Luc: Wild type CYP24 promoter–reporter construct. pRL-TK: Renilla luciferase control reporter vector pRSVhVDR: Human VDR construct pRSVhRXRα: Human RXRα construct pRSVhRXRα S206A: Human RXR mutant construct pCYP24mEBS (-298bp)-Luc: CYP24 promoter-reporter construct with mutated Ets binding site. pCMVERK1K71R: Dominant negative ERK1 mutant pSRαMEK5A: Dominant negative MEK5 mutant. pZIP-Ras(17N): Dominant negative Ras mutant.

142 5.2 Results:

5.2.1 1,25(OH)2D3 stimulated phosphorylation of ERK1/ERK2 in COS-1 cells.

We examined whether 1,25(OH)2D3 could stimulate the activation / phosphorylation of

ERK1/ERK2 in COS-1 cells. We first determined the optimum time for ERK1/ERK2 stimulation by 1,25(OH)2D3. COS-1 cells were incubated with 1,25(OH)2D3 (10-7 M) for 0, 1, 5, 15 and 30 min. Cells were also incubated for 5 min with PMA (10-7 M), a

-3 standard stimulator for ERK1/ERK2 (positive control), and H2O2 (10 M), which can stimulate ERK1/ERK2 activity in some but not all cell-types. However, H2O2 is also a standard activator of ERK5 (see below). Cells were lysed and lysates were western blotted for phosphorylated ERK1/ERK2. As expected, results from cells treated with

PMA showed increased phosphorylation of ERK1/ERK2 whereas no change in the phosphorylation of ERK1/ERK2 was detected in cells treated with H2O2 (Figure 5.1).

Cells treated with 1,25(OH)2D3 showed increased ERK1/ERK2 phosphorylation at 1, 5,

15 and 30 min treatment time. Maximum ERK1/ERK2 phosphorylation was observed at

5 min. Equal amounts of ERK1/ERK2 were present in each lane, (confirmed by stripping and reprobing with anti-ERK2 antibody (Figure 5.1, lower panel). This shows that the increase in the level of phosphorylated ERK1/ERK2 was not due to a greater amount of

ERK1/ERK2 being loaded.

143

Figure 5.1 Effects of different 1,25(OH)2D3 treatment times on the phosphorylation of ERK1/ERK2 in COS-1 cells. COS-1 cells were treated with 1,25(OH)2D3 (10-7 M) for 0, 1, 5, 15, 30 min. Cells were also stimulated with PMA (10-7 M) (positive control) and H2O2 (10-3 M) (negative control) for 5 min. Cells were lysed and samples western blotted using an anti-active ERK antibody to determine the amount of phosphorylated

ERK1/ERK2 (upper panel). The blot was stripped and reprobed with anti-ERK2 antibody to assess loading (lower panel). The experiment was repeated twice and similar results were observed.

144

We then tested whether phosphorylation of ERK1/ERK2 was dose-related. COS-1 cells were treated with the biologically inactive 25(OH)D3 (10-7 M), ethanol and 1,25(OH)2D3

(10 –10 M, 10 –9 M and 10 -7 M) for 5 min. Cells were lysed and lysates western blotted for phosphorylated ERK. The data in Figure 5.2 show that 25(OH)D3 did not affect the phosphorylation of ERK1/ERK2. Significant increases in ERK1/ERK2 phosphorylation were detected at 10 –9 M and 10 -7 M of 1,25(OH)2D3. However, no increase in

ERK1/ERK2 phosphorylation was observed in cells treated with 10 –10 M of

1,25(OH)2D3. These results demonstrate that 1,25(OH)2D3 stimulated the phosphorylation of ERK1/ERK2 in COS-1 cells at concentrations greater than 10–10 M and that maximum ERK1/ERK2 phosphorylation occurs at 5 min.

5.2.2 1,25(OH)2D3 stimulated the activity of ERK5 activity in COS-1 cells

We also investigated whether 1,25(OH)2D3 could stimulate ERK5 activity in COS-1 cells. We first investigated the kinetics of ERK5 activation by 1,25(OH)2D3. COS-1 cells were transfected with HA-ERK5 and treated with H2O2, a well known stimulator of

ERK5 (positive control), ethanol (control) or 1,25(OH)2D3 (10-7 M) for 2, 10, 15 and 30 min. After lysis, HA-ERK5 was immunoprecipitated with anti-HA antibody and kinase assay was performed using MBP as a substrate. ERK5 activity was quantified using an

Instant Imager. As expected, ERK5 activity was strongly stimulated in H2O2 treated cells

(Figure 6.3). Cells incubated with 1,25(OH)2D3 for 2 min did not show an increase in

145

pERK1 pERK2

ERK2 WB: αERK2

C M M

D3 M) M

0 -7 -9 -7 -1 0 10 10 10 (1 25(OH) 1,25(OH)2D3

Figure 5.2 Effects of different concentrations of 1,25(OH)2D3 on the phosphorylation of ERK1/ERK2 in COS-1 cells. COS-1 cells were treated with vehicle

Control, 25(OH)D3 (10 –7 M) or 1,25(OH)2D3 (10 –10 M, 10 -9 M and 10 -7 M). Cells were lysed and samples western blotted using an anti-ERK antibody (upper panel). The blot was stripped and reprobed with anti-ERK2 antibody (lower panel) to assess loading.

The experiment was repeated twice and similar results were observed.

146

MBP

12000

10000 y

t

i

v

i t

c

) 8000

A

m 5

p

c

K

(

R E

- 4000 A

H

2000

0 H2O2 C 2 10 15 30

1,25(OH)2D3 (minutes)

Figure 5.3 Kinetics of 1,25(OH)2D3-stimulated activation of ERK5 in COS-1 cells.

COS-1 cells were transfected with HA-ERK5 and stimulated with H2O2 (1 mM) for 15 min or with 1,25(OH)2D3 (10–7 M) for 2, 10, 15 and 30 min. Cells were lysed and HA-

ERK5 immunoprecipitated. Kinase activity was assayed as described in Materials and

Methods. Results are representative of three experiments.

147 ERK5 activity compared to control. However, significant increases in ERK5 activity were observed at 10, 15 and 30 min. Peak ERK5 activity was detected at 15 and 30 min (Figure

5.3).

We then investigated whether the stimulation of ERK5 activity by 1,25(OH)2D3 was dependent on the concentration of the hormone. COS-1 cells were transfected with HA-

ERK5 and treated with ethanol, 10-7 M of 25(OH)D3 or 10-10 M, 10-9 M and 10-7 M of

1,25(OH)2D3 for 15 min. As observed for ERK2 activity, ERK5 activity was not affected by 25(OH)D3. However, 1,25(OH)2D3 caused a dose-dependent increase in ERK5 activity (Figure 5.4), such that 10-7 M > 10-9 M > 10-10 M.

5.2.3 Dominant negative ERK1 and MEK5 inhibited 1,25(OH)2D3 mediated induction of the CYP24 promoter.

To investigate whether the ERK1 and ERK5 modules were involved in regulating the

1,25(OH)2D3-induction of the CYP24 promoter, COS-1 cells were transiently transfected with pRSVh VDR, pCYP24WT(-298bp)-Luc, dominant negative (DN)ERK1 (ERK1

K71R), DN MEK5 (MEK5A) or an empty vector. MEK5 is the immediate upstream activator of ERK5 (Sun et al., 2001). Cells were treated with 1,25(OH)2D3 (10-7 M) for

24 h. Results from these experiments (Figure 5.5), show that DN ERK1 significantly reduced the 1,25(OH)2D3 induction of the CYP24 promoter activity from approximately

26 fold to 6 fold. Note that the level of induction is greater than that seen in the results

148

MBP

12000

y y

t t

i i

v v i

i 8000

t t

c c

) )

A A

m m

5 5

p p

c c

K K

( (

R R

E E

- - A

A 4000

H H

0 E 2 1 1 1 5 0 tO 0 0 ( – – – H O 7 9 1 0 H M M ) M D 3

1,25(OH)2D3

Figure 5.4 Effects of varying concentrations of 1,25(OH)2D3 on ERK5 activity in

COS-1 cells. COS-1 cells were transfected with HA-ERK5 and treated with ethanol,

25(OH)D3 (10-7 M), or different concentrations of 1,25(OH)2D3 (10-7 M, 10–8 M, and 10-

10 M) for 15 min. Cells were lysed and HA-ERK5 immunoprecipitated. Kinase activity was assayed as described in Materials and Methods. Results are representative of three experiments.

149

30

25

n

tio )

n 20 duc io t

In * r c

e t 15

d Indu

ol F 4 promo ( 10

P2 ** **

CY 5

0 pCYP24WT(-298bp)-Luc + + + + ERK1 K71R - + - + MEK5A - - + +

Figure 5.5 Effects of dominant negative ERK1 and MEK5 mutants on 1,25(OH)2D3 mediated induction of the CYP24 promoter. COS-1 cells were co-transfected with 600 ng of pCYP24WT(-298bp)-Luc, 150 ng of pRL-TK and 500 ng of dominant negative

(DN) ERK1 (pCMVERK1 K71R), DN MEK5 (pSRαMEK5A) or a blank vector (-).

Cells were treated with 1,25(OH)2D3 (10-7 M) for 24 h. The results presented here are the mean ± S.D of three independent experiments. Significance of difference between control and (DN) ERK1 or/and DN MEK5. *p<0.05, **p<0.001 (ANOVA and Tukey’s multiple comparison test)

150 presented in the other chapters. This is because these cells were co-transfected with VDR.

DN MEK5 also suppressed CYP24 promoter activity from 26 fold to 14 fold. No further reduction in CYP24 promoter activity was observed in cells transfected with both DN

ERK1 and DN MEK5 compared to cells transfected with just DN ERK1 or DN MEK5.

These data demonstrate that both ERK1 and MEK5 played a role in 1,25(OH)2D3- mediated induction of the CYP24 promoter. Preliminary dose response studies have indicated that 500 ng of DN vectors produced the maximum degree of inhibition. The greater degree of inhibition by DN ERK1 than DN MEK5 suggests that the ERK1/ERK2 module may be more important than the MEK5-ERK5 module in the regulation of the

CYP24 promoter activity by 1,25(OH)2D3.

5.2.4 Dominant negative Ras inhibited the stimulation of ERK1/ERK2 activity by 1,25(OH)2D3 in COS-1 cells.

Ras is one of the upstream regulators of the ERK1/ERK2 modules (Wennerberg et al.,

2005) and it is not known whether 1,25(OH)2D3 utilizes Ras or other upstream regulators of the ERK1/ERK2 module, such as protein kinase C, to activate this module. We therefore investigated the effect of DN Ras (Ras 17N) on ERK1/ERK2 activity stimulated by 1,25(OH)2D3. COS-1 cells were cotransfected with HA-ERK2 ± DN Ras or empty vector. Cells were treated with 1,25(OH)2D3 (10-7 M) for 5 min, lysed and HA-ERK2 immunoprecipitated. Kinase assay was performed using myelin basic protein (MBP) as a

151 substrate and activity measured and quantified using an Instant Imager. Results from this set of experiments (Figure 5.6), show that 1,25(OH)2D3 increased HA-ERK2 activity when compared to untreated cells and that HA-ERK2 activity diminished in cells co- transfected with DN Ras. These results demonstrate that 1,25(OH)2D3 acted via Ras to stimulate the activity of ERK1/ERK2 in COS-1 cells.

5.2.5 Dominant negative Ras inhibited the stimulation of ERK5 activity by

1,25(OH)2D3 in COS-1 cells.

The data show that Ras was required for 1,25(OH)2D3 to stimulate ERK1/ERK2 activity

(Figure 5.6). We also investigated whether the stimulation of ERK5 activity by

1,25(OH)2D3 was dependent on Ras. COS-1 cells were co-transfected with HA-ERK5 ±

DN Ras (Ras17N) or empty vector and treated with ethanol or 1,25(OH)2D3 (10-7 M) for

15 min. The data in Figure 5.7 show that 1,25(OH)2D3 significantly increased ERK5 activity compared to ethanol-treated cells. However, 1,25(OH)2D3 did not cause an increase in ERK5 activity in cells transfected with DN Ras (Figure 5.7). This result demonstrates that Ras was required for the stimulation of ERK5 activity by 1,25(OH)2D3 in COS-1 cells.

152

MBP

10000

y 8000

t i

v

i

t

c )

A

m 6000

2

p

c K

(

R E -

A 4000

H

2000

0

1,25(OH)2D3 - + +

Ras17N - + -

Figure 5.6 Effect of dominant negative Ras on the stimulation of ERK1/ERK2 activity by 1,25(OH)2D3 in COS-1 cells. COS-1 cells were transfected with HA-ERK2

± DN Ras (Ras17N). After 24 h, the cells were treated with 1,25(OH)2D3 (10 –7 M) for 5 min, lysed and HA-ERK2 immunoprecipitated using an anti-HA antibody. Kinase activity was assayed as described in Materials and Methods. Results are representative of three experiments.

153

MBP

16000

y y

t t

i i

v v

i i t

t 12000

c c

) )

A A

m m

5 5

p p

c c

K K

( (

R R

E E -

- 8000

A A

H H

4000

0 1,25(OH)2D3 - + + Ras17N - + -

Figure 5.7 Effect of dominant negative Ras on 1,25(OH)2D3 stimulation of ERK5 activity in COS-1 cells. COS-1 cells were transfected with HA-ERK5 ± DN Ras

(Ras17N) or empty vector. Cells were treated with 1,25(OH)2D3 (10 –7 M) for 15 min, lysed and HA-ERK5 immunoprecipitated. Kinase activity was assayed as described in

Materials and Methods. Results are representative of three experiments.

154 5.2.6 Dominant negative Ras inhibited CYP24 promoter induction by

1,25(OH)2D3

Since dominant negative Ras blocked the activation of ERK1/ERK2 and ERK5 by

1,25(OH)2D3, we investigate whether Ras also regulate CYP24 promoter activation by

1,25(OH)2D3. To investigate the possible involvement of endogenous Ras activity, we have overexpressed an inhibitory dominant-negative Ras construct, Ras17N. The results show that expression of Ras17N strongly reduced 1,25(OH)2D3 induction of the CYP24 promoter activity from 26- to 3.5-fold at the highest concentration tested (Figure 5.8).

This results demonstrates that Ras plays an important role in 1,25(OH)2D3-dependent transactivation of CYP24 promoter in COS-1 cells.

5.2.7 ERK2 interacted with RXRα independently of 1,25(OH)2D3.

Protein-protein interaction plays a crucial role in cell biology, especially for enzyme- substrate relationships. We therefore investigated whether an interaction could be detected between the MAP kinases and their effectors such as transcription factors. An interaction between the ERKs and RXRα has not been reported before. COS-1 cells were transfected with HA-RXRα and treated with ethanol (control) or 1,25(OH)2D3 (10-7 M) for 0.5, 2 and 24 h. After cell lysis, HA-RXRα was immunoprecipitated using anti-HA antibody and western blot for the possible presence of co-immunoprecipitated ERK2 using anti-ERK2 antibody. The results show the presence of ERK2 in all lanes (Figure

5.9), demonstrating that ERK2 interacted with RXR independently of 1,25(OH)2D3 and

155

35

n 30

io t

) 25 n duc

In r ctio 20

e t 15 d Indu

ol F

( 10

5 CYP24 promo

0

Ras 17N (ng) 050 100 200 50000

Figure 5.8 Effect of dominant negative Ras on 1,25(OH)2D3 induction of CYP24 promoter in COS-1 cells. pCYP24WT(-298bp)-Luc were co-transfected into COS-1 cells together with increasing doses of Ras 17N expression clone (100-500 ng). Cells were treated with 10-7 M of 1,25(OH)2D3 for 24 h. The results presented here are the mean ± S.D of three independent experiments.

156

EtOH 0.5 2 24 (Hours) 1,25(OH)2D3

Figure 5.9 Effect of 1,25(OH)2D3 on the interaction between RXRα and ERK2 or

ERK5 in COS-1 cells. COS-1 cells were transfected with HA-RXRα and treated with ethanol or 1,25(OH)2D3 (10-7 M) for 0.5, 2 and 24 h. Cells were lysed, immunoprecipitated with anti-HA antibody and western blotted with anti-ERK2 antibody.

Blots were strip and reprobed for ERK5 using anti-ERK5 antibody. To assess protein loading, the blots were stripped and reprobed with anti-HA antibody. Results are representative of two experiments.

157 incubation time. To determine whether ERK5 interacted with HA-RXRα, the blots were stripped and reprobed with anti-ERK5 antibody. The western blots did not show the presence of ERK5 in the immunoprecipitates. This demonstrates that ERK5 did not interact with HA-RXRα. To confirm that similar amounts of HA-RXRα were present in all lanes, the blots were stripped and reprobed with anti-HA antibody. Results show that similar amounts of HA-RXRα were loaded in all lanes. These data imply that the docking domain of RXR is specific for ERK2.

5.2.8 Phosphorylation of RXRα was dependent on the interaction between activated ERK1/ERK2 and RXRα.

It has been previously reported that RXRα becames phosphorylated when keratinocytes are treated with 1,25(OH)2D3 (Solomon et al., 1999). To investigate whether

1,25(OH)2D3 also stimulates RXRα phosphorylation and whether phosphorylation of

RXR is dependent on ERK1/ERK2, COS-1 cells were cotransfected with HA-RXRα ±

DN ERK1 (ERK1K71R). Cells were treated with ethanol (control) or 1,25(OH)2D3 (10-7

M) for 2 h. Phosphorylation of RXRα was determined by western blotting with an anti- phosphoserine antibody. The data in Figure 5.10 show that 1,25(OH)2D3-treated cells contained higher levels of phosphorylated RXR when compared to control (lower panel, duplicate samples shown for 1,25(OH)2D3-treated cells). However, in cells cotransfected with DN ERK1, no difference in RXR phosphorylation was detected when compared to

158

Figure 5.10 Effects of 1,25(OH)2D3, dominant negative ERK1 and mutated RXRα on serine phosphorylation of RXRα in COS-1 cells. COS-1 cells were co-transfected with HA-RXRα and ERK1K71R or HA-RXRS260A before being stimulated with

1,25(OH)2D3 (10-7 M) for 2 h. After lysis, cell lysates were immunoprecipitated with anti-HA antibody and probed for RXR phosphorylation by western blotting with an anti- phosphoserine antibody. To assess the level of protein loading, blots were stripped and reprobed with anti-HA antibody. Results are representative of three experiments.

159 control. These results show that phosphorylation of RXR by 1,25(OH)2D3 required activated ERK1. We also examined whether mutation of RXRα at S260, by replacing the serine residue with alanine, (S260A), affected phosphorylation of RXR. The data show that in cells transfected with HA-RXRα S260A, 1,25(OH)2D3 did not increase the phosphorylation of RXR when compared to cells treated with ethanol (Figure 5.10). The above data demonstrate that ERK1/ERK2 mediated the phosphorylation of RXRα at

S260 in cells incubated with 1,25(OH)2D3. The data also suggest that 1,25(OH)2D3- induced activation of ERK1/ERK2 and phosphorylation of RXRα on S260 might be important for the activation of the CYP24 promoter by the hormone.

5.2.9 RXR phosphorylation played an important role in CYP24 induction by

1,25(OH)2D3.

To investigate the importance of RXR phosphorylation in 1,25(OH)2D3 induction of the

CYP24 promoter, COS-1 cells were transfected with pCYP24WT(-298bp)-Luc, pRSVhVDR, pRSVhRXRα, or pRSVhRXRα S206A. Cells were treated with

1,25(OH)2D3 (10-7 M) for 24 h. Results from these experiments (Figure 5.11) show that overexpression of wild type RXRα did not influence the CYP24 induction by

1,25(OH)2D3. However, the 1,25(OH)2D3 induction was inhibited in cells transfected with mutant RXRα S260A. These results demonstrate that phosphorylation of RXRα at

160 difference betweenhRXRandS260A results presentedherearem each ofpCYP24W promoter inCOS-1cells Figure 5.11EffectofmutantRXRS260A pRSVhRXR

α CYP24 promoter Induction RX S206A.Thecellsweretr

R S260A (Fold Induction) T RXR ( - 298bp)-Luc ande . Cellswereco-transfected ean ± S. - - - - D ofthreeindependent eated with1,25(OH)2D3(10 m : *p<0.01(Student’sunpairedt-test) pty vector,pRSVhVDR,pRSVhRXR + + - - on1,25(OH)2D3inductionofC with 150ngofpRL-TK,600 * + + - - experim e -7 nts. Significanceof

M ) for24h.The Y P24 16 α 1 ,

serine 260 by ERK1/ERK2 is required for RXRα functionality during 1,25(OH)2D3 induction of the CYP24 promoter.

5.2.10 Interaction between ERK5 and Ets-1 was 1,25(OH)2D3-dependent.

Since Ets-1 plays an important role in enhancing 1,25(OH)2D3 induction of the CYP24 promoter (Dwivedi et al., 2000), the following experiments were performed to examine whether ERK1/ERK2 or ERK5 interacted with the Ets-1 protein in the presence of

1,25(OH)2D3. To test this, COS-1 cells were transfected with HA-Ets-1 and treated with

1,25(OH)2D3 (10-7 M) for 2 and 24 h. The cells were lysed and HA-Ets-1 immunoprecipitated using the anti-HA antibody. Western blotting was performed to determine the presence of co-immunoprecipitated ERK2 and ERK5. The data in Figure

5.12 show that there was no interaction between ERK5 and Ets-1 in untreated cells.

However, a faint ERK5 band was detected in cells treated with 1,25(OH)2D3 for 2 h.

Strong interaction between ERK5 and Ets-1 was clearly observed in cells incubated with

1,25(OH)2D3 for 24 h. A weak interaction between ERK2 and Ets-1 was detected but this interaction was independent of 1,25(OH)2D3 treatment. To confirm that the lanes contained similar amounts of Ets-1 protein, the blots were stripped and reprobed with anti-Ets-1 antibody. Results show that there were no differences in Ets-1 protein loading between the lanes. The data from this set of experiments demonstrate that the interaction between ERK5 and Ets-1 transcription factor was regulated by 1,25(OH)2D3. In contrast,

162 EtOH 2 24 (Hours) 1,25(OH)2D3

Figure 5.12 Effects of 1,25(OH)2D3 on the interaction between Ets-1 and ERK2 or

ERK5 in COS-1 cells. COS-1 cells were transfected with HA-Ets-1 and treated with

1,25(OH)2D3 (10-7 M) for 2 or 24 h. Cells lysates were prepared and HA-Ets-1 immunoprecipitated with anti-HA-antibody and probed for the presence of ERK5 by western blotting using an anti-ERK5 antibody. Blots were stripped and reprobed for

ERK2 using an anti-ERK2 antibody. The levels of Ets-1 protein in the lanes were determined by stripping and reprobing with an anti-Ets-1 antibody. Results were representative of three experiments.

163 the weak interaction between ERK2 and Ets-1 was constitutive and not regulated by

1,25(OH)2D3.

5.2.11 1,25(OH)2D3 stimulated the phosphorylation of Ets-1 protein.

Analysis of the sequence of Ets-1 revealed that the transcription factor contains the MAP kinase docking site motif, LXL, where X is any amino acid. This docking site can be found both upstream and downstream of the consensus Ser/Thr-Pro (S/TP) MAP kinase phosphorylation site. Thus it is possible that 1,25(OH)2D3 could stimulate Ets-1 phosphorylation via ERK5. To investigate whether 1,25(OH)2D3 increased the phosphorylation of Ets-1, COS-1 cells were transfected with HA-Ets-1 and treated with

1,25(OH)2D3 (10-7 M) for 2, 4 and 24 h. Results based on western blotting using an anti- phosphothreonine antibody show a very low level of threonine phosphorylation of Ets-1 in ethanol treated cells. However, a gradual increase in the level of threonine phosphorylation was evident in cells incubated with 1,25(OH)2D3, with an increase being detectable at 2 h and this increased over a 24 hour period (Figure 5.13). No difference in

Ets-1 protein loading was detected in the lanes, demonstrating that the increase in threonine phosphorylation in 1,25(OH)2D3 was not due to higher levels of Ets-1 being loaded.

164

EtOH 2 4 24 (Hours) 1,25(OH)2D3

Figure 5.13 1,25(OH)2D3 stimulated the threonine phosphorylation of Ets-1 in

COS-1 cells. COS-1 cells were transfected with HA-Ets-1 and treated with ethanol or

1,25(OH)2D3 (10 -7 M) for 2, 4 and 24 h. After cell lysis, HA-Ets-1 was immunoprecipitated with anti-HA antibody and assessed for equal Ets-1 protein loading by western blotting with an anti-Ets-1 antibody. To determine the level of Ets-1 phosphorylation, blots were stripped and reprobed with anti-phosphothreonine antibody.

Results were representative of three experiments.

165 5.2.12 Phosphorylation of Ets-1 requires the activation of HA-ERK5 by

MEK5 and Ras.

In Figure 5.12, we demonstrated that unlike ERK2, ERK5 interacted with Ets-1 in a

1,25(OH)2D3-dependent manner. To determine whether Ets-1 is a substrate for ERK5,

COS-1 cells were co-transfected with HA-ERK5 or HA-ERK2 ± Ras17N or MEK5(A).

The cells were then treated with 1,25(OH)2D3 or ethanol (control) for 5 min (HA-ERK2) or 15min (HA-ERK5). After lysis, HA-tagged kinase was immunoprecipitated, and kinase activity was determined using His-Ets-1 as a substrate. Figure 5.14 shows that HA-

ERK5 from 1,25(OH)2D3-treated cells phosphorylated Ets-1. Dominant negative MEK5 and dominant negative Ras inhibited the phosphorylation of Ets-1 by HA-ERK5. In contrast, HA-ERK2 did not phosphorylate Ets-1 even in the presence of 1,25(OH)2D3.

Thus, the data show that ERK5 and its upstream activators play an important role in mediating the phosphorylation of Ets-1 whereas Ets-1 is not a target of ERK1/ERK2.

Ets-1 contains a putative ERK phosphorylating site at threonine 38. To investigate whether this is indeed an ERK5 phosphorylating residue, we mutated threonine 38 of Ets-

1 to alanine (GST-Ets-1T38A). Figure 5.15 show that ERK5 was not able to phosphorylate Ets-1 when threonine 38 of Ets-1 was substituted with alanine. This study demonstrates that threonine 38 of Ets-1 is an important target for ERK5.

166

HA-ERK5 HA-ERK2

His-Ets-1

4000

) 3000

(cpm y vit

ti c a 2000

se na i

K

1000

1,25(OH)2D3 + - + + - + MEK5(A) - - + - - -

Ras 17N - - - + - -

Figure 5.14 Phosphorylation of Ets-1 by ERK5 but not by ERK2 and the role of

MEK5 and Ras. COS-1 cells were co-transfected with HA- ERK5 or HA-ERK2 ± dominant negative MEK5 (MEK5A) or Ras (Ras17N) and treated with 1,25(OH)2D3 or ethanol for 5 (HA-ERK2) or 15 min (HA-ERK5). After cell lysis, HA-tagged kinase was immunoprecipitated, and kinase activity was assayed using His-Ets-1 as a substrate.

Results are representative of three experiments.

167

3

2 ion) tivity t * c ula

5 a m

RK E - old sti A F 1 H (

0 Ets-1 Ets-1T38A

5.15 Substituting threonine 38 for alanine in Ets-1T38A diminished the phosphorylation of Ets-1 by ERK5. COS-1 cells were transfected with HA-ERK5 and treated with 1,25(OH)2D3 for 15 min. After lysis, HA-ERK5 was immunoprecipitated with anti-HA antibody and the activity of ERK5 was assayed using GST-Ets-1 or GST-

Ets-1T38A as a substrate. The results presented here are mean ± S.D of three independent experiments. Significance of difference between Ets-1 and Ets-1T38A: *p<0.01

(Student’s unpaired t-test).

