molecular characterization of the von hippel-lindau … · molecular characterization of the von...

135
Molecular Characterization of the von Hippel-Lindau Tumour Suppressor Protein By Ryan Charles Russell A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Laboratory Medicine and Pathobiology University of Toronto © Copyright by Ryan Charles Russell 2009

Upload: ngoque

Post on 20-May-2018

220 views

Category:

Documents


0 download

TRANSCRIPT

Molecular Characterization of the von Hippel-Lindau Tumour Suppressor Protein

By

Ryan Charles Russell

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

Laboratory Medicine and Pathobiology University of Toronto

© Copyright by Ryan Charles Russell 2009

ii

Molecular Characterization of the von Hippel-Lindau

Tumour Suppressor Protein

Ryan Charles Russell

Doctor of Philosophy

Laboratory Medicine and Pathobiology University of Toronto

2009

Abstract

Inheritance of one mutant von Hippel-Lindau (VHL) allele gives rise to the development of the

autosomal dominant VHL disease, which affects approximately 1 in 36 000 individuals. The

VHL tumour suppressor protein plays a critical role in the E3 ubiquitin ligase-mediated

destruction of hypoxia-inducible factor (HIF) and the promotion of fibronectin extracellular

matrix assembly. A failure in either process is associated with oncogenic progression. Work

included in this thesis provides evidence that these tumour suppressor functions are mutually

exclusive. Additionally, post-translational modification of VHL by NEDD8 is shown to act as a

‘molecular switch’, altering VHL protein associations and providing a mechanism of pathway

segregation. As a result of HIF stabilization, the expression of a homophilic adhesion molecule

E-cadherin is significantly down-regulated in primary renal clear-cell carcinoma (RCC) upon

VHL loss. E-cadherin down-regulation is shown to increase the invasive potential and is of

prognostic value in RCC. Finally, VHL and SOCS1 are shown to dimerize and negatively

regulate the JAK2-STAT5 signalling cascade. Defects in this dimerization are shown to underlie

Chuvash polycythemia and provide a molecular understanding of the phenotypic observations

associated with VHL-related polycythemias.

iii

Acknowledgments

First, I would like to thank my supervisor, Dr. Michael Ohh, for his mentoring throughout my

PhD. His unquenchable enthusiasm for science and constant encouragement enabled me to push

my boundaries and become a better scientist. I would also like to thank my supervisory

committee, Dr. Eldad Zacksenhaus and Dr. Meredith Irwin, who have provided important

insight, advice and support that has been essential to my success. The members of the Ohh lab

have continually provided fruitful discussions that have impacted my studies. In particular I

would like to thank Dr. Olga Roche for her excellent collaboration in our joint study of E-

cadherin. I would also like to thank Roxana Sufan for her essential contributions to our joint

study of Chuvash Polycythemia. In addition, I would like to acknowledge the excellent

cooperation of Julie Metcalf in our joint study of many intriguing aspects of tumour biology, and

Stephanie Sybingco for her administrative prowess. I would also note our outstanding

collaborations with Drs. Andrew Evans, Kyle Furge and Bin Teh, which have given our research

a breadth and clinical significance that would be otherwise lacking.

I would like to thank my parents Tom and Laurie Russell for their constant support and

encouragement. Finally, I would like to thank my wife Kiely for her amazing patience,

understanding and support.

iv

Table of Contents

Abstract ........................................................................................................................................................................ii

Acknowledgments.......................................................................................................................................................iii

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

List of Abbreviations .................................................................................................................................................vii

List of Tables...............................................................................................................................................................xi

List of Figures ............................................................................................................................................................xii

Chapter 1 Introduction to the von Hippel-Lindau tumour

suppressor ....................................................................................................................................................................1

1.1 VHL disease ....................................................................................................................................................1

1.1.1 History ...................................................................................................................................................1

1.1.2 Haemangioblastoma in VHL disease.....................................................................................................1

1.1.3 Phaeochromocytoma in VHL disease....................................................................................................2

1.1.4 Renal clear cell carcinoma in VHL disease ...........................................................................................2

1.1.5 Classification of VHL disease ...............................................................................................................3

1.2 Molecular function of VHL .............................................................................................................................5

1.2.1 The VHL gene and tumour suppressor protein......................................................................................5

1.2.2 VHL containing E3 ubiquitin ligase ......................................................................................................6

1.2.2.1 Intracellular oxygen levels dictate HIFα stability ........................................................................9

1.2.3 Fibronectin/collagen IV matrix deposition ..........................................................................................12

1.2.4 Neddylation of VHL ............................................................................................................................12

1.2.5 Microtubule stability and ciliogenesis .................................................................................................13

1.2.6 Regulation of PHD3 in phaeochromocytoma ......................................................................................14

1.2.7 Regulation of early endosome fusion ..................................................................................................14

1.2.8 Maintenance of renal intracellular junctions........................................................................................15

1.2.9 E-cadherin in epithelial cancer ............................................................................................................16

1.3 Polycythemia in VHL disease........................................................................................................................16

1.3.1 Primary and secondary polycythemia..................................................................................................16

1.3.2 Chuvash polycythemia (CP) ................................................................................................................19

v

Chapter 2 VHL Promotes E2 Box-dependent E-cadherin

Transcription by HIF-mediated Regulation of SIP1 and Snail .............................................................................20

2.1 Rationale .......................................................................................................................................................21

2.2 MATERIALS AND METHODS .....................................................................................................................21

2.2.1 Cell Culture..........................................................................................................................................21

2.2.2 Antibodies............................................................................................................................................22

2.2.3 Plasmids...............................................................................................................................................22

2.2.4 Immunoprecipitation and immunoblotting ..........................................................................................22

2.2.5 Hypoxia treatment of cells ...................................................................................................................23

2.2.6 Immunohistochemical staining ............................................................................................................23

2.2.7 Subcellular fractionation......................................................................................................................24

2.2.8 Dual-luciferase assay ...........................................................................................................................24

2.2.9 Microarray analysis .............................................................................................................................25

2.2.10 siRNA-mediated VHL knockdown .................................................................................................25

2.2.11 Quantitative real-time PCR.............................................................................................................26

2.2.12 Chromatin Immunoprecipitation (ChIP) .........................................................................................27

2.3 RESULTS AND DISCUSSION ......................................................................................................................28

2.3.1 Expression of E-cadherin is down-regulated in RCC and correlates with VHL status. .......................28

2.3.2 ‘Knockdown’ of endogenous VHL results in dramatic attenuation of E-cadherin expression. ...........32

2.3.3 shRNA-mediated down-regulation of E-cadherin increases the invasive potential of RCC................35

2.3.4 VHL regulates E-cadherin expression via HIF-dependent mechanism. ..............................................38

2.3.5 VHL down-regulates E-cadherin-specific transcriptional repressors Snail and SIP1..........................43

2.3.6 Wild-type, but not RCC-causing mutant VHL, induces transcriptional activation of E-cadherin. ......48

2.3.7 E-cadherin expression is cell density-dependent. ................................................................................50

2.3.8 Discussion............................................................................................................................................52

Chapter3 NEDD8 defines tumour suppressor function of VHL

.....................................................................................................................................................................................56

3.1 Rationale .......................................................................................................................................................57

3.2 Materials and Methods..................................................................................................................................57

3.2.1 Cells .....................................................................................................................................................57

3.2.2 Antibodies and reagents.......................................................................................................................57

3.2.3 Plasmids...............................................................................................................................................58

3.2.4 Immunoprecipitation and immunoblotting ..........................................................................................58

3.2.5 Affinity Purification.............................................................................................................................59

3.2.6 Metabolic labeling ...............................................................................................................................59

3.2.7 Subcellular fractionation......................................................................................................................59

vi

3.2.8 Confocal microscopy ...........................................................................................................................60

3.2.9 siRNA ..................................................................................................................................................60

3.3 RESULTS AND DISCUSSION ......................................................................................................................60

3.3.1 ECV- and FN-associated functions of VHL are mutually exclusive ...................................................60

3.3.2 Disruption of NEDD8 pathway abrogates FN binding to VHL, but not ECV formation ....................65

3.3.3 Neddylation of VHL prevents ECV complex formation via steric hindrance .....................................69

3.3.4 Cul2 is excluded from the VHL/FN complex......................................................................................72

3.3.5 Discussion............................................................................................................................................75

Chapter 4 VHL/SOCS1 Heterocomplex Degrades JAK2........77

4.1 Rationale .......................................................................................................................................................78

4.2 Materials and Methods..................................................................................................................................79

4.2.1 Cells. ....................................................................................................................................................79

4.2.2 Antibodies............................................................................................................................................79

4.2.3 Plasmids...............................................................................................................................................80

4.2.4 Immunoprecipitation and immunoblotting. .........................................................................................80

4.2.5 Metabolic labeling. ..............................................................................................................................80

4.2.6 In vitro ubiquitylation assay. ...............................................................................................................81

4.2.7 Generation of phenylhydrazine-primed splenic erythroblasts. ............................................................81

4.2.8 Cytokine deprivation and stimulation of murine splenic erythroblasts................................................81

4.3 RESULTS ......................................................................................................................................................82

4.3.1 CP-VHL mutants have reduced capacity to form ECV .......................................................................82

4.3.2 VHL binds JAK2 in a proteasome-sensitive manner...........................................................................85

4.3.3 VHL promotes ubiquitin-mediated degradation of pJAK2..................................................................85

4.3.4 VHL binds and requires SOCS1 to promote pJAK2 degradation........................................................90

4.3.5 CP-VHL/SOCS1 association inhibits pJAK2 binding and degradation ..............................................95

4.3.6 pJAK2 and pSTAT5 are elevated in CP-mice .....................................................................................98

4.3.7 Discussion............................................................................................................................................98

Chapter 5 Conclusions and future directions .........................102

5.1 E-cadherin loss in RCC...............................................................................................................................102

5.2 Uncovering the mechanism of VHL mediated FN assembly .......................................................................105

5.3 Characterization of VHL mutation in additional haematopoietic malignancies.........................................106

References ................................................................................................................................................................108

vii

List of Abbreviations

AffPD affinity pull-down

ALPHA-MEM alpha modification Eagle's medium

aPKC atypical protein kinase C

APP-BP1 APP binding protein 1

AR autoradiography

BAC bacterial artificial chromosomes

Bcl-xl B-cell lymphoma-extra large

BFU-E burst forming units-erythroid

BNIP3L Bcl2/adenovirus E1B interacting protein 3L

CA9 carbonic anhydrase 9

CBP CREB binding protein

CDC53 coil domain containing 53

cDNA complementary DNA

CFU-E colony forming units-erythroid

ChIP chomatin immunoprecipitation

cHL classical Hodgkin lymphoma

CHO Chinese hamster ovary

Chr chromosome

CITED2 Cbp/p300-interacting transactivator, with Glu/Asp-rich carboxy-terminal domain, 2

CLL chronic lymphocytic leukemia

CMV cytomegalovirus

CNS central nervous system

CO2 carbon dioxide

CoCl2 Cobalt Chloride

COLIV collagen IV

CP Chuvash polycythemia

cp ferroxidase

CR chromophobe RCC

cRNA complementary RNA

C-SRC cellular-sarcoma

Ct cycle threshold

C-TAD carboxy-terminal transactivation domain

Cul cullin

CXCR4 chemokine (CXC motif) receptor 4

DAB diaminobenzidine

delta-ef1 eukaryotic translation elongation factor 1, delta

DFO deferoxamine

DMEM Dulbecco's modification Eagle's medium

DNA deoxyribonucleic acid

dNTP deoxynucleotide triphosphate

DTT dichlorodiphenyltrichloroethane

E-CAD epithelial-cadherin

ECM extracellular matrix

ECS elongins BC/Cul2 or 5/SOCS1

ECV elongins BC/Cul2/VHL

EDTA ethylenediaminetetraacetic acid

EGLN egg laying nine

EMT epithelial-mesenchymal transition

viii

ENO1 enolase 1

EPO erythropoietin

EPOR erythropoietin receptor

ER endoplasmic reticulum

ET essential thrombocythemia

FBS fetal bovine serum

FER FPS/FES related tyrosine kinase

FIH factor inhibiting HIF

FISH fluorescent in situ hybridization

FN fibronectin

FYN fibroblast src/yes novel gene

GFP green fluorescent protein

GLUT glucose transporter 1/3

GSK3 glycogen synthase kinase

H &E hematoxylin and eosin

HA hemagglutinin

HDAC histone deacetylase

HER2 human epidermal growth factor receptor 2

HGF hepatocyte growth factor

HIF hypoxia-inducible factor

HMOX1 heme oxygenase (decycling) 1

HRE hypoxia-responsive elements

IB immunoblot

IGFBP1 insulin-like growth factor binding proteins 1

IGFBP2 insulin-like growth factor binding proteins 2

IgGL immunoglobulin G, light chain

IHC immunohistochemistry

IP immunoprecipitation

JAK2 Janus kinase 2

KIF1B kinesin family member 1B

LEF leukocyte enhancer factor

LGL large granular lymphocyte

Log logarithm

Luc luciferase

MDM mouse double minute

MMM myelosclerosis with myeloid metaplasia

MMP matrix metalloproteinase

mRNA messenger RNA

mTOR mammalian target of rapamycin

NAE NEDD8 activating enzyme

NCE NEDD8 conjugating enzyme

NEDD8 neural precursor cell expressed developmentally downregulated protein 8

NEDP1 NEDD8 protease 1

NEM N-ethyl maleimide

NGF nerve growth factor

NLE NEDD8 ligating enzyme

O2 oxygen

ODD oxygen-dependent degradation domain

ON oncocytoma

Opti-MEM Optimal modification Eagle's medium

p27 protein of 27 kilodaltons

ix

p300 protein of 300 kilodaltons

p53 protein of 53 kilodaltons

p73 protein of 73 kilodaltons

PAGE polyacrylamide gel electrophoresis

PBS phosphate buffered saline

PCR polymerase chain reaction

PDF portable document format

PDGF platelet-derived growth factor

PhD doctor of philosophy

PHD prolyl hydroxylase domain

PHZ phenylhydrazine

PI3K phosphatidylinositol-3-kinase

pJAK2 phosphorylated JAK2

PMBL primary mediastinal B-cell lymphoma

PML promyelocytic leukemia

POLII polymerase II

pSTAT5 phosphorylated STAT5

PTEN phosphatase and tensin homolog

PV polycythemia vera

PVDF polyvinylidene difluoride

qPCR quantatative PCR

RAB5 Ras associated protein 5

RB retinoblastoma suspectibility protein

RBC red blood cell

RBX1 RING box 1

RCC renal clear-cell carcinoma

RLU relative luminescence units

RNA ribonucleic acid

RPMI Roswell Park Memorial Institute growth medium

RTK receptor tyrosine kinase

SCF Skp1/Cdc53/F-box protein complex

SCID severe combined immunodeficiency

SDF-1 stromal cell-derived factor 1

SDH succinate dehydrogenase

SDS sodium dodecyl sulfate

shRNA short hairpin RNA

SIP1 Smad-interacting protein 1

siRNA small interfering RNA

SKP-1 S phase kinase-associated protein 1

SNP single nucleotide polymorphism

SOCS suppressor of cytokine signalling

STAT5 signal transducer and activator of transcription 5

TCF T-cell factor

TGF transforming-growth factor

TIMP tissue inhibitor of metalloproteinase

TMA tissue microarray

TSC2 tuberous sclerosis complex 2

Ub ubiquitin

UBCH5A ubiquitin conjugating enzyme homolog 5a

VBC VHL/elongins B/C

VEGF vascular endothelial growth factor

VHL von Hippel-Lindau

x

WCE whole cell extract

WCP whole chromosome paint

WNT wingless type

WT wildtype

ZEB-2 zinc finger homeo box 1B

ZFHX1A Zinc finger homeodomain enhancer binding

xi

List of Tables

Chapter 1 Page

Table 1.1 Classification of VHL disease 4

xii

List of Figures

Chapter 1 Pages

Figure 1.1 Mutations across VHL open reading frame 4

Figure 1.2 Similarities between ECV and SCF ligases 8

Figure 1.3 Intracellular oxygen levels dictate HIFα stability 11

Figure 1.4 JAK2-STAT5 signalling 18

Chapter 2

Figure 2.1 Expression of E-cadherin is down-regulated in RCC and

correlates with VHL status

30

Figure 2.2 Loss of VHL results in down-regulation of E-cadherin 33

Figure 2.3 Down-regulation of E-cadherin increases the migration of

embryonic kidney cells and invasion of RCC cells

36

Figure 2.4 VHL regulation of E-cadherin is HIF-mediated 40-41

Figure 2.5 VHL-mediated transcription of E-cadherin is attenuated by

Snail and SIP1 via the conserved E2 boxes

45-46

Figure 2.6 VHL activity is required for E-cadherin transcription 49

Figure 2.7 Cell confluency influences E-cadherin expression 51

Figure 2.8 VHL gatekeeper’s pathway in renal epithelium 54

xiii

Chapter 3 Pages

Figure 3.1 ECV- and FN-associated functions of VHL are mutually

exclusive

62-63

Figure 3.2 Restriction of a dynamic NEDD8 pathway results in the

attenuation of VHL binding to FN

66-67

Figure 3.3 NEDD8 modification of VHL generates steric hindrance

blocking the formation of ECV

70

Figure 3.4 VHL/FN complex excludes ECV component Cul2 73

Chapter 4

Figure 4.1 CP-VHL exhibits altered binding to ECV components and

JAK2

83-84

Figure 4.2 VHL promotes ubiquitin-mediated destruction of pJAK2 87-88

Figure 4.3 VHL and SOCS1 cooperate to degrade pJAK2 in vivo 92-93

Figure 4.4 CP-VHL mutants are defective in pJAK2 degradation and

R200W/R200W CP mice exhibit elevated pJAK2 and

pSTAT5 levels

96-97

Figure 4.5 The ‘SOCS groove’ and the revised molecular model of CP 100

Chapter 5

Figure 5.1 Role of VHL in the regulation of E-cadherin and β-catenin 104

1

Chapter 1 Introduction to the von Hippel-Lindau tumour suppressor

1.1 VHL disease

1.1.1 History

In 1894, a British ophthalmologist E. Treacher Collins described bilateral retinal haemangiomas

in two siblings, representing the first report of von Hippel-Lindau (VHL) disease1. VHL disease

is a rare heritable disorder with an incidence of approximately 1/36000 live births2,3. The disease

derives its name from the German ophthalmologist Eugene von Hippel, who further described

kindred displaying retinal haemangiomas, and the Swedish neuropathologist Arvind Lindau who

recognized the common origin of the tumours in families who displayed cerebellar

haemangiomas and those with retinal haemangiomas4,5. The two neoplasms have subsequently

been shown to display similar histopathology and as a result are often collectively referred to as

haemangioblastomas6. In addition to haemangioblastoma, VHL patients are predisposed to

develop highly vascularized tumours in multiple organs including: renal clear cell carcinoma

(RCC), phaeochromocytoma, endolymphatic sac tumour, pancreatic islet tumour, epididymal

cystadenoma, and a variety of other malignant and benign tumours7-10. The cardinal

manifestations of VHL disease and the basis by which the disease is classified are

haemangioblastomas of the central nervous system (CNS) and retina, renal clear cell carcinoma,

and phaeochromocytoma.