168 5.2.13 CYP24 promoter induction by 1,25(OH)2D3 required the Ets binding site (EBS) and the ERK5 module

To investigate the role played by the EBS in 1,25(OH)2D3-mediated induction of the

CYP24 promoter and its functional relationship with ERK1/ERK2 and ERK5, COS-1 cells were co-transfected with pCYP24WT (-298bp)-Luc, pCYP24mEBS (-298bp)-Luc, pRL-TK, empty vector, ERK1 K71R and/or MEK5A. After 24 h, the cells were treated with 1,25(OH)2D3 (10-7 M) for 24 h and the luciferase activities assayed. The data in

Figure 5.15 show that 1,25(OH)2D3-stimulated CYP24 promoter activity was reduced by approximately 42% when the EBS was mutated. Promoter activity was reduced by a similar degree by the overexpression of the dominant negative MEK5. Co-transfection of the mutant EBS-Luc construct with dominant negative MEK5 did not produce a greater degree of suppression than that seen with one mutant alone. In contrast, expression of the dominant negative ERK1 reduced CYP24 promoter activity by approximately 80%

(Figure 5.16). This degree of inhibition was not enhanced by co-transfection of dominant negative ERK1 with dominant negative MEK5, mutant EBS or the combination of dominant negative MEK5 and mutant EBS. These results, together with the data on the phosphorylation of Ets-1 by ERK5, suggest that unlike ERK1/ERK2, the ERK5 module is specifically required for the functionality of the EBS binding protein in COS-1 cells. In contrast, ERK1/ERK2, functioning via RXRα, are essential for the induction of the

CYP24 promoter by 1,25(OH)2D3 via sites such as VDRE1 and VDRE2 which represent the principal signaling pathway in the induction process.

169 c b

a 30

25 n

tio d )

n 20 duc d In

r ctio e u t 15

d Ind ol F

( 10

CYP24 promo 5

0 pCYP24WT(-298bp)-Luc + + + + - - - -

pCYP24mEBS (-298bp)-Luc - - - - + + + + ERK1 K71R - + - + - + - + MEK5A - - + + - - + +

Figure 5.16 Roles of the EBS, ERK1/ERK2 and ERK5 in 1,25(OH)2D3-stimulated activation of the CYP24 promoter. COS-1 cells were transiently co-transfected with pCYP24WT (-298bp)-Luc (600 ng), pCYP24mEBS (-298bp)-Luc (600 ng), empty vector

(500 ng), ERK1 K71R (500 ng) and/or MEK5A (500 ng) as indicated, and pRL-TK (150 ng). After transfection, the cells were treated with 1,25(OH)2D3 (10-7 M) for 24 h, and the relative luciferase activity determined. Results are mean ± S.D of three independent experiments. Significance of difference between control and dominant negative

ERK1K71R: ap<0.001; between control and MEK5A: bp<0.01; between control and mEBS: cp<0.01; between ERK1K71R and a combination of ERK1K71R and MEK5A or mEBS + ERK1K71R + MEK5A: dp>0.05 (ANOVA and Tukey multiple comparison test).

170 5.3 Summary

Using western blotting and kinase activity assays, the data show that unlike 25(OH)D3,

1,25(OH)2D stimulated the activities /phosphorylation of ERK1/ERK2 and ERK5 and that the up-stream signaling molecule, Ras, played a crucial role in activating these MAP kinases in COS-1 cells. The data also established that both ERK1/ERK2 and MEK5 played a role in 1,25(OH)2D3-mediated induction of the CYP24 promoter. The

ERK1/ERK2 module was shown to play a greater role than the MEK5-ERK5 module in regulating CYP24 promoter induction by 1,25(OH)2D3.

Interestingly we found that ERK1/ERK2 but not ERK5 interacted with RXRα and that activation of ERK1 was required for the phosphorylation of RXRα by 1,25(OH)2D3.

Furthermore, phosphorylation of RXRα was required for CYP24 promoter induction by

1,25(OH)2D3, suggesting that ERK1/ERK2 acted via RXRα.

With respect to Ets-1, the data show that 1,25(OH)2D3 stimulated the binding of ERK5 to Ets-1. However, a weak interaction was observed between Ets-1 and ERK2. This interaction was constitutive and was not regulated by 1,25(OH)2D3. 1,25(OH)2D3 also stimulated the phosphorylation of Ets-1 on T38 and this was shown to be by ERK5.

Binding of Ets-1 to the EBS played a role in CYP24 promoter induction by

1,25(OH)2D3. These data provide strong evidence that the ERK5 module acted via Ets-1 and the EBS. Based on the data from this study, the following model is proposed (see

Figure 5.16). We proposed that 1,25(OH)2D3, acting on the plasma membrane VDR,

171 activates ERK1/ERK2 and ERK5 via up-stream Ras. Activated ERK1/ERK2 and ERK5 phosphorylate RXRα and Ets-1 which binds VDRE and EBS respectively. In addition,

1,25(OH)2D3 can also bind directly to the nuclear VDR to activate the CYP24 promoter.

RXR VDR P Ets-1 P

VDRE EBS CYP24

Figure 5.16 Regulation of the CYP24 promoter induction by 1,25(OH)2D3 via non- genomic and genomic action. (Dwivedi et al., 2002)

172

Chapter 6 Mechanism III

Role of MAP kinases, PI3K and PKC in mediating

CYP24 promoter induction by daunorubicin

hydrochloride and etoposide.

173 6.1 Introduction

Chemotherapeutics drugs such as daunorubicin hydrochloride, doxorubicin hydrochloride, etoposide, cisplatin and vincristine sulfate have been reported to stimulate the activity of PKCζ, PI3K and the MAP Kinases, ERK1/ERK2 and JNK, in cancer cells

(Mas et al., 2003; Brantley-Finley et al., 2003; Ohtsuka et al., 2003). Thus, chemotherapeutic drugs may stimulate CYP24 promoter activity through the activation of these signaling molecules.

6.2 Results

6.2.1 Dominant negative ERK1 and MEK5 prevent the induction of CYP24 promoter activity by daunorubicin hydrochloride in COS-1 cells.

In Chapter 5, we establish that ERK1/ERK2 and ERK5 plays an important role in

1,25(OH)2D3 induction of the CYP24 promoter in COS-1 cells. In this Chapter, we investigate whether ERK1/ERK2 and ERK5 are involved in the induction of the CYP24 promoter by daunorubicin hydrochloride and etoposide. Using kinase assay, we later confirm that the drugs can activate these signalling molecules.

174 COS-1 cells were co-transfected with pCYP24WT(-298bp)-Luc, pRL-Null-Luc and dominant negative (DN) ERK1 (ERK1K71R) or DN MEK5 (MEK5A) or a control vector using DOTAP. The cells were then treated with daunorubicin hydrochloride (2

µΜ) for 24 h. The relative luciferase activity of daunorubicin hydrochloride-treated cells was expressed as fold induction over ethanol-treated cells. The results (Figure 6.1) show that daunorubicin hydrochloride treatment increased CYP24 promoter activity to 9.6 ±

1.2 fold. However, in cells co-transfected with DN ERK1 or DN MEK5, stimulation of

CYP24 promoter activity by the drug was 4.5 ± 0.8 fold (53% reduction) and 4.0 ± 0.8 fold (59% reduction), respectively (Figure 6.1). These results demonstrate that daunorubicin hydrochloride-mediated induction of the CYP24 promoter in COS-1 cells required ERK1/ ERK2 and ERK5.

6.2.2 Daunorubicin hydrochloride stimulated the activities of ERK1/ERK2 and ERK5 in COS-1 cells.

To further confirm the above results that ERK1/ERK2 and ERK5 played a role in the daunorubicin hydrochloride-mediated induction of the CYP24 promoter, the activation of endogenous ERK1/ERK2 and ERK5 by daunorubicin hydrochloride was investigated.

175 12

10 on i 8 nduct I tion) 6 *

oter * prom

(Fold Induc 4 P24 CY 2

0 Control ERK1K71R MEKMEK5A

Figure 6.1 Dominant negative ERK1 and MEK5 inhibited CYP24 promoter induction by daunorubicin hydrochloride in COS-1 cells. COS-1 cells were co- transfected with pCYP24 WT (-298bp)-Luc, pRL-Null-Luc and dominant negative (DN)

ERK1 (ERK1K71R), DN MEK5 (MEK5A) or a control vector and treated with daunorubicin hydrochloride (2 µΜ) for 24 h. The results shown are mean ± SEM of three experiments, each conducted in triplicate determinations. Significance of difference between control and ERK1K71R or MEK5A: *p<0.01 (ANOVA followed by Dunnett multiple comparisons test).

176 COS-1 cells were treated with daunorubicin hydrochloride for 5 and 30 minutes for

ERK2 and ERK5 kinase assays, respectively. Results from preliminary data suggest that these were the optimum times for the stimulation of ERK1/ERK2 and ERK5 activities.

The data in Figure 6.2 demonstrate that daunorubicin hydrochloride increased

ERK1/ERK2 activity by 93% compared to EtOH treated cells. Daunorubicin hydrochloride also modestly increased ERK5 activity (21%) compared to EtOH treated cells (Figure 6.3). These results suggest that ERK1/ERK2 and ERK5 may be involved in mediating the stimulatory effect of daunorubicin hydrochloride on the CYP24 promoter.

6.2.3 Dominant negative MKK4 inhibited daunorubicin hydrochloride induction of CYP24 promoter activity in COS-1 cells.

Previous studies have shown that chemotherapeutic drugs could stimulate JNK activity in tumor cells (Mansat-de Mas et al., 1999; Hayakawa et al., 2003). JNK has also been shown to play a role in 1,25(OH)2D3 induction of the CYP24 promoter in HEK293T cells (Nutchy et al., 2005). We therefore investigated whether JNK was required for daunorubicin hydrochloride induction of the CYP24 promoter activity. To achieve this,

COS-1 cells were co-transfected with pCYP24WT(-298bp)-Luc, pRL-Null-Luc and dominant negative (DN) MKK4 (MKK4 K116R) or a control vector. MKK4 is one of the two direct upstream activators of JNK.

177 250 *

200 y )

l 150 o r ont

of c 100 % ( ERK1/ERK2 activit

50

0 EtOH DDauaunnororuubbiicincin

Treatment

Figure 6.2 Daunorubicin hydrochloride stimulated ERK1/ERK2 activity in COS-1 cells. COS-1 cells were incubated with daunorubicin hydrochloride (2 µΜ) for 5 minutes, lysed and the activities of ERK1/ERK2 assayed as described in Materials and Methods.

Results shown are mean ± SEM of three experiments. Significance of difference between

EtOH and daunorubicin hydrochloride treated cells: *p<0.01 (Student’s unpaired t-test).

178 150 *

100 ) y of control K5 activit R (% E 50

0 EtOH Daunorubicin

Treatment

Figure 6.3 Daunorubicin hydrochloride caused a modest increase in ERK5 activity in COS-1 cells. COS-1 cells were incubated with daunorubicin hydrochloride (2 µΜ) for

30 minutes, lysed and the activity of ERK5 assayed as described in Materials and

Methods. Results shown are mean ± SEM of three experiments. Significance of difference between EtOH and daunorubicin hydrochloride treated cells: *p<0.05

(Student’s unpaired t-test).

179 Cells were treated with daunorubicin hydrochloride (2 µΜ) for 24 h. The result from these experiments shows that DN MKK4 significantly inhibited the drug-induced CYP24 promoter activity (Figure 6.4). The data therefore show that the JNK pathway plays an important role in daunorubicin hydrochloride induction of the CYP24 promoter. This is in direct contrast to the data that showed that the JNK signaling pathway was not involved in 1,25(OH)2D3 induction of the CYP24 promoter in this cell-line (Dwivedi et al., 2002).

6.2.4 Daunorubicin hydrochloride stimulated JNK activity in COS-1 cells.

To further confirm that the JNK module was involved in mediating the effect of daunorubicin hydrochloride on CYP24 promoter activity in COS-1 cells, we also investigated whether daunorubicin hydrochloride could stimulate JNK activity. We first determined the optimal incubation time required for daunorubicin hydrochloride to activate JNK. COS-1 cells were incubated with daunorubicin hydrochloride for 5 minutes, 30 minutes, 1 h, 2 h and 6 h. The results show that daunorubicin hydrochloride only caused JNK activation at 6 h (Figure 6.5).

To study the dose-response relationship between daunorubicin hydrochloride and JNK activity, COS-1 cells were treated with 0.125 µΜ, 0.5 µΜ, 2 µΜ and 4 µΜ of daunorubicin hydrochloride for 6 h. A Significant increase in JNK activity was observed at 2 and 4 µΜ of daunorubicin hydrochloride. Stimulation of JNK activity was 1.5 fold

180 12

n 9 io ct ) Indu duction

n 6 omoter d I l r p 4 (Fo * 2 P

CY 3

0 Contrrolol MMKKK4K4 K116R

Figure 6.4 Effect of dominant negative MKK4 on CYP24 promoter induction by daunorubicin hydrochloride in COS-1 cells. COS-1 cells were co-transfected with pCYP24WT(-298bp)-Luc, pRL-Null-Luc and dominant negative MKK4 (MKK4 K116R) or a control vector and treated with daunorubicin hydrochloride (2 µΜ) for 24 h. The results shown are mean ± SEM of three experiments, each conducted in triplicate determinations. Significance of difference between control and MKK4: *p<0.001

(Student’s unpaired t-test).

181

Duration of treatment

6 hrs 5 min 30 min 1 hr 2 hrs 6 hrs

Phosphorylated GST-jun (1-79)

Control Daunorubicin hydrochloride (2µM)

Figure 6.5 Daunorubicin hydrochloride caused the activation of JNK in COS-1 cells. COS-1 cells were incubated with vehicle (control) or daunorubicin hydrochloride for 5 minutes, 30 minutes, 1 h, 2 h and 6 h, lysed and the activity of JNK was assayed as described in Materials and Methods. The experiment was repeated twice and similar results were obtained.

182 greater at 4 µΜ compared to 2 µΜ (Figure 6.6). These results demonstrated that the drug stimulated the activity of JNK in a time- and dose-dependent manner.

6.2.5 Dominant negative PKCζ inhibited daunorubicin hydrochloride induction of CYP24 promoter activity in COS-1 cells.

Previous studies have demonstrated that chemotherapeutic drugs can stimulate PKCζ activity in leukaemic cells (Mas et al., 2003) and recent studies from our laboratory have shown that PKCζ is involved in mediating the effect of 1,25(OH)2D3 on CYP24 promoter activity in HEK293T cells (Dwivedi et al., manuscript in preparation). We therefore investigated whether PKCζ was required for daunorubicin hydrochloride induction of the CYP24 promoter. To test this, COS-1 cells were co-transfected with pCYP24WT (-298bp)-Luc, pRL-Null-Luc and dominant negative (DN) PKCζ or a control vector. After transfection, the cells were treated with daunorubicin hydrochloride

(2 µΜ) for 24 h. Results from this experiment show that DN PKCζ significantly reduced the CYP24 promoter activity when compared to cells with control vector (Figure 6.7).

The data therefore show that PKCζ was involved in mediating the effect of daunorubicin hydrochloride on CYP24 promoter activity.

183

300 **

250 * 200 of control) 150 y (% t

100

JNK activi 50

0 012345 Daunorubicin (µΜ)

Figure 6.6 Effects of different concentrations of daunorubicin hydrochloride on

JNK activity in COS-1 cells. COS-1 cells were treated with 0, 0.125, 0.5, 2 or 4 µΜ of daunorubicin hydrochloride for 6 h, lysed and the activity of JNK was assayed as described in Materials and Methods. The data, expressed as % of control, are the mean ±

SEM of three experiments. Significance of difference between control cells and cells treated with daunorubicin hydrochloride: *p<0.05, **p<0.01 (ANOVA followed by

Dunnett multiple comparisons test).

184 12

10 on i 8 on) i Induct er

ot 6 om

old Induct * F ( 4 P24 pr CY 2

0 Control PKCζK281M

Figure 6.7 Dominant negative PKCζ inhibited CYP24 promoter induction by daunorubicin hydrochloride in COS-1 cells. COS-1 cells were co-transfected with pCYP24WT(-298bp)-Luc, pRL-Null-Luc and dominant negative PKCζ (PKCζ K281M) or a control vector and treated with daunorubicin hydrochloride (2 µΜ) for 24 h. The results shown are mean ± SEM of three experiments, each conducted in triplicate determinations. Significance of difference between control and PKCζ K281M: *p<0.05

(Student’s unpaired t-test).

185 6.2.6 Inhibitor of PI3K (LY294002) suppressed daunorubicin hydrochloride induction of CYP24 promoter activity in COS-1 cells.

PI3K had been shown to play an important role in daunorubicin hydrochloride stimulation of ERK1 and PKCζ (Mas et. al., 2003). Inhibition of PI3K by wortmannin was shown to partially inhibit ERK1 activation but significantly inhibit PKCζ in U937 cells (Mas et. al., 2003). We therefore investigated whether the specific pharmacological inhibitor of PI3K, LY294002, could inhibit CYP24 promoter activity. COS-1 cells were transiently co-transfected with pCYP24WT(-298bp)-Luc, pRL-Null-Luc and treated with

LY294002 (5 µΜ) or DMSO. The concentration of LY294002 used in this study was based on the concentration used by Versteeg et al. (2000) and Benini et al. (2004) to inhibit PI3K. Following 30 minutes treatment with LY294002, the cells were treated with daunorubicin hydrochloride (2 µΜ) for 24 h. The data in Figure 6.8 show that cells treated with LY294002 had a 71% reduction in CYP24 promoter activity compared to

DMSO-treated cells. These results demonstrate that PI3K played an important role in the induction of the CYP24 promoter by 1,25(OH)2D3 in COS-1 cells, most probably acting as an upstream regulator of the ERK1/ERK2 module and PKCζ.

186 10

8

6 ion) Induction

4 old Induct F ( * P24 promoter Y

C 2

0 LY 294002 - +

Daunorubicin + +

Figure 6.8 LY294002 inhibited CYP24 promoter induction by daunorubicin hydrochloride in COS-1 cells. COS-1 cell were transiently co-transfected with pCYP24WT(-298bp)-Luc, pRL-Null-Luc and treated with LY294002 (5 µΜ) for 30 minutes before being incubated with daunorubicin hydrochloride (2 µΜ) for 24 h. Results shown are mean ± SEM of three experiments, each conducted in triplicate determinations. Significance of difference between LY294002 treated and untreated:

*p<0.01. (Student’s unpaired t -test).

187 6.2.7 Inhibition of daunorubicin hydrochloride-mediated transcriptional activation of the CYP24 promoter activity by dominant negative ERK1 and

MEK5 in HEK293T cells.

The data in Chapter 3 demonstrate that daunorubicin hydrochloride alone was able to induce CYP24 promoter in a dose-dependent manner in COS-1 cells but not in HEK293T cells. However, the drug caused a greater enhancing effect on the stimulation of CYP24 promoter activity by 1,25(OH)2D3 in HEK293T cells than in COS-1 cells. Given these differences in the induction of the CYP24 promoter activity by daunorubicin hydrochloride between COS-1 and HEK293T cells, we also investigated the roles played by the ERK1/ERK2, ERK5 and JNK modules, and PKCζ in the induction of the CYP24 promoter in HEK293T cells. HEK293T cells were co-transfected with pCYP24WT(-

298bp)-Luc, pRL-Null-Luc and DN ERK1 (ERK1K71R), DN MEK5 (MEK5A) or a control vector. The cells were then incubated with daunorubicin hydrochloride (0.5 µΜ)

+ 1,25(OH)2D3 (10 nM) for 24 h. Co-treatment with 1,25(OH)2D3 and daunorubicin hydrochloride caused a synergistic activation of the CYP24 promoter in HEK293T cells

(Chapter 3). Results from these experiments (Figure 6.9) show that DN ERK1 or DN

MEK5 per se did not suppress the induction of CYP24 promoter activity by daunorubicin hydrochloride + 1,25(OH)2D3. However, when DN ERK1 and DN MEK5 were co- transfected together, a significant reduction in CYP24 promoter activity was observed

(11.4 ± 0.8 fold to 1.2 ± 0.3 fold). These data imply that unlike in COS-1 cells, the

ERK1/ERK2 and ERK5 signaling pathways in HEK293T cells were able to compensate

188 each other and thus one pathway was still able to transmit signals to the CYP24 promoter even when the other was inhibited. These results are consistent with recent data from our laboratory (Nutchey et al., 2005) which show that inhibition of ERK1 or MEK5 alone in

HEK293T cells did not suppress the 1,25(OH)2D3-mediated induction of the CYP24 promoter. However, Nutchey et al. (unpublished) demonstrated that dominant negative

Ras (Ras17N) inhibited the induction of the CYP24 promoter by 1,25(OH)2D3 in

HEK293T cells. Ras is located up-stream of ERK1/ERK2 and ERK5 and plays an important role in activating the ERK1/ERK2 and ERK5 modules and the CYP24 promoter (Chapter 5, Dwivedi et al., 2002)

189 16

12 n io ct ) on ndu i I

r e t

duct 8 n I omo 4 pr (Fold P2

CY 4 *

0 CCoontrol ERK1K71R MEK5A ERK1K71RK1K71R + MEK5A

Figure 6.9 Effects of dominant negative ERK1 and MEK5 on CYP24 promoter induction by daunorubicin hydrochloride and 1,25(OH)2D3 in HEK293T cells.

HEK293T cells were transiently co-transfected with pCYP24 WT (-298bp)-Luc, pRL-

Null-Luc and dominant negative (DN) ERK1 (ERK1K71R) and/or DN MEK5 (MEK5A) or a control vector and treated with daunorubicin hydrochloride (0.5 µΜ) + 1,25(OH)2D3

(10 nM) for 24 h. The results shown are mean ± SEM of three experiments, each conducted in triplicate determinations. Significance of difference between control and

ERK1K71R+MEK5A: *p<0.001 (ANOVA followed by Dunnett multiple comparisons test).

190 6.2.8 Daunorubicin hydrochloride stimulated ERK1/ERK2 and ERK5 activities in HEK293T cells.

The effects of daunorubicin hydrochloride on ERK1/ERK2 and ERK5 activities were investigated next. HEK293T cells were treated with daunorubicin hydrochloride (0.5

µΜ) for 5, 10 and 30 minutes (ERK2) and 6 h (ERK5) respectively. The data show that daunorubicin hydrochloride increased ERK2 activity by 1.4 fold compared to untreated cells (control) (Figure 6.10). In contrast, although 1,25(OH)2D3 showed a tendency to increase ERK1/ERK2 activity in HEK293T cells, this effect did not reach a statistically significant level (data not shown). This is consistent with the data of Nutchey et al.

(2005) who also found that 1,25(OH)2D3 did not significantly increase ERK1/ERK2 activity in HEK293T cells. These results are in direct contrast to the data in COS-1 cells in which 1,25(OH)2D3 stimulated the activity of ERK1/ERK2. (Dwivedi et al., 2002).

Similarly, in cells treated with daunorubicin hydrochloride alone and together with

1,25(OH)2D3, ERK5 activity was increased by 1.5 fold and 1.6 fold, respectively, compared to untreated cells (Figure 6.11). In these cells, 1,25(OH)2D3 per se did not stimulate or enhance the activation of ERK5 by daunorubicin hydrochloride although there was also a tendency for 1,25(OH)2D3 per se to cause a small increase in ERK5 activity which did not reach statistically significant levels.

191

160

* l)

140

(% of contro

y

it 120

100

ERK1/ERK2 activ

80

0102030

Time (minutes)

Figure 6.10 Daunorubicin hydrochloride stimulated ERK2 activity in HEK293T cells. HEK293T cells were incubated with daunorubicin hydrochloride (0.5 µΜ) for 5, 10 and 30 minutes, lysed and the activity of ERK1/ERK2 was assayed as described in

Materials and Methods. Results shown are mean ± SEM of three experiments.

Significance of difference between vehicle-treated (control) and daunorubicin hydrochloride-treated cells: *p<0.05 (ANOVA followed by Dunnett multiple comparisons test).

192 180 ** ) 120 of control

% ( y Activit 60 K 5 ER

0 1,25(OH)2D3 - + - + Daunorubicin - - + +

Figure 6.11 Daunorubicin hydrochloride stimulated ERK5 activity in HEK293T cells. HEK293T cells were incubated with daunorubicin hydrochloride (0.5 µΜ) for 6 h, lysed and the activity of ERK5 assayed as described in Materials and Methods. Results shown are mean ± SEM of three experiments. Significance of difference between cells treated with daunorubicin hydrochloride ± 1,25(OH)2D3 and untreated cells: *p<0.05

(ANOVA followed by Dunnett multiple comparisons test).

193 This is also in direct contrast to the ability of 1,25(OH)2D3 to stimulate ERK5 activity in

COS-1 cells (Chapter 5; Dwivedi et al., 2002). The above results show that daunorubicin hydrochloride could stimulate ERK1/ERK2 and ERK5 in COS-1 cells as well as in

HEK293T cells. In contrast, 1,25(OH)2D3 stimulated ERK1/ERK2 and ERK5 activities in COS-1 cells but not in HEK293T cells, thereby revealing subtle differences between the two different kidney cell lines which were derived from two different regions in the kidney and from two different species.

6.2.9 Effect of dominant negative MKK4 on daunorubicin hydrochloride- and 1,25(OH)2D3-mediated induction of CYP24 promoter activity in

HEK293T cells.

HEK293T cells were co-transfected with pCYP24WT(-298bp)-Luc, pRL-Null-Luc and

DN MKK4 (MKK4 K116R) or a control vector. After transfection, the cells were treated with daunorubicin hydrochloride (0.5 µΜ) + 1,25(OH)2D3 (10 nM) for 24 h.

The data in Figure 6.12 show that the ability of daunorubicin hydrochloride +

1,25(OH)2D3 to stimulate CYP24 promoter activity in cells co-transfected with DN

MKK4 was suppressed by approximately 54% compared to cells co-transfected with empty vector (DN MKK4: 3.8 ± 1.0 fold compared to empty vector: 8.2 ± 1.6 fold).

194 12

10

ion 8 duct n ion) I

er

duct 6 n * omot old I F ( 4 P24 pr CY 2

0 Control MKK4K116R

Figure 6.12 Effect of dominant negative MKK4 on CYP24 promoter induction by daunorubicin hydrochloride and 1,25(OH)2D3 in HEK293T cells. HEK293T cells were co-transfected with pCYP24 WT (-298bp)-Luc, pRL-Null-Luc and dominant negative (DN) MKK4 (MKK4 K116R) or a control vector and treated with daunorubicin hydrochloride (0.5 µΜ) + 1,25(OH)2D3 (10 nM) for 24 h. The results shown are mean ±

SEM of three experiments, each conducted in triplicate determinations. Significance of difference between control and MKK4 K116R: *p<0.05 (Student’s unpaired t-test).

195 6.2.10 Daunorubicin hydrochloride stimulated JNK activity in HEK293T cells.

To determine whether daunorubicin hydrochloride could stimulate JNK activity in

HEK293T cells, the cells were treated with daunorubicin hydrochloride (0.5 µΜ) for 6 h

(see Figure 6.5). The results show that daunorubicin hydrochloride caused a modest increase (1.2 fold) in JNK activity compared to vehicle-treated cells (Figure 6.13).

Neither 1,25(OH)2D3 nor daunorubicin hydrochloride alone significantly increased JNK activity (data not shown). These results show that the JNK signaling pathway also played a role in mediating the induction of the CYP24 promoter by daunorubicin hydrochloride and 1,25(OH)2D3 in HEK293T cells.

6.2.11 PKCζ was not required for the induction of the CYP24 promoter by daunorubicin hydrochloride and 1,25(OH)2D3 in HEK293T cells.

To investigate the role of PKCζ in the induction of the CYP24 promoter by daunorubicin hydrochloride and 1,25(OH)2D3 in HEK293T cells, the cells were co-transfected with pCYP24WT(-298bp)-Luc, pRL-Null-Luc and DN PKCζ (PKCζ K281M) or control

196 Phosphorylated GST-jun (1-79)

150 *

100

50 JNK activity (% of control)

0 EtOH Daunorubicin + 1,25(OH)2D3

Figure 6.13 Effect of daunorubicin hydrochloride and 1,25(OH)2D3 on JNK activity in HEK293T cells. HEK293T cells were treated with daunorubicin hydrochloride (0.5 µΜ) and 1,25(OH)2D3 (10 nM) for 6 h. Cells were lysed and the activity of JNK assayed as described in Materials and Methods. Results shown represent mean ± SEM of three experiments conducted in triplicate determinations. Significance of difference between EtOH-treated cells and daunorubicin hydrochloride + 1,25(OH)2D3- treated cells: *p<0.05 (Student’s unpaired t-test).

197 vector. Cells were then treated with daunorubicin hydrochloride (0.5 µΜ) +

1,25(OH)2D3 (10 nM) for 24 h. The data in Figure 6.14 show that DN PKCζ did not inhibit CYP24 promoter induction by daunorubicin hydrochloride + 1,25(OH)2D3. These data imply that PKCζ was not required for the induction of the CYP24 promoter by daunorubicin hydrochloride + 1,25(OH)2D3, unlike its role in COS-1 cells (Figure 6.7).

6.2.12 Role of MAP Kinases and PKCζ in the induction of CYP24 promoter by etoposide in COS-1 cells.

As with daunorubicin hydrochloride, etoposide has been shown to stimulate the activities of MAP Kinases in tumor cells in order to induce apoptosis. Besides this, etoposide also stimulated the transcriptional activation of the CYP24 promoter in COS-1 cells (Chapter

3). The following experiments were therefore conducted to investigate whether etoposide induction of the CYP24 promoter activity in COS-1 cells required the ERK1/ERK2,

ERK5, JNK and PKCζ signaling pathways.

198 12

10

8 duction n

6

(Fold Induction) 4 P24 promoter I CY 2

0 Control PKCζK281M

Figure 6.14 Dominant negative PKCζ did not affact CYP24 promoter induction by daunorubicin hydrochloride and 1,25(OH)2D3 in HEK293T cells. HEK293T cells were co-transfected with pCYP24 WT (-298bp)-Luc, pRL-Null-Luc and dominant negative (DN) PKCζ (PKCζ K281M) or a control vector and treated with daunorubicin hydrochloride (0.5 µΜ) + 1,25(OH)2D3 (10 nM) for 24 h. The results shown are mean ±

SEM of three experiments, each conducted in triplicate determinations. Significance difference between control and PKCζ K281M: *p>0.05 (Student’s unpaired t-test).

199 6.2.13 Dominant negative ERK1 and not MEK5 inhibited etoposide mediated induction of the CYP24 promoter activity in COS-1 cells.

Since CYP24 promoter induction by etoposide was greater in the -1.4Kb promoter than in the -298bp promoter (Chapter 3), the -1.4Kb CYP24 promoter was therefore used in these studies. COS-1 cells were co-transfected with pCYP24WT(-1.4Kb)-Luc, pRL-Null-

Luc and DN MEK5 (MEK5A) or DN ERK2 (ERK1K71R) or a control vector. Cells were then treated with etoposide (100 µΜ) for 24 h. Results from these experiments (Figure

6.15) show that DN MEK5 did not inhibit etoposide mediated induction of the CYP24 promoter. On the other hand, DN ERK1 severely suppressed CYP24 promoter activity. In fact, promoter activity was reduced by 94% below basal CYP24 promoter activity.

However, this was likely to be an overestimation of the effect because etoposide selectively caused an increase in renilla luciferase activity when the cells were transfected with DN ERK1. This effect of etoposide on the activity of renilla luciferase was not seen with the other DN mutants. Nevertheless, there was also a decrease in firefly luciferase activity in the presence of DN ERK1 compared to when an empty vector was transfected.

Thus, the ratio of firefly luciferase: renilla luciferase was exceptionally low when compared to control cells (please see Chapter 2.2.3.4 for data presentation). These results showed that ERK1/ERK2 but not ERK5 was involved in the induction of the CYP24 promoter by etoposide.

200 3 on 2 on) nducti i er I ot m

(Fold Induct 1 CYP24 Pro *

0 Control MEK5A ERKRK11KK771R1R

Figure 6.15 Effects of dominant negative MEK5 and ERK1 on CYP24 promoter induction by etoposide in COS-1 cells. COS-1 cells were co-transfected with pCYP24

WT (-1.4Kb)-Luc, pRL-null-Luc and dominant negative (DN) ERK1 (ERK1K71R), DN

MEK5 (MEK5A) or a control vector and treated with etoposide (100 µΜ) for 24 h.

Results shown are mean ± SEM of three experiments, each conducted in triplicate determinations. Significance of difference between control and ERK1K71R: *p<0.0001

(ANOVA followed by Dunnett’s multiple comparisons test).

201 6.2.14 Etoposide stimulated ERK1/ERK2 activity in COS-1 cells.

Since DN ERK1 inhibited etoposide-mediated induction of the CYP24 promoter activity, we investigated whether etoposide-stimulated the activity of ERK1/ERK2 in COS-1 cells. To test this, the cells were incubated with etoposide (100 µΜ) for 5 minutes. The results show that etoposide increased ERK1/ERK2 activity by 2.2 fold compared to untreated cells (Figure 6.16). Thus, these results demonstrate that the activation of

ERK1/ERK2 by etoposide played a role in the induction of the CYP24 promoter in COS-

1 cells.

6.2.15 Dominant negative MKK4 but not PKCζ inhibited etoposide-mediated induction of the CYP24 promoter activity in COS-1 cells.

To determine whether PKCζ and the JNK module played a role in the etoposide- mediated induction of the CYP24 promoter, COS-1 cells were co-transfected with DN

PKCζ (PKCζ K281M), or DN MKK4 (MKK4 K116R), pRL-Null-Luc and pCYP24WT(-1.4kb)-Luc. After transfection, the cells were treated with etoposide (100

µΜ) for 24 h. Results from this set of experiments (Figure 6.17) show that DN PKCζ did not suppress etoposide-mediated induction of the CYP24 promoter. However, DN MKK4 caused a significant reduction in CYP24 promoter induction (55% reduction compared to cells transfected with a empty vector). These results demonstrate that JNK

202 300

*

l) ro t

con 200 f

(% o ity

100

/ERK2 activ

ERK1

0 Control Etoposidetoposide Treatment

Figure 6.16 Effect of etoposide on ERK1/ERK2 activity in COS-1 cells. COS-1 cells were incubated with etoposide (100 µΜ) for 5 minutes, lysed and the activity of

ERK1/ERK2 was assayed as described in Materials and Methods. Results shown are mean ± SEM of three experiments. Significance of difference between vehicle-treated cells (control) and etoposide-treated cells: *p<0.05 (Student’s unpaired t-test).

203 4

3 n o nducti er I

ot 2 m

old Induction) * (F

CYP24 Pro 1

0 Control PKCζK281M MMKK4K116RKK4K116R

Figure 6.17 Effects of dominant negative PKCζ and MKK4 on CYP24 promoter induction by etoposide in COS-1 cells. COS-1 cells were co-transfected with pCYP24

WT (-1.4Kb)-Luc, pRL-Null-Luc and dominant negative (DN) PKCζ (PKCζ K281M) or

DN MKK4 (MKK4 K116R) or a control vector and treated with etoposide (100 µΜ) for

24 h. Results shown are mean ± SEM of three experiments, each conducted in triplicate determinations. Significance of difference between control and MKK4 K116R: *p<0.05

(ANOVA followed by Dunnett’s multiple comparisons test).

204 and not PKCζ was involved in the induction of the CYP24 promoter by etoposide. This contrasts with a role for PKCζ in the action of daunorubicin hydrochloride (Figure 6.7).

6.2.16 Etoposide stimulated JNK activity in COS-1 cells.

To further confirm the role of the JNK module in the action of etoposide, we examined whether etoposide could stimulate JNK activity in COS-1 cells. First we investigated the optimal incubation time for JNK activation in response to etoposide treatment. COS-1 cells were incubated with etoposide for 0.5, 2 and 6 h. The data show that etoposide significantly increased JNK activity at 6 hr although a small increase was also evident at

2 h (Figure 6.18). We then investigated whether the stimulation of JNK activity by etoposide was concentration-dependent. COS-1 cells were treated with 4, 20 and 100 µΜ of etoposide for 6 h. The results show that at all three concentrations of etoposide tested,

JNK activity was significantly increased (Figure 6.19). This is consistent with the data on

CYP24 promoter activity in that the lowest concentration of etoposide tested, 6.25 µΜ was also effective at stimulating CYP24 promoter activity (Figure 3.5).

205

Duration of Treatment (Hours) 6 hr 0.5 hr 2 hr 6 hr

Phosphorylated GST-Jun (1-79)

Control Etoposide

Figure 6.18 Effects of different etoposide exposure times on JNK activity in COS-1 cells. COS-1 cells were incubated with etoposide for 0.5, 2 and 6 h. Cells were lysed and the activity of JNK assayed as described in Materials and Methods. The experiment was repeated twice and similar results were obtained.

206 * 300 * * 250

200

150

100 JNK activity (% of control) 50

0 0014 20 10000 Etoposide (µΜ)

Figure 6.19 Effects of different concentrations of etoposide on JNK activity in COS-

1 cells. COS-1 cells were treated with 0, 4, 20, 100 µΜ of etoposide for 6 h. Cells were lysed and the activities of JNK was assayed as described in Materials and Methods.

Results shown represent mean ± SEM of three experiments, each conducted in triplicate determinations. Significance of difference between control and cells treated with etoposide: *p<0.05 (ANOVA followed by Dunnett’s multiple comparisons test).

207 6.3 Summary

In summary, the above data demonstrate that ERK1, ERK2, ERK5, JNK, PI3K or PKCζ were involved in regulating CYP24 promoter induction by daunorubicin hydrochloride or etoposide. The extent to which each of these signaling molecules was involved in mediating the action of the drugs depended on the kidney cell-line used as well as the drug involved.

208

Chapter 7

Effects of chemotherapeutic drugs on endogenous CYP24 mRNA expression in

kidney cells

209 7.1 Introduction

In the preceding chapters, we demonstrated that chemotherapeutic drugs can up-regulate the activity of the CYP24 promoter in kidney cells. We therefore aimed to investigate the effects of these drugs on endogenous CYP24 expression. Since it is not possible to measure the levels of CYP24 protein due to the unavailability of commercial anti-CYP24 antibody at present, we therefore investigated the effects of daunorubicin hydrochloride and etoposide on endogenous CYP24 mRNA expression in COS-1 cells.

7.2 Methods

COS-1 cells were treated with daunorubicin hydrochloride or etoposide in the presence or absence of 1,25(OH)2D3 for 24 h. Cells were lysed using Trizol reagent and RNA extracted according to manufacturer’s instructions. RNA (1µg) was reverse transcribed using Random hexamers and SuperScript III reverse transcriptase. All real-time PCR reactions were carried out in duplicate in a final volume of 10 µl. The cDNA samples were amplified using iQ SYBR Green Supermix primers (250 nM) specific for CYP24 and GAPDH cDNA. Relative expression between samples was calculated using the comparative cycle threshold (CT) method (∆CT) (24).

210 7.3 Results

7.3.1 Daunorubicin hydrochloride and etoposide inhibited the expression of endogenous CYP24 by 1,25(OH)2D3.

As expected, the results from this study showed that 1,25(OH)2D3 increased the levels of

CYP24 mRNA in COS-1 cells. Unexpectedly, daunorubicin hydrochloride and etoposide abrogated the 1,25(OH)2D3 induction of CYP24 mRNA expression whereas daunorubicin hydrochloride and etoposide per se did not increase the CYP24 mRNA level in COS-1 cells (Figure 7.1).

7.3.2 Pre-treatment of 1,25(OH)2D3 prevented etoposide from suppressing the induction of CYP24 mRNA expression by 1,25(OH)2D3.

Daunorubicin Hydrochloride and etoposide has been shown to inhibit topoisomerase II enzyme which has an important role in DNA replication and transcription (Mao et al.,

2001). It is possible that the inhibition of the endogenous CYP24 induction by

1,25(OH)2D3 is due to the inhibition of the uncoiling of DNA by both daunorubicin hydrochloride and etoposide. To test this hypothesis, COS-1 cells were incubated with

1,25(OH)2D3 for 6 h before adding etoposide which had a stronger synergistic action than daunorubicin hydrochloride. The level of CYP24 mRNA was determined 12 h later.

As shown in Figure 7.2, delaying the exposure of the cells to etoposide permitted the unhindered stimulation of CYP24 expression by 1,25(OH)2D3. However, there was no enhancement of the effect of 1,25(OH)2D3 by the drug. This result suggests that etoposide inhibited the initiation of CYP24 mRNA transcription by 1,25(OH)2D3.

211

0.05

*

0.04

0.03 : GAPDH mRNA

0.02

P24 mRNA *#* *#* CY 0.01

0 1,25(OH)2D3 - + - - + +

Daunorubicin - - + - + - Etoposide - - - + - +

Figure 7.1 Effects of daunorubicin hydrochloride and etoposide on endogenous

CYP24 mRNA expression in COS-1 cells. COS-1 cells were incubated with 2 µM of daunorubicin hydrochloride or 100 µM of etoposide ± 1,25(OH)2D3 for 24 h. Total RNA was extracted and the level of CYP24 mRNA was determined by quantitative real time

PCR. Significance of difference between control and drug or 1,25(OH)2D3 treatment:

*p<0.001; significance of difference between 1,25(OH)2D3 and 1,25(OH)2D3 + drug:

#p<0.001 (ANOVA followed by Tukey-Kramer multiple comparisons test).

212

0.016 * ***

0.012

mRNA

0.008 : GAPDH

P24 mRNA 0.004 CY

0 EtOH + + - - Me SO + - + - 2 1,25(OH)2D3 - - + + Etoposide - + - +

Figure 7.2 Etoposide did not suppress the ability of 1,25(OH)2D3 to stimulate

CYP24mRNA expression in cells which had been preincubated with 1,25(OH)2D3.

COS-1 cells were incubated with vehicle or 1,25(OH)2D3 for 6 h. Etoposide or vehicle was then added and the cells were incubated for a further 12 h. RNA was extracted and the level of CYP24 mRNA was determined. The results are expressed as a ratio of

CYP24 mRNA:GAPDH mRNA. Significance of difference between vehicle and

1,25(OH)2D3: *p<0.05.

213

3.3.3 Etoposide stimulated the degradation of CYP24 mRNA

The above data also suggest that etoposide might have another action in addition to its role as a topoisomerase inhibitor because it did not enhance the effect of 1,25(OH)2D3

(Fig 7.2) or cause the which would have been expected from the CYP24 promoter assays in the preceeding chapters. accumulation of CYP24 mRNA on its own (Figs 7.1 and 7.2),

We therefore investigated whether the drug could also cause the degradation of CYP24 mRNA. To investigate this possibility, COS-1 cells were incubated with 1,25(OH)2D3 for 6 h, washed and actinomycin D was added to stop transcription. The data in Figure

7.3 demonstrate that etoposide accelerated the decline in CYP24 mRNA over the period of study.

214

Figure 7.3 Etoposide accelerated CYP24 mRNA degradation

COS-1 cells in serum deficient media were incubated with 1,25(OH)2D3 (10 nM) for 6 h, washed and the cells were then incubated with actinomycin D (5 µg/ml) and etoposide (

■ ) or Me2SO ( ♦ ). At the indicated times, triplicate wells from each group were removed, total RNA extracted and CYP24 mRNA levels were determined by qRT-PCR.

Significance of difference between etoposide and Me2SO: *p<0.05 (Student’s unpaired t- test). Similar results were obtained in a replicate experiment.

215 7.4 Summary

The results from this study show that daunorubicin hydrochloride and etoposide alone did not cause an increase in CYP24 mRNA in kidney cells as would have been expected from the promoter studies in the preceeding chapters. Incubation of daunorubicin hydrochloride or etoposide together with 1,25(OH)2D3 resulted in the inhibition of

1,25(OH)2D3 induction of CYP24 mRNA expression. This inhibition could be abrogated with 1,25(OH)2D3 pre incubation suggesting the chemotherapeutic drugs inhibited the transcriptional machinery of the endogenous CYP24 promoter. This is most likely to be at the level of topoisomerase II as both drugs inhibit this enzyme. In addition, we demonstrated that etoposide increased the degradation of CYP24 mRNA. These results therefore demonstrate disparate effects of the drugs on transfected CYP24 promoter activity and endogenous CYP24 mRNA expression. The reason for this discrepancy and implications of these finding are discussed in the next chapter.

216

Chapter 8

Discussion

217 8.1 Discussion

8.1.1 Anti-cancer drugs and CYP24 promoter activity

The data demonstrate that a number of chemotherapeutic drugs representing several classes of the drugs were able to stimulate the transcriptional activity of the CYP24 promoter and/or amplify the effect of 1,25(OH)2D3 on the promoter in two kidney cell- lines which had been transiently transfected with one of two constructs that carried the

CYP24 promoter fused to the firefly luciferase gene. One promoter-reporter construct contained a -298 bp promoter sequence (relative to the transcription initiation site) whereas the other contained a -1.4 kb sequence which encompasses the 298 bp sequence.

These promoter sequences contained VDRE1 and VDRE2 (Kerry et al., 1996), an EBS

(Ets-binding site, Dwivedi et al., 2000), a VSE (vitamin D response element, Nutchey et al., 2005), a GC box (Dwivedi et al., manuscript in preparation) and perhaps other as yet unidentified elements. The drugs which were tested included the anthracycline, daunorubicin hydrochloride, the vinca , vincristine sulphate, the alkylating agent, cisplatin, and etoposide, a podophyllotoxin. Between the two kidney cell-lines examined, some subtle differences in drug response were detected. Thus, when the -298 bp promoter was used, daunorubicin hydrochloride per se stimulated promoter activity in COS-1 but not HEK293T cells despite causing a synergistic activation of the promoter with

1,25(OH)2D3 in both cell-lines. The synergistic activation of the -298 bp promoter was greater in the HEK293T than in the COS-1 cells. The difference in drug response between the two kidney cell lines could be attributed to the difference in their origin.

218 While HEK293T is a human cell-line, derived from embryonic kidney and has characteristics of proximal tubular cells, COS-1 came from the kidney of an African green monkey. Despite this difference, both cell-lines responded to 1,25(OH)2D3 with enhanced CYP24 promoter activity and both have been used previously as models to investigate the regulation of CYP24 promoter activity by 1,25(OH2D3 (Dwivedi et al.,

2000, Nutchey et al., 2005). When the -1.4 kb sequence was used, daunorubicin hydrochloride, vincristine sulphate and etoposide per se also stimulated the transcriptional activation of the construct. In contrast, cisplatin alone did not stimulate

CYP24 promoter activity in COS-1 or HEK293T cells, either with the -298 bp or the -

1.4kb promoter construct. All of the four above listed drugs caused a synergistic enhancement in promoter activity when the cells were treated concurrently with a drug and 1,25(OH)2D3 although this effect was smaller with the -1.4 kb than the -298 bp construct in case of daunorubicin hydrochloride. Comparing the data obtained from the two promoter constructs, it can be concluded that the actions of the drugs were predominantly channelled through sequences in the -298 bp promoter.

In contrast to daunorubicin hydrochloride, the hydroxylated derivative, doxorubicin hydrochloride, was inactive on its own and did not amplify the action of 1,25(OH)2D3 on promoter induction, suggesting that not all chemotherapeutic drugs will affect vitamin D metabolism. Consistent with this view, ifosamide, a drug that is used to treat solid tumors in children, has been found not to affect the plasma levels of 1,25(OH)2D3 (de Schepper et al., 2004). In addition, we have also examined whether dexamethasone, a glucocorticoid which is frequently given to patients to prevent chemotherapy-induced

219 nausea and vomiting (Jordan et al., 2005) as well as having direct anti-tumour activity for some cancer-types (Child et al., 2005), could affect CYP24 promoter activity since are known to cause bone loss (Sivagurunathan et al., 2005). Our data (not shown) demonstrate that dexamethasone neither stimulated nor enhanced the effect of

1,25(OH)2D3 on CYP24 promoter activity. Thus, these data exclude an effect of dexamethasone on CYP24 promoter activity and reinforce the view that not all chemotherapeutic drugs will affect vitamin D metabolism.

Very few agents, apart from 1,25(OH)2D3 and calcitonin, have consistently been demonstrated to induce CYP24 expression (Omdahl et al., 2002). However, agents such as insulin (Armbrecht et al., 1996) and PMA (Chen et al., 1993; Nutchey et al., 2005), although they do not increase CYP24 expression by themselves, enhance the action of

1,25(OH)2D3 on CYP24 induction. Thus, our findings place chemotherapeutic drugs in a class of agents that can both stimulate the transcriptional activation of the CYP24 promoter as well as amplify the effects of 1,25(OH) 2D3, as has previously been reported for calcitonin (Gao et al., 2004).

8.1.2 Cytotoxic effects of the drugs in HL60 cells but not in kidney cells

The tested drugs were found to readily kill the human promyelocytic leukaemic HL60 cells at relatively low doses but the drugs did not affect the viability of the kidney cell- lines, even at relatively high but pharmacologically achievable doses. Primary rat

220 hepatocytes also showed substantial resistance to the cytotoxic actions of the drugs (data not shown). Thus, it is unlikely that toxicity to kidney or liver cells and hence blockade of the bioactivation of 1,25(OH)2D3 could contribute in any significant manner to the reduction in plasma level 25(OH)D3 and 1,25(OH)2D3 levels in the patients.

8.1.3 Lack of a role for the VDR and VDREs in the actions of the drugs

The actions of 1,25(OH)2D3 are mediated via the VDR and overexpression of the nuclear

VDR in COS-1 cells greatly facilitates the ability of the hormone to stimulate CYP24 promoter activity (Dwivedi et al., 2002) and this was confirmed in the studies undertaken in this thesis. To investigate whether the actions of the drugs required the VDR, the cells were transiently transfected with the nuclear VDR and transcriptional activation of the

CYP24 promoter was investigated. The data suggest that the stimulatory effect of daunorubicin hydrochloride per se on CYP24 promoter activity was unlikely to have involved the VDR because VDR overexpression not only did not augment the action of the drug but it suppressed promoter induction by the drug. Consistent with this conclusion, mutating VDRE1 and VDRE2 within the CYP24 promoter did not affect promoter induction by daunorubicin hydrochloride. In contrast, overexpression of the

VDR augmented the effect of 1,25(OH)2D3 whereas mutating the VDREs greatly diminished the stimulatory effect of 1,25(OH)2D3. To demonstrate that the drugs did not directly affect VDR activity, we used a construct which contained a consensus VDRE of the DR3-type (two AGGTCA motifs separated by 3 bp), cloned immediately upstream of

221 the thymidine kinase promoter. When transfected in COS-1 cells, our data show that

1,25(OH)2D3 but not daunorubicin hydrochloride stimulated the activity of this reporter construct. The drug also did not alter the response to 1,25(OH)2D3. These data clearly demonstrate that the VDR and VDREs do not mediate the actions of drugs such as daunorubicin hydrochloride.

8.1.4 Daunorubicin and etoposide increased the expression of the VDR

Despite the VDR-independent action of daunorubicin hydrochloride per se on promoter activity, the VDR could be involved in mediating the synergistic effect of the drug and

1,25(OH)2D3. One likely scenario was that daunorubicin hydrochloride could increase the expression of the VDR. By analogy to the VDR-overexpression experiments in which transient transfection with the nuclear VDR augmented the response to 1,25(OH)2D3, a drug-induced enhancement in the expression of the VDR could have a similar effect.

Indeed, the data demonstrate that treatment of the cells with daunorubicin hydrochloride increased the expression of the VDR. This effect was also seen with etoposide. It has previously been demonstrated that etoposide could increase VDR mRNA and protein expression in leukaemic cells. Studies by Torres et al. (2000) have demonstrated that treating HL-60 and U-937 cells with 0.15 mM of etoposide for 72 h significantly increased VDR protein expression and 1,25(OH)2D3 binding activity. These authors showed that the increase in VDR mRNA expression by 0.15 mM etoposide was time dependent. Together, these data demonstrate that etoposide and daunorubicin can up- regulate the expression of the VDR in some cell-types, due primary to an increase in

222 VDR transcription (Torres et al., 2000). Although the effects of the other drugs on VDR expression were not investigated, the data strongly suggest that while the actions of the drugs per se did not require the VDR, drug-induced increase in VDR expression could be involved, at least in part, in mediating the synergistic effects of the drugs and

1,25(OH)2D3.

8.1.5 Reactive oxygen species and the promoter inducing effect of daunorubicin and etoposide

The anti-cancer actions of several anti-cancer drugs are dependent on the generation of reactive oxygen species (Gervasoni et al., 2004; Huang et al., 2003; Wu et al., 2002).

Thus, it was of interest to investigate whether the promoter-activating effects of daunorubicin and etoposide required the generation of reactive oxygen species. We used

N-acetylcysteine which is a precursor in the formation of the antioxidant glutathione in the body. The thiol (sulfhydryl) group confers antioxidant effects and is able to reduce free radicals. Contrary to the requirements for reactive oxygen species in the cytotoxic action of anthracyclines (Wu et al., 2002), the data demonstrate that reactive oxygen species were not required for daunorubicin hydrochloride to stimulate the activity of the

CYP24 promoter. Thus, N-acetylcysteine did not affect the action of daunorubicin hydrochloride. Similarly, N-acetylcysteine also did not affect the action of etoposide on the CYP24 promoter activity. Interestingly, treatment of the cells with N-acetylcysteine increased CYP24 promoter activity in response to 1,25(OH)2D3. Oxygen is a co-

223 substrate for CYP24 enzyme activity. The observation that the anti-oxidant enhanced the ability of 1,25(OH)2D3 to induce CYP24 promoter activity is most likely due to decreased activity of endogenous CYP24 activity with a concomitant increase in intracellular 1,25(OH)2D3. The redox state of the cell is likely to be altered by

1,25(OH)2D3 since the hormone has previously been reported to deplete cellular reduced glutathione. This has been suggested as a mechanism by which the hormone exerts its anti-cancer action (Ravid and Koren, 2003). Thus, our data with N-acetylcysteine are not only consistent with 1,25(OH)2D3 altering the redox state of the cells but also reveal that oxidative processes initiated by 1,25(OH)2D3 exert a negative impact on the induction of

CYP24 promoter by the hormone.