1.1.2 Haemangioblastoma in VHL disease

Haemangioblastomas arise upon VHL inactivation in the retina, cerebellum, and spinal chord6.

There are also rare occurrences of VHL haemangioblastomas in the pituitary, hypothalamus,

optic nerve, corpus callosum, and other areas of the brain 11-13. These tumours are often cystic

and highly angiogenic due to the secretion of growth factors such as VEGF and PDGF, which

are required for the stimulation and stabilization of the blood vessels within the tumour14. The

cell of origin for haemangioblastoma is poorly understood. It is currently believed that the

tumour arises from a developmentally arrested angioblast that maintains expression of EPO

receptor15. Interestingly, the tumours themselves secrete EPO, which acts along with TGF-α in

an autocrine loop to stimulate tumour cell growth16-19. Collectively, these observations suggest

2

that the alteration in growth factor secretion plays a key role in the genesis of

haemangioblastomas.

1.1.3 Phaeochromocytoma in VHL disease

Phaeochromocytoma is a tumour of the adrenal gland. Tumours arise from the chromaffin cells

of the sympathetic nervous system20. These lesions are usually benign; however, they often

cause a drastic increase in circulating hormones including norepinephrine, epinephrine,

dopamine, and metanephrines21. Elevation of these hormones can cause a variety of symptoms

including hypertension, nausea, headaches, or heart failure22. Phaeochromocytoma was

classically considered at ‘10% tumour’, where 10% were considered to be hereditary, bilateral,

malignant, or extra-adrenal. These numbers are now considered inaccurate and the percentage of

familial phaeochromocytoma is estimated to be between 15 and 25%23. Other clinical

syndromes with predisposition to phaeochromocytoma include: multiple endocrine neoplasia

Types 2A and 2B and neurofibromatosis Type 1. Additionally, mutations in the mitochondrial

complex II, namely succinate dehydrogenase subunits B-D, have recently been described to give

rise to familial phaeochromocytoma23. Recent work has proposed that a common defect in

embryonic culling of sympathetic neurons may represent a generalized defect present in all

hereditary syndromes displaying phaeochromocytoma (for details see section 1.26)24.

1.1.4 Renal clear cell carcinoma in VHL disease

The primary cause of morbidity and mortality in VHL kindred is due to RCC, which is resistant

to conventional chemo and radio theraputics25,26. Similar to haemangioblastomas RCC secretes

EPO, VEGF, PDGF, and TGFα27-29. Prior to RCC formation, VHL kindred develop renal cysts

that display loss of the remaining wild type VHL allele by immunohistochemistry30,31. It is

unclear, however, if RCC must first develop from a renal cyst, and the origin of RCC is still a

matter of debate32-34. It is commonly thought that RCC arises from renal tubular epithelial cells

and tumour cells display markers of both proximal and distal tubules. One hypothesis by

Maxwell and colleagues suggests that loss of VHL in the distal tubule is responsible for the

acquisition of proximal tubule markers and the loss of some distal tubule markers32. In support

3

of the distal tubule as the origin of RCC, cell division is described to be greatly enhanced upon

VHL loss in the distal tubule when compared to the proximal tubule32. Inactivation of the

remaining wildtype allele in the renal tubule does not initiate tumourigenesis, IHC analysis

shows that VHL loss occurs in numerous pre-neoplastic lesions that have not yet gained the

requirements for tumourigenesis32,35. This has lead to the commonly held believe that additional

mutations are required to drive the formation of RCC. It has recently been shown that VHL loss

can induce cellular senescence in murine fibroblasts via the dephosphorylation of RB in a p27

dependent manner36. It has been postulated that for tumourigenesis to occur RB-dependent

senescence must be overcome, although these observations have not been observed in human

renal cells36.

1.1.5 Classification of VHL disease

VHL disease can be divided into two subcategories depending on the risk of developing

phaeochromocytoma (see table 1). Individuals with Type 1 VHL disease are not predisposed to

develop phaeochromocytoma, while Type 2 patients have an increased propensity to develop

phaeochromocytoma37. Type 2 VHL disease is further subdivided into Type 2A, 2B, and 2C:

Type 2B patients also develop renal clear cell carcinoma (RCC) and Type 2C patients

exclusively develop phaeochromocytoma38,39 In addition, VHL patients with Types 1, 2A, and

2B have an increased predisposition to develop the two principal features of the disease, retinal

and CNS haemangioblastomas. Interestingly, a heritable polycythemic disorder, Chuvash

polycythemia, has recently been described as a VHL-related disorder that exists without an

increased cancer predisposition40. Due to the distinct nature of Chuvash polycythemia it can be

considered Type 3 VHL disease.

4

Table 1.1: Classification of VHL disease

Figure 1.1: Mutations across VHL open reading frame. Adapted from compilation of

mutational data obtained from Universal VHL-Mutation Database. Height of bars represents

numbers of afflicted families (ranging from 1-52). See text for additional details.

5

1.2 Molecular function of VHL

1.2.1 The VHL gene and tumour suppressor protein

In 1988, Seizinger and colleagues mapped the locus of the putative VHL tumour suppressor gene

to a narrow region of chromosome 3p. In accord, deletions in this region have been observed in

RCC41. In 1993, Latif and colleagues identified and cloned the gene defective in VHL patients42.

Homozygous deletion of VHL in murine embryonic stem cells results in an embryonic lethal

phenotype43. Defects in extra embryonic vasculogenesis result in death in utero between days

10.5 to 12.5 of gestation43. VHL kindred inherit one mutant copy of the VHL gene and tumours

in this setting arise from the mutational inactivation, gene silencing, or loss of the remaining

wild-type VHL allele, in keeping with Knudson’s ‘Two-hit’ model of tumourigenesis.

The VHL gene contains three exons that produce a 4.5 kb mRNA (see figure 1.1). Two

translation products are observed from the human VHL gene: A full-length VHL of 213 amino

acids with a molecular weight of 30KDa (VHL30), and an internally translated shorter 160 amino

acid VHL of 19 kDa (VHL19), which results from an alternative start site at codon 5444,45. While

both VHL30 and VHL19 are functional tumour suppressors, there are subtle differences between

the two isoforms. For example, VHL19 is equally distributed in the nucleus and cytoplasm, while

VHL30 is found primarily in the cytoplasm with minor fractions localized in the nuclear and

membrane compartments44. VHL30 has the ability to shuttle between the nucleus and the

cytoplasm46. Recently it has been shown that under acidic conditions VHL30 and VHL19 can be

sequestered to nuclear loci, and upon reinstatement of neutral pH, VHL30 and VHL19 return to the

cytoplasm47. Unless otherwise specified, VHL will henceforth refer to both VHL19 and VHL30.

The crystal structure of VHL was determined in 1999 and showed that VHL contains two

functional domains48. The α domain (so named for the α helices that form this domain) was

predicted to function as a binding site for the adaptor elongin C. The β domain (so named for the

β-pleated sheets that form this domain) was predicted to function as a protein-protein interaction

interface. Tumour-derived mutations frequently occur on the surface residues within the α and β

domains, suggesting a significance for these regions in the tumour suppressor function of VHL48.

While the alpha and beta domains of VHL are considered mutational ‘hotspots’, mutations that

6

give rise to VHL disease have been found over the entire open reading frame of the VHL gene

(see figure 1.1). Despite the heterogeneity of these mutations, a phenotypic pattern that

corresponds to specific mutations has arisen. Type 1 VHL disease is often associated with

mutations resulting in gross truncations or even a complete loss of VHL. Mutations that cause

Type 2 VHL disease are frequently missense mutations. These differences in susceptibility to

develop phaeochromocytoma suggest a gain-of-function mutation, or that complete loss of VHL

function is not permissible for the development of phaeochromocytoma. Mutations that give rise

to Type 2B have been described to cause a more profound defect in VHL function compared to

Type 2A, providing an explanation for the difference in susceptibility to RCC49. Unlike Types 1

and 2, Type 3 disease has been described to be caused by inheritance of two point mutations and

therefore disease presents much earlier than Type 1 and 250,51.

1.2.2 VHL containing E3 ubiquitin ligase

As the initial discovery of the VHL gene sequence did not contain any known domains or give

any clues of possible functions, efforts to find VHL-associated proteins were made with the

supposition that these interacting proteins will have known functions or contain motifs with

predicted functions. It is now known that VHL forms a multiprotein complex with elongin C,

elongin B, Rbx1, and Cul252. The VHL complex (ECV) has high structural similarity with a

yeast multiprotein complex called SCF (Skp1/Cdc53/F-box protein) (see figure 1.2). Cul2 and

Cdc53 are members of the cullin family. Elongin C is an orthologue of yeast Skp1, and both

ECV and SCF contain a ring-finger protein Rbx1. SCF is a known E3 ubiquitin ligase complex

that targets substrates recruited via the F-box protein for ubiquitylation (see figure 1.2).

Ubiquitylation represents a common scheme for targeting proteins for rapid degradation by the

26S proteasome. Ubiquitylation of proteins is accomplished by the concerted action of a

common ubiquitin-activating enzyme (E1), ubiquitin-conjugating enzyme (E2) and ubiquitin-

ligating enzyme (E3 often referred to as E3 ligase)53. The ECV acts as an E3 ligase targeting the

HIFα family of transcription factors for polyubiquitylation (see section 1.2.2.1 for a detailed

description of HIFα mediated transcription and oxygen dependent post translational

modification). VHL acts as the substrate recognition component of the ECV binding directly to

HIFα via the β-domain. Interactions with elongins B and C act to dock VHL to Cul2, a

7

scaffolding component that brings also recruits the E2-ubiquitin conjugating enzyme UbcH5a54.

Thus, disruption of either ECV nucleation (α domain mutation) or HIFα binding (β domain

mutation) results in the stabilization of the HIFα transcription factors. Tumours devoid of VHL

show an upregulation of many hypoxia-responsive genes55.

8

Figure 1.2 Similarities between ECV and SCF ligases. See text for details.

9

1.2.2.1 Intracellular oxygen levels dictate HIFα stability

Hypoxia-inducible factor (HIF) is a major regulator mediating the adaptive response to changing

oxygen tension55. HIF is a heterodimeric transcriptional activator, composed of constitutively

stable HIF-β subunit and labile HIF-α subunit, which is stabilized under hypoxia. Thus, the

activity of HIF is conferred by the oxygen-dependent stability of HIF-α. There are three

members of the HIF-α family: HIF-1α, -2α, and -3α55. Under hypoxia, HIF dimers bind to the

hypoxia-responsive elements (HRE) contained in the promoter/enhancer regions of many

hypoxia-responsive genes including VEGF, GLUT1, and EPO56.

As predicted from the VHL crystal structure, the β domain of VHL is necessary and sufficient to

bind the α subunit of HIF. However, this interaction is strictly dependent on oxygen tension57.

That is, under normal oxygen levels VHL recognizes HIF-α, while under reduced oxygen level

VHL fails to recognize HIF-α, explaining why HIF-α is no longer degraded under hypoxia58.

The selectivity of VHL binding of HIF-α was determined to be dependent on the hydroxylation

of conserved proline residues (402 and 564 based on HIF-1α sequence) within the LAPYIXMD

motif found within and near the oxygen-dependent degradation (ODD) domain of HIF-α59-61.

Prolyl-hydroxylation is carried out by a newly identified class of enzymes called prolyl

hydroxylases (PHD) 1, 2, and 3 in the presence of oxygen62. Interestingly, it was recently shown

that PHDs are upregulated during hypoxia63. The current explanation this upregulation of PHDs

is that upon recovering oxygen homeostasis there will be an abundance of PHDs ready to rapidly

hydroxylate HIF-α for subsequent ubiquitin-mediated destruction, thus curtailing the hypoxic

response.

HIF-α is also regulated in a VHL-independent manner through the C-terminal transactivation

domain (C-TAD), which is present in HIF-1α and HIF-2α, but not HIF-3α64. C-TAD interacts

with co-activators p300/CBP to effectively induce the transcription of hypoxia-inducible genes

via HRE55. Recently, it was shown that C-TAD was subjected to hydroxylation at an asparagine

residue at position 803 (based on HIF-1α sequence) under normoxia by Factor Inhibiting HIF-1

(FIH)65,66. Importantly, asparaginyl-hydroxylation of C-TAD prevented the recruitment of

10

p300/CBP via steric hindrance. This represents an added preventive mechanism to suppress the

transcriptional activity of HIF under normoxia, preventing triggering of hypoxia response during

normal oxygen levels (see Fig. 1.3).

Upon VHL inactivation in RCC, HIFα is constitutively stable and inappropriately activates the

hypoxic program under normoxic conditions. The constitutive overexpression of hypoxia

responsive genes such as VEGF and PDGF likely explains the angiogenic phenotype of VHL-

associated tumours, but also supports the notion that constitutive stabilization of HIF-α is

causally linked to tumourigenesis. In support, Kaelin and colleagues have shown that forced

stable expression of HIF-2α in RCC cells ectopically expressing wild-type VHL overrides the

tumour suppressor capacity of VHL and restores the tumourigenic potential of RCC cells in an

animal xenograft system 67. Conversely, shRNA-mediated knockdown of HIF-2α is sufficient to

suppress the tumourigenic capacity of RCC cells devoid of VHL 68. Notably, all RCC-causing

VHL mutants tested-to-date have shown a failure in either assembling of ECV complex or

binding to HIFα 69,70. However, the critical event(s) downstream of HIF that causes neoplastic

transformation of renal tubular epithelial cell is unclear.

11

Figure 1.3 Intracellular oxygen levels dictate HIFαααα stability. See text for details.

12

1.2.3 Fibronectin/collagen IV matrix deposition

VHL binds fibronectin (FN) and this physical interaction is critical for the promotion of proper

extra cellular matrix (ECM) assembly71. All tumour-causing VHL mutants tested-to-date show a

striking failure in either binding to and/or assembly of FN69,70. It was recently shown that an

intact ECM attenuates angiogenesis of RCC lines by impeding the formation of new blood

vessels72. This inhibition of blood vessel formation was shown to be VHL-dependent, but HIF-

independent72. In addition, mice with a conditional knockout of VHL in endothelial cells

showed defects in vasculogenesis, which was correlated to a defect in ECM deposition. Addition

of exogenous FN can partially restore normal vascular phenotype of VHL-null cells73. These

two studies have helped show that the promotion of a normal ECM by VHL not only helps to

impede tumour formation but also represents a ubiquitous pathway necessary for embryonic

vascularization. The conservation of this pathway is highlighted by genomic clustering studies in

Caenorhabditis elegans that identified a discrete HIF-independent role of VHL in ECM

function74. Recently, VHL was shown to also interact with collagen IV (ColIV) to promote its

deposition in the extracellular space75,76. These findings expand the role of VHL-mediated ECM

assembly beyond FN. However, the mechanisms by which VHL promotes matrix deposition

remain poorly understood.

1.2.4 Neddylation of VHL

Analogous to the ubiquitin pathway, ubiquitin-like NEDD8 modification of proteins involves the

concerted actions of a common NEDD8-activating enzyme (E1 or NAE) a specific NEDD8-

conjugating enzyme (E2 or NCE) and a NEDD8-ligating enzyme or E3 ligase (E3 or NLE). The

classic targets of NEDD8 are the cullins, the scaffolding component of E3 ubiquitin ligases SCF

and ECV, where NEDD8 modification has been shown to affect E3 formation and activity 54,77.