8.1.6 Non-genomic actions of 1,25(OH)2D3 involving intracellular signalling molecules

Whereas the genomic actions of 1,25(OH)2D3 are well established, recent studies have strongly suggested that the hormone also has non-genomic actions. The best examples are the observations that some of the actions of 1,25(OH)2D3 occur at kinetics which are incompatible with a genomic action and these include the activation of intracellular signalling molecules. Thus, we and others have demonstrated that 1,25(OH)2D3 not only caused the activation of PKC and the MAP kinases (Nutchey et al., 2005; Dwivedi et al.,

2002; Chae et al., 2002; Bikle et al., 2001) but the action of 1,25(OH)2D3 on CYP24 promoter induction was also shown to be mediated by the ERK1/ERK2 and ERK5

224 modules in COS-1 cells (Dwivedi et al., 2002). A recent study in HEK293T cells has demonstrated that PKCβ and JNK were also required for promoter activation by the hormone (Nutchey et al., 2005). In this study, PMA and 1,25(OH)2D3 were found to cause a synergistic activation of the CYP24 promoter. This well known phenomenon was found to be protein kinase C-dependent. Whereas ERK1/ERK2 was not activated by

1,25(OH)2D3 alone, its activation by PMA was potentiated by 1,25(OH)2D3 also. The importance of ERK1/2 for transcriptional synergy was demonstrated by transfection of the dominant-negative ERK1(K71R) mutant, which resulted in a reduced level of synergy on CYP24 promoter activity. JNK, in addition to being necessary for promoter activity incuced by 1,25(OH)2D3, was also shown to be required for synergy (Nutchey et al.,

2005). Whereas the likely physiological target of ERK1/ERK2 in COS-1 cells was

RXRα, ERK5 targeted Ets-1. We were able to demonstrate a direct interaction between

ERK1/ERK2 and RXRα, ERK1/ERK2 and Ets-1, and ERK5 and Ets-1. Whereas the interaction between ERK1/ERK2 and RXRα and between ERK1/ERK2 and Ets-1 was constitutive, the interaction between ERK5 and Ets-1 was induced by 1,25(OH)2D3.

Interestingly, Waas and Dalby (2001) reported that a His-tagged Ets-1 (1-138) was a good substrate for ERK2 in cell-free assays. Although we could not demonstrate such phosphorylation by ERK1 in cell-free assays using full length GST-Ets-1, our finding that Ets-1 showed constitutive binding to ERK1/ERK2 suggests a possible functional relationship between the MAP kinase and Ets-1 in intact cells. The precise reason for the difference between our data and those of Waas and Dalby (2001) is not known but could be due to different assay conditions. We used full length Ets-1 whereas Waas and Dalby

225 (2001, 2002, 2003) used a truncated Ets-1; we used 10 mM Mg2+ in our buffers whereas

Waas and Dalby used 20 mM Mg2+ and 100 mM K+.

As described in Chapter 1, RXRα and VDR form a complex which binds to the VDREs.

This is responsible for mediating the genomic action of the hormone. Our data demonstrate that S260 on RXRα is the ERK1/ERK2 phosphorylation site. On the other hand, Ets-1 binds to the EBS (5'-AGAGGATGGA-3' containing the core 5'-GGAT-3' consensus sequence) at -128/-119 (Dwivedi et al., 2000). The likely residue on Ets-1 which was phosphorylated by ERK5 was shown to be T38 since a T→A mutation at this residue rendered the protein non-phosphorylatable by ERK5. The function of Ets-1 was also lost when this mutation was introduced into the transcription factor. The studies by

Nutchey et al, (2005) demonstrate that JNK targets an unidentified transcription factor which interacts with the VSE (vitamin D stimulatory element) at -132/-126. Data obtained in HEK293T cells by Dwivedi et al., (manuscript in preparation) have also demonstrated that PI3 kinase and PKCζ target the transcription factor, Sp1, which binds to a GC box (5’-GGGGCGGGT-3’) at -114/-101. In this case, Sp1, acting via the GC box

(Gao et al., 2004), was demonstrated to regulate basal as well as 1,25(OH)2D3-mediated induction of the CYP24 promoter activity. These data demonstrate that a number of intracellular signalling molecules are involved in regulating the transcriptional activation of the CYP24 promoter by 1,25(OH)2D3 with Sp1 regulating the basal transcriptional activity.

226 8.1.7 Anti-cancer drugs and intracellular signalling molecules in CYP24 promoter activation

Many chemotherapeutic drugs, including daunorubicin hydrochloride and etoposide, have been found to stimulate the activities of the above signalling molecules. This is confirmed in this thesis. We have now presented data which demonstrate that the signalling molecules, ERK1/ERK2, ERK5, PI3 kinase and JNK were involved in mediating the effects of daunorubicin hydrochloride and etoposide per se on promoter activity in COS-

1 cells. Thus, promoter induction by the drugs was inhibited when the cells were transiently transfected with the dominant negative mutant of a target kinase or its upstream kinase, or incubated with an inhibitor of the target kinase. Consistent with a role for the MAP kinases in promoter induction, the drugs also stimulated the activities of

ERK1/ERK2, ERK5 and JNK in COS-1 cells. The lower than expected fold stimulation for some of the kinases was probably due to a high baseline kinase activity in the kidney cell-line compared to other cell-lines. The MAP kinases were also required for the synergistic activation of the CYP24 promoter in HEK293T cells by the drugs and

1,25(OH)2D3. Although previous studies have reported that anthracyclines and etoposide can stimulate the activities of ERK1/ERK2 and JNK (Boldt et al., 2002; Tang et al.,

2002; Brantley-Finley et al., 2003; Mas et al., 2003), their effects on ERK5 have not been investigated. Our data therefore extends the number of MAP kinases that these chemotherapeutic drugs can activate. The precise targets of the kinases within the -298 promoter remain to be identified but we can exclude the VDREs and GC-box as targets.

227 Our recent studies in HEK293T cells have demonstrated that 1,25(OH)2D3 can also stimulate the activity of PKCζ (Dwivedi et al, manuscript in preparation). With respect to this hormone, PKCζ appears to target Sp1 and the GC box. In contrast to the MAP kinases, our data suggest that PKCζ played a more selective role, depending on the drug and cell-line tested. Thus, while the dominant negative PKCζ mutant blocked the action of daunorubicin hydrochloride in COS-1 cells, it did not affect the response to etoposide in this cell-line. The PKCζ mutant did not affect drug-induced promoter activation in

HEK293T cells, suggesting that this PKC isozyme did not play any roles in mediating

CYP24 promoter induction by the drugs in HEK293T cells. These results again highlight the subtle differences between COS-1 and HEK293T kidney cells.

An interesting finding was that the hydroxylated form, doxorubicin hydrochloride, does not induce CYP24 promoter activity, despite being reported to stimulate the activity of

ERK1/ERK2 in a variety of cell cancer lines (Brantley-Finley et al., 2003; Yeh et al.,

2004). Although doxorubicin hydrochloride has been shown to be more active against solid tumour compared to daunorubicin hydrochloride (Weiss., 1992), overall, there are very little differences between doxorubicin hydrochloride and daunorubicin hydrochloride in terms of toxicity, pharmacokinetics and stability (Dine et al.,1992;

Maniez-Devos et al., 1986; Heijn et al., 1999). Like daunorubicin hydrochloride, doxorubicin hydrochloride has been shown to kill cancer cell by binding to DNA and causing the DNA to break (Swift et al., 2006). Furthermore, doxorubicin has been shown to inhibit topoisomerase II enzyme thus disrupting transcription and replication (Swift et al., 2006). Doxorubicin can also bind to cell membranes and sarcoplasmic reticulum and

228 interrupts the ion transport (Earm et al., 1994). It also cause the generation of oxygen free radicals (Tsang et al., 2003). Our data therefore suggest that activation of ERK1/ERK2 per se is insufficient for the induction of the CYP24 promoter activity.

Overall our findings are consistent with a model for daunorubicin hydrochloride or etoposide induction that involves the phosphorylation and activation by ERK1/ERK2,

ERK5, JNK and/or PKCζ/PI3 kinase of specific transcription factors associated with the first –298bp of the CYP24 promoter. Transcription factors known to enhance basal activity of the first –298bp of promoter and thus possible targets for the ERKs include

Ets-1, Sp1, and a VSE and CCAATT box binding proteins (Gao et al., 2004; Nutchey et al., 2005, Dwivedi et al., in preparation). A key issue is the question of how the chemotherapeutic drugs activate the kinases. Mas et al, (2003) reported that the ability of daunorubicin hydrochloride to stimulate the activity of ERK1/ERK2 in U937 cells was blocked by kinase dead DN PKCζ, demonstrating that this PKC isozyme is involved in regulating the activity of ERK1/ERK2. In cardiac myocytes, the stimulation of

ERK1/ERK2 activity by the anthracyclines was demonstrated to involve the small G protein, p21ras (Zhu et al., 1999). It is possible that p21ras could also be involved in the activation of not just ERK1/ERK2 but also ERK5 by the drugs in COS-1 cells since we and others have previously demonstrated this p21ras-ERK relationship under different conditions (English et al., 1999; Dwivedi et al., 2002). Unlike ERK activation by

1,25(OH)2D3 that most likely involves binding of the hormone to VDR located at the cytoplasmic membrane, our work eliminates the involvement of VDR and VDRE in the induction of the CYP24 promoter by daunorubicin hydrochloride or etoposide per se

229 although its ability to up-regulate the expression of endogenous VDR could be responsible, at least in part, for enhancing the effect of 1,25(OH)2D3 on promoter induction.

8.1.8 Suppression of CYP24 mRNA induction by daunorubicin and etoposide

Lastly, we also measured the levels of endogenous CYP24 mRNA in COS-1 cells using real time quantitative PCR. Unexpectectly, we found that daunorubicin hydrochloride and etoposide did not increase CYP24 mRNA expression. Furthermore, the drugs inhibited

1,25(OH)2D3 induction of CYP24 mRNA expression. Pre-incubating the cells with

1,25(OH)2D3 for 6 h before etoposide treatment resulted in the unhindered stimulation of

CYP24 expression by 1,25(OH)2D3. These results suggest that etoposide may be inhibiting the initiation of the transcriptional activation by 1,25(OH)2D3. One well- characterised action of daunorubicin hydrochloride and etoposide as anti-cancer drugs is the inhibition of topoisomerase II activity, thereby inhibiting DNA transcription (Minotti et al., 2004; Shimada et al., 2003). Topoisomerase enzyme plays a crucial role in gene transcription by helping DNA to uncoil (Mao et al., 2001). The inhibition of topoisomerase II by etoposide may explain the attenuation of 1,25(OH)2D3 induction of

CYP24 mRNA expression in the presence of etoposide. Since plasmid DNA do not require uncoiling for transcription to take place, daunorubicin hydrochloride and etoposide were able to stimulate the CYP24 promoter and hence resulting in an increase

230 in luciferase activity. However, the lack of an enhancing effect of etoposide also suggests that the drug may be increasing the degradation of the CYP24 mRNA. Using actinomycin

D to stop transcription, we were able to confirm that etoposide facilitates the degradation of CYP24 mRNA. This unique ability of etoposide to up-regulate CYP24 promoter while enhancing degradation of its mRNA has also been observed previously with parathyroid hormone (PTH). Earlier studies by Zierold et al, 2000 have shown that PTH enhanced the action of 1,25(OH)2D3 induction of the CYP24 promoter in AOK-B50 cells. However, further studies by Zierold et al, 2002 demonstrated that PTH reduces the half-life of

CYP24 mRNA by 4.2-fold. Whether the increase in degradation of the CYP24 mRNA caused by the drugs translates to a decrease in 24-hydroxylase enzyme and therefore a reduction in the degradation of 1,25(OH)2D3 is yet to be determine.

8.1.9 Clinical implications of this work and future directions

Apart from the classical role of 1,25(OH)2D3 in Ca2+ homeostasis, 1,25(OH)2D3 and analogues of 1,25(OH)2D3 have attracted immense interest as a novel therapy to treat malignancies. 1,25(OH)2D3 and analogues have anti-proliferative actions and they sensitize cancer cells to chemotherapeutic drugs (Trump et al., 2004). The idea that chemotherapeutic drugs could used in conjunction with 1,25(OH)2D3 to treat advanced tumours have already attracted immense interest. In novel approaches to treat advanced prostate cancers, several groups have focussed on combination therapy using

1,25(OH)2D3 and conventional drugs. This is based on the observations that

231 1,25(OH)2D3 and analogues can enhance, either synergistically or additively, the anti- tumor activities of several classes of anti-neoplastic agents, including platinum compounds, docetaxel, etoposide and paclitaxel (Beer, 2005; Moffatt et al., 1999).

Sevaral clinical trials are currently underway to evaluate the efficacy of combining

1,25(OH)2D3 and drugs such as docetaxel (taxotere) in the treatment of androgen- independent prostate cancer (Beer, 2005; Vijayakumar et al., 2005). How such combinations enhance the anti-tumor activity of each other is poorly understood.

Nevertheless, our data demonstrate that some anti-neoplastic agents cause CYP24 mRNA instability and prevent 1,25(OH)2D3 from increasing the level of CYP24, despite stimulating the promoter activity in transfection assays. There are possible important medical repercussions for anti-cancer therapy. In combination chemotherapy involving

1,25(OH)2D3 and an anti-neoplastic agent, relatively high doses of 1,25(OH)2D3 are generally required with the concomitant increase in risk of adverse side effects. Indeed, a major hurdle to achieving a favourable clinical response is the need for doses that are just below those which cause hypercalcaemia and hypercaluria (Beer, 2005). Moreover,

CYP24 is regarded as an oncogene and in some cancers, its expression is elevated compared to normal tissues (Albertson et al., 2000). This is likely to provide an escape mechanism for these cancer cells to avoid the cytotoxic and anti-proliferative effects of

1,25(OH)2D3. Our findings that anti-cancer drugs increased the expression of the VDR and prevented the expression of CYP24 suggest that it is possible to reduce the dose of

1,25(OH)2D3 to achieve a favorable clinical response because increased VDR expression would increase the sensitivity as well as responsiveness of cancer cells to 1,25(OH)2D3

232 while the inability of 1,25(OH)2D3 to induce CYP24 expression would maintain the bioavailability of circulating 1,25(OH)2D3 over a longer period.

While our data on CYP24 mRNA expression have provided some novel insights into how it may be possible to increase the efficacy of 1,25(OH)2D3 in anti-cancer treatments, more studies are required to expand this field and elucidate the mechanism by which the drugs promote CYP24 mRNA degradation. Future studies could investigate whether other anti-cancer agents can also promote CYP24 mRNA degradation and stimulate VDR expression. These studies could also explore the effect of chemotherapeutic drugs on the rate of 1,25(OH)2D3 degradation, either alone or in combination with other agents. Other studies could explore the effects of anti-cancer drugs on CYP24 expression in cancer cells especially those that express high levels of CYP24 e.g. DU145 prostate cancer cells

(Krishnan et al., 2003). In addition to kidney and cancer cells, other cell-types such as bone cells and monocytes are also known to express CYP24. Thus, it would be important to conduct similar studies using these cell-types. Finally, animal models with established tumours could be used to investigate the effects of chemotherapeutic drugs on circulating levels of 1,25(OH)2D3 and the levels of 1,25(OH)2D3 within tumour cells. These can be investigated in the nude mice model which have been transplanted with human tumours or the Transgenic Adenocarcinoma of the Mouse Prostate (TRAMP) model which spontaneously develops prostate cancer.

233 8.2 Summary

In summary, the data in the present study demonstrate, for the first time, that certain anti- cancer drugs could stimulate the activity of the CYP24 promoter and enhance the induction of promoter activity by 1,25(OH)2D3. In addition, this study also demonstrate that ERK1/ERK2, ERK5, PKCζ and PI3K played an important role in the promoter induction by the anti-cancer drugs. These results are summarised in Table 8.1 and Table

8.2. However, the increase in promoter activity by chemotherapeutic drugs did not translate to an increase in CYP24 mRNA expression. Furthermore, daunorubicin hydrochloride and etoposide inhibited the expression of CYP24 mRNA by 1,25(OH)2D3.

The inhibitory effects were attributed to the inhibition of transcriptional activation and enhanced degradation of CYP24 mRNA by etoposide. Thus, although our data cannot explain why children undergoing chemotherapy had reduce levels of 1,25(OH)2D3, these results, nevertheless, open new avenues for the use of chemotherapy drugs in conjunction with 1,25(OH)2D3 to treat cancer and provide further justification for such use.

234 Table 8.1 Summary of the effects of chemotherapeutic drugs on the -298bp and -

1.4kb CYP24 promoter constructs in the presence or absence of 1,25(OH)2D3 in kidney cell lines

Cell Type COS-1 HEK293T

CYP24 298 bp 1.4 Kbp 298 bp 1.4 Kbp promoter Treatment Type EtOH 1,25D EtOH 1,25D EtOH 1,25D EtOH 1,25D

Daunorubicin + +++ + ++ + +++ + - Hydrochloride

Etoposide + +++ + +++ ND ND + +++

Vincristine + - + +++ ND ND - +++ Sulfate

Cisplatin - +++ - +++ ND ND - -

Doxorubicin - - ND ND ND ND ND ND Hydrochloride

+ = Increase in CYP24 promoter activity ++ = Enhance CYP24 promoter activity with 1,25(OH)2D3 +++ = Synergistic increase in CYP24 promoter activity with 1,25(OH)2D3 ND = Not determined

235 Table 8.2 Summary of the effects of dominant negative ERK1, MEK5, MEK4,

PKCζ mutants and inhibitor of PI3K on CYP24 promoter induction by daunorubicin hydrochloride and etoposide in kidney cell lines.

Cell Line COS-1 HEK293T

Promoter 298 bp 1.4 Kb 298 bp

Treatment Daunorubicin Etoposide Daunorubicin Hydrochloride Hydrochloride

ERK1K71R - MEK5A - -

ERK1K71R + ND ND MEK5A

MKK4K116R

PKCζK281M - -

LY294002 ND ND

= Inhibition of CYP24 promoter activity

= No changes in CYP24 promoter activity

ND = Not Determined

236 References

Abrams SA, O’Brien KO. (2004) Calcium and bone mineral metabolism in children with chronic illnesses. Annu Rev Nutr. 24: 13-32.

Adorini L. (2003) Tolerogenic dendritic cells induced by vitamin D receptor ligands enhance regulatory T cells inhibiting autoimmune diabetes. Ann N Y Acad Sci. 987: 258-

61.

Akeno N, Matsunuma A, Maeda T, Kawane T, Horiuchi N. (2000) Regulation of vitamin

D-1alpha-hydroxylase and -24-hydroxylase expression by dexamethasone in mouse kidney. J Endocrinol. 164:339-48.

Akhter J, Lu Y, Finlay I, Pourgholami MH, Morris DL. (2001) 1alpha,25-

Dihydroxyvitamin D3 and its analogues, EB1089 and CB1093, profoundly inhibit the in vitro proliferation of the human hepatoblastoma cell line HepG2. ANZ J Surg. 7: 414-7.

al-Aqeel A, Ozand P, Sobki S, Sewairi W, Marx S. (1993) The combined use of intravenous and oral calcium for the treatment of vitamin D dependent rickets type II

(VDDRII). Clin Endocrinol (Oxf). 39: 229-37.

237 Alberola-Ila J, Hernandez-Hoyos G (2003) The Ras/MAPK cascade and the control of positive selection. Immunol Rev. 191: 79-96.

Albertson DG, Ylstra B, Segraves R, Collins C, Dairkee SH, Kowbel D, Kuo WL, Gray

JW, Pinkel D. (2000) Quantitative mapping of amplicon structure by array CGH identifies CYP24 as a candidate oncogene. Nat Genet. 25: 144-6.

Allegretto EA, Shevde N, Zou A, Howell SR, Boehm MF, Hollis BW, Pike JW. (1995)

Retinoid X receptor acts as a hormone receptor in vivo to induce a key metabolic enzyme for 1,25-dihydroxyvitamin D3. J Biol Chem. 270: 23906-9.

Amling M, Priemel M, Holzmann T, Chapin K, Rueger JM, Baron R, Demay MB. (1999)

Rescue of the skeletal phenotype of vitamin D receptor-ablated mice in the setting of normal mineral ion homeostasis: formal histomorphometric and biomechanical analyses.

Endocrinology. 140: 4982-7.

Anderson DM, Maraskovsky E, Billingsley WL, Dougall WC, Tometsko ME, Roux ER,

Teepe MC, DuBose RF, Cosman D, Galibert L. 1997. A homologue of the TNF receptor and its ligand enhance T-cell growth and dendritic-cell function. Nature 390: 175-179.

238 Anderson PH, O'loughlin PD, May BK, Morris HA. (2005) Modulation of CYP27B1 and

CYP24 mRNA expression in bone is independent of circulating 1,25(OH)(2)D(3) levels.

Bone. [Epub ahead of print]

Andersson S, Davis DL, Dahlback H, Jornvall H, Russell DW. (1989) Cloning, structure, and expression of the mitochondrial cytochrome P-450 sterol 26-hydroxylase, a bile acid biosynthetic enzyme. J Biol Chem. 264: 8222-9.

Andreka P, Zang J, Dougherty C, Slepak TI, Webster KA, Bishopric NH. (2001)

Cytoprotection by Jun kinase during nitric oxide-induced cardiac myocyte apoptosis.

Circ Res. 88: 305-12.

Andress DL, Ozuna J, Tirschwell D, Grande L, Johnson M, Jacobson AF, Spain W

(2002). Antiepileptic drug-induced bone loss in young male patients who have seizures.

Arch Neurol. 59:781-6.

Arbour NC, Prahl JM, DeLuca HF. (1993) Stabilization of the vitamin D receptor in rat osteosarcoma cells through the action of 1,25-dihydroxyvitamin D3. Mol Endocrinol.

7:1307-12.

Arikoski P, Komulainen J, Riikonen P, Voutilainen R, Knip M, Kroger H. (1999a)

Alterations in bone turnover and impaired development of bone mineral density in newly

239 diagnosed children with cancer: a 1-year prospective study. J Clin Endocrinol Metab.

84:3174-81.

Arikoski P, Kroger H, Riikonen P, Parviainen M, Voutilainen R, Komulainen J (1999b)

Disturbance in bone turnover in children with a malignancy at completion of chemotherapy. Medical and Pediatric Oncology. 33:455-461

Armbrecht HJ, Boltz MA, Hodam TL, Kumar VB. (2001) Differential responsiveness of intestinal epithelial cells to 1,25-dihydroxyvitamin D3--role of protein kinase C. J

Endocrinol. 169:145-51.

Armbrecht HJ, Hodam TL, Boltz MA (2003) Hormonal regulation of 25-hydroxyvitamin

D3-1alpha-hydroxylase and 24-hydroxylase gene transcription in opossum kidney cells.

Arch Biochem Biophys. 409: 298-304.

Armbrecht HJ, Hodam TL, Boltz MA, Partridge NC, Brown AJ, Kumar VB. (1998)

Induction of the vitamin D 24-hydroxylase (CYP24) by 1,25-dihydroxyvitamin D3 is regulated by parathyroid hormone in UMR106 osteoblastic cells. Endocrinology.

139:3375-81.

Armbrecht HJ, Wongsurawat VJ, Hodam TL, Wongsurawat N. (1996) Insulin markedly potentiates the capacity of parathyroid hormone to increase expression of 25-

240 hydroxyvitamin D3-24-hydroxylase in rat osteoblastic cells in the presence of 1,25- dihydroxyvitamin D3. FEBS Lett. 393:77-80.

Atkinson SA, Halton JM, Bradley C, Wu B, Barr RD. (1998) Bone and mineral abnormalities in childhood acute lymphoblastic leukemia: influence of disease, drugs and nutrition. Int J Cancer Suppl. 11:35-9.

Aubin J, Heersche J (1997) Vitamin D and osteoblast. In Feldman D, Glorieux F, Pike J

(eds): “Vitamin D.” London: Academic Press Limited, pp 313-328.

Baier R, Bondeva T, Klinger R, Bondev A, Wetzker R. (1999) Retinoic acid induces selective expression of phosphoinositide 3-kinase gamma in myelomonocytic U937 cells.

Cell Growth Differ. 10:447-56.

Balogh G, Boland R, de Boland AR. (2000) 1,25(OH)(2)-vitamin D(3) affects the subcellular distribution of protein kinase C isoenzymes in rat duodenum: influence of aging. J Cell Biochem. 79:686-94.

Balsan S, Garabedian M, Larchet M, Gorski AM, Cournot G, Tau C, Bourdeau A, Silve

C, Ricour C. (1986) Long-term nocturnal calcium infusions can cure rickets and promote normal mineralization in hereditary resistance to 1,25-dihydroxyvitamin D. J Clin Invest.

77:1661-7.

241 Baran DT, Quail JM, Ray R, Leszyk J, Honeyman T. (2000) Annexin II is the membrane receptor that mediates the rapid actions of 1alpha,25-dihydroxyvitamin D(3). J Cell

Biochem. 78:34-46.

Bareis P, Bises G, Bischof MG, Cross HS, Peterlik M. (2001) 25-hydroxy-vitamin d metabolism in human colon cancer cells during tumor progression. Biochem Biophys Res

Commun. 285:1012-7.

Bareis P, Kallay E, Bischof MG, Bises G, Hofer H, Potzi C, Manhardt T, Bland R, Cross

HS. (2002) Clonal differences in expression of 25-hydroxyvitamin D(3)-1alpha- hydroxylase, of 25-hydroxyvitamin D(3)-24-hydroxylase, and of the vitamin D receptor in human colon carcinoma cells: effects of epidermal growth factor and 1alpha,25- dihydroxyvitamin D(3). Exp Cell Res. 276:320-7.

Barletta F, Dhawan P, Christakos S. (2004) Integration of hormone signaling in the regulation of human 25(OH)D3 24-hydroxylase transcription. Am J Physiol Endocrinol

Metab. 286: E598-608.

Barletta F, Freedman LP, Christakos S. (2002) Enhancement of VDR-mediated transcription by phosphorylation: correlation with increased interaction between the VDR and DRIP205, a subunit of the VDR-interacting protein coactivator complex. Mol

Endocrinol. 16: 301-14.

242 Barr RK, Bogoyevitch MA. (2001) The c-Jun N-terminal protein kinase family of mitogen-activated protein kinases (JNK MAPKs). Int J Biochem Cell Biol. 33:1047-63.

Beer TM. (2005) ASCENT: the androgen-independent prostate cancer study of calcitriol enhancing taxotere. BJU Int. 96:508-13.

Beer TM, Munar M, Henner WD. (2001) A Phase I trial of pulse calcitriol in patients with refractory malignancies: pulse dosing permits substantial dose escalation. Cancer.

91:2431-9.

Belkowski SM, Levine JE, Prystowsky MB. (1999) Requirement of PI3-kinase activity for the nuclear transport of prolactin in cloned murine T lymphocytes. J Neuroimmunol.

94:40-7.

Benini S, Manara MC, Cerisano V, Perdichizzi S, Strammiello R, Serra M, Picci P,

Scotlandi K. (2004) Contribution of MEK/MAPK and PI3-K signaling pathway to the malignant behavior of Ewing's sarcoma cells: therapeutic prospects. Int J Cancer.

108:358-66.

Beno DW, Brady LM, Bissonnette M, Davis BH. (1995) Protein kinase C and mitogen- activated protein kinase are required for 1,25-dihydroxyvitamin D3-stimulated Egr induction. J Biol Chem. 270:3642-7.

243 Berry JL, Davies M, Mee AP. (2002) Vitamin D metabolism, rickets, and osteomalacia.

Semin Musculoskelet Radiol. 6:173-82.

Bianchi K, Rimessi A, Prandini A, Szabadkai G, Rizzuto R. (2004) Calcium and mitochondria: mechanisms and functions of a troubled relationship. Biochim Biophys

Acta. 1742:119-31

Bikle DD, Ng D, Tu CL, Oda Y, Xie Z. (2001) Calcium- and vitamin D-regulated keratinocyte differentiation. Mol Cell Endocrinol. 177:161-71.

Bisogno G, Ferrari A, Bergeron C, Scagnellato A, Prete A, Alaggio R, Casanova M,

D'Angelo P, Di Cataldo A, Carli M. (2005) The IVADo regimen--a pilot study with ifosfamide, vincristine, actinomycin D, and doxorubicin in children with metastatic soft tissue sarcoma: a pilot study of behalf of the European pediatric Soft tissue sarcoma

Study Group. Cancer. 103:1719-24.

Bjorkhem I., and Moberg K. (1995) Inborn errors in bile acid biosynthesis and storage of steroids other than cholesterol. In Scriver C, beandet A., Sly W., Valk D (eds) “The

Metabolic basis of Inherited Disease”. New York: McGraw-Hill, pp 2073-2099

244 Bjorling-Poulsen M, Issinger OG. (2003) cDNA array analysis of alterations in gene expression in the promyelocytic leukemia cell line, HL-60, after apoptosis induction with etoposide. Apoptosis. 8:377-88.

Bleyer WA (1990) The impact of childhood cancer on the United States and the world.

CA Cancer J Clin. 40:355–367

Bliziotes M, Yergey AL, Nanes MS, Muenzer J, Begley MG, Viera NE, Kher KK,

Brandi ML, Marx SJ (1988) Absent intestinal response to calciferols in hereditary resistance to 1,25-dihydroxyvitamin D: documentation and effective therapy with high dose intravenous calcium infusions. J Clin Endocrinol Metab 66:294–300

Bogoyevitch MA, Court NW. (2004) Counting on mitogen-activated protein kinases--

ERKs 3, 4, 5, 6, 7 and 8. Cell Signal. 16:1345-54.

Boldt S, Weidle UH, Kolch W (2002). The role of MAPK pathways in the action of chemotherapeutic drugs. Carcinogenesis, 11:1831-8.

Bollinger Bollag W, Bollag RJ. (2001) 1,25-Dihydroxyvitamin D(3), phospholipase D and protein kinase C in keratinocyte differentiation. Mol Cell Endocrinol. 177:173-82.

245 Bouillon R., Van Cromphaut S., Carmeliet G. (2003) Intestinal calcium absorption:

Molecular vitamin D mediated mechanisms. J. Cell Biochem. 88, 2, 332-339.

Boulton TG, Yancopoulos GD, Gregory JS, Slaughter C, Moomaw C, Hsu J, Cobb MH.

(1990) An insulin-stimulated protein kinase similar to yeast kinases involved in cell cycle control. Science. 249:64-7.

Bourlon PM, Faure-Dussert A, Billaudel B. (1997) Modulatory role of 1,25 dihydroxyvitamin D3 on pancreatic islet insulin release via the cyclic AMP pathway in the rat. Br J Pharmacol. 121:751-8.

Boyan BD, Schwartz Z. (2004) Rapid vitamin D-dependent PKC signaling shares features with estrogen-dependent PKC signaling in cartilage and bone. Steroids. 69:591-

7.

Boyan BD, Sylvia VL, Dean DD, Del Toro F, Schwartz Z. (2002) Differential regulation of growth plate chondrocytes by 1alpha,25-(OH)2D3 and 24R,25-(OH)2D3 involves cell- maturation-specific membrane-receptor-activated phospholipid metabolism. Crit Rev

Oral Biol Med. 13:143-54.

246 Boyan BD, Sylvia VL, Dean DD, Schwartz Z. (2002) Membrane mediated signaling mechanisms are used differentially by metabolites of vitamin D(3) in musculoskeletal cells. Steroids. 67:421-7.

Brandlin I, Eiseler T, Salowsky R, Johannes FJ. (2002) Protein kinase C(mu) regulation of the JNK pathway is triggered via phosphoinositide-dependent kinase 1 and protein kinase C(epsilon). J Biol Chem. 277:45451-7.

Brantley-Finley C, Lyle CS, Du L, Goodwin ME, Hall T, Szwedo D, Kaushal GP,

Chambers TC. (2003) The JNK, ERK and p53 pathways play distinct roles in apoptosis mediated by the antitumor agents vinblastine, doxorubicin, and etoposide. Biochem

Pharmacol. 66, 3:459-69.

Brasitus TA, Dudeja PK, Eby B, Lau K. (1986) Correction by 1-25- dihydroxycholecalciferol of the abnormal fluidity and lipid composition of enterocyte brush border membranes in vitamin D-deprived rats. J Biol Chem. 261:16404-9.

Brenza HL, DeLuca HF. (2000) Regulation of 25-hydroxyvitamin D3 1alpha- hydroxylase gene expression by parathyroid hormone and 1,25-dihydroxyvitamin D3.

Arch Biochem Biophys. 381:143-52.

247 Brock K, Wilkinson M, Cook R, Lee S, Bermingham M. (2004) Associations with

Vitamin D deficiency in "at risk" Australians. J Steroid Biochem Mol Biol. 89-90:581-8

Brodie MJ, Boobis AR, Hillyard CJ, Abeyasekera G, Stevenson JC, MacIntyre I, Park

BK. (1982) Effect of and isoniazid on vitamin D metabolism. Clin Pharmacol

Ther. 32:525-30.

Bronner F. (2003) Mechanisms of intestinal calcium absorption. J Cell Biochem. 2003

88:387-93.

Bronner F. (2003) Mechanisms of intestinal calcium absorption. J. Cell Biochem. 88, 2,

387-393.

Buitrago C, Gonzalez Pardo V, de Boland AR. Nongenomic action of 1 alpha,25(OH)(2)- vitamin D3. (2002) Activation of muscle cell PLC gamma through the tyrosine kinase c-

Src and PtdIns 3-kinase. Eur J Biochem. May;269:2506-15.

Buitrago CG, Pardo VG, de Boland AR, Boland R. (2003) Activation of RAF-1 through

Ras and protein kinase Calpha mediates 1alpha,25(OH)2-vitamin D3 regulation of the mitogen-activated protein kinase pathway in muscle cells. J Biol Chem. 278:2199-205.

248 Burgos-Trinidad M., DeLuca H.F. (1991): Kinetic properties of 25-hydroxyvitamin D- and 1,25-dihydroxyvitamin D-24-hydroxylase from chick kidney. Biochim. Biophys. Acta

1078:226–30.

Burne TH, McGrath JJ, Eyles DW, Mackay-Sim A. (2005) Behavioural characterization of vitamin D receptor knockout mice. Behav Brain Res. 157:299-308.

Burns DJ, Bell RM. (1991) Protein kinase C contains two phorbol ester binding domains.

J Biol Chem. 266:18330-8.

Cairo MS, Davenport V, Bessmertny O, Goldman SC, Berg SL, Kreissman SG, Laver J,

Shen V, Secola R, van de Ven C, Reaman GH. (2005) Phase I/II dose escalation study of recombinant human interleukin-11 following ifosfamide, carboplatin and etoposide in children, adolescents and young adults with solid tumours or lymphoma: a clinical, haematological and biological study. Br J Haematol. 128:49-58.

Cantorna MT, Zhu Y, Froicu M, Wittke A. (2004) Vitamin D status, 1,25- dihydroxyvitamin D3, and the immune system. Am J Clin Nutr. 80:1717S-20S

Cantorna,M.T., Hullet, D.A., Redaelli, C., Brandt,C.R., Humpal-Winter,J.,

Sollinger,H.W., and DeLuca, H.F.(1998) 1,25-Dihydroxyvitamin D3 prolongs graft

249 survival without compromising host resistance to infection or bone mineral density.

Transplantation 66: 828-831.

Cantrell DA. (2001) Phosphoinositide 3-kinase signalling pathways. J Cell Sci.

114:1439-45.

Cao LP, Bolt MJ, Wei M, Sitrin MD, Chun Li Y. (2002) Regulation of calbindin-D9k expression by 1,25-dihydroxyvitamin D(3) and parathyroid hormone in mouse primary renal tubular cells. Arch Biochem Biophys. 400:118-24.

Capiati D, Benassati S, Boland RL. (2002) 1,25(OH)2-vitamin D3 induces translocation of the vitamin D receptor (VDR) to the plasma membrane in skeletal muscle cells. J Cell

Biochem. 86(1):128-35.

Capiati DA, Vazquez G, Boland RL. (2001) Protein kinase C alpha modulates the Ca2+ influx phase of the Ca2+ response to 1alpha,25-dihydroxy-vitamin-D3 in skeletal muscle cells. Horm Metab Res. 33(4):201-6.

Carlberg,C and Polly,P (1998) Gene regulation by vitamin D3. Crit.Rev. Eukaryot. Gene

Express. 8, 19-42

250 Carnevale V, Modoni S, Pileri M, Di Giorgio A, Chiodini I, Minisola S, Vieth R,

Scillitani A. (2001) Longitudinal evaluation of vitamin D status in healthy subjects from southern Italy: seasonal and gender differences. Osteoporos Int. 12(12):1026-30.

Cavigelli M, Dolfi F, Claret FX, Karin M. (1995) Induction of c-fos expression through

JNK-mediated TCF/Elk-1 phosphorylation. EMBO J. 14(23):5957-64.

Cenni V, Maraldi NM, Ruggeri A, Secchiero P, Del Coco R, De Pol A, Cocco L,

Marmiroli S. (2004) Sensitization of multidrug resistant human ostesarcoma cells to

Apo2 Ligand/TRAIL-induced apoptosis by inhibition of the Akt/PKB kinase. Int J

Oncol. 25(6):1599-608.

Chae HJ, Jeong BJ, Ha MS, Lee JK, Byun JO, Jung WY, Yun YG, Lee DG, Oh SH,

Chae SW, Kwak YG, Kim HH, Lee ZH, Kim HR. (2002) ERK MAP Kinase is required in 1,25(OH)2D3-induced differentiation in human osteoblasts. Immunopharmacol

Immunotoxicol. 24(1):31-41.

Chan ED, Winston BW, Uh ST, Wynes MW, Rose DM, Riches DW. (1999) Evaluation of the role of mitogen-activated protein kinases in the expression of inducible by IFN-gamma and TNF-alpha in mouse macrophages. J Immunol. 162:415-22.

251 Chang L, Karin M. (2001) Mammalian MAP kinase signalling cascades. Nature.

410(6824):37-40.

Chao TH, Hayashi M, Tapping RI, Kato Y, Lee JD. (1999) MEKK3 directly regulates

MEK5 activity as part of the big mitogen-activated protein kinase 1 (BMK1) signaling pathway. J Biol Chem. 274(51):36035-8.

Chen A, Davis BH, Bissonnette M, Scaglione-Sewell B, Brasitus TA. (1999) 1,25-

Dihydroxyvitamin D(3) stimulates activator protein-1-dependent Caco-2 cell differentiation. J Biol Chem. 274(50):35505-13.

Chen KS and DeLuca HF. (1995) Cloning of the human 1 alpha,25-dihydroxyvitamin D-

3 24-hydroxylase gene promoter and identification of two vitamin D responsive elements.

Biochim Biophys Acta 1263:1 –9,

Chen ML, Boltz MA, Armbrecht HJ. 1993. Effects of 1,25-dihydroxyvitamin D3 and phorbol ester on 25-hydroxyvitamin D3 24-hydroxylase cytochrome P450 messenger ribonucleic acid levels in primary cultures of rat renal cells. Endocrinology 132:1782–88

Chen SF. (1995) 1 alpha, 25-Dihydroxyvitamin D3 decreased ICAM-1 and ELAM-1 expressions on pulmonary microvascular endothelial cells and neutrophil motivation. J

Steroid Biochem Mol Biol. 52(1):67-70.

252 Chen TC, Wang L, Whitlatch LW, Flanagan JN, Holick MF. (2003) Prostatic 25- hydroxyvitamin D-1alpha-hydroxylase and its implication in prostate cancer. J Cell

Biochem. 88(2):315-22.

Chen, J., Thomas, H. F., and Sodek, J. (1996). Regulation of bone sialoprotein and osteopontin mRNA expression by dexamethasone and 1,25-dihydroxyvitamin D3 in rat bone organ cultures. Connect. Tissue Res. 34, 41–51.

Child JA, Russell N, Sonneveld P, Schey S. (2005). Future directions in multiple myeloma treatment. Acta Haematol.114:8-13.

Chou MM, Hou W, Johnson J, Graham LK, Lee MH, Chen CS, Newton AC,

Schaffhausen BS, Toker A. (1998) Regulation of protein kinase C zeta by PI 3-kinase and

PDK-1. Curr Biol. 8(19):1069-77.

Christakos S, Dhawan P, Liu Y, Peng X, Porta A. (2003) New insights into the mechanisms of vitamin D action. J Cell Biochem. 88(4):695-705.

Christakos S., Raval-Pandya M., Wernyj R P., and Yang W. (1996) Genomic mechanisms involved in the pleiotropic actions of 1,25-dihydroxyvitamin D3. Biochem.

J. 316, 361-371.

253 Chung S, Ahn C. (1994) Effects of anti-epileptic drug therapy on bone mineral density in ambulatory epileptic children. Brain Dev. 16(5):382-5.

Clerk A, Aggeli IK, Stathopoulou K, Sugden PH. (2005) Peptide growth factors signal differentially through protein kinase C to extracellular signal-regulated kinases in neonatal cardiomyocytes. Cell Signal. [Epub ahead of print]

Cohen ML, Douvdevani A, Chaimovitz C, Shany S. (2001) Regulation of TNF-alpha by

1alpha,25-dihydroxyvitamin D3 in human macrophages from CAPD patients. Kidney Int.

59(1):69-75.

Cohen MS, Gray TK. (1984) Phagocytic cells metabolize 25-hydroxyvitamin D3 in vitro.

Proc Natl Acad Sci U S A. 81(3):931-4.

Corna G, Santambrogio P, Minotti G, Cairo G. (2004) Doxorubicin paradoxically protects cardiomyocytes against iron-mediated toxicity: role of reactive oxygen species and ferritin. J Biol Chem. 279:13738-45

Costa EM, Hirst MA, Feldman D. (1985) Regulation of 1,25-dihydroxyvitamin D3 receptors by vitamin D analogs in cultured mammalian cells. Endocrinology.

117(5):2203-10.

254 Cuenda A. (2000) Mitogen-activated protein kinase kinase 4 (MKK4). Int J Biochem Cell

Biol. 32(6):581-7.

Dahlback H, Wikvall K. (1987) 25-hydroxylation of vitamin D3 in rat liver: roles of mitochondrial and microsomal cytochrome P-450. Biochem Biophys Res Commun.

142(3):999-1005.

Dahlback H, Wikvall K. (1988) 25-Hydroxylation of vitamin D3 by a cytochrome P-450 from rabbit liver mitochondria. Biochem J. 252(1):207-13.

Daniel C, Schlauch T, Zugel U, Steinmeyer A, Radeke HH, Steinhilber D, Stein J. (2005)

22-ene-25-oxa-vitamin D: a new vitamin D analogue with profound immunosuppressive capacities. Eur J Clin Invest. 35(5):343-9.

Dardenne O, Prud’homme J, Arabian A, Glorieux FH, St-Arnaud R (2001): Targeted inactivation of the 25-hydroxyvitamin D-3-1 alpha-hydroxylase gene (CYP27B1) creates an animal model of pseudovitamin D – deficiency rickets. Endocrinology. 142: 3135-

3141.

Dardenne O, Prud'homme J, Glorieux FH, St-Arnaud R. (2004) Rescue of the phenotype of CYP27B1 (1alpha-hydroxylase)-deficient mice. J Steroid Biochem Mol Biol. 89-90(1-

5):327-30

255 Dardenne O, Prudhomme J, Hacking SA, Glorieux FH, St-Arnaud R. (2003) Rescue of the pseudo-vitamin D deficiency rickets phenotype of CYP27B1-deficient mice by treatment with 1,25-dihydroxyvitamin D3: biochemical, histomorphometric, and biomechanical analyses. J Bone Miner Res. 18(4):637-43.

Dartsch D.C, Schaefer A, Boldt S., Kolch W. and Marquardt H. (2002) Comparison of anthracycline-induced death of human leukemia cells: Programmed cell death versus necrosis. Apoptosis. 7: 537-548.

Davies JH, Evans BA, Jenney ME, Gregory JW. (2003) Effects of chemotherapeutic agents on the function of primary human osteoblast-like cells derived from children. J

Clin Endocrinol Metab. 88(12):6088-97.

De Mas-Mansat V, Hernandez H, Plo I, Bezombes C, Maestre N, Quillet-Mary A,

Filomenko R, Demur C, Jaffrezou JP and Laurent G. (2002) Protein Kinase C ζ mediated

Raf-1/ extracellular- regulated kinase activation by daunorubicin. Blood –2002-05-1585

de Schepper J, Hachimi-Idrissi S, Louis O, Maurus R, Otten J. (1994) Bone metabolism and mineralisation after cytotoxic chemotherapy including ifosfamide. Arch Dis Child.

71: 346-8.

256 Deacon K, Blank JL. (1997) Characterization of the mitogen-activated protein kinase kinase 4 (MKK4)/c-Jun NH2-terminal kinase 1 and MKK3/p38 pathways regulated by

MEK kinases 2 and 3. MEK kinase 3 activates MKK3 but does not cause activation of p38 kinase in vivo. J Biol Chem. 272(22):14489-96.

Deacon K, Blank JL. (1999) MEK kinase 3 directly activates MKK6 and MKK7, specific activators of the p38 and c-Jun NH2-terminal kinases. J Biol Chem. 274(23):16604-10.

Del Prete A, Vermi W, Dander E, Otero K, Barberis L, Luini W, Bernasconi S, Sironi M,

Santoro A, Garlanda C, Facchetti F, Wymann MP, Vecchi A, Hirsch E, Mantovani A,

Sozzani S. (2004) Defective dendritic cell migration and activation of adaptive immunity in PI3Kgamma-deficient mice. EMBO J. 23(17):3505-15.

DeLuca H. and Cantorna MT. (2001) Vitamin D: its role and uses in immunology.

FASEB J. 15, 2579-2585.

DeLuca HF. (2004) Overview of general physiologic features and functions of vitamin D.

Am J Clin Nutr. 80(6 Suppl):1689S-96S

Demay, M. B., Gerardi, J. M., DeLuca, H. F., and Kronenberg,H. M. (1990). DNA sequences in the rat osteocalcin gene that bind the 1,25-dihydroxyvitamin D3 receptor and confer responsive to 1,25-dihydroxyvitamin D3. Proc. Natl. Acad. Sci. USA, 87, 369–

257 373.

Deutsch A, Chaimovitzch C, Nagauker-Shriker O, Zlotnik M, Shany S, Levy R. (1995)

Elevated superoxide generation in mononuclear phagocytes by treatment with 1 alpha hydroxyvitamin D3: changes in kinetics and in oxidase cytosolic factor p47. J Am Soc

Nephrol. 6(1):102-9.

DeVries TA, Neville MC, Reyland ME (2002) Nuclear import of PKCdelta is required for apoptosis: identification of a novel nuclear import sequence. EMBO J. 21(22):6050-

60.

Dine T, Cazin JC, Gressier B, Luyckx M, Brunet C, Cazin M, Goudaliez F, Mallevais

ML, Toraub I. (1992) Stability and compatibility of four anthracyclines: doxorubicin, epirubicin, daunorubicin and pirarubicin with PVC infusion bags. Pharm Weekbl Sci.

14:365-9.

Dong C, Davis RJ, Flavell RA. (2002) MAP kinases in the immune response. Annu Rev

Immunol. 20:55-72.

258 Dreyer Z, Blatt J, Bleyer A. Late effects of childhood cancer and its treatment. In: Pizzo

PA , Poplack DG , editors. Principles and practice of pediatric oncology, 4th ed.

Philadelphia: Lippincott, Williams and Wilkins, 2002: 1431-1461.

Drissi H, Pouliot A, Koolloos C, Stein JL, Lian JB, Stein GS, van Wijnen AJ. (2002)

1,25-(OH)2-vitamin D3 suppresses the bone-related Runx2/Cbfa1 gene promoter. Exp

Cell Res. 274(2):323-33.

Dueland S., Holmberg I., Berg T., and Pedersen J. I. (1981) Uptake and 25-

Hydroxylation of Vitamin D3 by Isolated Rate Liver Cells. The Journal of Biological

Chemistry 256. 20, 10430-10434

Dusso AS, Brown AJ, Slatopolsky E. (2005) Vitamin D. Am J Physiol Renal Physiol.

289(1):F8-28.

Dwivedi PP, Gao XH, Tan J, Ferrante A, Omdahl JL, Morris HA, May BK, Hii CS.

Inverted GC and CCAAT boxes are important for the Regulation of Basal and 1,25- dihydroxyvitamin D3 mediated Expression of Rat CYP24-Promoter in kidney cells: Role of PI3 kinase-PKC zeta signalling module in the functionality of the GC box. (IN

PREPARATION)

259 Dwivedi PP, Hii CS, Ferrante A, Tan J, Der CJ, Omdahl JL, Morris HA, May BK. (2002)

Role of MAP kinases in the 1,25-dihydroxyvitamin D3-induced transactivation of the rat cytochrome P450C24 (CYP24) promoter. Specific functions for ERK1/ERK2 and ERK5.

J Biol Chem. 277(33):29643-53.

Dwivedi PP, Muscat GE, Bailey PJ, Omdahl JL, May BK. (1998). Repression of basal transcription by vitamin D receptor: evidence for interaction of unliganded vitamin D receptor with two receptor interaction domains in RIP13delta1. J. Mol. Endocrinol.

20:327–35

Dwivedi PP, Omdahl JL, Kola I, Hume DA, May BK. (2000) Regulation of rat cytochrome P450C24 (CYP24) gene expression. Evidence for functional cooperation of

Ras-activated Ets transcription factors with the vitamin D receptor in 1,25- dihydroxyvitamin D(3)-mediated induction. J Biol Chem. 275(1): 47-55.

Dziegiel P., Suder E., Surowiak P., Jethon Z., Rabczynski J., Januszewska L., Sopel M., and Zabel M. (2002) Role of exogenous melatonin in reducing the nephrotoxic effect of daunorubicin and doxorubicin in the rat. J. Pineal Res. 33, 2, 95-100.

Earm YE, Ho WK, So I. (1994) Effects of adriamycin on ionic currents in single cardiac myocytes of the rabbit. J Mol Cell Cardiol. 26:163-72.

260 Einhorn LH. (2002) Curing metastatic testicular cancer. Proc Natl Acad Sci U S A. 99(7):

4592-5.

English JM, Pearson G, Hockenberry T, Shivakumar L, White MA, Cobb MH. (1999)

Contribution of the ERK5/MEK5 pathway to Ras/Raf signaling and growth control. J

Biol Chem. 274: 31588-92.

Erben RG, Soegiarto DW, Weber K, Zeitz U, Lieberherr M, Gniadecki R, Moller G,

Adamski J, Balling R. (2002) Deletion of deoxyribonucleic acid binding domain of the vitamin D receptor abrogates genomic and nongenomic functions of vitamin D. Mol

Endocrinol. 16(7):1524-37.

Etzioni A, Hochberg Z, Pollak S, Meshulam T, Zakut V, Tzehoval E, Keisari Y, Aviram

I, Spirer Z, Benderly A. (1989) Defective leukocyte fungicidal activity in end-organ resistance to 1,25-dihydroxyvitamin D. Pediatr Res. 25(3):276-9.

Evans TR, Colston KW, Lofts FJ, Cunningham D, Anthoney DA, Gogas H, de Bono JS,

Hamberg KJ, Skov T, Mansi JL. (2002) A phase II trial of the vitamin D analogue

Seocalcitol (EB1089) in patients with inoperable pancreatic cancer. Br J Cancer. 86:680-

5.

261 Farhan H, Wahala K, Cross HS. (2003) Genistein inhibits vitamin D hydroxylases

CYP24 and CYP27B1 expression in prostate cells. J Steroid Biochem Mol Biol.

84(4):423-9.

Fleet JC. (1999) Vitamin D receptors: not just in the nucleus anymore. Nutr Rev.

57(2):60-2.

Fleet JC. (2004) Rapid, membrane-initiated actions of 1,25 dihydroxyvitamin D: what are they and what do they mean? J Nutr. 134:3215-8.

Flicker L, Mead K, MacInnis RJ, Nowson C, Scherer S, Stein MS, Thomasx J, Hopper

JL, Wark JD. (2003) Serum vitamin D and falls in older women in residential care in

Australia. J Am Geriatr Soc. 2003 Nov;51(11):1533-8.

Frappaz D, Michon J, Hartmann O, Bouffet E, Lejars O, Rubie H, Gentet JC, Chastagner

P, Sariban E, Brugiere L. (1992) Etoposide and carboplatin in neuroblastoma: a French

Society of Pediatric Oncology phase II study. J Clin Oncol. 10(10):1592-601.

Fratzl-Zelman N, Glantschnig H, Rumpler M, Nader A, Ellinger A, Varga F. (2003) The expression of matrix metalloproteinase-13 and osteocalcin in mouse osteoblasts is related to osteoblastic differentiation and is modulated by 1,25-dihydroxyvitamin D3 and thyroid hormones. Cell Biol Int. 27(6):459-68.

262 Fruman DA, Meyers RE, Cantley LC. (1998) Phosphoinositide kinases. Annu Rev

Biochem. 67:481-507.

Fry MJ. (1994) Structure, regulation and function of phosphoinositide 3-kinases. Biochim

Biophys Acta. 1226(3):237-68.

Fukasawa H, Kato A, Fujigaki Y, Yonemura K, Furuya R, Hishida A. (2001)

Hypercalcemia in a patient with B-cell acute lymphoblastic leukemia: a role of proinflammatory cytokine. Am J Med Sci. 322(2):109-12.

Gao XH, Dwivedi PP, Omdahl JL, Morris HA, May BK. (2004) Calcitonin stimulates expression of the rat 25-hydroxyvitamin D3-24-hydroxylase (CYP24) promoter in HEK-

293 cells expressing calcitonin receptor: identification of signaling pathways. J Mol

Endocrinol. 32(1):87-98.

Gao Y, Shimizu M, Yamada S, Ozaki Y, Aso T. (1993) The effect of chemotherapy including cisplatin on vitamin D metabolism. Endocrine Journal. 40: 737-742.

Gascon-Barre M., Demers C., Ghrab O., Theodoropoulos C., Lapointe R., Jones G.,

Valiquette L., Menard D. (2001) Expression of CYP27A, a gene encoding a vitamin D-

25 hydroxylase in human liver and kidney. Clin. Endocrinol. 54, 107-115

263 Gentili C, Morelli S, de Boland AR. (2004) 1alpha,25(OH)2D3 and parathyroid hormone

(PTH) signaling in rat intestinal cells: activation of cytosolic PLA2. J Steroid Biochem

Mol Biol. 89-90(1-5):297-301.

Gervasoni JE Jr, Hindenburg AA, Vezeridis MP, Schulze S, Wanebo HJ, Mehta S.

(2004) An effective in vitro antitumor response against human pancreatic carcinoma with paclitaxel and daunorubicin by induction of both necrosis and apoptosis. Anticancer Res.

24:2617-26.

Ghijsen WE, Van Os CH, Heizmann CW, Murer H. (1986) Regulation of duodenal Ca2+ pump by calmodulin and vitamin D-dependent Ca2+-binding protein. Am J Physiol.

251(2 Pt 1):G223-9.