In the case of Cul2, RBX1, a common component of cullin containing E3 ligases acts as the E3

for neddylation. Recently, evidence has begun to emerge for the role of NEDD8 in the

suppression of cancer through its recently identified targets VHL, p53, breast cancer-associated

protein 3, and p73. In each case NEDD8 modification has been shown to modulate the activity

of these genes, often impacting their viability as a tumour suppressor78-82. MDM2 acts as the E3

13

NEDD8 ligase for p53 and p73. Recently, we have shown that VHL is covalently modified by

NEDD8 on lysyl residues; however, the NEDD8 ligase for VHL has not yet been identified78.

NEDD8-conjugation, like many other ubiquitin like modifications occurs with a rapid turnover.

Endogenously neddylated VHL comprises less than 5% of total VHL at any given time within

the cell under physiologic conditions78,83. There are 3 lysyl residues on VHL (K159, K171, and

K196) of which K159 is the major acceptor site of NEDD8. Neddylation-defective VHL mutant

with lysine (K) to arginine (R) substitutions retains ‘wild-type’ level of ECV activity78.

However, neddylation-defective VHL mutant showed dramatic attenuation in binding FN. In

addition, RCC cells ectopically expressing neddylation-defective VHL consequently exhibited

reduced extracellular FN fibrillar array and more importantly, despite having ‘normal’ HIF

profile, grew as tumours in SCID mouse xenograft assay78, underscoring the importance of this

minor fraction of NEDD8-modified VHL in renal oncogenesis.

1.2.5 Microtubule stability and ciliogenesis

In addition to the development of RCC, VHL patients are predisposed to develop renal cysts84.

Development of renal cysts is often linked to defects in primary cilia, a sensory appendage with a

core of microtubules capable of measuring both biochemical and mechanical stimuli85-87.

Interestingly, RCC cell lines devoid of VHL do not display primary cilia. Stable re-constitution

of wild type VHL in this background is capable to rescuing native cilia formation85-87. Type 1

and Type 2A mutations of VHL display defects in the ability to promote ciliogenesis; however,

Type 2B mutations retain the ability to maintain cilia88. Interestingly, Type 2B mutations and

not Type 2A result in RCC, raising the possibility that defects in ciliogenesis may not be

necessary for RCC development and that RCC may develop independently of renal cysts. While

the exact mechanism by which VHL promotes ciliogenesis remains unclear, the ability to

maintain cilia appears to correlate with the ability to interact and stabilize microtubules at the

cell periphery87. The ability of VHL to stabilize microtubules is inhibited by phosphorylation on

two serines in the N-terminus. First, phosphorylation of VHL by casein kinase I ‘primes’ VHL

for phosphorylation on serine 69 by glycogen synthase kinase 3 (GSK3). Interestingly, these

phosphorylation events disrupt microtubule stability without disrupting the interaction of VHL

with tubules88. The exact mechanism by which GSK3-mediated phosphorylation inhibits VHL

14

function is unclear, but may be partially explained by the observation that N-terminally

phosphorylated VHL has a reduced capacity to bind HIFα.

1.2.6 Regulation of PHD3 in phaeochromocytoma

Sporadic mutations of VHL are common in RCC, haemangioblastomas and in other tumours

afflicting VHL kindred89. Phaeochromocytoma is a notable exception, where VHL mutations are

not a common cause of the sporadic form of this neoplasm90. This paradox has been attributed to

a loss or gain of function of VHL that must exert itself during embryogenesis in the neuronal

population of cells that give rise to phaeochromocytoma, setting the stage for disease. During

development an excess of sympathetic neurons are produced24,91. Following this expansion of

neurons a developmental cell death program is initiated as the availability of nerve growth factor

(NGF) becomes limiting91.

Molecularly, inadequate NGF levels initiate a JUN-dependent apoptotic program which depends

on PHD3 and downstream KIF1B-beta for apoptosis92. Type 2 VHL mutation results in an

increase in atypical protein kinase C (aPKC), which in turn elevates the levels of JUNB.

Accumulation of JUNB antagonizes c-JUN and inhibits the apoptotic signalling initiated by NGF

withdrawal24. Mutations or deletions that give rise to Type 1 VHL disease are also defective for

aPKC activity, but do not cause phaeochromocytoma. This may be due to a more drastic

stabilization of HIFα, which accompanies these mutations. PHD3 levels are increased by HIFα

and Type 1 mutations seem to recover enough PHD3 activity to allow apoptosis upon NGF

withdrawal93. Thus, the determining factor of whether a particular VHL mutation will give rise

to phaeochromocytoma is the extent to which the negative (through aPKC) and positive (through

HIFα) impact of VHL loss add up to affect PHD3 activity.

Mutation of VHL and other genes that give rise to familial phaeochromocytoma, such as NF1,

confer a resistance to NGF withdrawal94. It is currently believed that a common defect in

embryonic culling of sympathetic neurons is responsible for the survival of a unique subset of

cells that give rise to phaeochromocytoma in these disorders.

1.2.7 Regulation of early endosome fusion

VHL loss in renal cells has been described to increase the levels or activity of multiple receptor

tyrosine kinases (RTK)95,96. Receptor tyrosine kinases are cell surface receptors that bind a

15

variety of hormones, growth factors and cytokines to initiate receptor dimerization and

intracellular signalling. The negative regulation of RTK signalling can take place by

dephosphorylation, ubiquitin mediated degradation or endocytosis followed by lysosomal

destruction97,98. It has recently been shown that VHL loss reduces the rate of RTK turnover

through a general repression of the endocytic pathway99.

The rab family of proteins are responsible for the progression of the endocytic cycle100.

Perturbations of rab proteins severely alter the rate and progression of endosomes by regulation

of endosome fusion and cycling back to the plasma membrane101. The decreased rate of

endocytosis in the case of VHL loss was shown to be a result of transcriptional repression of

rabaptin-5, a Rab5 activator, critical for early endosome fusion99. Rabaptin-5 has an HRE

element in its promoter and the stabilization of HIFα that occurs upon VHL inactivation results

in a repression of rabaptin-5 expression in a HIFα dependent manner. Moreover, it was also

shown that rabaptin-5 is transcriptionally down-regulated in other solid tumours including breast

cancer and oncocytoma99. Thus, HIFα stabilization resulting from genetic alteration, as is the

case in tumours of VHL disease, or by limiting intracellular oxygen results in a generalized

increase in RTK-mediated signalling via the inhibition of early endosome fusion.

1.2.8 Maintenance of renal intracellular junctions

Loss of VHL in RCC cells leads to a loss of cell polarity and hallmarks of differentiation102,103.

These observations have been linked to the loss of both tight and adheren junctions upon VHL

loss104-106. Adheren junctions are known to participate in the signalling from extracellular cues.

One of the most important and well characterized mediators of these signalling cascades is

β-catenin107,108. β-catenin links cadherins to the actin cytosketelon and is important for the

integrity of the junction. When released from the cadherins β-catenin can activate the

transcription of genes by association with the transcription factor TCF (see discussion for

details)107. Disruption of intracellular junctions is associated with loss of epithelial

characteristics and an increase in cell migration and invasion109. The loss of cell-cell junctions

upon VHL inactivation has been hypothesized to occur through both HIF-dependent and

independent mechanisms104-106. It is currently proposed that perturbations in ECM, aPKC,

β -catenin ubiquitylation and or transcriptional regulation of junction members may all play a

role in the loss of adheren or tight junctions upon VHL loss104-106.

16

1.2.9 E-cadherin in epithelial cancer

The transmembrane protein E-cadherin is a major constituent of adherent cell-cell junctions,

which forms homophilic associations via its extracellular cadherin repeats110. The cytoplasmic

tail of E-cadherin associates with β-catenin, which links to α-catenin and actin forming a

dynamic junction108. Loss of E-cadherin is a hallmark of epithelial-mesenchymal transition

(EMT)111. EMT is a process essential in development for various morphogenic events allowing

programmed migration and invasion of cells during normal embryogenesis112. Paradoxically, a

similar program is seen in many cancers of the epithelial origin in which E-cadherin expression

is frequently lost113,114. However, the loss of expression is rarely due to germline or sporadic

mutations in E-cadherin gene, but rather by epigenetic alterations (e.g., CpG island methylation)

or upregulation of E-cadherin-specific transcriptional repressors114,115. The latter has been shown

to play a role in EMT observed several cancer types, including ovarian, breast, prostate, and

gastric cancers111,116-118.

1.3 Polycythemia in VHL disease

In recent years a unique subset of VHL kindred have been identified who do not develop the

classic tumour types associated with Types 1 and 2 VHL disease. Rather, these patients develop

a unique polycythemic disorder that has characteristics of both primary and secondary

polycythemia, caused by the inheritance of two VHL point mutations.

1.3.1 Primary and secondary polycythemia

Polycythemia is a condition characterized by a net increase in the total number of blood cells,

primarily red blood cells (RBCs) resulting in elevated haematocrit, and is generally categorized

as primary or secondary119. Primary polycythemia, the most common form of which is

polycythemia vera (PV), is defined by excessive erythrocytosis arising from an intrinsic defect in

erythroid progenitors rendering them hypersensitive to or independent of EPO stimulation 119.

Secondary polycythemia is defined as excessive erythrocytosis arising from increased production

of EPO 119. For example, perturbation of the oxygen-sensing pathway due to mutations in PHD2

and HIF2α has been identified in individuals with congenital secondary polycythemia 120-122.

17

Polycythemia can also develop secondary to increased EPO production by some renal tumours or

in mice with constitutive expression of HIF2α50,123. Recently, JAK2 mutations, predominated by

V617F, have been identified in the vast majority of PV patients that encode constitutively active

JAK2 124-128. JAK2 binds most prominently STAT5 transcription factors, which, upon

phosphorylation by JAK2, dimerize and translocate to the nucleus to regulate expression of

genes that control proliferation, differentiation and survival of haematopoietic cells (see Fig. 1.4) 129. STAT5 also triggers a negative feedback mechanism by transactivating the expression of

SOCS family members, which bind and inhibit activated JAKs 130. Notably, SOCS1 directly

binds and targets phosphorylated JAK2 for ubiquitin-mediated degradation via E3 ubiquitin

ligase ECS (Elongins BC/Cul2 or 5/SOCS1) 131,132. In addition, colony-forming units-erythroid

(CFU-E) cells from the fetal livers of SOCS1-/- mice were shown to be hyper-responsive to EPO 133. Moreover, JAK2(V617F) mutation induces PV phenotype in mouse bone marrow

transplantation assays, and the introduction of JAK2(V617F) into cytokine-dependent cell lines

promotes cytokine-independent signalling 134-137. JAK2(V617F) is constitutively phosphorylated

at Y1007, which is required for JAK2 activation 124-128,138. Regardless of JAK2(V617F) status,

high STAT5 phosphorylation is detected in bone marrow biopsies of PV patients 139. These lines

of evidence suggest that constitutive activation of JAK2-STAT5 signalling is a major causative

determinant of PV, and that increases in JAK2-STAT5 signalling represents a common

mechanism for the development of primary polycythemia.

18

Figure 1.4 JAK2-STAT5 signalling. See text for details.

19

1.3.2 Chuvash polycythemia (CP)

CP has features of both primary and secondary polycythemia40,140. Homozygous or compound

heterozygous germline mutations of VHL has recently been shown to cause of CP(REFs). The

best characterized of these mutations, R200W is carried with a particularly high frequency in the

Chuvash Autonomous Republic of the Russian Federation, causing an endemic polycythemia

disorder140. Additional mutations in the extreme C-terminus of VHL (i.e. H191D) have also

been described. Since its discovery the R200W mutation has been found in diverse ethnic

populations including anther endemic population in Italy141. The development of CP appears

distinct from the tumour Types 1 and 2 of VHL disease, and as such, CP-patients are not

afflicted with an increased risk of VHL-related tumours. However, Type 3 kindred have a

reduced lifespan due to polycythemia-related complications such as cerebral vascular events and

thrombosis. Current treatment is limited to Aspirin or phlebotomy. Retrospective analysis has

shown no increase in overall survival for either treatment. A mouse model of CP has been

developed by insertion of two alleles of R166W, the equivalent of the R200W mutation in

humans. To avoid confusion from this point the R166W mutation will be referred to as R200W,

in keeping with the human nomenclature of the original report142. CP-patients and

R200W/R200W mice that faithfully recapitulate the human CP condition have high EPO levels

and an intrinsic hypersensitivity to EPO displayed by burst forming units-erythroid (BFU-E)

cells, prominent features of secondary and primary polycythemia, respectively 40,142. The

secondary polycythemic feature was previously explained by Ang et al. who showed diminished

capacity of CP-VHL(R200W) to bind HIFα, resulting in mild HIFα stabilization and elevated

levels of EPO 40. However, HIF has not been associated with hypersensitivity of erythroid

progenitors to EPO and thus, the molecular mechanism underlying primary polycythemic

features of CP remains unknown and unexplained by the currently established functions of VHL,

which infer an additional yet-to-be-defined role(s) of VHL.

20

Chapter 2 VHL Promotes E2 Box-dependent E-cadherin Transcription by

HIF-mediated Regulation of SIP1 and Snail

This work is now published:

Andrew J. Evans*, Ryan C. Russell*, Olga Roche Losada*, T. Nadine Burry, Jason E. Fish, William Y. Kim, Mindy A. Maynard, Michelle L. Gervais, Roxana I. Sufan, Andrew M. Roberts, Leigh A. Wilson, Mark Betten, Cindy Vandewalle, Geert Berx, Philip A. Marsden, Meredith S. Irwin, Bin T. Teh, Michael A.S. Jewett, and Michael Ohh. 2007. VHL Promotes E2 Box-dependent E-cadherin Transcription by HIF-mediated Regulation of SIP1 and Snail. Mol Cell Biol 27(1): 157-169.

* These authors contributed equally to this work

21

2.1 Rationale

Proper regulation of cell-cell adhesion is vital during cell growth, differentiation, and tissue

development. Loss of cell-cell adhesion is frequently associated with tumour progression,

metastasis, and poor prognosis 143. Major constituents of the cell junctions in polarized epithelial

cells are E-cadherins, homophilic adhesion molecules, and their associated catenins 143.

Increased expression of E-cadherin is associated with the differentiation of mesenchymal cells

into tubular epithelial cells of the adult nephron. Conversely, the loss of E-cadherin is associated

with the progression of numerous carcinoma types 143. In addition, forced expression of E-

cadherin suppresses tumour development and invasion in various in vitro and in vivo tumour

model systems, establishing E-cadherin as a critical tumour suppressor of the epithelium 143.

Here, we show that the expression of E-cadherin is significantly down-regulated in human

primary RCC. siRNA-mediated knockdown of endogenous VHL or functional hypoxia resulted

in dramatic attenuation of E-cadherin expression. Importantly, re-introduction of wild-type

VHL, but not RCC-causing VHL mutant incapable of promoting HIF-α degradation, in RCC

(VHL-/-) cells fully restored E-cadherin transcription, in part, via HIF-dependent regulation of

transcriptional repressors Snail and SIP1 (Smad-interacting protein-1; also known as ZEB-2) and

the engagement of RNA Polymerase II on endogenous E-cadherin promoter/gene. These

findings reveal a potentially critical molecular pathway governing the development and

aggressive nature of RCC upon the loss of VHL function.

2.2 MATERIALS AND METHODS

2.2.1 Cell Culture

HEK293A embryonic kidney cells, U2OS osteosarcoma, and 786-O (VHL-/-) renal clear cell

carcinoma cell lines were obtained from the American Type Culture Collection (Rockville, MD)

and maintained in Dulbecco’s modified Eagle’s medium (DMEM) supplemented with 10% heat-

inactivated fetal bovine serum (Sigma) at 37°C in a humidified 5% CO2 atmosphere. 786-O

subclones ectopically expressing wild-type HA-VHL (786-VHL) or mutant HA-VHL(C162F) or

22

(L188V) were previously described 70,144. RCC4 (VHL-/-) renal clear cell carcinoma subclones

stably expressing HA-VHL (RCC4-VHL) or empty plasmid (RCC4-MOCK) were previously

described 57. 786-VHL stably expressing HIF-2α(P531A) (786-VHL+HIF-2α) or empty control

(786-VHL+MOCK) via retrovirus were previously described 67 and generously provided by Dr.

William G. Kaelin. 786-O subclones stably expressing pRetroSUPER-empty or pRetroSUPER-

HIF2α shRNA were previously described 68.

2.2.2 Antibodies

Monoclonal anti-hemagglutinin (HA) antibody (12CA5) was obtained from Roche Molecular

Biochemicals. Monoclonal anti-VHL antibody (IG32) was as previously described 145. Anti-β-

catenin, anti-Lamin A/C and anti-α-tubulin antibodies were obtained from Santa Cruz (Santa

Cruz, CA), Abcam (Cambridge, MA), and Sigma-Aldrich (Oakville, Ontario, Canada),

respectively. Anti-E-cadherin antibody was obtained from BD Transduction Labs (Mississauga,

Canada). Anti-HIF-2α antibody was obtained from Novus Biologicals Inc. (Littleton, CO).

2.2.3 Plasmids

Mammalian expression plasmid pRc-CMV-HA-VHL(WT) was described previously 144. E-

cadherin core promoter (–308/+21)-luciferase reporter plasmids (WT and mut E2, which

contains inactivating mutations in both E2 boxes) and expression plasmid encoding SIP1 were

previously described 146. Expression plasmid encoding Snail was generously provided by Dr.

Paul Hamel.