Gjerset RA, Lebedeva S, Haghighi A, Turla ST, Mercola D. (1999) Inhibition of the Jun kinase pathway blocks DNA repair, enhances p53-mediated apoptosis and promotes gene amplification. Cell Growth Differ. 10(8):545-54.

Glorieux FH. (1990) Calcitriol treatment in vitamin D-dependent and vitamin D-resistant rickets. Metabolism 39:10-2.

264 Gomez-Lechon MJ, O'Connor E, Castell JV, Jover R. (2002) Sensitive markers used to identify compounds that trigger apoptosis in cultured hepatocytes. Toxicol Sci. 65(2):299-

308.

Gommans IM, Vlak MH, de Haan A, van Engelen BG. (2002) Calcium regulation and muscle disease. J Muscle Res Cell Motil. 23(1):59-63.

Gouaze V, Mirault ME, Carpentier S, Salvayre R, Levade T, Andrieu-Abadie N. (2001)

Glutathione peroxidase-1 overexpression prevents ceramide production and partially inhibits apoptosis in doxorubicin-treated human breast carcinoma cells. Mol Pharmacol.

60:488-96.

Grant WB. (2002) An ecologic study of dietary and solar ultraviolet-B links to breast carcinoma mortality rates. Cancer. 94(1):272-81.

Grant WB. (2004) Geographic variation of prostate cancer mortality rates in the United

States: Implications for prostate cancer risk related to vitamin D. Int J Cancer.

111(3):470-1.

Green DM (2002) Paediatric update. Eur J Cancer 38:1251–1253

265 Griffin MD, Lutz WH, Phan VA, Bachman LA, McKean DJ, Kumar R. (2000) Potent inhibition of dendritic cell differentiation and maturation by vitamin D analogs. Biochem

Biophys Res Commun. 270(3):701-8.

Griffin, M. D., Xing, N. and Kumar, R. (2003) Vitamin D and its analogs as regulators of immune activation and antigen presentation. Annu. Rev. Nutr. 23, 117-145

Gschwendt M. (1999) Protein kinase C delta. Eur J Biochem. 259(3):555-64.

Gurlek A, Pittelkow MR, Kumar R. (2002) Modulation of growth factor/cytokine synthesis and signaling by 1alpha,25-dihydroxyvitamin D(3): implications in cell growth and differentiation. Endocr Rev. 23(6):763-86.

Guyton K.Z., Kensler T.W., and Posner G.H. (2003) Vitamin D and Vitamin D Analogs as Cancer Chemopreventive Agents. Nutrition Reviews 61, 7, 227-238.

Gysemans CA, Cardozo AK, Callewaert H, Giulietti A, Hulshagen L, Bouillon R, Eizirik

DL, Mathieu C. (2005) 1,25-Dihydroxyvitamin D3 modulates expression of chemokines and cytokines in pancreatic islets: implications for prevention of diabetes in nonobese diabetic mice. Endocrinology. 146(4):1956-64.

266 Hahn CN, Kerry DM, Omdahl JL, May BK. (1994) Identification of a vitamin D responsive element in the promoter of the rat cytochrome P450(24) gene. Nucleic Acids

Res. 22(12):2410-6.

Halton JM, Atkinson SA, Fraher L, Webber C, Gill GJ, Dawson S, Barr RD. (1996)

Altered mineral metabolism and bone mass in children during treatment for acute lymphoblastic leukemia. J Bone Miner Res. 11(11):1774-83.

Hansen CM, Binderup L, Hamberg KJ, Carlberg C. (2001) Vitamin D and cancer: effects of 1,25(OH)2D3 and its analogs on growth control and tumorigenesis. Front Biosci.

6:D820-48.

Hanukoglu I, Feuchtwanger R, Hanukoglu A. (1990) Mechanism of corticotropin and cAMP induction of mitochondrial cytochrome P450 system enzymes in adrenal cortex cells. J Biol Chem. 265(33):20602-8.

Harras A. Cancer rates and risks, 4th ed. NIH publication no. 96-691. Bethesda: National

Cancer Institute, 1996

Harrell RM, Lyles KW, Harrelson JM, Friedman NE, Drezner MK. (1985) Healing of bone disease in X-linked hypophosphatemic rickets/osteomalacia. Induction and maintenance with phosphorus and calcitriol. J Clin Invest. 75(6):1858-68.

267 Harrington CR. (1990) Lowry protein assay containing sodium dodecyl sulfate in microtiter plates for protein determinations on fractions from brain tissue. Anal Biochem.

186(2):285-287.

Harrison, J. R., Petersen, D. N., Lichtler, A. C., Mador, A. T., Rowe, D. W., and Kream,

B. E. (1989). 1,25-Dihydroxyvitamin D3 inhibits transcription of type I collagen genes in the rat osteosarcoma cell line ROS 17/2.8. Endocrinology 125, 327–333

Haugen JD, Pittelkow MR, Zinsmeister AR, Kumar R. (1996) 1 alpha,25- dihydroxyvitamin D3 inhibits normal human keratinocyte growth by increasing transforming growth factor beta 2 release. Biochem Biophys Res Commun. 229(2):618-

23.

Hayakawa J, Depatie C, Ohmichi M, Mercola D. (2003) The activation of c-Jun NH2- terminal kinase (JNK) by DNA-damaging agents serves to promote drug resistance via activating transcription factor 2 (ATF2)-dependent enhanced DNA repair. J Biol Chem.

278, 23:20582-92.

Hayakawa J, Ohmichi M, Kurachi H, Ikegami H, Kimura A, Matsuoka T, Jikihara H,

Mercola D, Murata Y. (1999) Inhibition of extracellular signal-regulated protein kinase or c-Jun N-terminal protein kinase cascade, differentially activated by cisplatin, sensitizes human ovarian cancer cell line. J Biol Chem. 274(44):31648-54.

268 Hayashi M, Lee JD. (2004) Role of the BMK1/ERK5 signaling pathway: lessons from knockout mice. J Mol Med. 82(12):800-8.

Hayes C.E., Nashold F.E., Spach K.M., and Pedersen L.B. (2003) The immunological functions of the vitamin D endocrine system. Cell. Mol. Biol. 49. 2, 277-300

Healy KD, Frahm MA, DeLuca HF. (2005) 1,25-Dihydroxyvitamin D3 up-regulates the renal vitamin D receptor through indirect gene activation and receptor stabilization. Arch

Biochem Biophys. 433(2):466-73.

Heijn M, Roberge S, Jain RK. (1999) Cellular membrane permeability of anthracyclines does not correlate with their delivery in a tissue-isolated tumor. Cancer Res. 59:4458-63.

Henry H L. (1997) The 25-hydroxyvitamin D 1a-hydroxylase. In: Feldman D., Glorieux

FH., Pike JW. (Eds), Vitamin D. Academic Press, San Diego, pp 57-68

Henry H. L. (2001) The 25(OH)D3/1a,25(OH)2D3-24R-hydroxylase: a catabolic or biosynthetic enzyme? Steroids 66: 391-398.

Henry HL, Norman AW. (1984) Vitamin D: metabolism and biological actions. Annu

Rev Nutr. 4:493-520.

269 Hershberger PA, McGuire TF, Yu WD, Zuhowski EG, Schellens JH, Egorin MJ, Trump

DL, Johnson CS. (2002) Cisplatin potentiates 1,25-dihydroxyvitamin D3-induced apoptosis in association with increased mitogen-activated protein kinase kinase kinase 1

(MEKK-1) expression. Mol Cancer Ther. 1(10):821-9.

Hewison M., Zehnder D., Bland R., and Stewart P M. (2000) 1a-Hydroxylase and the action of vitamin D. Journal of Molecular Endocrinology. 25, 141-148

Higashimoto Y, Ohata M, Iwamoto Y, Fujimoto H, Uetani K, Suruda T, Nakamura Y,

Nakai S, Funasako M. (1995) Stimulatory effect of 1 alpha,25-dihydroxyvitamin D3 on mouse alveolar macrophage tumor necrosis factor-alpha production in vitro: involvement of protein kinase C and Ca2+/calmodulin-dependent kinase. Respiration. 62(2):89-94.

Hii CS, Ferrante A, Edwards YS, Huang ZH, Hartfield PJ, Rathjen DA, Poulos A,

Murray AW. (1995) Activation of mitogen-activated protein kinase by in rat liver epithelial WB cells by a protein kinase C-dependent mechanism. J Biol Chem.

270:4201-4.

Hii CS, Huang ZH, Bilney A, Costabile M, Murray AW, Rathjen DA, Der CJ, Ferrante

A. (1998) Stimulation of p38 phosphorylation and activity by arachidonic acid in HeLa cells, HL60 promyelocytic leukemic cells, and human neutrophils. Evidence for cell type- specific activation of mitogen-activated protein kinases. J Biol Chem. 273:19277-82.

270 Hijiya N, Gajjar A, Zhang Z, Sandlund JT, Ribeiro RC, Rubnitz JE, Jeha S, Liu W,

Cheng C, Raimondi SC, Behm FG, Rivera GK, Relling MV, Pui CH. (2004) Low-dose oral etoposide-based induction regimen for children with acute lymphoblastic leukemia in first bone marrow relapse. Leukemia. 18(10):1581-6.

Hilliard GM 4th, Cook RG, Weigel NL, Pike JW. (1994) 1,25-dihydroxyvitamin D3 modulates phosphorylation of serine 205 in the human vitamin D receptor: site-directed mutagenesis of this residue promotes alternative phosphorylation. Biochemistry.

33(14):4300-11.

Hirai S, Katoh M, Terada M, Kyriakis JM, Zon LI, Rana A, Avruch J, Ohno S. (1997)

MST/MLK2, a member of the mixed lineage kinase family, directly phosphorylates and activates SEK1, an activator of c-Jun N-terminal kinase/stress-activated protein kinase. J

Biol Chem. 272(24):15167-73.

Hmama Z, Nandan D, Sly L, Knutson KL, Herrera-Velit P, Reiner NE. (1999) 1alpha,25- dihydroxyvitamin D(3)-induced myeloid cell differentiation is regulated by a vitamin D receptor-phosphatidylinositol 3-kinase signaling complex. J Exp Med. 190(11):1583-94.

Hmama Z, Sendide K, Talal A, Garcia R, Dobos K, Reiner NE. (2004) Quantitative analysis of phagolysosome fusion in intact cells: inhibition by mycobacterial

271 lipoarabinomannan and rescue by an 1alpha,25-dihydroxyvitamin D3-phosphoinositide

3-kinase pathway. J Cell Sci. 117(Pt 10):2131-40.

Hochberg Z, Tiosano D, Even L. (1992) Calcium therapy for calcitriol-resistant rickets. J

Pediatr. 121:803–808

Hoenderop JG, Dardenne O, Van Abel M, Van Der Kemp AW, Van Os CH, St -Arnaud

R, Bindels RJ. (2002) Modulation of renal Ca2+ transport protein genes by dietary Ca2+ and 1,25-dihydroxyvitamin D3 in 25-hydroxyvitamin D3-1alpha-hydroxylase knockout mice. FASEB J. 16(11):1398-406.

Holick M. F. (1994) The Liver: Biology and Pathobiology (Arias I. M., Boyer J. L.,

Fausto N., Jakoby W. B., Schachter D., and Shafritz D. A., Eds.), pp. 543-562, Raven

Press, New York.

Holick MF. (2004) Sunlight and vitamin D for bone health and prevention of autoimmune diseases, cancers, and cardiovascular disease. Am J Clin Nutr. 80(6 Suppl):

1678S-88S

Holick, M. F. (1997) Vitamin D (Feldman D. Glorieux, F.H., and Pike, J.W. eds) pp 624-

639, Academic Press, San Diego

272 Holick, M. F. (2003) Vitamin D: a millenium perspective. J. Cell. Biochem. 88, 296-307

Holtrop ME, Cox KA, Carnes DL, Holick MF. (1986) Effects of serum calcium and phosphorus on skeletal mineralization in vitamin D-deficient rats. Am J Physiol. 251(2 Pt

1):E234-40.

Hosogane N., Shinki T., Kasuga H., Taketomi S., Toyama Y., and Suda T. (2003)

Mechanisms for the reduction of 24,25-dihydroxyvitamin D3 levels and bone mass in 24- hydroxylase transgenic rats. FASEB J. 17, 737-739.

Hosseinpour F, Wikvall K. (2000) Porcine microsomal vitamin D(3) 25-hydroxylase

(CYP2D25). Catalytic properties, tissue distribution, and comparison with human

CYP2D6. J Biol Chem. 275(44):34650-5.

Hsieh JC, Jurutka PW, Galligan MA, Terpening CM, Haussler CA, Samuels DS, Shimizu

Y, Shimizu N, Haussler MR. (1991) Human vitamin D receptor is selectively phosphorylated by protein kinase C on serine 51, a residue crucial to its trans-activation function. Proc Natl Acad Sci U S A. 88(20):9315-9.

Huang HL, Fang LW, Lu SP, Chou CK, Luh TY, Lai MZ. (2003) DNA-damaging reagents induce apoptosis through reactive oxygen species-dependent Fas aggregation.

Oncogene. 22:8168-77.

273 Huang Y, Ishizuka T, Miura A, Kajita K, Ishizawa M, Kimura M, Yamamoto Y, Kawai

Y, Morita H, Uno Y, Yasuda K. (2002) Effect of 1 alpha,25-dihydroxy vitamin D3 and vitamin E on insulin-induced glucose uptake in rat adipocytes. Diabetes Res Clin Pract.

55(3):175-83.

Hullett, D.A., Cantorna,M., Redaelli,C., Humpal-Winter,J., Hayes, C.E., Sollinger,H.W., and DeLuca, H.F.(1998) Prolongation of allograft survival by 1,25-dihydroxyvitamin D3.

Transplantation 66, 824-828.

Ichijo H. (1999) From receptors to stress-activated MAP kinases. Oncogene.

18(45):6087-93.

Ichikawa F, Sato K, Nanjo M, Nishii Y, Shinki T, Takahashi N, Suda T. (1995) Mouse primary osteoblasts express vitamin D3 25-hydroxylase mRNA and convert 1 alpha- hydroxyvitamin D3 into 1 alpha,25-dihydroxyvitamin D3. Bone. 16(1):129-35.

Iida K, Taniguchi S, Kurokawa K. (1993) Distribution of 1,25-dihydroxyvitamin D3 receptor and 25-hydroxyvitamin D3-24-hydroxylase mRNA expression along rat nephron segments. Biochem Biophys Res Commun. 194(2):659-64.

274 Iioka Y, Mishima K, Azuma N, Tsuchida A, Takagi Y, Aoki T, Saito I. (2005)

Overexpression of protein kinase Cdelta enhances cisplatin-induced cytotoxicity correlated with p53 in gastric cancer cell line. Pathobiology. 72(3):152-9.

Ito Y, Mishra NC, Yoshida K, Kharbanda S, Saxena S, Kufe D. (2001) Mitochondrial targeting of JNK/SAPK in the phorbol ester response of myeloid leukemia cells. Cell

Death Differ. 8(8):794-800.

James SY, Mackay AG, Colston KW. (1996) Effects of 1,25 dihydroxyvitamin D3 and its analogues on induction of apoptosis in breast cancer cells. J Steroid Biochem Mol

Biol. 58(4):395-401.

Johansen C, Kragballe K, Henningsen J, Westergaard M, Kristiansen K, Iversen L.

(2003) 1alpha,25-dihydroxyvitamin D3 stimulates activator protein 1 DNA-binding activity by a phosphatidylinositol 3-kinase/Ras/MEK/extracellular signal regulated kinase

1/2 and c-Jun N-terminal kinase 1-dependent increase in c-Fos, Fra1, and c-Jun expression in human keratinocytes. J Invest Dermatol. 120(4):561-70.

Jones G, Strugnell SA, DeLuca HF. (1998) Current understanding of the molecular actions of vitamin D. Physiol Rev. 78(4):1193-231.

275 Jordan P, Carmo-Fonseca M. (2000) Molecular mechanisms involved in cisplatin cytotoxicity. Cell Mol Life Sci. 57(8-9):1229-35.

Jordan SC, Toyoda M, Prehn J, Lemire JM, Sakai R, Adams JS. (1989) 1,25-

Dihydroxyvitamin-D3 regulation of interleukin-2 and interleukin-2 receptor levels and gene expression in human T cells. Mol Immunol. 26(10):979-84.

Jung M, Sabat R, Kratzschmar J, Seidel H, Wolk K, Schonbein C, Schutt S, Friedrich M,

Docke WD, Asadullah K, Volk HD, Grutz G. (2004) Expression profiling of IL-10- regulated genes in human monocytes and peripheral blood mononuclear cells from psoriatic patients during IL-10 therapy. Eur J Immunol. 34(2):481-93.

Jurutka PW, Hsieh JC, Nakajima S, Haussler CA, Whitfield GK, Haussler MR. (1996)

Human vitamin D receptor phosphorylation by casein kinase II at Ser-208 potentiates transcriptional activation. Proc Natl Acad Sci U S A. 93(8):3519-24.

Jurutka PW, MacDonald PN, Nakajima S, Hsieh JC, Thompson PD, Whitfield GK,

Galligan MA, Haussler CA, Haussler MR. (2002) Isolation of baculovirus-expressed human vitamin D receptor: DNA responsive element interactions and phosphorylation of the purified receptor. J Cell Biochem. 85(2):435-57.

276 Jurutka PW, Whitfield GK, Hsieh JC, Thompson PD, Haussler CA, Haussler MR. (2001)

Molecular nature of the vitamin D receptor and its role in regulation of gene expression.

Rev Endocr Metab Disord. 2(2):203-16.

Kallay E, Adlercreutz H, Farhan H, Lechner D, Bajna E, Gerdenitsch W, Campbell M,

Cross HS. (2002) Phytoestrogens regulate vitamin D metabolism in the mouse colon: relevance for colon tumor prevention and therapy. J Nutr. 132(11 Suppl): 3490S-3493S.

Kamakura S, Moriguchi T, Nishida E. (1999) Activation of the protein kinase

ERK5/BMK1 by receptor tyrosine kinases. Identification and characterization of a signaling pathway to the nucleus. J Biol Chem. 274(37):26563-71.

Kaste SC. (2004) Bone-mineral density deficits from childhood cancer and its therapy. A review of at-risk patient cohorts and available imaging methods. Pediatr Radiol.

34(5):373-8

Kasuga H., Hosogane N., Matsuoka K., Mori I., Sakura Y., Shimakawa K., Shinki T.,

Suda T., and Taketomi S. (2002) Characterization of transgenic rats canstitutively expressing vitamin D-24-hydroxylase gene. Biochem. Biophys. Res. Comm. 297, 1332-

1338.

277 Kato Y, Nishimura SI, sakura N, and Ueda K. (2003) Pharmacokinetics of etoposide with intravenous drug administration in children and adolescents. Pediatrics International, 45:

74-79

Kerner, S. A., Scott, R. A., and Pike, J. W. (1989). Sequence elements in the human osteocalcin gene confer basal activation and inducible response to hormonal vitamin D3.

Proc. Natl. Acad. Sci. USA 86, 4455–4459.

Kerry DM, Dwivedi PP, Hahn CN, Morris HA, Omdahl JL, May BK. (1996)

Transcriptional synergism between vitamin D-responsive elements in the rat 25- hydroxyvitamin D3 24-hydroxylase (CYP24) promoter. J Biol Chem. 271(47): 29715-21.

Kim HJ, Abdelkader N, Katz M, McLane JA. (1992) 1,25-Dihydroxy-vitamin-D3 enhances antiproliferative effect and transcription of TGF-beta1 on human keratinocytes in culture. J Cell Physiol. 151(3):579-87.

Kim J, Freeman MR (2003) JNK/SAPK mediates doxorubicin-induced differentiation and apoptosis in MCF-7 breast cancer cells. Breast Cancer Res Treat. 79, 3:321-8.

Kitazawa S, Kajimoto K, Kondo T, Kitazawa R. (2003) Vitamin D3 supports osteoclastogenesis via functional vitamin D response element of human RANKL gene promoter. J Cell Biochem. 89(4):771-7.

278 Kizaki M, Norman AW, Bishop JE, Lin CW, Karmakar A, Koeffler HP. (1991) 1,25-

Dihydroxyvitamin D3 receptor RNA: expression in hematopoietic cells. Blood.

77(6):1238-47.

Kobayashi T, Hashimoto K, Yoshikawa K. (1993) Growth inhibition of human keratinocytes by 1,25-dihydroxyvitamin D3 is linked to dephosphorylation of retinoblastoma gene product. Biochem Biophys Res Commun. 196(1):487-93.

Kohno M, Pouyssegur J. (2003) Pharmacological inhibitors of the ERK signaling pathway: application as anticancer drugs. Prog Cell Cycle Res. 5:219-24.

Kondo T, Kitazawa R, Maeda S, Kitazawa S. (2004) 1 alpha,25 dihydroxyvitamin D3 rapidly regulates the mouse osteoprotegerin gene through dual pathways. J Bone Miner

Res. 19(9):1411-9.

Kong XF, Zhu XH, Pei YL, Jackson DM, Holick MF. (1999) Molecular cloning, characterization, and promoter analysis of the human 25-hydroxyvitamin D3-1alpha- hydroxylase gene. Proc Natl Acad Sci U S A. 96(12):6988-93.

Kong YY, Yoshida H, Sarosi I, Tan HL, Timms E, Capparelli C, Morony S, Oliveira- dos-Santos AJ, Van G, Itie A, Khoo W, Wakeham A, Dunstan CR, Lacey DL, Mak TW,

279 Boyle WJ, Penninger JM. (1999) OPGL is a key regulator of osteoclastogenesis, lymphocyte development and lymph-node organogenesis. Nature. 397(6717):315-23.

Koren R, Rocker D, Kotestiano O, Liberman UA, Ravid A. (2000) Synergistic anticancer activity of 1,25-dihydroxyvitamin D(3) and immune cytokines: the involvement of reactive oxygen species. J Steroid Biochem Mol Biol. 73(3-4):105-12.

Koyama H, Inaba M, Nishizawa Y, Ohno S, Morii H. (1994) Protein kinase C is involved in 24-hydroxylase gene expression induced by 1,25(OH)2D3 in rat intestinal epithelial cells. J Cell Biochem. 55(2):230-40.

Krishnan A V., Peehl D M., and Feldman D (2003) Inhibition of prostate cancer growth by vitamin D: Regulation of target gene expression. J. Cell Biochem. 88 (2), 363-371

Kurahashi I, Matsunuma A, Kawane T, Abe M, Horiuchi N. (2002) Dexamethasone enhances vitamin D-24-hydroxylase expression in osteoblastic (UMR-106) and renal

(LLC-PK1) cells treated with 1alpha,25-dihydroxyvitamin D3. Endocrine. 17:109-18.

Kurz EU, Douglas P, Lees-Miller SP. (2004) Doxorubicin activates ATM-dependent phosphorylation of multiple downstream targets in part through the generation of reactive oxygen species. J Biol Chem. 279:53272-81.

280 Kyriakis JM, Avruch J. (2001) Mammalian mitogen-activated protein kinase signal transduction pathways activated by stress and inflammation. Physiol Rev. 81(2):807-69.

Lacey DL, Timms E, Tan HL, Kelley MJ, Dunstan CR, Burgess T, Elliott R, Colombero

A, Elliott G, Scully S, Hsu H, Sullivan J, Hawkins N, Davy E, Capparelli C, Eli A, Qian

YX, Kaufman S, Sarosi I, Shalhoub V, Senaldi G, Guo J, Delaney J, Boyle WJ. (1998)

Osteoprotegerin ligand is a cytokine that regulates osteoclast differentiation and activation. Cell. 93(2):165-76.

Lamberg-Allardt CJ, Outila TA, Karkkainen MU, Rita HJ, Valsta LM. (2001) Vitamin D deficiency and bone health in healthy adults in Finland: could this be a concern in other parts of Europe? J Bone Miner Res. 16(11):2066-73.

Lau K, Langman CB, Gafter U, Dudeja PK, Brasitus TA. (1986) Increased calcium absorption in prehypertensive spontaneously hypertensive rat. Role of serum 1,25- dihydroxyvitamin D3 levels and intestinal brush border membrane fluidity. J Clin Invest.

78(4):1083-90.

Lawler S, Fleming Y, Goedert M, Cohen P. (1998) Synergistic activation of

SAPK1/JNK1 by two MAP kinase kinases in vitro. Curr Biol. 8(25):1387-90.

281 Lechner D, Cross HS. (2003) Phytoestrogens and 17beta-estradiol influence vitamin D metabolism and receptor expression-relevance for colon cancer prevention. Recent

Results Cancer Res.164:379-91.

Levy R, Klein J, Rubinek T, Alkan M, Shany S, Chaimovitz C. (1990) Diversity in peritoneal macrophage response of CAPD patients to 1,25-dihydroxyvitamin D3. Kidney

Int. 37(5):1310-5.

Li YC, Amling M, Pirro AE, Priemel M, Meuse J, Baron R, Delling G, Demay MB.

1998. Normalization of mineral ion homeostasis by dietary means prevents hyperparathyroidism, rickets, and osteomalacia, but not alopecia in vitamin D receptor- ablated mice. Endocrinology 139: 4391-4396.

Lissoos TW, Beno DW, Davis BH. (1993) 1,25-Dihydroxyvitamin D3 activates Raf kinase and Raf perinuclear translocation via a protein kinase C-dependent pathway. J

Biol Chem. 268(33):25132-8.

List AF, Kopecky KJ, Willman CL, Head DR, Persons DL, Slovak ML, Dorr R, Karanes

C, Hynes HE, Doroshow JH, Shurafa M, Appelbaum FR. (2001) Benefit of cyclosporine modulation of drug resistance in patients with poor-risk acute myeloid leukemia: a

Southwest Oncology Group study. Blood. 98(12):3212-20.

282 Liu M, Lee MH, Cohen M, Bommakanti M, Freedman LP. (1996) Transcriptional activation of the Cdk inhibitor p21 by vitamin D3 leads to the induced differentiation of the myelomonocytic cell line U937. Genes Dev. 10(2):142-53.

Liu Q, Ning W, Dantzer R, Freund GG, Kelley KW. (1998) Activation of protein kinase

C-zeta and phosphatidylinositol 3'-kinase and promotion of macrophage differentiation by insulin-like growth factor-I. J Immunol. 160(3):1393-401.

Liu WS, Heckman CA. (1998) The sevenfold way of PKC regulation. Cell Signal.

10(8):529-42.

Livak KJ, Schmittgen TD. (2001) Analysis of relative gene expression data using real- time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods. 25:402-8.

Lokeshwar BL, Schwartz GG, Selzer MG, Burnstein KL, Zhuang SH, Block NL,

Binderup L. (1999) Inhibition of prostate cancer metastasis in vivo: a comparison of 1,23- dihydroxyvitamin D (calcitriol) and EB1089. Cancer Epidemiol Biomarkers Prev.

8(3):241-8.

Lorente F, Fontan G, Jara P, Casas C, Garcia-Rodriguez MC, Ojeda JA. (1976) Defective neutrophil motility in hypovitaminosis D rickets. Acta Paediatr Scand. 65(6):695-9.

283 Lowe L, Hansen CM, Senaratne S, Colston KW. (2003) Mechanisms implicated in the growth regulatory effects of vitamin D compounds in breast cancer cells. Recent Results

Cancer Res.164:99-110.

Lutz W, Burritt MF, Nixon DE, Kao PC, Kumar R. (2000) Zinc increases the activity of vitamin D-dependent promoters in osteoblasts. Biochem Biophys Res Commun. 271: 1-7.