2.2.4 Immunoprecipitation and immunoblotting

Immunoprecipitation and Western blotting were performed as described previously 71. In brief,

cells were lysed in EBC buffer (50 mM Tris [pH 8.0], 120 mM NaCl, 0.5% NP-40)

supplemented with a cocktail of protease and phosphatase inhibitors (Roche, Laval, Canada).

23

Immunoprecipitates immobilized on protein A-Sepharose beads (Amersham Biosciences,

Piscataway, NJ) were washed five times with NETN buffer (20 mM Tris [pH 8.0], 120 mM

NaCl, 1 mM ETDA, 0.5% NP-40), eluted by boiling in sodium dodecyl sulfate (SDS)-containing

sample buffer, and size-fractionated by SDS-polyacrylamide gel electrophoresis (PAGE).

Resolved proteins were then electro-transferred onto PVDF membrane (Bio-Rad Laboratories,

Hercules, CA), immunoblotted with the various antibodies, and visualized by

chemiluminescence (Amersham Biosciences, Piscataway, NJ).

2.2.5 Hypoxia treatment of cells

Cells were maintained at 1% O2 for indicated times in a ThermoForma (Marietta, OH) hypoxia

chamber (5% CO2, 10% H2, 85% N2). Cell lysates were prepared in the chamber in hypoxic

environment prior to further experimentation.

2.2.6 Immunohistochemical staining

Formalin-fixed paraffin-embedded sections from 13 nephrectomy specimens with renal cell

carcinoma of clear cell type (RCC) were obtained from the files of The Department of Pathology

and Laboratory Medicine at The University Health Network (Toronto, Canada). These tissue

blocks were used and processed in accordance with a University Health Network Research

Ethics Board-approved protocol concerning gene expression in renal cell carcinoma. Tissues

were fixed in 10% neutral buffered formalin for 24-36 h. Representative sections of tumour with

adjacent non-tumour renal parenchyma, 3-4 mm in thickness, were embedded in paraffin and 5-

micron sections were cut and placed on coated slides for light microscopy. Tumour morphology

and classification were assessed using standard hematoxylin and eosin (H&E) staining. The

tumours were classified as RCC according to criteria described in the World Health Organization

classification of renal tumours147. Immunohistochemical staining for E-cadherin and VHL was

performed manually using a standard avidin-biotin-peroxidase complex method. Sections were

incubated overnight in a humidified chamber with either unlabeled mouse anti-human E-

cadherin or mouse anti-human VHL antibodies, each at a 1:2000 dilution, following microwave

pretreatment for antigen retrieval. The sections were then incubated with a biotinylated

secondary antibody (horse anti-mouse IgG, 1:200 dilution) and the avidin-peroxidase complex.

24

The color reaction was visualized using diaminobenzidine (DAB) as the chromagen. The tissue

was then lightly counterstained with hematoxylin.

2.2.7 Subcellular fractionation

Cells were resuspended in Buffer A (10 mM HEPES [pH 7.9], 10 mM KCl, 1.5 mM MgCl2, 0.34

M sucrose, 10% glycerol) supplemented with protease and phosphatase inhibitors (Roche, Laval,

Canada) and 1mM DTT and subsequently lysed with 0.1% Triton X-100. Samples were

incubated 7 min on ice and centrifuged. While the supernatant was recovered (cytoplasmic

fraction), the pellet was washed with Buffer A and resuspended in Buffer B (0.2 mM EGTA [pH

8], 3 mM EDTA [pH 8]) supplemented with protease and phosphatase inhibitors (Roche, Laval,

Canada) and 1 mM DTT. After 30 min incubation, samples were centrifuged and the resulting

supernatant was isolated (nuclear fraction).

2.2.8 Dual-luciferase assay

U2OS osteosarcoma cells grown on 6-well plates were transfected with a total of 2.5 µg of

expression plasmids using Fugene 6 (Roche). E-cad prom-luc WT or mutE2 (0.9 µg per

transfection) was used to measure E-cadherin core promoter-mediated transcription and 0.1 µg of

the renilla luciferase plasmid, pRL-SV40 (Promega), was used as a transfection control. An

empty pcDNA3.1 plasmid (Invitrogen) was used to maintain a constant final amount of

transfected DNA. Cells were lysed 48 h after transfection and luciferase assays performed using

the Dual-Luciferase Reporter Assay system (Promega) and the relative light units (RLUs)

measured using the lumat LB9507 luminometer (Berthold Technologies). Firefly luciferase

RLUs were normalized against Renilla luciferase RLUs and standardized to the result of the E-

cad prom-luc WT only transfection, which was arbitrarily set to 1.0. Experiments and

transfections were performed in triplicate with one representative experiment presented. Error

bars represent standard deviations.

25

2.2.9 Microarray analysis

We have established a large gene expression profiling database of renal tumours, some of which

have previously been published 148,149. For this study, we selected a total of 105 renal tumours of

clear-cell type and 12 normal kidney tissue samples. The Affymetrix HGU133 Plus 2.0

GeneChip oligonucleotide arrays were used for all 117 cases. The HGU133 Plus 2.0 arrays

contain 54,675 probe sets, representing approximately 47,000 transcripts and variants. The

manufacturer’s recommended protocol (GeneChip Expression Analysis Technical Manual,

Affymetrix, April 2003) was followed for expression profiling. Briefly, for oligonucleotide

expression profiling, 5-20 µg of total RNA was used to prepare antisense biotinylated RNA. A

subset of cases was spiked with external poly-A RNA positive controls (Affymetrix, CA).

Synthesis of complementary DNA was performed with the use of T7-oligo (dT) primer. In vitro

transcription was performed using Enzo Bioarray Transcript Labelling Kit (Enzo, NY). The

biotinylated cRNA was subsequently fragmented, and 15 ug was hybridized to each array at

45°C for 16 h. Scanning was performed in a GeneChip 3000 scanner. Quality assessment was

performed in GeneChip Operating System (GCOS) 1.4 (Affymetrix) using global scaling to a

target signal of 500. Quality assessment was also performed using denaturing gel

electrophoresis. Median background was 73, median scaling factor was 3.06 and median

GADPH 3’/5’ratio was 1.03, indicative of a high overall array and RNA quality.

Statistical analyses were performed in the statistical environment R 2.2, utilizing packages from

the Bioconductor project. The MAS 5 algorithm was used to perform pre-processing of the CEL

files, including background adjustment, quartile normalization and summarization. The means

and the standard errors for E-cadherin gene expressions were calculated for each of the group of

samples. A two-tailed Student’s t test was used to determine statistically significant differences

between various groups.

2.2.10 siRNA-mediated VHL knockdown

siGENOME SMARTpool targeted to VHL was used (Dharmacon, Austin, TX). A non-targeting

scrambled siRNA duplex was used as a negative control (5’-CCAUUCCGAUCCUGAUCCG-

3’). HEK293A (VHL+/+) cells grown on 6 well tissue culture plates were transfected with

26

scrambled and VHL siRNA at a final concentration of 200 nM. Briefly, 8 µL of Oligofectamine

(Invitrogen) was incubated with 48 µL of Opti-MEM I (Gibco/Invitrogen) for 8 min. The

oligofectamine mixture was added to the siRNA diluted in 175 µL of Opti-MEM I and incubated

for 20 min before adding to 800 µL of Opti-MEM I into the wells. After 3 h, 300 µL of DMEM

containing 30% heat-inactivated FBS (Sigma) was added to the plates. RNA was extracted 48 h

after transfection using the RNeasy kit (Qiagen, Mississauga, ON) treated with RNA-free DNase

(Ambion, TX, USA) and first-strand cDNA synthesis was performed.

2.2.11 Quantitative real-time PCR

First-strand cDNA synthesis: 1 µL of oligo(dT)23 primer (Sigma) was incubated with 5 µg of

RNA and dH2O (total reaction volume of 20 µL) for 10 min at 70°C in a thermal cycler (MJ

Research, Boston, MA). The mixture was cooled to 4°C at which time 4 µL of 5x 1st strand

reaction buffer, 2 µL of 0.1 M DTT, 1 µL of 10 mM dNTPs, and 1 µL Superscript II reverse

transcriptase (Invitrogen) were added. cDNA synthesis was performed for 1.5 h at 42°C,

followed by 15 min at 70°C in the thermal cycler. Human genomic DNA standards (human

genomic DNA was obtained from Roche, Mannheim, Germany) or cDNA equivalent to 20 ng of

total RNA were added to the qPCR reaction in a final volume of 10 µL containing 1x PCR buffer

(without MgCl2), 3 mM MgCl2, 0.25 units of Platinum Taq DNA polymerase, 0.2 mM dNTPs,

0.3 µL SYBR Green I, 0.2 µL ROX reference dye, and 0.5 µM each primer (Invitrogen).

Amplification conditions were performed as follows: 95°C (3 min), 40 cycles of 95°C (10 s),

65°C (15 s), 72°C (20 s), 95°C (15 s). qPCR was performed using the ABI Prism 7900HT

Sequence Detection System (Applied Biosystems, Foster City, CA). Gene-specific

oligonucleotide primers designed using Primer Express (Applied Biosystems) were as follows:

Snail primer set (5’-TTCAACTGCAAATACTGCAACAAG-3’ and 5’-

CGTGTGGCTTCGGATGTG-3’), SIP1 primer set (5’-CCACACTTCGCGGCTTCTT-3’ and

5’-CGATCTGCGAAGTCTTGTTTGT-3’), E-cadherin primer set (5’-

GTCATCCAACGGGAATGCA-3’ and 5’-TGATCGGTTACCGTG ATCAAAA-3’), GLUT-1

primer set (5'-CACCACCTCACTCCTGT-TACTT-3' and 5'-

CAAGCATTTCAAAACCATGTTTCTA-3'), VEGF primer set (5'-

CTCTCTCCCTCATCGGTGACA-3' and 5'-GGAGGGCAGAGCTGAGTGTTAG-3'), and

27

U1AsnRNP1 primer set (5’-CAACGACAGCCGAGACATGTA-3’ and 5’-

AGCCTCCATCAAATACCCATTC-3’). SYBR Green I fluoresces during each cycle of the

qPCR by an amount proportional to the quantity of amplified cDNA (the amplicon) present at

that time. The point at which the fluorescent signal is statistically significant above background

is defined as the cycle threshold (Ct). Expression levels of the various transcripts were

determined by taking the average Ct value for each cDNA sample performed in triplicate and

measured against a standard plot of Ct values from amplification of serially diluted human

genomic DNA standards. Since the Ct value is inversely proportional to the log of the initial

copy number, the copy number of an experimental mRNA can be obtained from linear regression

of the standard curve. A measure of the fold difference in copy number was determined for each

mRNA. Values were normalized to expression of U1AsnRNP1 mRNA and expressed relative to

scrambled siRNA samples (arbitrarily set to 1.0) and represented as the mean value of three

independent experiments performed in triplicate ± standard deviations.

2.2.12 Chromatin Immunoprecipitation (ChIP)

ChIP was performed as published previously using the Upstate ChIP assay kit 150. 5 µg of anti-

RNA Polymerase II (N-20) antibody (Santa Cruz) was added to sheared, formaldehyde cross-

linked chromatin preparations from 1 x 106 cells, and immunoprecipitation was performed

overnight at 4oC. A control immunoprecipitation without the addition of antibody was also

performed in parallel. An 18 µL aliquot (of 1800 µL total) of chromatin was removed prior to

immunoprecipitation to serve as an input control. The cross-links were reversed by addition of 2

µL of 5M NaCl, and the sample was diluted 1 in 10 before real-time PCR was performed.

Immune complexes were collected with protein A-agarose beads and, after extensive washing,

immune complexes were released, formaldehyde cross-links were reversed and DNA was

purified by phenol-chloroform extraction. Following ethanol precipitation, DNA was

resuspended in 30 µL of water. Real-time was performed on 2 µL of anti-Pol II

immunoprecipitated DNA, 2 µL of no antibody control and 2 µL of the diluted input sample.

Real-time PCR was performed in triplicate using SYBR green chemistry. Copies of the target

gene were determined using genomic DNA as a standard curve (where 1 ng of genomic DNA =

300 copies of a single copy gene). Immunoprecipitated DNA (IP DNA) was determined by

28

subtracting the number of copies from the no antibody control from the anti-Pol II

immunoprecipitated DNA and dividing by the number of copies in the diluted input sample.

Primers were designed to amplify the human E-cadherin promoter; forward: 5’-CCACGC

ACCCCCTCTCAGT-3’ and reverse: 5’-GAGCGGGCTGGAGTCTGAAC-3’, human E-

cadherin exon 10; forward: 5’-CCGTGGATGTGCTGGATGTGA-3’ and reverse: 5’-

TGGGCAGTGTAGGATGTGATTTC-3’ and the human Cyclophilin A promoter; forward: 5’-

CCTCATGTGTCGTCCCCATCA-3’ and reverse: 5’-CGCCCGTTTTATACCACGTTCG-3’.

2.3 RESULTS AND DISCUSSION

2.3.1 Expression of E-cadherin is down-regulated in RCC and correlates

with VHL status.

We have established a large gene expression profiling database of renal tumours, some of which

have previously been published 148,149. For this study, we selected a total of 105 human renal

tumours of clear-cell type and 12 normal kidney tissue samples. Using the Affymetrix HGU133

Plus 2.0 GeneChip oligonucleotide arrays for all 117 cases, we found that the expression of E-

cadherin transcripts was significantly down-regulated in RCC (Fig. 2.1a). This is consistent with

immunohistochemical studies that showed reduced E-cadherin staining in the vast majority of

RCC tumour samples and cell lines tested 151,152. However, the molecular mechanism that

accounts for the frequent loss of E-cadherin in RCC is unknown.

To date, VHL is the most frequently mutated gene in RCC and biallelic inactivation of the VHL

locus is associated with the development of greater than 80% of sporadic RCC. Thus, we asked

whether the expression of E-cadherin is associated with the VHL status. Hematoxylin and eosin

staining of representative sections from nephrectomy specimens from 13 patients confirmed the

characteristic morphologic features of RCC including nests of cells with abundant, optically

clear cytoplasm and delicate cell membranes surrounded by a network of small, thin-walled

blood vessels (data not shown). Each section studied by immunohistochemistry contained

normal renal parenchyma including core convoluted tubules within the renal cortex (Fig. 2.1b,

left lower half of the micrograph) adjacent to RCC (right upper half of the micrograph).

Membranous anti-E-Cadherin staining (upper panel) and cytoplasmic/membranous anti-VHL

staining (lower panel) shown by core convoluted tubules was used as an internal positive control

29

on each slide. Cells in this representative tumour sample showed correlative staining for E-

cadherin and VHL, where negative staining for VHL observed in RCC corresponded with

markedly reduced staining of E-cadherin (Fig. 2.1b, compare upper and lower panels).

To further validate the positive correlation between VHL and E-cadherin expression, TMAs

consisting of 56 RCC cores in quadruplicate were generated. Thirty-nine of the RCC samples

met the quality standard criteria (see Materials and Methods) and were analyzed for E-cadherin

and VHL protein expression patterns. While only 33% (5/15) of the tumours that stained

negative for VHL (15/39) stained positive for E-cadherin, the majority (67% or 16/24) of

tumours that stained positive for VHL (24/39) also stained positive for E-cadherin (Fig 2.1C).

However, a positive stain for VHL does not formally indicate the presence of a wild-type VHL,

as, for example, a subtle point mutation will likely produce a positive staining signal. Thus,

additional mutational analysis will be required to generate a more precise E-cadherin:VHL

correlation index.

30

Figure 2.1. Expression of E-cadherin is down-regulated in RCC and correlates with VHL

status. (A) 105 RCC tumour samples and 12 normal kidney tissue samples were analyzed using

Affymetrix HGU133 Plus 2.0 GeneChip oligonucleotide arrays. Mean E-cadherin expression

31

and standard error were calculated and two-tailed Student’s t test was used to determine

statistical significance between the two groups. (B) Immunohistochemical staining of a

representative RCC with anti-E-cadherin (top panel) and anti-VHL (bottom panel) antibodies.

Note the negative staining of the tumour cells (upper right in each image) with each marker, in

contrast to the positive staining shown by core tubule epithelium in the adjacent non-tumour

renal cortex (lower left in each image) (50x original magnification).

32

2.3.2 ‘Knockdown’ of endogenous VHL results in dramatic attenuation

of E-cadherin expression.

Reconstitution of 786-O (VHL-/-; HIF-1α-/-) or RCC4 (VHL-/-) renal carcinoma cells with HA-

VHL dramatically restored the expression of E-cadherin protein and mRNA, as measured by

Western blotting and quantitative real-time PCR, respectively (Fig. 2.2a and b). In addition,

siRNA-mediated knockdown of endogenous VHL in HEK293A embryonic kidney epithelial

cells resulted in marked down-regulation of E-cadherin expression (Fig. 2.2c). Microarray (Fig.

2.1a) and real-time PCR data strongly suggest that E-cadherin regulation by VHL is at the pre-

translational level. The cytoplasmic domain of E-cadherin is in a complex with β-catenin,

implicating a potential ‘outside-in’ signalling where a loss of E-cadherin would release β-catenin

to associate with the leukocyte enhancer factor (LEF)/T cell factor (TCF) to regulate the

transcription of cell cycle- (e.g., Cyclin D1) or invasion-related genes (e.g., metalloproteinase

matrilysin and FN) 143. Interestingly, increased level of Cyclin D1 has been observed in RCC

cells devoid of VHL at high cell density 153 and cells expressing tumour-causing VHL mutants

fail to assemble proper extracellular FN matrices 71,78. However, both the overall expression and

subcellular localization of β-catenin remained unaffected by VHL (Fig. 2.2a and d), suggesting

that β-catenin-mediated transcription is likely not involved in potential ‘outside-in’ signalling via

the loss of E-cadherin in the context of RCC.

33

Figure 2.2. Loss of VHL results in down-regulation of E-cadherin. (A) VHL-/- 786-O and

RCC4 cells stably expressing wild-type VHL or empty plasmid (MOCK) were lysed, equal

amounts of total cellular lysates separated on SDS-PAGE, and immunoblotted with the indicated

antibodies. Anti-α-tubulin immunoblot was performed as an internal loading control. (B)

Expression of E-cadherin was measured by quantitative real-time PCR in 786-MOCK and 786-

VHL cells and normalized to U1AsnRNP1 mRNA level. E-cadherin level in 786-VHL cells was

arbitrarily set to 1.0. Error bars represent standard deviations of the fold-changes between the

34

indicated cell types over three independent experiments. (C) Endogenous VHL in HEK293A

cells was knocked-down using VHL-specific siRNA or scrambled non-targeting control siRNA.

RNA was then extracted for cDNA synthesis and endogenous transcript levels of VHL, E-

cadherin, and U1AsnRNP1 measured. Error bars represent standard deviations of the fold-

changes between the expression of the indicated mRNA relative to its expression using control

siRNA (arbitrarily set to 1.0) over three independent experiments. (D) 786-MOCK and 786-

VHL cells were biochemically fractionated (see Materials and Methods) into cytoplasmic (C)

and nuclear (N) fractions. 100µg of each fraction were resolved on SDS-PAGE and

immunoblotted with anti-β-catenin (upper panel), anti-Lamin A/C (nuclear protein control;

middle panel) and anti-α-tubulin (cytoplasmic protein control; lower panel) antibodies. IB:

immunoblot.

35

2.3.3 shRNA-mediated down-regulation of E-cadherin increases the

invasive potential of RCC

The role of E-cadherin in modulating the migration and invasion properties of epithelial cells

is well established. However, it is not known whether E-cadherin has similar biological

effects in the context of kidney epithelial cells or RCC. Although E-cadherin expression can

be predictably determined by manipulating the status of VHL, altering the expression level of

VHL has other consequences that can influence the motility and invasion properties of RCC

(see discussion below). Thus, we used an shRNA approach to specifically down-regulate the

endogenous expression level of E-cadherin in HEK293A embryonic kidney epithelial cells,

which resulted in a significant enhancement of migration as measured by percent wound

closure (61.2% ± 4.8%) compared to cells expressing scrambled shRNA (43.0% ± 3.0%)

(Fig 2.3A and B). The change in motility was noticeable from the early time points,

suggesting that the effect of modulating the expression of E-cadherin is not only potent but

also immediate (Fig 2.3C). shRNA-mediated down-regulation of E-cadherin consistently

increased the motility of 786-VHL cells in a similar wound assay but was not statistically

significant (data not shown). However, the invasion potential was increased (2.4 ± 0.2%)-

fold in comparison to 786-VHL cells expressing the scrambled shRNA, as measured on the

standard matrigel invasion chambers (Fig. 2.3D and E). It should be noted that the changes

in motility and invasion are likely underestimated due to the incomplete knockdown of E-

cadherin (Fig. 2.3A and D). Nevertheless, these results suggest that the diminution of E-

cadherin expression would promote the invasive property of RCC.

36

Figure 2.3. Down-regulation of E-cadherin increases the migration of embryonic kidney

cells and invasion of RCC cells. (A) HEK293A cells were transiently transfected with a

plasmid encoding the scrambled shRNA or a cocktail of four E-cadherin-specific shRNAs.

Equal amounts of the whole-cell lysates were immunoprecipitated with an anti-E-cadherin

37

antibody, resolved by SDS-PAGE and immunoblotted with an anti-γ-tubulin antibody. E-

cadherin signal intensities were quantified using a Kodak Image Station 2000R densitometer

and normalized against the corresponding γ-tubulin signals; values are indicated in the

parentheses. (B) Wounds were created 48h post-transfection with the indicated plasmids.

Percent wound closure was determined by measuring the migration of cells from the wound

edge 25h post-wound scrape. Each wound measurement was taken in triplicate, and the

experiment was repeated three times. (C) Line graph representing early migration profile, as

indicated by percent wound closure, as measured in the experiment shown in panel B, of

HEK293A cells transfected with the indicated shRNA plasmids. (D) 786-VHL cells were

transiently transfected with a plasmid encoding the scrambled shRNA or a cocktail of four E-

cadherin-specific shRNAs. Equal amounts of the whole-cell lysates were

immunoprecipitated with an anti-E-cadherin antibody, resolved by SDS-PAGE and

immunoblotted with an anti-hnRNP antibody. E-cadherin signal intensities were quantified

using a Kodak Image Station 2000R densitometer and normalized against the corresponding

hnRNP signals; results are given in parentheses. (E) 786-VHL cells were transiently

transfected with the indicated plasmids as shown in panel D. Cells were counted 72h

postransfection, and 2.5 x 104 cells were seeded into BD Matrigel Invasion Chambers and

incubated for 22h. The invading cells were stained with 0.1% crystal violet, and images were

captured under an inverted light microscope. Cells were counted from photographs of the

membrane, and each experiment was repeated twice. The relative change in invasion was

determined by counting the number of invading cells transfected with E-cadherin-specific

shRNA and normalizing the value against the number of invading cells transfected with the

scrambled shRNA (arbitrarily set at 1.0). Anti-E-cad, anti-E-cadherin; IP,

immunoprecipitaion; IB, immunoblot; Anti-Tub, anti- γ-tubulin; shE-cad, E-cadherin-

specific shRNA; shScram, scrambled shRNA; T, time.

38

2.3.4 VHL regulates E-cadherin expression via HIF-dependent

mechanism.

We next asked whether the regulation of E-cadherin expression by VHL is mediated through

the activity of HIF. RCC4-VHL and 786-VHL cells were maintained under normoxic (21%)

or hypoxic (1%) conditions for 16 h and analyzed by Western blotting (Fig. 2.4a and b). The

expression of E-cadherin was dramatically reduced under hypoxia while preserving the

expression status of VHL (Fig. 2.4a). The effect of hypoxic treatment was confirmed by the

increase in HIF-2α expression (Fig. 2.4a, upper panel). Although this result suggests that

hypoxia-induced stabilization of HIF results in repression of E-cadherin expression, it is

formally possible that VHL, independent of HIF, regulates the expression of E-cadherin in an

oxygen-dependent manner. Therefore, we examined various VHL mutants that have retained

or lost the ability to regulate HIF. Certain non-RCC-associated VHL mutants have been

shown to retain the ability to regulate HIF activity 69,70,78. For example, The L188V mutation

allows proper oxygen-dependent degradation of HIF-α and is associated with a sub-class of

VHL disease (Type 2C), which is clinically characterized by the exclusive development of

phaeochromocytoma 70. In contrast, invariably all RCC-associated VHL mutations test-to-

date, such as C162F, result in a complete loss of ability to mediate the destruction of HIF-α

via the ubiquitin-proteasome pathway 69,70. 786-O cells ectopically expressing VHL(C162F)

showed negligible expression of E-cadherin, while those expressing VHL(L188V) showed

higher detectable levels of E-cadherin, albeit at a lower level than observed in cells

expressing VHL(WT) (Fig. 2.4c). In addition, 786-O (VHL-/-; HIF-1α-/-) cells stably

expressing wild-type VHL (786-VHL) infected with retroviruses that express functional and

stable HIF-2α(P531A; escapes VHL recognition) demonstrated reduced level of E-cadherin

relative to 786-VHL cells infected with ‘empty’ retrovirus (Fig. 2.4d, compare lanes 1 and 2).

Notably, the level of E-cadherin was inversely proportional to the level of HIF-2α (Fig. 4d).

Conversely, 786-O subclones infected with retroviruses that express HIF-2α-specific shRNA

demonstrated markedly increased level of E-cadherin relative to 786-O cells infected with

‘empty’ retrovirus (Fig. 2.4e). In addition, the activity of the exogenous E-cadherin

promoter-driven luciferase reporter was much higher in 786-O cells reconstituted with wild-

type VHL (786-WT; low HIF activity) than in 786-MOCK (high HIF activity) cells (Fig.

39

2.4f). Taken together, these results strongly suggest that HIF negatively regulates E-cadherin

expression, at a minimum, at the level of transcription.

40

41

Figure 2.4. VHL regulation of E-cadherin is HIF-mediated. (A) RCC4-VHL cells were

maintained under normoxia (N; 21% O2) or hypoxia (H; 1% O2) for 16h, lysed, resolved on

SDS-PAGE, and immunoblotted with anti-HIF-2α (top panel), anti-E-cadherin (middle panel),

and anti-HA (bottom panel) antibodies. Asterisk denotes non-specific bands and illustrates equal

loading of total cellular extracts between lanes. (B) 786-MOCK and 786-VHL cells were

maintained under normoxia (N; 21% O2) or hypoxia (H; 1% O2) for 16h and lysed. Cell lysates

were equilibrated following Bradford protein assay and immunoprecipitated with anti-E-cadherin

antibody and resolved on SDS-PAGE. 100µg of cell lysates were also resolved on SDS-PAGE.

Proteins separated on SDS-PAGE were transferred onto PVDF membrane and immunoblotted

with anti-E-cadherin (top panel) and anti-HA (bottom panel) antibodies. IP:

immunoprecipitation; IB: immunoblot. (C) 786-O cells stably expressing HA-VHL(WT), HA-

VHL(C162F), or HA-VHL(L188V) were lysed, resolved on SDS-PAGE, and immunoblotted

with the indicated antibodies, where α-tubulin served as an internal loading control. (D) 786-

MOCK and 786-VHL cells infected with ‘empty’ retrovirus (786-VHL+EMPTY) or retrovirus

expressing constitutively stable and functional HIF-2α(P531A) were lysed, resolved on SDS-

PAGE, and immunoblotted with the indicated antibodies, where α-tubulin served as an internal

loading control. (E) 786-O (VHL -/-) subclones stably expressing pRetroSUPER-empty or

42

pRetroSUPER-HIF2α shRNA were lysed, comparable amounts of whole cell extracts

immunoprecipitated and immunoblotted with an anti-E-cadherin antibody (top panel). Equal

amounts of the whole cell extracts were also resolved on SDS-PAGE and immunoblotted with

anti-HIF-2α (middle panel) and anti-actin (bottom panel) antibodies. (F) Dual-luciferase assays

were performed in 786-MOCK and 786-VHL cells transfected with the firefly luciferase

construct (E-cad prom-luc) driven by the human E-cadherin promoter sequence. CMV-driven

renilla luciferase was used as an internal transfection control and the firefly luciferase relative

light units (RLUs) were normalized against Renilla luciferase RLUs. Experiments and

transfections were performed in triplicate with one representative experiment presented. Error

bars represent standard deviations.

43

2.3.5 VHL down-regulates E-cadherin-specific transcriptional repressors

Snail and SIP1.

Mutational analyses have shown that biallelic somatic inactivating mutations of E-cadherin are

rare 154,155, and emerging evidences suggest that the loss or reduction in E-cadherin expression in

cancer cells primarily occurs at the level of transcription 113,156-158. The two major regulators of

E-cadherin transcription are the zinc finger transcriptional repressors Snail and SIP1 (Smad-

interacting protein-1; also known as ZEB-2), which bind evolutionarily conserved E2 boxes

located within the E-cadherin core promoter resulting in the inhibition of E-cadherin

transcription 111,146,159. Moreover, hypoxic treatment of ovarian carcinoma cells was shown to

attenuate the expression of E-cadherin via the upregulation of Snail 116.

Quantitative real-time PCR analysis showed a significant attenuation of both Snail and SIP1 in

RCC 786-O cells restored with VHL (Fig. 2.5a). As expected, restoration of VHL reduced the

expression of HIF-target genes, VEGF and GLUT-1, and increased E-cadherin expression (Fig.

2.5a and as shown in Fig. 2.2). These results suggest the possibility that VHL may increase the

expression of E-cadherin by down-regulating the transcriptional repressors Snail and SIP1. In a

complementary experiment, we tested the ability of VHL in the transactivation of E-cadherin

promoter-driven luciferase reporter (Fig. 2.5b). As expected, E-cadherin promoter containing

both E2 boxes had lower basal transcriptional activity relative to E-cadherin promoter with

mutations in the E2 boxes that abrogate Snail/SIP1 binding (Fig. 2.5b). Importantly, the addition

of VHL markedly increased the wild-type E-cadherin promoter-driven luciferase transcription,

but had insignificant effect on the E2 mutant E-cadherin promoter (Fig. 2.5b). Moreover, the

increase in VHL-mediated transactivity of wild-type E-cadherin promoter-luciferase was

dampened by the addition of SIP1 or/and Snail in a dosage-dependent manner (Fig. 2.5c).

Coexpression analysis indicated that Snail or SIP1 had negligible effect on the steady-state level

of VHL (Fig. 2.5d). These results demonstrate that the E2 boxes are functionally important in

upregulating E-cadherin transcription by VHL, in part, via the down-regulation of SIP1 and/or

Snail. However, neither SIP1 nor Snail individually or in combination achieved a complete

inhibition of E-cadherin promoter-driven reporter activity. This suggests the existence of other

yet-to-be-defined VHL-HIF-mediated E-cadherin-specific transcriptional repressors or that full

repression requires the concerted actions of multiple repressors and involves, in addition to the

44

E2 boxes, other elements within the E-cadherin promoter. These results, however, do

demonstrate that the E2 boxes are functionally important in upregulating E-cadherin

transcription by VHL, in part, via the down-regulation of SIP1 and/or Snail.

45

46

Figure 2.5. VHL-mediated transcription of E-cadherin is attenuated by Snail and SIP1 via

the conserved E2 boxes. (A) Expression of E-cadherin, Snail, SIP1, VEGF, GLUT-1 were

measured by quantitative real-time PCR in 786-MOCK and 786-VHL cells and normalized to

U1AsnRNP1 mRNA expression. Solid bars represent expression of the indicated mRNA in 786-

47

MOCK cells relative to its expression in 786-VHL cells, which was arbitrarily set to 1.0. (B)

Dual-luciferase assays were performed in U2OS cells transfected with the indicated expression

plasmids. The firefly luciferase construct (E-cad prom-luc) was driven by the human E-cadherin

promoter sequence (WT) or the promoter with deletion of both E2 boxes (mut E2). CMV-driven

renilla luciferase was used as an internal transfection control and the firefly luciferase relative

light units (RLUs) were normalized against Renilla luciferase RLUs. Experiments and

transfections were performed in triplicate with one representative experiment presented. Error

bars represent standard deviations. (C) Performed as in B with increasing concentrations of

Snail/SIP-1 mixed into HA-VHL transfection reactions at a ratio of 1:2, 2:2, and 4:2 (Snail/SIP-

1:HA-VHL). Fold induction was standardized to E-cad prom-luc activity in the absence of

exogenous VHL. (D) U2OS cells were transfected with the expression plasmid encoding HA-

VHL alone or in combination with a plasmid encoding Snail in increasing amounts (represented

as a triangle). Equal amounts of the whole cell extracts were immunoprecipitated with an anti-

Snail antibody, resolved on SDS-PAGE and immunoblotted with an anti-Snail antibody (top

panel). Equal amounts of the remaining whole cell extracts were resolved on SDS-PAGE and

immunoblotted with an anti-VHL (middle panel) or anti-α-tubulin (bottom panel) antibody.

48

2.3.6 Wild-type, but not RCC-causing mutant VHL, induces

transcriptional activation of E-cadherin.

Principal mechanism by which transcriptional repressors attenuate the rate of transcription is by

blocking the engagement of RNA Polymerase II (Pol II) and associated factors to the promoter.

Notably, Snail has been shown to repress E-cadherin expression through the binding of

Sin3A/histone deacetylase 1 (HDAC1)/HDAC2 complex 160, which is thought to impair

recruitment of Pol II and transcriptional initiation via repressive changes in chromatin structure.

Repressors can also act at a post-initiation step of transcription including inhibition of

phosphorylation of the Pol II holoenzyme, which represents a key step in promoter escape and

cessation in elongation 161. Therefore, we asked whether the engagement of Pol II on the

endogenous E-cadherin promoter/gene is influenced by VHL. Chromatin immunoprecipitation

(ChIP) assays at the E-cadherin locus was performed using antibodies recognizing the N-

terminus of Pol II in 786-O cells expressing wild-type or HIF regulation-defective mutant

VHL(C162F) (Fig. 2.6a). The total amount of Pol II at the E-Cadherin promoter and exon 10

was dramatically decreased in the absence of VHL functional activity (Fig. 2.6a). This was in

contrast to the promoter of the housekeeping gene, Cyclophilin A, where binding of Pol II was

similar in wild-type and mutant VHL cells (Fig. 2.6a). This suggests that VHL functional

activity to negatively regulate HIF is necessary for transcriptional activation of E-cadherin. To

further establish a functional role of VHL/HIF in E-cadherin gene transcription, 786-O cells

stably expressing wild-type VHL were exposed to hypoxia. Assessment of Pol II binding to

genomic regions corresponding to coding regions of E-cadherin (exon 10) revealed that hypoxia

decreased E-cadherin transcription (Fig. 2.6b). A similar decrease in Pol II binding was

demonstrated at the promoter of E-cadherin (data not shown). As expected, hypoxia elicited a

time-dependent increase in VEGF mRNA expression (Fig. 2.6b). Taken together, these results

suggest that VHL activity, specifically E3 ligase function to negatively regulate HIF, is required

for the transcription of the E-cadherin gene. Conversely, cellular hypoxia or loss of HIF-

associated function of VHL results in the activation of HIF and disengagement of Pol II from the

E-cadherin promoter, resulting in the down-regulation of E-cadherin transcription. However, it

is not formally known whether HIF-mediated engagement of Pol II on E-cadherin promoter is

SIP1/Snail-dependent.