Ly LH, Zhao XY, Holloway L, Feldman D. (1999) Liarozole acts synergistically with

1alpha,25-dihydroxyvitamin D3 to inhibit growth of DU 145 human prostate cancer cells by blocking 24-hydroxylase activity. Endocrinology. 140(5):2071-6.

Lyman D. (2005) Undiagnosed vitamin D deficiency in the hospitalized patient. Am Fam

Physician. 71(2):299-304.

Ma JF, Nonn L, Campbell MJ, Hewison M, Feldman D, Peehl DM. (2004) Mechanisms of decreased Vitamin D 1alpha-hydroxylase activity in prostate cancer cells. Mol Cell

Endocrinol. 221(1-2):67-74.

Mahonen A, Pirskanen A, Keinanen R, Maenpaa PH. (1990) Effect of 1,25(OH)2D3 on its receptor mRNA levels and osteocalcin synthesis in human osteosarcoma cells.

Biochim Biophys Acta. 1048(1):30-7.

284 Majewski S, Szmurlo A, Marczak M, Jablonska S, Bollag W. (1993) Inhibition of tumor cell-induced angiogenesis by retinoids, 1,25-dihydroxyvitamin D3 and their combination.

Cancer Lett. 75(1):35-9.

Malloy PJ, Xu R, Peng L, Peleg S, Al-Ashwal A, Feldman D. (2004) Hereditary 1,25- dihydroxyvitamin D resistant rickets due to a mutation causing multiple defects in vitamin D receptor function. Endocrinology. 145(11):5106-14.

Mandla S, Boneh A, Tenenhouse HS. (1990) Evidence for protein kinase C involvement in the regulation of renal 25-hydroxyvitamin D3-24-hydroxylase. Endocrinology.

127(6):2639-47.

Mandla S, Tenenhouse HS. (1992) Inhibition of 25-hydroxyvitamin D3-24-hydroxylase by forskolin: evidence for a 3',5'-cyclic adenosine monophosphate-independent mechanism. Endocrinology. 130(4):2145-51.

Maniez-Devos DM, Baurain R, Lesne M, Trouet A. (1986) Doxorubicin and daunorubicin plasmatic, hepatic and renal disposition in the rabbit with or without enterohepatic circulation. J Pharmacol. 17:1-13.

Manning G, Whyte DB, Martinez R, Hunter T, Sudarsanam S. (2002) The protein kinase complement of the . Science. 298(5600):1912-34.

285 Mansat-de Mas V, Bezombes C, Quillet-Mary A, Bettaieb A, D'orgeix AD, Laurent G,

Jaffrezou JP. (1999) Implication of radical oxygen species in ceramide generation, c-Jun

N-terminal kinase activation and apoptosis induced by daunorubicin. Mol Pharmacol.

56(5):867-74.

Mao Y, Desai SD, Ting CY, Hwang J, Liu LF. (2001) 26 S proteasome-mediated degradation of topoisomerase II cleavable complexes. J Biol Chem. 276(44):40652-8.

Marcinkowska E, Wiedlocha A, Radzikowski C. (1997) 1,25-Dihydroxyvitamin D3 induced activation and subsequent nuclear translocation of MAPK is upstream regulated by PKC in HL-60 cells. Biochem Biophys Res Commun. 241(2):419-26.

Marcinkowska E, Wiedlocha A, Radzikowski C. (1998) Evidence that phosphatidylinositol 3-kinase and p70S6K protein are involved in differentiation of HL-

60 cells induced by calcitriol. Anticancer Res. 18(5A):3507-14.

Marcinkowska E. (2001) Evidence that activation of MEK1,2/erk1,2 signal transduction pathway is necessary for calcitriol-induced differentiation of HL-60 cells. Anticancer

Res. 21(1A):499-504.

286 Marie PJ, Travers R, Glorieux FH. (1982) Healing of bone lesions with 1,25- dihydroxyvitamin D3 in the young X-linked hypophosphatemic male mouse.

Endocrinology. 111(3):904-11.

Marina NM, Cochrane D, Harney E, Zomorodi K, Blaney S, Winick N, Bernstein M,

Link MP. (2002) Dose escalation and pharmacokinetics of pegylated liposomal doxorubicin (Doxil) in children with solid tumors: a pediatric oncology group study. Clin

Cancer Res. 8(2):413-8.

Markose, E. R., Stein, J. L., Stein, G. S., and Lian, J. B. (1990). Vitamin D-mediated modifications in protein-DNA interactions at two promoter elements of the osteocalcin gene. Proc. Natl. Acad. Sci. USA. 87, 1701–1705.

Martinez, M. E, Willett, W. C. (1998) Calcium, vitamin D, and colorectal cancer: a review of the epidemiologic evidence. Cancer Epidemiol. Biomarkers Prev. 7:163-168.

Mas VM, Hernandez H, Plo I, Bezombes C, Maestre N, Quillet-Mary A, Filomenko R,

Demur C, Jaffrezou JP, Laurent G. (2003) Protein kinase Czeta mediated Raf-

1/extracellular-regulated kinase activation by daunorubicin. Blood. 101, 4:1543-50.

Masquelier M, Vitols S. (2004) Drastic effect of cell density on the cytotoxicity of daunorubicin and cytosine arabinoside. Biochem Pharmacol. 67(9):1639-46.

287 Masuda S, Byford V, Arabian A, Sakai Y, Demay MB, St-Arnaud R, Jones G. (2005)

Altered pharmacokinetics of 1alpha,25-dihydroxyvitamin D3 and 25-hydroxyvitamin D3 in the blood and tissues of the 25-hydroxyvitamin D-24-hydroxylase (Cyp24a1) null mouse. Endocrinology. 146(2):825-34.

Masuda S, Kaufmann M, Byford V, Gao M, St-Arnaud R, Arabian A, Makin HL,

Knutson JC, Strugnell S, Jones G. (2004) Insights into Vitamin D metabolism using cyp24 over-expression and knockout systems in conjunction with liquid chromatography/mass spectrometry (LC/MS). J Steroid Biochem Mol Biol. 89-90(1-

5):149-53.

Matsukawa J, Matsuzawa A, Takeda K, Ichijo H. (2004) The ASK1-MAP kinase cascades in mammalian stress response. J Biochem 136(3):261-5.

Mazue G., Iatropoulos M., Imondi A., et al (1995) Antracyclines: a review of general and specific toxicity studies. Int.J.Oncol. 7, 713-726.

McGuire TF, Trump DL, Johnson CS. (2001) Vitamin D(3)-induced apoptosis of murine squamous cell carcinoma cells. Selective induction of caspase-dependent MEK cleavage and up-regulation of MEKK-1. J Biol Chem. 276(28):26365-73.

288 Meadows AT, Hobbie WL (1986) The medical consequences of cure. Cancer 58:524–

528

Mee AP; Hoyland JA; Braidman IP; Freemont AJ; Davies M; Mawer EB. (1996)

Demonstration of vitamin D receptor transcripts in actively resorbing osteoclasts in bone sections. Bone. 18(4):295-9.

Mehta PB, Jenkins BL, McCarthy L, Thilak L, Robson CN, Neal DE, Leung HY. (2003)

MEK5 overexpression is associated with metastatic prostate cancer, and stimulates proliferation, MMP-9 expression and invasion. Oncogene. 22(9):1381-9.

Menaa C, Barsony J, Reddy SV, Cornish J, Cundy T, Roodman GD. (2000) 1,25-

Dihydroxyvitamin D3 hypersensitivity of osteoclast precursors from patients with Paget's disease. J Bone Miner Res. 15(2):228-36.

Miao L, Yi P, Wang Y, Wu M. (2003) Etoposide upregulates Bax-enhancing tumour necrosis factor-related apoptosis inducing ligand-mediated apoptosis in the human hepatocellular carcinoma cell line QGY-7703. Eur J Biochem. 270(13):2721-31.

Miettinen S, Ahonen MH, Lou YR, Manninen T, Tuohimaa P, Syvala H, Ylikomi T.

(2004) Role of 24-hydroxylase in vitamin D3 growth response of OVCAR-3 ovarian cancer cells. Int J Cancer. 108(3):367-73.

289 Miller GJ, Stapleton GE, Hedlund TE, Moffat KA. (1995) Vitamin D receptor expression, 24-hydroxylase activity, and inhibition of growth by 1alpha,25- dihydroxyvitamin D3 in seven human prostatic carcinoma cell lines. Clin Cancer Res.

1(9):997-1003.

Minderman H, Linssen P, van der Lely N, Wessels J, Boezeman J, de Witte T, Haanen C.

(1994) Toxicity of idarubicin and doxorubicin towards normal and leukemic human bone marrow progenitors in relation to their proliferative state. Leukemia. 8(3):382-7.

Minotti G, Menna P, Salvatorelli E, Cairo G, Gianni L. (2004) Anthracyclines: molecular advances and pharmacologic developments in antitumor activity and cardiotoxicity.

Pharmacol Rev. 56(2):185-229.

Mizwicki MT, Bishop JE, Olivera CJ, Huhtakangas J, Norman AW. (2004) Evidence that annexin II is not a putative membrane receptor for 1alpha,25(OH)2-vitamin D3. J Cell

Biochem. 91(4):852-63.

Moffatt KA, Johannes WU, Miller GJ. (1999) 1Alpha,25dihydroxyvitamin D3 and platinum drugs act synergistically to inhibit the growth of prostate cancer cell lines. Clin

Cancer Res. 5:695-703.

290 Morales O, Samuelsson MK, Lindgren U, Haldosen LA. (2004) Effects of 1alpha,25- dihydroxyvitamin D3 and growth hormone on apoptosis and proliferation in UMR 106 osteoblast-like cells. Endocrinology. 145(1):87-94.

Morelli S, Buitrago C, Boland R, de Boland AR. (2001) The stimulation of MAP kinase by 1,25(OH)(2)-vitamin D(3) in skeletal muscle cells is mediated by protein kinase C and calcium. Mol Cell Endocrinol. 173(1-2):41-52.

Mori M, Kitamura K, Masuda M, Hotta T, Miyazaki T, Miura AB, Mizoguchi H, Shibata

A, Saito H, Matsuda T, Masaoka T, Harada M, Niho Y, Takaku F. (2005) Long-term results of a multicenter randomized, comparative trial of modified CHOP versus THP-

COP versus THP-COPE regimens in elderly patients with non-Hodgkin's lymphoma. Int

J Hematol. 81(3):246-54.

Muller K, Haahr PM, Diamant M, Rieneck K, Kharazmi A, Bendtzen K. (1992) 1,25-

Dihydroxyvitamin D3 inhibits cytokine production by human blood monocytes at the post-transcriptional level. Cytokine. 4(6):506-12.

Muller K, Odum N, Bendtzen K. (1993) 1,25-dihydroxyvitamin D3 selectively reduces interleukin-2 levels and proliferation of human T cell lines in vitro. Immunol Lett.

35(2):177-82.

291 Murayama A, Takeyama K, Kitanaka S, Kodera Y, Kawaguchi Y, Hosoya T, Kato S.

(1999) Positive and negative regulations of the renal 25-hydroxyvitamin D3 1alpha- hydroxylase gene by parathyroid hormone, calcitonin, and 1alpha,25(OH)2D3 in intact animals. Endocrinology. 140(5):2224-31.

Narayanan R, Sepulveda VA, Falzon M, Weigel NL. (2004) The functional consequences of cross-talk between the vitamin D receptor and ERK signaling pathways are cell- specific. J Biol Chem. 279(45):47298-310.

Nemere I, Dormanen MC, Hammond MW, Okamura WH, Norman AW. (1994)

Identification of a specific binding protein for 1 alpha,25-dihydroxyvitamin D3 in basal- lateral membranes of chick intestinal epithelium and relationship to transcaltachia. J Biol

Chem. 269(38):23750-6.

Nishina H, Wada T, Katada T. (2004) Physiological roles of SAPK/JNK signaling pathway. J Biochem (Tokyo). 136(2):123-6.

Nishizuka Y. (1995) Protein kinase C and lipid signaling for sustained cellular responses.

FASEB J. 9(7):484-96.

Noda, M., Vogel, R. L., Craig, A. M., Prahl, J., DeLuca, H. F., and Denhardt, D. (1990).

Identification of a DNA sequence responsible for binding of the 1,25-dihydroxyvitamin

292 D3 receptor and 1,25-dihydroxyvitamin D3 enhancement of mouse secreted phosphoprotein 1 (Spp-1 or osteopontin) gene expression. Proc. Natl. Acad. Sci. USA. 87,

9995–9999

Norman AW, Henry HL, Bishop JE, Song XD, Bula C, Okamura WH. (2001) Different shapes of the steroid hormone 1alpha,25(OH)(2)-vitamin D(3) act as agonists for two different receptors in the vitamin D endocrine system to mediate genomic and rapid responses. Steroids. 66(3-5):147-58.

Nowak G. (2002) Protein kinase C-alpha and ERK1/2 mediate mitochondrial dysfunction, decreases in active Na+ transport, and cisplatin-induced apoptosis in renal cells. J Biol Chem. 277, 45:43377-88.

Nowosielska A, Marinus MG. (2005) Cisplatin induces DNA double-strand break formation in Escherichia coli dam mutants. DNA Repair (Amst). 4(7):773-81.

Nowson CA, Diamond TH, Pasco JA, Mason RS, Sambrook PN, Eisman JA. (2004)

Vitamin D in Australia. Issues and recommendations. Aust Fam Physician. 33(3):133-8.

Nutchey BK, Kaplan JS, Dwivedi PP, Omdahl JL, Ferrante A, May BK, Hii CS. (2005)

Molecular action of 1,25-dihydroxyvitamin D 3 and phorbol ester on the activation of the rat cytochrome P450C24 (CYP24)promoter: role of MAP kinase activities and

293 identification of an important transcription factor binding site. Biochem J. [Epub ahead of print]

Ohyama Y, Ozono K, Uchida M, Yoshimura M, Shinki T, Suda T, Yamamoto O. (1996)

Functional assessment of two vitamin D-responsive elements in the rat 25- hydroxyvitamin D3 24-hydroxylase gene. J Biol Chem. 271(48):30381-5.

Okano T, Tsugawa N, Morishita A, Kato S. (2004) Regulation of gene expression of epithelial calcium channels in intestine and kidney of mice by 1alpha,25- dihydroxyvitamin D3. J Steroid Biochem Mol Biol. 89-90(1-5):335-8.

Olson JA. (1985) Calcium requirements and recommended dietary allowances. J Nutr Sci

Vitaminol. 31 Suppl:S1-6.

Omdahl JL, Morris HA, May BK. (2002) Hydroxylase enzymes of the vitamin D pathway: expression, function, and regulation. Annu Rev Nutr. 22:139-66.

Ono Y, Fujii T, Igarashi K, Kuno T, Tanaka C, Kikkawa U, Nishizuka Y. (1989) Phorbol ester binding to protein kinase C requires a cysteine-rich zinc-finger-like sequence. Proc

Natl Acad Sci U S A. 86(13):4868-71.

294 Ordonez-Moran P, Larriba MJ, Pendas-Franco N, Aguilera O, Gonzalez-Sancho JM,

Munoz A. (2005) Vitamin D and cancer: an update of in vitro and in vivo data. Front

Biosci. 10:2723-49.

Ory S, Morrison DK. (2004) Signal transduction: implications for Ras-dependent ERK signaling. Curr Biol. 14(7):R277-8.

Overbergh L, Decallonne B, Waer M, Rutgeerts O, Valckx D, Casteels KM, Laureys J,

Bouillon R, Mathieu C. (2000) 1alpha,25-dihydroxyvitamin D3 induces an autoantigen- specific T-helper 1/T-helper 2 immune shift in NOD mice immunized with GAD65

(p524-543). Diabetes. 49(8):1301-7.

Owen TA, Aronow MS, Barone LM, Bettencourt B, Stein GS, Lian JB. (1991)

Pleiotropic effects of vitamin D on osteoblast gene expression are related to the proliferative and differentiated state of the bone cell phenotype: dependency upon basal levels of gene expression, duration of exposure, and bone matrix competency in normal rat osteoblast cultures. Endocrinology. 128(3):1496-504.

Owen TA, Bortell R, Yocum SA, Smock SL, Zhang M, Abate C, Shalhoub V, Aronin N,

Wright KL, van Wijnen AJ. (1990) Coordinate occupancy of AP-1 sites in the vitamin D- responsive and CCAAT box elements by Fos-Jun in the osteocalcin gene: model for phenotype suppression of transcription. Proc Natl Acad Sci U S A. 87(24):9990-4.

295 Paine-Murrieta GD, Taylor CW, Curtis RA, Lopez MH, Dorr RT, Johnson CS, Funk CY,

Thompson F, Hersh EM. (1997) Human tumor models in the severe combined immune deficient (scid) mouse. Cancer Chemother Pharmacol. 40(3):209-14.

Panda DK, Al Kawas S, Seldin MF, Hendy GN, Goltzman D. (2001) 25-hydroxyvitamin

D 1alpha-hydroxylase: structure of the mouse gene, chromosomal assignment, and developmental expression. J Bone Miner Res. 16(1):46-56.

Panda DK, Miao D, Tremblay ML, Sirois J, Farookhi R, Hendy GN, Goltzman D. (2001)

Targeted ablation of the 25-hydroxyvitamin D 1alpha -hydroxylase enzyme: evidence for skeletal, reproductive, and immune dysfunction. Proc Natl Acad Sci U S A. 98(13):7498-

503.

Pandey SK, Theberge JF, Bernier M, Srivastava AK. (1999) Phosphatidylinositol 3- kinase requirement in activation of the ras/C-raf-1/MEK/ERK and p70(s6k) signaling cascade by the insulinomimetic agent vanadyl sulfate. Biochemistry. 38(44):14667-75.

Parisi MT, Fahmy JL, Kaminsky CK, et al (1999) Complications of cancer therapy in children: a radiologist's guide. Radiographics 19: 283–297

Parker PJ, Murray-Rust J. (2004) PKC at a glance. J Cell Sci. 117:131-2.

296 Parker SL, Tong T, Bolden S, Wingo PA (1997). Cancer statistics 1997. CA Cancer J

Clin.; 47: 5-27

Pascussi JM, Robert A, Nguyen M, Walrant-Debray O, Garabedian M, Martin P, Pineau

T, Saric J, Navarro F, Maurel P, Vilarem MJ. (2005) Possible involvement of pregnane X receptor-enhanced CYP24 expression in drug-induced osteomalacia. J Clin Invest.

115(1):177-86.

Pearson G, Robinson F, Beers Gibson T, Xu BE, Karandikar M, Berman K, Cobb MH.

(2001) Mitogen-activated protein (MAP) kinase pathways: regulation and physiological functions. Endocr Rev. 22(2):153-83.

Peng B, English MW, Boddy AV, Price L, Wyllie R, Pearson AD, Tilby MJ, Newell DR.

(1997) Cisplatin pharmacokinetics in children with cancer. Eur J Cancer. 33(11):1823-8.

Pettifor JM. (2004) Nutritional rickets: deficiency of vitamin D, calcium, or both? Am J

Clin Nutr. 80(6 Suppl):1725S-9S.

Phadnis R, Nemere I. (2003) Direct, rapid effects of 25-hydroxyvitamin D3 on isolated intestinal cells. J Cell Biochem. 90(2):287-93.

297 Pinto AL, Lippard SJ. (1985) Binding of the antitumor drug cis- diamminedichloroplatinum(II) (cisplatin) to DNA. Biochim Biophys Acta. 780(3):167-80.

Polly P, Herdick M, Moehren U, Baniahmad A, Heinzel T, Carlberg C. (2000). VDR- alien: a novel, DNA-selective vitamin D3 receptor-corepressor partnership. FASEB J.

14:1455–63

Portale AA, Miller WL. (2000) Human 25-hydroxyvitamin D-1alpha-hydroxylase: cloning, mutations, and gene expression. Pediatr Nephrol. 14(7):620-5.

Potapova O, Basu S, Mercola D, Holbrook NJ. (2001) Protective role for c-Jun in the cellular response to DNA damage. J Biol Chem. 276(30):28546-53.

Prince RL, Glendenning P. (2004) Disorders of bone and mineral other than osteoporosis.

Med J Aust. 180(7):354-9.

Prosser DE, Jones G. (2004) Enzymes involved in the activation and inactivation of vitamin D. Trends Biochem Sci. 29(12):664-73.

Provot S, Schipani E. (2005) Molecular mechanisms of endochondral bone development.

Biochem Biophys Res Commun. 328(3):658-65.

298 Pryke AM, Duggan C, White CP, Posen S, Mason RS. (1990) Tumor necrosis factor- alpha induces vitamin D-1-hydroxylase activity in normal human alveolar macrophages.

J Cell Physiol. 142(3):652-6.

Puri KD, Doggett TA, Huang CY, Douangpanya J, Hayflick JS, Turner M, Penninger J,

Diacovo TG. (2005) The role of endothelial PI3Kgamma activity in neutrophil trafficking. Blood. 106(1):150-7.

Qi X, Pramanik R, Wang J, Schultz RM, Maitra RK, Han J, DeLuca HF and Chen G

(2002). The p38 and JNK pathways cooperate to trans-activate vitamin D receptor via c- jun/AP1 and sensitize human breast cancer cells to vitamin D3 growth inhibition. J Biol

Chem. 277: 25884 – 25892.

Quilliam LA, Kato K, Rabun KM, Hisaka MM, Huff SY, Campbell-Burk S, Der CJ.

(1994) Identification of residues critical for Ras(17N) growth-inhibitory phenotype and for Ras interaction with guanine nucleotide exchange factors. Mol Cell Biol. 14:1113-21.

Rameh LE, Cantley LC. (1999) The role of phosphoinositide 3-kinase lipid products in cell function. J Biol Chem. 274(13):8347-50.

299 Raval-Pandya M, Dhawan P, Barletta F, Christakos S. (2001) YY1 represses vitamin D receptor-mediated 25-hydroxyvitamin D(3)24-hydroxylase transcription: relief of repression by CREB-binding protein. Mol Endocrinol. 15(6): 1035-46.

Ravid A, Koren R. (2003); The role of reactive oxygen species in the anticancer activity of vitamin D. Recent Results Cancer Res. 164:357-67.

Ravid A, Rocker D, Machlenkin A, Rotem C, Hochman A, Kessler-Icekson G, Liberman

UA, Koren R. (1999) 1,25-dihydroxy- vitamin D3 enhances the susceptibility of breast cancer cells to doxorubicin-induced oxidative damage, Cancer Res. 59 862-867.

Rebsamen MC, Sun J, Norman AW, Liao JK. (2002) 1alpha,25-dihydroxyvitamin D3 induces vascular smooth muscle cell migration via activation of phosphatidylinositol 3- kinase. Circ Res. 91(1):17-24.

Regan CP, Li W, Boucher DM, Spatz S, Su MS, Kuida K. (2002) Erk5 null mice display multiple extraembryonic vascular and embryonic cardiovascular defects. Proc Natl Acad

Sci U S A. 99(14):9248-53.

Reginato AJ, Coquia JA. (2003) Musculoskeletal manifestations of osteomalacia and rickets. Best Pract Res Clin Rheumatol. 17(6):1063-80.

300 Reichel H, Koeffler HP, Barbers R, Norman AW. (1987) Regulation of 1,25- dihydroxyvitamin D3 production by cultured alveolar macrophages from normal human donors and from patients with pulmonary sarcoidosis. J Clin Endocrinol Metab.

65(6):1201-9.

Repasky GA, Chenette EJ, Der CJ. (2004) Renewing the conspiracy theory debate: does

Raf function alone to mediate Ras oncogenesis? Trends Cell Biol. 14(11):639-47.

Reuther GW, Der CJ. (2000) The Ras branch of small GTPases: Ras family members don't fall far from the tree. Curr Opin Cell Biol. 12:157-65.

Richardson JP. (2005) Vitamin D deficiency--the once and present epidemic. Am Fam

Physician. 71(2):241-2.

Rosen H., Reshef A., Maeda N., Lippoldt A., Shpizen S., Triger L., Eggertsen G.,

Bjorkhem I., and Leitersdorf E. (1998) Markedly reduced bile acid synthesis but maintained levels of cholesterol and vitamin D metabolites in mice with disrupted sterol

27-hydroxylase gene. J Biol Chem. 273, 14805-14812

Rosenberg JE, Carroll PR, Small EJ. (2005) Update on chemotherapy for advanced bladder cancer. J Urol. 174(1):14-20.

301 Rosseland CM, Wierod L, Oksvold MP, Werner H, Ostvold AC, Thoresen GH, Paulsen

RE, Huitfeldt HS, Skarpen E. (2005) Cytoplasmic retention of peroxide-activated ERK provides survival in primary cultures of rat hepatocytes. Hepatology. 42(1):200-7.

Roulston A, Reinhard C, Amiri P, Williams LT. (1998) Early activation of c-Jun N- terminal kinase and p38 kinase regulate cell survival in response to tumor necrosis factor alpha. J Biol Chem. 273(17):10232-9.

Sakaki T., Sawada N., Komai K., Shiozawa S., Yamada S., Yamamoto K., Ohyama Y., and Inouye K. (2000) Dual metabolic pathway of 25-hydroxyvitamin D3 catalyzed by human CYP24. Eur. J. Biochem. 267, 6158-6165.

Salmeron A, Ahmad TB, Carlile GW, Pappin D, Narsimhan RP, Ley SC. (1996)

Activation of MEK-1 and SEK-1 by Tpl-2 proto-oncoprotein, a novel MAP kinase kinase kinase. EMBO J. 15(4):817-26.

Sambrook J, Fritsch EF, Maniatis T (1989): “Molecular Cloning- A Laboratory Manual.”

New York: Cold Spring Harbor Laboratory.

Santillan GE, Boland RL. (1998) Studies suggesting the participation of protein kinase A in 1, 25(OH)2-vitamin D3-dependent protein phosphorylation in cardiac muscle. J Mol

Cell Cardiol. 30(2):225-33.

302 Schedl HP, Ronnenberg W, Christensen KK, Hollis BW. (1994) Vitamin D and enterocyte brush border membrane calcium transport and fluidity in the rat. Metabolism.

43(9):1093-103.

Schu PV, Takegawa K, Fry MJ, Stack JH, Waterfield MD, Emr SD. (1993)

Phosphatidylinositol 3-kinase encoded by yeast VPS34 gene essential for protein sorting.

Science. 260(5104):88-91.

Schuster I., Egger H., Reddy G.S., and Vorisek G. 2003 Combination of vitamin D metabolites with selective inhibitors of vitamin D metabolism. Recent Results Cancer

Res. 164, 169-188.

Schwartz Z, Ehland H, Sylvia VL, Larsson D, Hardin RR, Bingham V, Lopez D, Dean

DD, Boyan BD. (2002) 1alpha,25-dihydroxyvitamin D(3) and 24R,25-dihydroxyvitamin

D(3) modulate growth plate chondrocyte physiology via protein kinase C-dependent phosphorylation of extracellular signal-regulated kinase 1/2 mitogen-activated protein kinase. Endocrinology. 143(7):2775-86.

Schweyer S, Soruri A, Heintze A, Radzun HJ, Fayyazi A. (2004a) The role of reactive oxygen species in cisplatin-induced apoptosis in human malignant testicular germ cell lines. Int J Oncol. 25(6):1671-6.

303 Schweyer S, Soruri A, Meschter O, Heintze A, Zschunke F, Miosge N, Thelen P, Schlott

T, Radzun HJ, Fayyazi A. (2004) Cisplatin-induced apoptosis in human malignant testicular germ cell lines depends on MEK/ERK activation. Br J Cancer. 91, 3:589-98.