49

Figure 2.6. VHL activity is required for E-cadherin transcription. (A) Chromatin

immunoprecipitation (ChIP) using anti-RNA polymerase II antibody was performed on

sheared chromatin from 786-O cell lines (VHL-/-) that had been stably transfected with wild-

type VHL (open bar) or mutant VHL(C162F) (solid bar). IP DNA was determined for the

promoter and exon 10 of E-cadherin and the promoter of Cyclophilin A using real-time PCR,

and the value in VHL(WT) cells was arbitrarily set to 1.0. (B) RNA polymerase II ChIPs

were performed in 786-VHL(WT) cells exposed to 4 or 20h of hypoxia (1% oxygen). IP

DNA for exon 10 of E-cadherin was normalized to the IP DNA for the Cyclophilin A

promoter (left graph). Normoxia was arbitrarily set to 1.0. Expression of VEGF was

assessed by real-time PCR as internal control for hypoxia treatment (right graph).

50

2.3.7 E-cadherin expression is cell density-dependent.

E-cadherin expression in RCC cells is also cell density-dependent as measured by Western

blotting and quantitative real-time PCR (Fig 2.7a and b). Interestingly, the expression of VHL is

strictly regulated by cell density where the steady-state amount of VHL in human renal proximal

tubule epithelial cells was shown to increase more than 100-fold in dense cultures relative to

sparse cultures 162. In addition, other components of the VHL E3 ligase complex showed a

similar cell density-dependent regulation 163. Importantly, HIF-2α level was elevated in sparsely

growing cells with low levels of VHL and significantly reduced or undetectable in confluent

cells containing abundant VHL 163. Moreover, the ability of VHL to shuttle between the nucleus

and the cytoplasm is also regulated by cell density 46, which in turn may influence the ability of

VHL to regulate HIF activity 164. Thus, cell density-dependent expression of E-cadherin may be

due to a corresponding cell density-dependent regulation of VHL stability/function.

51

Figure 2.7. Cell confluency influences E-cadherin expression. (A) 786-MOCK, 786-VHL,

RCC4-MOCK, and RCC4-VHL were grown to varying levels of confluency, lysed, equal

amounts (150µg) of total cell lysates separated on SDS-PAGE, and immunoblotted for E-

cadherin protein expression. (B) Representative experiment showing mRNA expression of E-

cadherin assayed using quantitative real-time PCR after RNA isolation from 786-VHL and 786-

MOCK cells that were grown to the indicated confluencies. Days grown past 100% confluency

are noted in parentheses. E-cadherin expression was normalized to U1AsnRNP1 mRNA

expression.

52

2.3.8 Discussion

VHL is a direct oxygen-dependent negative regulator of HIF-α via the ubiquitin pathway 89.

Loss of VHL or VHL mutations associated with the development of RCC invariably results in

the accumulation/hyper-activation of HIF due to a failure in VHL’s ability to either bind or

ubiquitylate HIF-α. Here, we propose that HIF - stabilized by hypoxia in the presence of wild-

type VHL - or upon mutation/loss of VHL activates the transcriptional repressors SIP1 and Snail

(likely via HIF-engagement to the HRE element (5’-GCGTG-3’) found in the Snail promoter at

position –86 to –82; SIP1 promoter/enhancer has not been defined), preventing PolII engagement

on E-cadherin promoter and resulting in the down-regulation of E-cadherin expression (see Fig.

2.8). There is, however, an alternate pathway to consider (described below).

Increased transforming-growth factor (TGF)-β signalling and expression of Snail and SIP-1, as

well as the loss of E-cadherin expression, have all been correlated with epithelial to

mesenchymal transition (EMT) process that occurs during normal development and acquisition

of invasive phenotype in epithelial cancers 111,146,159,165. Smad-mediated signalling by TGF-β has

been shown to induce the expression of the repressors Snail and SIP-1 146,166,167. HIF-1 has been

shown to upregulate the expression of the members of TGF-β family in a transcription-

dependent manner under hypoxic conditions. HIF-1 and Smad proteins cooperate in regulating

the expression of several hypoxia and TGF-β-regulated genes, including the expression of TGF-

β2 in human umbilical vein endothelial cells 168. In addition, HIF-1 has been shown to directly

bind to the TGF-β3 promoter and upregulate its expression under hypoxia during placental and

epithelial development processes 169,170. Therefore, increased transcriptional activity of HIF by

the functional loss of VHL in RCC may result in an upregulation of TGF-β signalling, resulting

in a Smad-mediated induction of SIP1 and Snail and subsequent loss of E-cadherin (Fig. 2.8).

Notably, VHL has also been shown to repress the expression of TGF-β1 via regulating its mRNA

stability (Fig. 2.8, dashed line) 171. Whether this process is HIF-mediated is currently unknown.

In the current work, we demonstrated that the loss of VHL leads to the dramatic down-regulation

of E-cadherin in RCC. VHL-dependent transactivation of E-cadherin was dependent on the

conserved E2 boxes known to recruit transcriptional repressors Snail and SIP1 to the promoter of

E-cadherin. Re-introduction of VHL in RCC cells devoid of VHL showed a reduction in the

53

expression of both Snail and SIP1 and thereby explaining, at least in part, the resulting

restoration of E-cadherin expression. Transcriptional repressors principally block transcription

by inhibiting the engagement of PolII to the promoter. In support, Snail has been shown to

repress E-cadherin expression through the binding of histone deacetylase, promoting repressive

changes in chromatin structure and thereby impairing the recruitment of PolII and transcription 160. Consistent with this view, wild-type VHL enhanced the recruitment of PolII to the E-

cadherin promoter/gene. However, hypoxia or tumour-causing VHL mutant with a failure in

targeting HIF-α for ubiquitin-mediated destruction dramatically decreased the association of

PolII with the E-cadherin gene. Thus, VHL directly affects PolII engagement on E-cadherin

DNA via HIF-dependent regulation of E-cadherin-specific transcriptional repressors, revealing a

previously unrecognized regulation of a major epithelial tumour suppressor E-cadherin.

54

Figure 2.8. VHL gatekeeper’s pathway in renal epithelium. See text for details.

55

Although the loss of VHL-HIF-mediated regulation of E-cadherin likely provides an important

biological basis for the malignant nature of RCC, as well as the epithelial-to-mesenchymal

transition, there is one notable VHL-dependent event bearing on RCC progression to consider.

CXCR4 is a chemokine receptor that aids in the metastasis of tumour cells to organs abundant in

CXCR4-specific ligand, stromal cell-derived factor-1α (SDF-1α). Staller and colleagues

showed that the expression of chemokine receptor CXCR4 increases upon the loss of VHL,

suggesting a potential mechanism of RCC metastasis 172. In addition, Zagzag and colleagues

recently demonstrated that RCC and haemangioblastoma cells devoid of VHL overexpress not

only CXCR4, but also its ligand SDF-1α 173. These findings suggest that loss-of-function of

VHL can establish an autocrine signalling pathway providing selective survival advantage and

increased tendency for metastasis. The impact of the individual events (i.e., VHL-mediated E-

cadherin versus CXCR4/SDF-1α regulation) in RCC development/progression is not yet

established, but nevertheless remains an important question to address.

There are other examples of cancer-causing mutations (aside from inactivating mutations on

VHL) that often increase the expression of HIF-1α and provide mechanistic explanation for the

highly vascular tumours including RCC that develop in the absence of VHL mutations.

Mutations in TSC2 tumour suppressor gene increase the level of HIF-1α via the mammalian

target of Rapamycin (mTOR)-dependent and -independent mechanisms that may involve

chromatin remodelling 174. Loss of PTEN, which has been observed in the brain tumour

glioblastoma multiforme, results in increased HIF-1α levels via the activation of the Akt/protein

kinase B signalling cascade 175. The increased expression of HER2 receptor tyrosine kinase in

breast cancer and the loss of p53 in various tumours enhance HIF-1-dependent transcription,

often correlating with tumour aggressiveness 176,177. Although these examples support the notion

that there are multiple important regulators of HIF to ultimately promote oncogenic

transformation, whether non-VHL-associated HIF activation likewise results in the down-

regulation of E-cadherin via the activation of SIP1/Snail family of transcriptional repressors is an

important question that remains to be resolved.

56

Chapter3 NEDD8 defines tumour suppressor function of VHL

This work is now published: Ryan C. Russell and Michael Ohh. 2008. NEDD8 acts as a

‘molecular switch’ defining the functional selectivity of VHL. EMBO Reports 9(5):486-91

57

3.1 Rationale

Although the significance of the both the HIF and fibronectin functions of VHL in tumour

suppression has been well established, the mechanisms that determine the specific tumour

suppressive effects of VHL remain a mystery. It was recently shown that covalent modification

of VHL by NEDD8 is required for physical interaction with FN. Further, it was shown that

neddylation of VHL is not required for HIF function, providing the first hints at a molecular

determinant for the demarcation of these pathways. Here, we show that NEDD8 modification of

VHL acts as a ‘molecular switch’ where its covalent conjugation to VHL precludes ECV

function and concomitantly allows interaction with fibronectin, and thus providing the first

mechanistic step in the definition of functional selectivity of VHL.

3.2 Materials and Methods

3.2.1 Cells

786-O RCC, U2OS osteosarcoma and HEK293A embryonic kidney cell lines were obtained

from the American Type Culture Collection (Rockville, MD) and maintained in Dulbecco’s

modified Eagle’s medium supplemented with 10% heat-inactivated fetal bovine serum (Sigma,

Milwaukee, WI) at 37°C in a humidified 5% CO2 atmosphere. 786-O subclones ectopically

expressing wild-type VHL (786-WT) or empty plasmid (786-MOCK), RCC4 cells ectopically

expressing wild-type VHL (RCC4-WT) or empty plasmid (RCC4-MOCK), and ts41 Chinese

hamster ovary cells were as previously described69,71,178.

3.2.2 Antibodies and reagents

Monoclonal anti-haemagglutinin (HA) (12CA5) and anti-HIF1α antibodies were obtained from

58

Boehringer Ingelheim (Laval, QC) and Novus Biological (Littleton, CO), respectively.

Monoclonal anti-T7 antibody was obtained from Novagen (Madison, WI). Monoclonal anti-

vinculin, tubulin, and hnRNP were obtained from Abcam (Cambridge, MA). Polyclonal anti-

GLUT1 and anti-Cul2 antibodies were obtained from Alpha Diagnostics (San Antonio, TX) and

Zymed (San Francisco, CA), respectively. Monoclonal anti-VHL antibody (IG32) was as

previously described71. Polyclonal anti-Col IV, anti-luminal Calnexin, and anti-cyto Calnexin

were obtained from Abcam (Cambridge, MA). MG132 and NEDP1 were obtained from Boston

Biochem (Boston, MA). N-ethyl maleimide (NEM), cobalt chloride (CoCl2) and desferroxamine

(DFO) were obtained from Sigma (Oakville, ON). Trypsin was obtained from Invitrogen

(Burlington, ON).

3.2.3 Plasmids

Mammalian expression plasmids pRc-CMV-HA-VHL(WT), pRc-CMV-HA-VHL(C162F), pRc-

CMV-HA-VHL(RRR), pRc-CMV-HA-Cul2, and pRc-CMV-T7-VHL were described

previously71,78,144,179,180. pcDNA3-NEDD8 was generated by PCR from a human fetal brain

library using primers 5-ATGGATCCATGCTAATTAAAGTGAAGACGCTGAC-3 and 5-

TGAATTCGCTGCCTAAGACCACCTCCT-3. The PCR product was then ligated into the

BamHI and EcoRI sites in pcDNA3(-). All plasmids were confirmed by direct DNA sequencing.

3.2.4 Immunoprecipitation and immunoblotting

Immunoprecipitation and Western blotting were performed as described previously71. In brief,

cells were lysed in EBC buffer (50 mM Tris [pH 8.0], 120 mM NaCl, 0.5% NP-40)

supplemented with a cocktail of protease and phosphatase inhibitors (Roche, Laval, Canada).

Immunoprecipitates immobilized on protein A-Sepharose beads (Amersham Biosciences,

Piscataway, NJ) were washed five times with NETN buffer (20 mM Tris [pH 8.0], 120 mM

NaCl, 1 mM ETDA, 0.5% NP-40), eluted by boiling in sodium dodecyl sulfate (SDS)-containing

sample buffer, and size-fractionated by SDS-polyacrylamide gel electrophoresis (PAGE).

Resolved proteins were then electro-transferred onto PVDF membrane (Bio-Rad Laboratories,

59

Hercules, CA), immunoblotted with the various antibodies, and visualized by

chemiluminescence (Amersham Biosciences, Piscataway, NJ).

3.2.5 Affinity Purification

Gelatin-Sepherose beads (Amersham Pharmaceuticals, Piscataway, NJ) were used to affinity

purify FN from whole cell extracts by rocking at 4°C for 3 hours. FN complexes were eluted in

250mM Arginine in PBS, rocking for 10 min at 22°C, as previously described181.

3.2.6 Metabolic labeling

Metabolic labelling was performed as described previously. In brief, 786-O cells were

maintained in methionine-free Dulbecco's modified Eagle's medium for 45 min then

supplemented with 35S-methionine (100 µCi/ml of medium; Amersham Biosciences,

Buckinghamshire, United Kingdom) and 2% dialyzed fetal bovine serum for 3 h at 37 °C in a

humidified 5% CO2 atmosphere.

3.2.7 Subcellular fractionation

Cells were resuspended in Buffer A (10 mM HEPES [pH 7.9], 10 mM KCl, 1.5 mM MgCl2, 0.34

M sucrose, 10% glycerol) supplemented with protease and phosphatase inhibitors (Roche, Laval,

Canada) and 1mM DTT and subsequently lysed with 0.1% Triton X-100. Samples were

incubated 7 min on ice and centrifuged. The supernatant was recovered (cytoplasmic fraction),

and the pellet was washed with Buffer A and resuspended in Buffer B (0.2 mM EGTA [pH 8], 3

mM EDTA [pH 8]) supplemented with protease and phosphatase inhibitors (Roche, Laval,

Canada) and 1 mM DTT. After a 30 min incubation period, samples were centrifuged and the

resulting supernatant isolated (nuclear fraction).

60

3.2.8 Confocal microscopy

Confocal images were acquired on a Zeiss LSM 510 Meta Laser Scanning confocal system (Carl

Zeiss, Thornwood, NJ, USA) with a 100× Plan-Apochromat 1.4 NA oil-immersion objective.

3.2.9 siRNA

SMARTpool Cul2-specific siRNA was obtained from Dharmacon (Chicago, IL) using

Oligofectamine (Invitrogen, Burlington, ON) or EXTREMEGENE (Roche, Laval, QU).

3.3 RESULTS AND DISCUSSION

3.3.1 ECV- and FN-associated functions of VHL are mutually exclusive

The ability of VHL to regulate HIF is dependent on its ability to form a functional ECV.

Therefore, we asked whether the ability of VHL to bind FN was likewise dependent of ECV.

VHL-associated FN was immunoprecipitated following siRNA-mediated knockdown of Cul2, a

scaffold component on ECV, in 35S-radiolabelled VHL-null RCC4 renal carcinoma cells

ectopically expressing HA-VHL(WT) or empty plasmid (MOCK). The level of FN co-

precipitating with VHL did not diminish despite marked reduction in Cul2 expression (Fig. 3.1

and Fig. 3.1d). As expected, siRNA-mediated knockdown of Cul2 attenuated VHL-dependent

ubiquitylation of HIF1αODD (data not shown). Furthermore, biotinylated HIF1αODD-OH

peptides co-precipitated ECV components without the presence of FN as compared to FN co-

precipitated from 35S-radiolabelled 786-VHL RCC cells using anti-HA antibody directed against

HA-VHL (Fig. 3.1c, compare lanes 2 and 3). Although formally possible that HIF1αODD-OH

peptides could have displaced FN via competition for VHL, hypoxia or hypoxia mimetic

(desferroxamine or CoCl2) treatment of 786-VHL cells did not increase VHL/FN interaction

(Fig. 1e). These results argue against the notion that VHL binding to FN is influenced by

competition with HIFα, and that ECV complex is not necessarily required for VHL to bind FN.

The identity of the high molecular weight protein co-precipitating with VHL was confirmed as

FN via immunoblotting and subsequent autoradiography of the same PVDF membrane (Fig. 3f).

FN is known to bind ColIV. However, affinity purified FN complex containing VHL generated

61

from 786-WT cells failed to show the presence of ColIV (data not shown), suggesting that

VHL/FN interaction is independent of ColIV.

62

63

FN

Cul2

HA-VHL

Elongin BElongin C

KDa

175

86

47

26

19

Anti-HA:IP/AR

AR: :IB

1 2

1 2 1 2

E

Anti-HA

Anti-FN

Anti-Vinculin

Anti-GLUT1

WCE1 2 3 4 5

IB:

INPUT

AffPD6 7 8 9

Anti-HA

Anti-FN

IB:

F

Figure 3.1. ECV- and FN-associated functions of VHL are mutually exclusive.

(A) HEK293A cells were treated with increasing amounts (10-100nM) of Cul2 siRNA (lanes 2-

4) or scrambled siRNA (lanes 5-7) or transfect reagent alone (MOCK; lane 1). Equalized whole

cell lysates were resolved on SDS-PAGE and immunoblotted with anti-Cul2 (lower panel) or

anti-vinculin (upper panel) antibodies. (B) RCC4 cells stably transfected with HA-VHL or empty

plasmid (MOCK) were radiolabelled (metabolically labelled with 35S-methionine). Cells were

treated with (+) or without (-) Cul2 siRNA as indicated. Cell lysates were immunoprecipitated

with anti-HA antibody and the resolved proteins were visualized by autoradiography. (C) 786-

64

WT and 786-MOCK cells were metabolically labelled with 35S-methionine, lysed and

immunoprecipitated with an anti-HA antibody (lanes 1 and 2) or pulled-down with synthetic

HIF1α[ODD]-OH peptides (lanes 3 and 4). Bound proteins were resolved and visualized by

autoradiography. (D) Extracts prepared from RCC4-VHL or RCC4-MOCK cells treated with

Cul2-specific or scrambled siRNA in the presence of MG132 were immunoprecipitated with

anti-HA antibody and the resolved proteins were visualized by immunoblotting with the

indicated antibodies. (E) 786-MOCK and 786-WT cells were radiolabelled with 35S-Methionine

and lysates were immunoprecipitated with anti-HA antibody. Bound proteins were resolved by

SDS-PAGE, transferred to PVDF membrane and visualized by autoradiography (left panels).

The upper portion of the membrane (above 130KDa) was also immunoblotted with an anti-FN

antibody (right panel). (F) Equal amounts of whole cell extracts generated from 786-WT cells

treated with hypoxia mimetics DFO and CoCl2 or hypoxic conditions (1% Oxygen) were

resolved on SDS-PAGE and immunoblotted with the indicated antibodies (left panel).

Fibronectin complexes were affinity purified from the indicated whole cell extracts, bound

proteins resolved and immunoblotted with anti-FN (right, top panel) or anti-HA (right, bottom

panel) antibody. AffPd: FN affinity pull-down; IB: immunoblot; IP: immunoprecipitation; AR:

autoradiography; WCE: whole cell extract * denotes non-specific protein bands.

65

3.3.2 Disruption of NEDD8 pathway abrogates FN binding to VHL, but

not ECV formation

Recently, we have shown that mutations in VHL that disrupt NEDD8 conjugation lead to a

failure in binding FN78. To further address whether VHL binding to FN is dependent on the

NEDD8 pathway independent of ECV complex formation, Chinese hamster ovary (CHO) ts41

cells with a temperature-sensitive APP-BP1 (a component of NEDD8-activating enzyme;

NAE)178 were transiently transfected with plasmids encoding HA-VHL and GFP-FN. VHL

binding to FN was dramatically decreased in cells maintained under restrictive temperature as

compared to cells under permissive temperature (Fig. 3.2a), suggesting that an intact NEDD8

pathway is critical for promoting VHL binding to FN. As expected, neddylation of VHL and

Cul2 was curtailed under non-permissive temperature (Fig. 3.2b and c). Importantly, an intact

ECV capable of binding HIF1α was observed under both temperatures conditions (Fig. 3.2c),

suggesting that the ability of VHL to form an ECV is not sufficient for binding FN. Furthermore,

non-neddylatable VHL(RRR)78, while showing similar subcellular distribution pattern as

VHL(WT) has compromised ability to bind FN78 (data not shown). These results suggest that

neddylation of VHL does not promote FN binding by altering the subcellular localization of

VHL, which binds to the cytosol-exposed region of FN in ER/Golgi (data not shown).

Neddylated substrates are often de-conjugated post-lysis making detection of the neddylated

species technically challenging. Therefore, we performed our purification of the VHL/FN

complex in the presence of a deneddylase inhibitor NEM and show that VHL in complex with

FN is exclusively unmodified VHL (Fig. 3.2e). As a control, neddylated Cul2 was preserved in

the presence of NEM, even in the presence of a well-established purified deneddylase NEDP1

(Fig. 3.2d). The exclusive presence of unneddylated VHL in complex with FN suggests that the

neddylation of VHL is an intermediary step, which is proceeded by dennedylation of VHL

allowing unhindered association with FN.

66

67

Figure 3.2. Restriction of a dynamic NEDD8 pathway results in the attenuation of VHL

binding to FN. (A) Ts41 cells were transfected with plasmids encoding HA-VHL(WT) and

GFP-FN. Cells were grown at permissive or non-permissive temperature for 15 h and

radiolabelled. Cells lysates were immunoprecipitated with an anti-HA antibody and resolved

proteins were visualized by autoradiography. (B) Ts41 CHO cells were transfected with plasmids

encoding HA-VHL(WT) and NEDD8. Cells were grown at permissive (P; 33°C) or non-

permissive (NP; 39°C) temperature for 15 h, lysed, immunoprecipitated and immunoblotted with

an anti-HA antibody. (C) ts41 cells were transfected with plasmids encoding HA-VHL and

HIF1α. Cells were grown at permissive or non-permissive temperature for 15 h, then lysed, and

immunoprecipitated with anti-HIF1α antibody. Resolved proteins were visualized by

immunoblotting with anti-HIF1α (upper panel), anti-Cul2 (middle panel) or anti-HA (lower

panel) antibodies. (D) 786-WT cellular extracts were incubated with or without a purified

68

deneddylase NEDP1 in the presence or absence of a deneddylase inhibitor NEM for 30min at

37°C. Reaction mixtures were then resolved on SDS-PAGE and visualized by immunoblotting

with the indicated antibodies. (E) FN complexes were affinity purified from the indicated whole

cell extracts in the presence of excess NEM. Bound proteins were competitively eluted and

immunoprecipitated with anti-HA antibody and immunoblotted with anti-FN or anti-HA

antibody (lanes 1-3, top and bottom panel, respectively). Equal amounts of whole cell extracts

were also immunoblotted with anti-FN or anti-vinculin antibody (lanes 4-6, top and bottom

panel, respectively).* denotes uncharacterized protein band; IP: immunoprecipitation; IB:

immunoblot; AR: autoradiography; AffPd: FN affinity pull-down; Open arrow denotes predicted

molecular weight of NEDD8-conjugated HA-VHL.

69

3.3.3 Neddylation of VHL prevents ECV complex formation via steric

hindrance

VHL contains two major functional domains; α and β48. The β domain is required for binding

substrates48,179 and the α domain is required for binding Elongin C48, which serves as a bridge

connecting VHL to the rest of the ECV components. Residues spanning 158-172 (Elongin B/C-

box) within the α domain has been shown to be necessary and sufficient for binding Elongin C182

and K159 has been shown to be the major acceptor site of NEDD878. Structurally and

functionally, ECV is analogous to SCF. Although ECV has not been crystallized, SCF183 and the

VHL/Elongins B/C (VBC) complex48,184 have been solved. To determine possible effects of

NEDD8 conjugation to VHL, we superimposed the VBC complex against SCF. In particular,

Skp1 and its orthologue Elongin C polypeptide backbones were aligned within 1.3Å, giving

confidence that Cul1 would be positioned similarly to Cul2 in the context of ECV. Based on the

composite VBC-Cul1 structure, NEDD8 conjugation of VHL at K159 would create significant

steric hindrance that would prohibit the incorporation of Cul2 or possibly Elongin C to VHL

(Fig. 3.3a and b). Based on this prediction, a VHL mutant incapable of binding Elongin C due to

a mutation within the Elongin B/C-box (excluding K159 and K171) would be more accessible

for NEDD8 modification. HEK293A cells were transfected with plasmids encoding T7-NEDD8

in combination with plasmids encoding HA-VHL(WT), non-neddylatable HA-VHL(RRR), and

HA-VHL(C162F), a well-established α domain mutant incapable of binding Elongin C179,182. As

expected, HA-VHL(WT) generated a slower migrating T7-NEDD8-conjugated HA-VHL, while

HA-VHL(RRR) failed to generate a NEDD8-modified isoform (Fig. 3.3c). Consistent with the

steric hindrance model, HA-VHL(C162F) was neddylated to a greater extend in comparison to

VHL(WT) (Fig. 3.3c). Moreover, while the neddylated VHL comprises a minor fraction of total

VHL, significantly less neddylated VHL was found in complex with Cul2 (Fig. 3.3d, compare

lanes 1 and 2), suggesting an exclusion of neddylated VHL in ECV complex. These results

strongly suggest that neddylation of VHL generates a steric clash preventing its association into

the ECV complex.

70

Figure 3.3. NEDD8 modification of VHL generates steric hindrance blocking the formation

of ECV. (A) VBC (VHL/Elongins B/C) crystallized with HIF1αODD peptide was visualized

using DeepView/Swiss-PdbViewer v3.7. Complex was viewed with side chains and showing van

der Waals forces. Lysine 159, the primary site of neddylation, has been highlighted. (B) The

backbone of Elongin C in the VBC (1LM8.pdb) was overlaid with the backbone of Elongin C-

orthologue Skp1 in the SCF complex (1LDK.pdb). Using the iterative Magic Fit function of

71

DeepView/Swiss-PdbViewer, a fit was generated with an overlap consisting of 99 residues

between Elongin C and Skp1 with a RMS of 1.26 Å. (C) U2OS cells were transfected with

plasmids encoding HA-VHL(WT), HA-VHL(RRR), HA-VHL(C162F), T7-NEDD8, or empty

vector (MOCK). Cells were then lysed, immunoprecipitated and immunoblotted with anti-HA

antibody. (D) HEK293A cells were transfected with the indicated combination of plasmids

encoding HA-Cul2, T7-VHL, and NEDD8. Immunoprecipitation with anti-HA (lanes 2 and 3) or

anti-T7 (lanes 1 and 4) antibodies were performed on pooled lysates. Resolved proteins were

immunoblotted with an anti-Cul2 (top panel), anti-T7 (middle panel), or anti-VHL (bottom

panel) antibody. A long exposure of anti-T7 immunoblot was taken to better visualize neddylated

VHL. * denotes uncharacterized protein band; IP: immunoprecipitation; IB: immunoblot.

72

3.3.4 Cul2 is excluded from the VHL/FN complex

The NEDD8-induced steric hindrance model would predict an exclusion of one or more ECV

components, which may be necessary for the promotion of FN-mediated function. To directly

determine whether ECV components are excluded from the VHL-FN complex, we performed

affinity purification of intracellular FN from 786-MOCK, 786-WT and 786-C162F cells. FN-

containing complexes were then competitively eluted from the sepharose beads and

immunoprecipitated with an anti-HA antibody selecting for the FN complexes associated with

HA-VHL. As expected, HA-VHL(WT) was present in the affinity purified FN complex and co-

precipitated FN, while a diseasing-causing HA-VHL(C162F) mutant, which has an intrinsic

defect in FN binding78, was absent in the affinity purified FN complex (Fig. 3.4a, lanes 4 and 5).

Equal amounts of whole cell extracts were separated on SDS-PAGE and immunoblotted for total

FN and HA-VHL, which indicated the presence of FN in all of the indicated cell types (Fig. 3.4a,

lanes 1-3). Notably, FN is known to bind ColIV, which has been shown recently to interact with

VHL. However, the affinity purified FN co-precipitated via HA-VHL did not contain ColIV

(data not shown), suggesting that VHL binds FN independently of ColIV. Next, anti-HA

immunoprecipitations were performed on the whole cell extracts or affinity purified intracellular

FN complexes generated from 786-MOCK, 786-WT and 786-C162F cells (Fig. 3.4b). While,

HA-VHL(WT) co-precipitated Cul2 from the whole cell extracts as expected, HA-VHL(WT) in

the FN complex did not co-precipitate Cul2 (Fig. 3.4b, compare lanes 2 and 5). In parallel, an

anti-HA immunoprecipitation from the affinity purified FN complex from 35S-radiolabelled 786-

VHL cells showed an absence of Cul2, but a clear presence of Elongins B and C, in the HA-

VHL/FN complex (Fig. 3.4c). These results demonstrate that Cul2 is excluded from the VHL/FN

complex. NEDD8 conjugation to VHL precludes VHL from entering the ECV; we therefore

reasoned that additional means of removing Cul2 might rescue the FN binding of non-

neddylatable VHL. Immunoprecipitated HA-VHL was washed under high-salt/detergent buffer

condition and then mixed with whole cell extracts prepared from 786-MOCK cells radiolabelled

with 35S-Methionine. Post-lysis complexes were then washed, resolved by SDS-PAGE and

visualised by audioradiography (Fig. 3.4d, right panel vs. left panel). Notably, the exclusive

presence of unneddylated VHL in complex with FN (Fig. 3.4, Fig. 3.2e) also suggests that the

neddylation of VHL represents an intermediary step that prohibits Cul2 engagement, which is

proceeded by deneddylation of VHL allowing unhindered association with FN.

73

FN

HA-VHL

A

1 2 3 4 5 6

INPUT

AffPD

Anti-FN

IB:

Anti-HA

Anti-HA:IP

WCE

B

AffPD WCE

HA-VHLAnti-HA

Anti-Cul2 Cul2

1 2 3 4 5 6

IB:

Anti-HA:IP

IgG light

Cul2

HA-VHL

Elongin B

AffPD WCE1 2 3 4 5 6

Elongin C

*

Anti-HA:IP/AR

C

FN

5 642 31WCE-Anti-HA:IB

WASH-Anti-HA:IP/AR

HA-VHL

*

INPUT

FN*

HA-VHL

Anti-HA:IP/AR

Anti-HA:IP/AR

D

Figure 3.4. VHL/FN complex excludes ECV component Cul2. (A) Equal amounts of whole

cell extracts generated from 786-MOCK, 786-WT, and 786-C162F cells were resolved and

immunoblotted with anti-FN (top panel) or anti-HA (bottom panel) antibodies (lanes 1-3). FN

complexes were affinity purified from the indicated whole cell extracts, bound proteins

competitively eluted and immunoprecipitated with anti-HA antibody (lanes 4-6). Resolved

proteins were immunoblotted with anti-FN (top panel), or anti-HA (bottom panel) antibody. (B)

HA-VHL was immunoprecipitated with anti-HA antibody from either purified FN complexes

(lanes 1-3) or whole cell extract (lanes 4-6) generated from pooled cell lysate from the indicated

74

cell lines. Bound proteins were resolved and immunoblotted with anti-Cul2 (top panel) or anti-

HA (bottom panel) antibody. (C) Cells were radiolabelled and prepared as in B. Resolved

proteins were visualized by autoradiography. (D) Left: 786-O cells stably expressing the

indicated VHL mutants were metabolically labelled with 35S-Methionine, lysed and

immunoprecipitated with anti-HA antibody. Bound proteins were resolved on SDS-PAGE and

visualized by autoradiography. Right: HA-VHL was immunoprecipitated from the indicated cell

lysates and washed under high-salt/detergent buffer condition and equilibrated with PBS. HA-

VHL-bound beads were then mixed with whole cell extracts prepared from 786-MOCK cells

radiolabelled with 35S-Methionine. Post-lysis complexes were then washed, resolved by SDS-

PAGE and visualized by autoradiography (top panel). Whole cell extracts of the indicated cells

were also resolved on SDS-PAGE and immunoblotted using anti-HA antibody (bottom panel).

IP: immunoprecipitation; IB: immunoblot; AR: autoradiography.

75

3.3.5 Discussion

In keeping with the prediction based on the composite VBC-Cul1 structure (see Fig. 3), NEDD8

modification of VHL prevents Cul2 engagement and thus excluded from the ECV complex. This

‘freed’, albeit, minor pool of VHL binds FN, representing a requisite step in the eventual

assembly of the extracellular FN matrix. VHL in complex with FN is unmodified, which

suggests that VHL is transiently modified by a dynamic neddylation and deneddylation process.

In support, the inhibition of NEDD8 pathway or the ablation of the NEDD8-conjugation sites on

VHL markedly attenuated the ability of VHL to interact with FN while preserving ECV

integrity. The requirement of this dynamic process also explains why the non-neddylatable

VHL(RRR) mutant is defective in FN binding and assembly.

The preclusion of Cul2 in the VHL/FN complex also suggests that the physical presence of Cul2

may be inhibitory in the engagement of FN to VHL. In support of this notion, a near-complete

knockdown of Cul2 increased the amount of FN-bound to VHL (Fig 1d). In a complementary

experiment, HA-VHL from 786-WT and 786-RRR cells was immunoprecipitated and washed

under higher stringency salt and detergent conditions to strip away VHL-associated proteins. The

‘stripped’ VHL was then mixed with radiolabelled VHL-null 786-O cell lysates and re-

immunoprecipitated. Under such condition, VHL(RRR)’s ability to bind de novo FN was

restored to a level comparable to VHL(WT). VHL(C162F) was still incapable of binding FN

(Fig 4d), consistent with the notion that Elongins B/C are required for VHL/FN interaction.