Sedletska Y, Giraud-Panis MJ, Malinge JM. (2005) Cisplatin Is a DNA-Damaging

Antitumour Compound Triggering Multifactorial Biochemical Responses in Cancer

Cells: Importance of Apoptotic Pathways. Curr Med Chem Anti-Canc Agents. 5(3):251-

65.

Seo EG, Einhorn TA, Norman AW. (1997) 24R,25-dihydroxyvitamin D3: an essential vitamin D3 metabolite for both normal bone integrity and healing of tibial fracture in chicks. Endocrinology. 138(9):3864-72.

Shaulian E, Karin M. (2001) AP-1 in cell proliferation and survival. Oncogene. 20:2390-

400.

Shimada K, Nakamura M, Ishida E, Kishi M, Yonehara S, Konishi N. (2003) c-Jun NH2- terminal kinase-dependent Fas activation contributes to etoposide-induced apoptosis in p53-mutated prostate cancer cells. Prostate. 55, 4:265-80

Shimizu T, Taira N, Senou M, Takeda K. (2002) Involvement of diverse protein kinase C isoforms in the differentiation of ML-1 human myeloblastic leukemia cells induced by

304 the vitamin D3 analogue KH1060 and the phorbol ester TPA. Cancer Lett. 186(1):67-74.

Shingu T, Yamada K, Hara N, Moritake K, Osago H, Terashima M, Uemura T,

Yamasaki T, Tsuchiya M. (2003) Synergistic augmentation of antimicrotubule agent- induced cytotoxicity by a phosphoinositide 3-kinase inhibitor in human malignant glioma cells. Cancer Res. 63(14):4044-7.

Siddhanti SR, Quarles LD. (1994) Molecular to pharmacologic control of osteoblast proliferation and differentiation. J Cell Biochem. 55(3):310-20.

Siddik ZH. (2003) Cisplatin: mode of cytotoxic action and molecular basis of resistance.

Oncogene. 22(47):7265-79.

Singh RA, Zhang JZ. (2004) Differential activation of ERK, p38, and JNK required for

Th1 and Th2 deviation in myelin-reactive T cells induced by altered peptide ligand. J

Immunol. 173(12):7299-307.

Sinha BK, Mimnaugh EG. (1990) Free radicals and anticancer drug resistance: oxygen free radicals in the mechanisms of drug cytotoxicity and resistance by certain tumors.

Free Radic Biol Med. 8(6):567-81.

305 Smalley KS. (2003) A pivotal role for ERK in the oncogenic behaviour of malignant melanoma. Int J Cancer. 104(5):527-32.

Smith SJ, Green LM, Hayes ME, Mawer EB. (1999) Prostaglandin E2 regulates vitamin

D receptor expression, vitamin D-24-hydroxylase activity and cell proliferation in an adherent human myeloid leukemia cell line (Ad-HL60). Prostaglandins Other Lipid

Mediat. 57(2-3):73-85.

Soh JW, Lee EH, Prywes R, Weinstein IB. (1999) Novel roles of specific isoforms of protein kinase C in activation of the c-fos serum response element. Mol Cell Biol.

19:1313-24.

Sohn SJ, Sarvis BK, Cado D, Winoto A. (2002) ERK5 MAPK regulates embryonic angiogenesis and acts as a hypoxia-sensitive repressor of vascular endothelial growth factor expression. Biol Chem. 277(45):43344-51.

Solomon C, White JH, Kremer R. (1999) Mitogen-activated protein kinase inhibits 1,25- dihydroxyvitamin D3-dependent signal transduction by phosphorylating human retinoid

X receptor alpha. J Clin Invest. 103(12): 1729-1735

306 Song LN. (1996) Demonstration of vitamin D receptor expression in a human megakaryoblastic leukemia cell line: regulation of vitamin D receptor mRNA expression and responsiveness by forskolin. J Steroid Biochem Mol Biol. 57(5-6):265-74.

Sooy K, Sabbagh Y, Demay MB. (2005) Osteoblasts lacking the vitamin D receptor display enhanced osteogenic potential in vitro. J Cell Biochem. 94(1):81-7.

Staal A, van den Bemd GJ, Birkenhager JC, Pols HA, van Leeuwen JP. (1997)

Consequences of vitamin D receptor regulation for the 1,25-dihydroxyvitamin D3- induced 24-hydroxylase activity in osteoblast-like cells: initiation of the C24-oxidation pathway. Bone. 20(3):237-43.

Stains JP, Civitelli R. (2005) Cell-to-cell interactions in bone. Biochem Biophys Res

Commun. 328(3):721-7.

St-Arnaud R, Arabian A, Travers R, Barletta F, Raval-Pandya M, Chapin K, Depovere J,

Mathieu C, Christakos S, Demay MB, Glorieux FH. (2000) Deficient mineralization of intramembranous bone in vitamin D-24-hydroxylase-ablated mice is due to elevated

1,25-dihydroxyvitamin D and not to the absence of 24,25-dihydroxyvitamin D.

Endocrinology. 141(7):2658-66.

307 St-Arnaud R. (1999) Novel findings about 24,25-dihydroxyvitamin D: an active metabolite? Curr Opin Nephrol Hypertens. 8(4):435-41.

St-Arnaud R. and Glorieux F.H. (1997) Vitamin D and bone development. In Feldman D,

Glorieux, Pike JW (eds): “Vitamin D.” London UK: Academic Press Limited, pp 293-

303.

Stone AA, Chambers TC. (2000) Microtubule inhibitors elicit differential effects on

MAP kinase (JNK, ERK, and p38) signaling pathways in human KB-3 carcinoma cells.

Exp Cell Res. 254(1):110-9.

Styczynski J, Kurylak A, Wysocki M. (2005) Cytotoxicity of cortivazol in childhood acute lymphoblastic leukemia. Anticancer Res. 25:2253-8.

Suda T, Ueno Y, Fujii K, Shinki T. (2003) Vitamin D and bone. J Cell Biochem.

88(2):259-66.

Sun W, Kesavan K, Schaefer BC, Garrington TP, Ware M, Johnson NL, Gelfand EW,

Johnson GL. (2001) MEKK2 associates with the adapter protein Lad/RIBP and regulates the MEK5-BMK1/ERK5 pathway. J Biol Chem. 276(7):5093-100.

308 Swift LP, Rephaeli A, Nudelman A, Phillips DR, Cutts SM. (2006) Doxorubicin-DNA adducts induce a non-topoisomerase II-mediated form of cell death. Cancer Res.

66:4863-71.

Takeda S, Yoshizawa T, Nagai Y, Yamato H, Fukumoto S, Sekine K, Kato S, Matsumoto

T, Fujita T. (1999) Stimulation of osteoclast formation by 1,25-dihydroxyvitamin D requires its binding to vitamin D receptor (VDR) in osteoblastic cells: studies using VDR knockout mice. Endocrinology. 140(2):1005-8.

Takeyama K, Kitanaka S, Sato T, Kobori M, Yanagisawa J, Kato S. (1997) 25-

Hydroxyvitamin D3 1alpha-hydroxylase and vitamin D synthesis. Science.

277(5333):1827-30.

Tallman MS, Gilliland DG, Rowe JM. (2005) Drug therapy of acute myeloid leukemia.

Blood. 0: 200501017

Tang D, Wu D, Hirao A, Lahti JM, Liu L, Mazza B, Kidd VJ, Mak TW, Ingram AJ.

(2002) ERK activation mediates cell cycle arrest and apoptosis after DNA damage independently of p53. J Biol Chem. 277, 15:12710-7.

Taniguchi T, Eto TA, Shiotsuki H, Sueta H, Higashi S, Iwamura T, Okuda KI, Setoguchi

T. (2001) Newly established assay method for 25-hydroxyvitamin D3 24-hydroxylase

309 revealed much lower Km for 25-hydroxyvitamin D3 than for 1alpha,25- dihydroxyvitamin D3. J Bone Miner Res. 16(1):57-62.

Taniguchi T, Eto TA, Shiotsuki H, Sueta H, Higashi S, Iwamura T, Okuda KI, Setoguchi

T. (2002) Newly established assay method for 25-hydroxyvitamin D3 24-hydroxylase revealed much lower Km for 25-hydroxyvitamin D3 than for 1alpha,25- dihydroxyvitamin D3. J Bone Miner Res. 17(1), 179-81.

Tanner MM, Grenman S, Koul A, Johannsson O, Meltzer P, Pejovic T, Borg A, Isola JJ.

(2000) Frequent amplification of chromosomal region 20q12-q13 in ovarian cancer. Clin

Cancer Res. 6(5):1833-9.

Thompson PD, Jurutka PW, Haussler CA, Whitfield GK, Haussler MR. (1998)

Heterodimeric DNA binding by the vitamin D receptor and retinoid X receptors is enhanced by 1,25-dihydroxyvitamin D3 and inhibited by 9-cis-retinoic acid. Evidence for allosteric receptor interactions. J Biol Chem. 273(14):8483-91.

Tillmann V, Darlington AS, Eiser C, et al (2002) Male sex and low physical activity are associated with reduced spine bone mineral density in survivors of childhood acute lymphoblastic leukemia. J Bone Miner Res 17:1073–1080

310 Toker A, Cantley LC. (1997) Signalling through the lipid products of phosphoinositide-3-

OH kinase. Nature. 387(6634):673-6.

Toker A, Meyer M, Reddy KK, Falck JR, Aneja R, Aneja S, Parra A, Burns DJ, Ballas

LM, Cantley LC. (1994) Activation of protein kinase C family members by the novel polyphosphoinositides PtdIns-3,4-P2 and PtdIns-3,4,5-P3. J Biol Chem. 269(51):32358-

67.

Toker A, Newton AC. (2000) Cellular signaling: pivoting around PDK-1. Cell.

103(2):185-8.

Tokuda N, Kano M, Meiri H, Nomoto K, Naito S. (2000) Calcitriol therapy modulates the cellular immune responses in hemodialysis patients. Am J Nephrol. 20(2):129-37.

Torres R, Calle C, Aller P, Mata F. (2000) Etoposide stimulates 1,25-dihydroxyvitamin

D3 differentiation activity, hormone binding and hormone receptor expression in HL-60 human promyelocytic cells. Mol Cell Biochem. 208: 157-62.

Toyota N, Sakai H, Takahashi H, Hashimoto Y, Iizuka H. (1996) Inhibitory effect of 1 alpha,25-dihydroxyvitamin D3 on mast cell proliferation and A23187-induced histamine release, also accompanied by a decreased c-kit receptor. Arch Dermatol Res.

288(11):709-15.

311 Trump DL, Hershberger PA, Bernardi RJ, Ahmed S, Muindi J, Fakih M, Yu WD,

Johnson CS. (2004) Anti-tumor activity of calcitriol: pre-clinical and clinical studies. J

Steroid Biochem Mol Biol. 89-90: 519-26.

Tsang WP, Chau SP, Kong SK, Fung KP, Kwok TT. (2003) Reactive oxygen species mediate doxorubicin induced p53-independent apoptosis. Life Sci. 73:2047-58.

Tsuchiya R, Suzuki K, Ichinose Y, Watanabe Y, Yasumitsu T, Ishizuka N, Kato H.

(2005) Phase II trial of postoperative adjuvant cisplatin and etoposide in patients with completely resected stage I-IIIa small cell lung cancer: the Japan Clinical Oncology Lung

Cancer Study Group Trial (JCOG9101). J Thorac Cardiovasc Surg. 129(5):977-83.

Underwood JL, DeLuca HF. (1984) Vitamin D is not directly necessary for bone growth and mineralization. Am J Physiol. 246:E493-8.

Usui E, Noshiro M, Ohyama Y, Okuda K. (1990) Unique property of liver mitochondrial

P450 to catalyze the two physiologically important reactions involved in both cholesterol catabolism and vitamin D activation. FEBS Lett. 274(1-2):175-7.

Van Cromphaut S.J., Dewerchin M., Hoenderop J.G., Stockmans I., Herck E., Kato S.,

Bindels R.J., Collen D., Carmeliet P., Bouillon R. and Carmeliet G. (2001) Duodenal

312 calcium absorption in vitamin D receptor-knockout mice: functional and molecular aspects. Proc. Natl. Acad. Sci. U.S.A. 98, 13324–13329.

van Dam H, Wilhelm D, Herr I, Steffen A, Herrlich P, Angel P. (1995) ATF-2 is preferentially activated by stress-activated protein kinases to mediate c-jun induction in response to genotoxic agents. EMBO J. 14(8):1798-811.

Van Leeuwen BL., Kamps WA., Jansen HW., Hoekstra HJ., (2000) The effect of chemotherapy on the growing skeleton. Cancer Treat Rev. 26: 363-76

van Leeuwen JP, van Driel M, van den Bemd GJ, Pols HA. (2001) Vitamin D control of osteoblast function and bone extracellular matrix mineralization. Crit Rev Eukaryot Gene

Expr. 11(1-3):199-226.

Vanchieri C. (2004) Studies shedding light on vitamin D and cancer. J Natl Cancer Inst.

96(10):735-6.

Vanhaesebroeck B, Alessi DR. (2000) The PI3K-PDK1 connection: more than just a road to PKB. Biochem J. 346 Pt 3:561-76.

313 Veldman, C.M., Cantorna, M. T., and DeLuca, H.F.(2000) Expression of 1,25- dihydroxyvitamin D3 receptor in the immune system. Arch.Biochem. Biophys. 374, 334-

338.

Venezio FR, Kozeny GA, DiVincenzo CA, Hano JE. (1988)Effects of 1,25 dihydroxyvitamin D3 on leukocyte function in patients receiving chronic hemodialysis. J

Infect Dis. 158(5):1102-5.

Venning G. (2005) Recent developments in vitamin D deficiency and muscle weakness among elderly people. BMJ. 330(7490):524-6.

Versteeg HH, Evertzen MW, van Deventer SJ, Peppelenbosch MP. (2000) The role of phosphatidylinositide-3-kinase in basal mitogen-activated protein kinase activity and cell survival. FEBS Lett. 465(1):69-73.

Vertino AM, Bula CM, Chen JR, Almeida M, Han L, Bellido T, Kousteni S, Norman

AW, Manolagas SC. (2005) Nongenotropic, Anti-Apoptotic Signaling of

1{alpha},25(OH)2-Vitamin D3 and Analogs through the Ligand Binding Domain of the

Vitamin D Receptor in Osteoblasts and Osteocytes: MEDIATION BY Src,

PHOSPHATIDYLINOSITOL 3-, AND JNK KINASES. J Biol Chem. 280(14):14130-7.

314 Vijayakumar S, Mehta RR, Boerner PS, Packianathan S, Mehta RG. (2005) Clinical trials involving vitamin D analogs in prostate cancer. Cancer J. 11:362-73.

Vojtek AB, Der CJ. (1998) Increasing complexity of the Ras signaling pathway. J Biol

Chem. 273(32):19925-8.

Waas WF, Dalby KN. (2001) Purification of a model substrate for transcription factor phosphorylation by ERK2. Protein Expr Purif. 23: 191-7.

Waas WF, Dalby KN. (2002) Transient protein-protein interactions and a random- ordered kinetic mechanism for the phosphorylation of a transcription factor by extracellular-regulated protein kinase 2. J Biol Chem. 277: 12532-40.

Waas WF, Dalby KN. (2003) Physiological concentrations of divalent magnesium ion activate the serine/threonine specific protein kinase ERK2. Biochemistry 42: 2960-70.

Wali RK, Kong J, Sitrin MD, Bissonnette M, Li YC. (2003) Vitamin D receptor is not required for the rapid actions of 1,25-dihydroxyvitamin D3 to increase intracellular calcium and activate protein kinase C in mouse osteoblasts. J Cell Biochem. 88(4):794-

801.

315 Walters JR, Howard A, Lowery LJ, Mawer EB, Legon S. (1999) Expression of genes involved in calcium absorption in human duodenum. Eur J Clin Invest. 29(3):214-9.

Wang G, Reed E, Li QQ. 2004 Molecular basis of cellular response to cisplatin chemotherapy in non-small cell lung cancer (Review). Oncol Rep. 12(5):955-65.

Wang JC. (2002) Cellular roles of DNA topoisomerases: a molecular perspective. Nat

Rev Mol Cell Biol. 3(6):430-40.

Wang LG, Liu XM, Kreis W, Budman DR. (1999) The effect of antimicrotubule agents on signal transduction pathways of apoptosis: a review. Cancer Chemother Pharmacol.

44(5):355-61.

Wang Q, Salman H, Danilenko M, Studzinski GP. (2005) Cooperation between antioxidants and 1,25-dihydroxyvitamin D(3) in induction of leukemia HL60 cell differentiation through the JNK/AP-1/Egr-1 pathway. J Cell Physiol. [Epub ahead of print]

Wang Q, Wang X, Studzinski GP. (2003) Jun N-terminal kinase pathway enhances signaling of monocytic differentiation of human leukemia cells induced by 1,25- dihydroxyvitamin D3. J Cell Biochem. 89(6):1087-101.

316 Wang X, Merritt AJ, Seyfried J, Guo C, Papadakis ES, Finegan KG, Kayahara M, Dixon

J, Boot-Handford RP, Cartwright EJ, Mayer U, Tournier C. (2005) Targeted deletion of mek5 causes early embryonic death and defects in the extracellular signal-regulated kinase 5/myocyte enhancer factor 2 cell survival pathway. Mol Cell Biol. 25(1):336-45.

Warner JT, Bell W, Webb DK, Gregory JW. (1998) Daily energy expenditure and physical activity in survivors of childhood malignancy. Pediatr Res. 43:607–613

Wasserman RH, Chandler JS, Meyer SA, Smith CA, Brindak ME, Fullmer CS, Penniston

JT, Kumar R. (1992) Intestinal calcium transport and calcium extrusion processes at the basolateral membrane. J Nutr. 122:662-71.

Wasserman RH, Fullmer CS. (1983) Calcium transport proteins, calcium absorption, and vitamin D. Annu Rev Physiol. 45:375-90.

Wasserman RH, Fullmer CS. (1995) Vitamin D and intestinal calcium transport: facts, speculations and hypotheses. J Nutr. 125:1971S-1979S.

Weinstein RS, Underwood JL, Hutson MS, DeLuca HF. (1984) Bone histomorphometry in vitamin D-deficient rats infused with calcium and phosphorus. Am J Physiol.

246:E499-505.

317 Weisman Y, Bab I, Gazit D, Spirer Z, Jaffe M, Hochberg Z (1987) Long-term intracaval calcium infusion therapy in end-organ resistance to 1,25-dihydroxyvitamin D. Am J Med

83:984–990.

Weiss MM, Snijders AM, Kuipers EJ, Ylstra B, Pinkel D, Meuwissen SG, van Diest PJ,

Albertson DG, Meijer GA. (2003) Determination of amplicon boundaries at 20q13.2 in tissue samples of human gastric adenocarcinomas by high-resolution microarray comparative genomic hybridization. J Pathol. 200(3):320-6.

Weldon CB, Scandurro AB, Rolfe KW, Clayton JL, Elliott S, Butler NN, Melnik LI,

Alam J, McLachlan JA, Jaffe BM, Beckman BS, Burow ME. (2002) Identification of mitogen-activated protein kinase kinase as a chemoresistant pathway in MCF-7 cells by using gene expression microarray. Surgery. 132(2):293-301.

Wennerberg K, Rossman KL, Der CJ. (2005) The Ras superfamily at a glance. J Cell Sci.

118: 843-6.

Westwick JK, Cox AD, Der CJ, Cobb MH, Hibi M, Karin M, Brenner DA. (1994)

Oncogenic Ras activates c-Jun via a separate pathway from the activation of extracellular signal-regulated kinases. Proc Natl Acad Sci U S A. 91(13):6030-4.

Wharton B, Bishop N. (2003) Rickets. Lancet. 362(9393):1389-400.

318 Whitfield GK, Hsieh JC, Jurutka PW, Selznick SH, Haussler CA, MacDonald PN,

Haussler MR. (1995) Genomic actions of 1,25-dihydroxyvitamin D3. J Nutr. 125:1690S-

1694S.

Wiese RJ, Uhland-Smith A, Ross TK, Prahl JM, DeLuca HF. (1992) Up-regulation of the vitamin D receptor in response to 1,25-dihydroxyvitamin D3 results from ligand-induced stabilization. Biol Chem. 267(28):20082-6.

Wikvall K. (2001) Cytochrome P450 enzymes in the bioactivation of vitamin D to its hormonal form. Int. J. Mol. Med. 7, 201-209

Wolter H, Gottfried HW, Mattfeldt T. (2002) Genetic changes in stage pT2N0 prostate cancer studied by comparative genomic hybridization. BJU Int. 89(3):310-6.

Wu CH, Gordon J, Rastegar M, Ogretmen B, Safa AR. (2002) Proteinase-3, a serine protease which mediates doxorubicin-induced apoptosis in the HL-60 leukemia cell line, is downregulated in its doxorubicin-resistant variant. Oncogene. 21: 5160-74.

Yamanaka Y, Matsuo H, Mochizuki S, Nakago S, Yoshida S, Maruo T. (2003) Effects of estriol on cell viability and 1,25-dihydroxyvitamin D3 receptor mRNA expression in cultured human osteoblast-like cells. Gynecol Endocrinol. 17(6):455-61.

319 Yang J, Boerm M, McCarty M, Bucana C, Fidler IJ, Zhuang Y, Su B. (2000) Mekk3 is essential for early embryonic cardiovascular development. Nat Genet. Mar 24(3):309-13.

Yang W, Friedman PA, Kumar R, Omdahl JL, May BK, Siu-Caldera ML, Reddy GS,

Christakos S. (1999) Expression of 25(OH)D3 24-hydroxylase in distal nephron: coordinate regulation by 1,25(OH)2D3 and cAMP or PTH. Am J Physiol. 276:E793-805.

Yeh PY, Chuang SE, Yeh KH, Song YC, Chang LL, Cheng AL. (2004) Phosphorylation of p53 on Thr55 by ERK2 is necessary for doxorubicin-induced p53 activation and cell death. Oncogene. 23:3580-8.

Yeh PY, Chuang SE, Yeh KH, Song YC, Ea CK, Cheng AL. (2002) Increase of the resistance of human cervical carcinoma cells to cisplatin by inhibition of the MEK to

ERK signaling pathway partly via enhancement of anticancer drug-induced NF kappa B activation. Biochem Pharmacol. 63:1423-30.

Zehnder D, Bland R, Williams MC, McNinch RW, Howie AJ, Stewart PM, Hewison M.

(2001) Extrarenal Expression of 25-Hydroxyvitamin D3-1a-Hydroxylase. J. Clin

Endocrinol Metab. 86, 888-894

Zehnder D, Hewison M. (1999) The renal function of 25-hydroxyvitamin D3-1a- hydroxylase. Molecular and Cellular Endocrinology. 151, 213-220

320 Zhang AB, Zheng SS, Jia CK, Wang Y. (2004) Role of 1,25-dihydroxyvitamin D3 in preventing acute rejection of allograft following rat orthotopic liver transplantation. Chin

Med J (Engl). 117:408-12.

Zhang MY, Wang X, Wang JT, Compagnone NA, Mellon SH, Olson JL, Tenenhouse

HS, Miller WL, Portale AA. (2002) Dietary phosphorus transcriptionally regulates 25- hydroxyvitamin D-1alpha-hydroxylase gene expression in the proximal renal tubule.

Endocrinology. 143:587-95.

Zhao XY, Ly LH, Holloway L, Feldman D. (1999). Liarozole acts synergistically with

1,25-dihydroxyvitamin D3 to inhibit DU-145 human prostate cancer cell growth by blocking 24-hydroxylase activity. Endocrinology 140: 2071-2076.

Zheng W, Xie Y, Li G, Kong J, Feng JQ, Li YC. (2004) Critical role of calbindin-D28k in calcium homeostasis revealed by mice lacking both vitamin D receptor and calbindin-

D28k. J Biol Chem. 279:52406-13.

Zhou JY, Norman AW, Chen DL, Sun GW, Uskokovic M, Koeffler HP. (1990) 1,25-

Dihydroxy-16-ene-23-yne-vitamin D3 prolongs survival time of leukemic mice. Proc

Natl Acad Sci U S A. 87:3929-32.

321 Zhu W, Zou Y, Aikawa R, Harada K, Kudoh S, Uozumi H, Hayashi D, Gu Y, Yamazaki

T, Nagai R, Yazaki Y, Komuro I. (1999) MAPK superfamily plays an important role in daunomycin-induced apoptosis of cardiac myocytes. Circulation. 100: 2100-7.

Zhuo X, Zheng N, Felix CA, Blair IA. (2004) Kinetics and regulation of cytochrome

P450-mediated etoposide metabolism. Drug Metab Dispos. 32:993-1000.

Zierold C, Darwish HM, DeLuca HF. (1994) Identification of a vitamin D-response element in the rat calcidiol (25-hydroxyvitamin D3) 24-hydroxylase gene. Proc Natl

Acad Sci U S A. 91: 900-902.

Zierold C, Mings JA, DeLuca HF. (2001) Parathyroid hormone regulates 25- hydroxyvitamin D(3)-24-hydroxylase mRNA by altering its stability. Proc Natl Acad Sci

U S A. 98: 13572-6.

Zierold C, Mings JA, DeLuca HF. (2003) Regulation of 25-hydroxyvitamin D3-24- hydroxylase mRNA by 1,25-dihydroxyvitamin D3 and parathyroid hormone. J Cell

Biochem. 88:234-7.

Zierold C, Reinholz GG, Mings JA, Prahl JM, DeLuca HF. (2000) Regulation of the procine 1,25-dihydroxyvitamin D3-24-hydroxylase (CYP24) by 1,25-dihydroxyvitamin

D3 and parathyroid hormone in AOK-B50 cells. Arch Biochem Biophys. 381: 323-7.

322 Zierold, C., Darwish, H.M. and Deluca, H.F. (1995) Two vitamin D response elements function in the rat 1,25-dihydroxyvitamin D 24-hydroxylase promoter. J. Biol. Chem.

270, 1675-1678

Zinzani PL, Gherlinzoni F, Storti S, Zaccaria A, Pavone E, Moretti L, Gentilini P,

Guardigni L, De Renzo A, Fattori PP, Falini B, Lauta VM, Mannina D, Zaja F, Mazza P,

Volpe E, Lauria F, Aitini E, Ciccone F, Tani M, Stefoni V, Alinari L, Baccarani M, Tura

S. (2002) Randomized trial of 8-week versus 12-week VNCOP-B plus G-CSF regimens as front-line treatment in elderly aggressive non-Hodgkin's lymphoma patients. Ann

Oncol. 13:1364-9.

Zoidis E, Gosteli-Peter M, Ghirlanda-Keller C, Meinel L, Zapf J, Schmid C. (2002) IGF-I and GH stimulate Phex mRNA expression in lungs and bones and 1,25-dihydroxyvitamin

D(3) production in hypophysectomized rats. Eur J Endocrinol. 146:97-105.

Zou A, Elgort MG, Allegretto EA (1997) Retinoid X receptor (RXR) ligands activate the human 25-hydroxyvitamin D3-24-hydroxylase promoter via RXR heterodimer binding to two vitamin D-responsive elements and elicit additive effects with 1,25- dihydroxyvitamin D3. J Biol Chem. 272:19027-34.

Zoubek A, Holzinger B, Mann G, Peters C, Emminger W, Perneczky-Hintringer E,

Gadner H, Mostbeck G, Horcher E, Dobrowsky W. (1994) High-dose cyclophosphamide,

323 adriamycin, and vincristine (HD-CAV) in children with recurrent solid tumor. Pediatr

Hematol Oncol. 11:613-23.

324