Moreover, a direct interaction between FN and VHL was shown to not require Cul2 as VBC

complex lacking Cul2 was sufficient to bind FN70. These results suggest that the deficiency of

non-neddylatable VHL lies in the inability to disengage Cul2 or Cul2-associated inhibitory

factor(s) in the absence of dynamic NEDD8 processing. Elongins B/C are likely providing

stability to the unstable tertiary structure of VHL48,185 within the VHL/FN complex. The

structural requirement provided by Elongins B/C perhaps explains why α domain VHL mutants

including C162F fail to bind FN. In this regard, analogous to HIFα, FN binding by VHL requires

both direct physical interaction and the association of Elongins.

VHL plays a critical role in ECV-mediated destruction of HIFα and the assembly of FN ECM.

Neddylation of VHL prohibits the engagement of Cul2 and concomitantly activates the

association with FN. Thus, NEDD8 acts as a molecular switch that defines the functional

76

selectivity of VHL and provides the first mechanistic demarcation of the HIF-dependent and

HIF-independent pathways.

77

Chapter 4 VHL/SOCS1 Heterocomplex Degrades JAK2

Ryan C. Russell*, Roxana I. Sufan*, Olga Roche, Terri D. Richmond, Dwayne L. Barber,

Meredith S. Irwin, and Michael Ohh. VHL and SOCS1 cooperate to degrade JAK2:

implications for polycythemia. In preparation.

*Authors contributed equally to this work.

78

4.1 Rationale

Mutations of VHL leading to CP cause a disease with features of both primary and secondary

polycythemia. CP patients and the CP mouse have high EPO levels and an intrinsic

hypersensitivity to EPO 40,142. Secondary polycythemic features have been explained by a

diminished capacity of CP-VHL(R200W) to bind HIFα resulting in mild HIFα stabilization and

elevation of EPO levels 40. However, HIFα stabilization has not been associated with

hypersensitivity of erythroid progenitors to EPO. Therefore, we biochemically characterized

mutants that give rise to autosomal recessive polycythemia in order to address the molecular

mechanism underlying primary polycythemic features of CP. Here, we reveal that the wild-type

VHL and tumor-causing VHL mutants form a complex with SOCS1 to target phosphorylated

JAK2 for ubiquitin-mediated destruction. We further show that a select cluster of VHL mutants

including CP-VHL(R200W and H191D) mutants form a defective heterodimer with SOCS1,

severely compromising JAK2 degradation and consequently enhancing JAK2-STAT5 signalling

pathway. These findings provide the mechanism underlying primary polycythemic features of

CP and introduce VHL as a novel regulator of JAK2.

79

4.2 Materials and Methods

4.2.1 Cells.

786-O RCC and HEK293A cells were obtained from the American Type Culture Collection

(Rockville, MD, USA) and maintained in Dulbecco’s modified Eagle’s medium supplemented

with 10% heat-inactivated fetal bovine serum (FBS) (Sigma, Milwaukee, WI, USA) at 37°C in a

humidified 5% CO2 atmosphere. 786-O subclones ectopically expressing HA-VHL(WT), HA-

VHL(C162F) or empty plasmid were previously described 38. 786-O subclones ectopically

expressing HA-VHL(R200W) and HA-VHL(H191D) were generated as previously described 38.

Ba/F3 pro B cells were obtained from the American Type Culture Collection (Rockville, MD,

USA) and maintained in RPMI 1640 supplemented with 10% FBS and 0.5 U/ml recombinant

human EPO (Janssen Ortho, Toronto, ON, Canada).

4.2.2 Antibodies.

Rabbit antibodies against JAK2, VHL, pJAK2, and pSTAT5 were obtained from Cell Signalling

Technologies (Danvers, MA, USA). Polyclonal antibodies against ubiquitin, Elongin B and

HIF2α antibodies were obtained from DAKO Canada (Mississauga, ON, Canada), Santa Cruz

Biotechnology (Santa Cruz, CA, USA) and Novus Biologicals (Littleton, CO, USA),

respectively. Monoclonal antibodies against HA (12CA5), T7 and VHL(IG32) were obtained

from Boehringer Ingelheim (Laval, QC, Canada), Novagen (Madison, WI, USA) and BD

Biosciences (Mississauga, ON, Canada), respectively. Monoclonal anti-α-tubulin antibody was

obtained from Abcam (Cambridge, MA, USA). Polyclonal anti-Cul2 and anti-SOCS1 antibodies

were obtained from Invitrogen (Burlington, ON, Canada) and Novus Biologicals (Littleton, CO,

USA), respectively. MG132 proteasome inhibitor was obtained from Boston Biochem

(Cambridge, MA, USA).

80

4.2.3 Plasmids.

Plasmid encoding HA-SOCS1 was generously provided by Dr. Robert Rottapel (Ontario Cancer

Institute, Toronto, ON, Canada). T7-VHL and HA-VHL(WT, R64P, V74G, Y98H, S111H,

Y112H, Y112N, F119S, L128F, L158S, K159E, C162F, L188V) were previously described 38,78,186,187. HA-VHL(R200W, H191D) were generated using QuikChange site-directed

mutagenesis kit (Stratagene, La Jolla, CA, USA) and mutations verified by direct DNA

sequencing.

4.2.4 Immunoprecipitation and immunoblotting.

Immunoprecipitation and Western blotting were performed as described previously 188. In brief,

cells were lysed in EBC buffer (50 mM Tris, pH 8.0; 120 mM NaCl; and 0.5% NP-40)

supplemented with protease and phosphatase inhibitors (Roche, Laval, Canada). Cell lysates

were immunoprecipitated with indicated antibodies in the presence of Protein-A agarose beads

(Waltham, MA, USA). Bound proteins were washed five times with NETN buffer (20 mM Tris,

pH 8.0; 120 mM NaCl; 1 mM EDTA; and 0.5% NP-40), eluted by boiling in sodium dodecyl

sulfate (SDS)-containing sample buffer, and resolved by SDS polyacrylamide gel electrophoresis

(PAGE).

4.2.5 Metabolic labeling.

Metabolic labeling was performed as described previously 188. In brief, 786-O cells were

maintained in methionine-free Dulbecco's modified Eagle's medium for 45 min then

supplemented with 35S-methionine (100 µCi/ml of medium; Amersham Biosciences,

Buckinghamshire, United Kingdom) and 2% dialyzed fetal bovine serum for 3 h at 37 °C in a

humidified 5% CO2 atmosphere.

81

4.2.6 In vitro ubiquitylation assay.

T7-JAK2 and T7-pJAK2 were purified on Protein-A agarose beads (Waltham, MA, USA) with

anti-T7 antibody from HEK293 cells transfected with T7-JAK2 and EPOR stimulated with or

without EPO. Ubiquitylation reaction was then performed as described previously 189 on JAK2 or

pJAK2 bound on beads.

4.2.7 Generation of phenylhydrazine-primed splenic erythroblasts.

Mice were injected intraperitoneally with 50 mg/kg phenylhydrazine hydrochloride (Sigma-

Aldrich, Oakville, ON, Canada) in PBS on days 1 and 2, as previously described 142. Mice were

sacrificed and spleens removed on day 4 under sterile conditions. Single-cell suspensions were

generated using a 70-µm cell strainer for further analysis.

4.2.8 Cytokine deprivation and stimulation of murine splenic erythroblasts.

Cells were washed twice in PBS, starved in α-MEM supplemented with 2% FCS for 4 hrs at

37°C, and then stimulated with various concentrations of EPO for 15 min at 37°C. Cells were

pelleted at 6000 rpm for 1 min and lysed in 1% Triton X-100 lysis buffer supplemented with

phosphatase inhibitors (Laval, Canada) supplemented with 20mM Na3P2O2, 10mM NaF and

1mM Na3VO4.

82

4.3 RESULTS

4.3.1 CP-VHL mutants have reduced capacity to form ECV

VHL(R200W and H191D) mutants showed diminished association with Elongins B/C and Cul2,

the core components of ECV, when expressed in human embryonic kidney epithelial cells

HEK293 or RCC 786-O(VHL-/-) cells (Fig. 4.1a, b and c). Tumor-associated VHL(C162F)

mutant, which is known to be defective in forming an ECV complex 190, served as control. Thus,

in addition to the previously reported defect in HIFα binding, CP-VHL mutants are

compromised in ECV assembly, which is also likely to contribute to HIFα stabilization.

Furthermore, all tumour-associated VHL mutants tested-to-date have invariably shown a failure

in binding to FN and formation of FN fibrillar array in the extracellular space 188. In contrast,

CP-VHL mutants, but as expected not VHL(C162F), showed intact interaction with FN and

robust extracellular FN matrix deposition (Fig. 4.1d). Thus, VHL(R200W) and VHL(H191D)

are the first naturally occurring VHL mutants exhibiting proper FN matrix deposition, which is

consistent with the absence of cancer predisposition in individuals with CP.

83

84

Figure 4.1. CP-VHL exhibits altered binding to ECV components and JAK2.

(A) HEK293 cells transfected with the indicated plasmids were lysed, immunoprecipitated with

anti-HA antibody and immunoblotted with indicated antibodies. (B,C) 35S-radiolabelled 786-O

subclones stably expressing indicated HA-VHL were immunoprecipitated with anti-HA

antibody, resolved by SDS-PAGE and visualized by autoradiography. (D) 786-O subclones

stably expressing the indicated HA-VHL were grown on glass coverslips and immunostained for

FN (red) and visualized by fluorescent microscopy. DAPI (blue) staining indicates nuclei.

(E) HEK293 cells transfected with the indicated combination of plasmids were treated with (+)

or without (-) MG132. Equal amounts of cell lysates were immunoprecipitated with anti-VHL

antibody and immunoblotted with the indicated antibodies. (F) HEK293 cells transfected with

the indicated plasmids were lysed in the absence of MG132, immunoprecipitated with anti-HA

antibody and immunoblotted with the indicated antibodies. WCE: whole cell extract; IP:

immunoprecipitation; IB: immunoblot; AR: autoradiography. Asterisk denotes non-specific

protein bands.

85

4.3.2 VHL binds JAK2 in a proteasome-sensitive manner

In addition to reduced Cul2 binding, 35S-metabolic labelling of 786-O cells stably expressing

VHL(R200W or H191D) revealed an associated protein of 120kDa in the absence of proteasome

inhibitor (Fig. 4.1c). JAK2 is approximately 120KDa and aberrant JAK2-STAT5 signalling has

been reported to cause hypersensitivity in BFU-E cells to EPO in PV patients 191. Similarly,

BFU-E cells of CP patients are also hypersensitive to EPO 40 and thus, we asked whether VHL

interacts with JAK2. HEK293 cells transfected with plasmids encoding HA-VHL(WT) and T7-

JAK2 were treated with or without proteasome inhibitor MG132 and immunoprecipitated with

anti-VHL antibody. HA-VHL(WT) co-precipitated JAK2 preferentially in the presence of

MG132 (Fig. 4.1e). HA-VHL(R200W) and HA-VHL(H191D) showed increased association

with JAK2 in comparison to VHL(WT) in the absence of MG132 (Fig. 4.1f). These results

identify JAK2 as a novel substrate of VHL and suggest that CP-VHL mutants have a diminished

capacity to promote proteasome-dependent degradation of JAK2.

4.3.3 VHL promotes ubiquitin-mediated degradation of pJAK2

The level of total JAK2 remained unaffected by ectopic expression of VHL (Fig. 4.1e and f,

bottom panels) suggesting that VHL promotes diminution of a select population of JAK2 upon

engagement. We asked whether VHL promoted degradation of activated JAK2, which is defined

by Y1007/1008 phosphorylation 192. Introduction of HA-VHL(WT) in HEK293 cells resulted in

a dramatic loss of phosphorylated JAK2 (pJAK2) (Fig. 4.2a). We next asked whether the loss of

pJAK2 was due to VHL-mediated ubiquitylation of pJAK2. HEK293 cells were co-transfected

with plasmids encoding T7-JAK2 and EPOR and stimulated with EPO (20U/ml) for 15 min to

generate robust levels of pJAK2 (Fig. 4.2b, left panel), which was subsequently isolated via anti-

T7 immunoprecipitation. The enriched T7-pJAK2 was then subjected to an in vitro

ubiquitylation reaction using S100 extracts devoid of or reconstituted with VHL(WT) (Fig. 4.2b,

right panel). While the total JAK2 levels were unaffected, the level of pJAK2 decreased

dramatically in the presence of VHL(WT), which was accompanied by the appearance of pJAK2

polyubiquitylation (Fig. 4.2b, lane 4). Notably, the low level of VHL-dependent ubiquitylation

observed in the absence of EPO is likely due to limited spontaneous JAK2 autophosphorylation,

86

commonly observed upon ectopic JAK2 expression (Fig. 4.2b, right panel, lane 3). These results

demonstrate that VHL promotes pJAK2 ubiquitylation.

We next investigated the effect of CP-VHL mutants on pJAK2 stability and observed that while

VHL(WT) co-precipitated negligible levels of pJAK2 in the absence of MG132, both CP-VHL

mutants R200W and H191D co-precipitated higher levels of pJAK2, supporting the notion that

CP-VHL mutants have a diminished capacity to promote pJAK2 degradation (Fig. 4.2c).

VHL patients rarely develop polycythemia despite harboring VHL mutations that abolish HIFα

degradation 142. Thus, tumor-causing VHL mutants incapable of binding or ubiquitylating HIFα

are predicted to retain the ability to promote ubiquitin-mediated destruction of pJAK2. A panel

of VHL substitution mutants spanning the open reading frame were tested for their ability to

degrade pJAK2. Consistent with our prediction, expression of VHL mutants, with the exception

of F119S and L128F (discussed below), resulted in negligible levels of pJAK2 in the absence of

MG132 (Fig. 4.2d). In addition, a panel of tumor-causing VHL mutants retained binding to

JAK2 in the presence of MG132 (Fig. 4.2e). Notably, well established α domain VHL mutants

C162F and L158S, which cannot form an ECV 186,190, decreased pJAK2 levels comparable to

that of VHL(WT). These results infer that a novel, ECV-independent, mechanism is responsible

for VHL-mediated pJAK2 degradation.

87

88

Figure 4.2. VHL promotes ubiquitin-mediated destruction of pJAK2.

(A) HEK293 cells transfected with the indicated plasmids were lysed and immunoblotted with

the indicated antibodies. (B) HEK293 cells transfected with the indicated plasmids were treated

with (+) or without (-) EPO and pJAK2 was isolated via anti-T7 immunoprecipitation (left

panel), which was then added to an in vitro ubiquitylation reaction containing proteasome-

depleted S100 fractions containing (+) or not containing (-) VHL (right panels). Reaction

mixtures were then re-immunoprecipitated with anti-T7 antibody, resolved by SDS-PAGE and

immunoblotted with the indicated antibodies. (C) HEK293 cells transfected with the indicated

plasmids were lysed in the absence of MG132, immunoprecipitated with anti-HA antibody and

immunoblotted with the indicated antibodies. (D) HEK293 cells transfected with the indicated

plasmids encoding various tumor-causing HA-VHL mutants were lysed, equal amount of whole

cell extracts resolved by SDS-PAGE and immunoblotted with the indicated antibodies. WCE:

whole cell extract; IP: immunoprecipitation; IB: immunoblot; AR: autoradiography. Asterisk

denotes non-specific protein bands. (E) HEK293 cells transfected with the indicated plasmids

encoding various tumor-causing HA-VHL mutants in combination with T7-JAK2 were treated

with MG132. Equal amounts of cell lysates were immunoprecipitated with anti-HA antibody

89

and immunoblotted with the indicated antibodies. IP: immunoprecipitation; IB: immunoblot;

WCE: whole cell extract.

90

4.3.4 VHL binds and requires SOCS1 to promote pJAK2 degradation

The F-box protein SOCS1 is the principal negative regulator of pJAK2 via ubiquitin-mediated

degradation. VHL, as well as other F-box proteins that confer substrate specificity, have been

shown to homodimerize 193-197. Moreover, homodimerization of entire E3 enzymes such as the

SCF (Skp1/Cdc53 or Cul1/F-box protein) has been shown to increase the efficiency of

ubiquitylation by improving spatial orientation of substrate to active site 197. We asked whether

SOCS1 interacts with VHL to promote ECV-independent degradation of pJAK2. T7-VHL co-

precipitated HA-SOCS1 when ectopically expressed in HEK293 cells (Fig. 4.3a, left panel), and

similar results were obtained by reciprocal immunoprecipitation (Fig. 4.3a, right panel). We

then asked whether VHL/SOCS1 interaction occurred under physiologic conditions. BaF3 cells

that stably express EPOR were treated with or without MG132. Cell lysates were

immunoprecipitated with anti-VHL or isotype-matched control antibody and bound proteins

were visualized by Western blot analysis, which showed endogenous VHL co-precipitating

SOCS1 in the presence of MG132 (Fig. 4.3b). Notably, VHL/SOCS1 interaction was

significantly reduced in the absence of MG132, suggesting perhaps that the complex is sensitive

to proteasomal degradation. The endogenous binding of VHL and SOCS1 will be repeated using

affinity purification for hydroxylated HIF ODD-OH and phosphor tyrosine JAK2 peptides to

improve clarity of band visualization by removing immunoglobulin light chain.

VHL(F119S) and VHL(L128F) mutants are capable of forming an intact ECV and targeting

HIFα for degradation (Fig. 4.3c and 3d), but fail to promote pJAK2 degradation (Fig. 4.3e and

see Fig. 4.2d) and thus, supporting again the notion that the defect in pJAK2 regulation is

independent of ECV. One possibility is that the failure in pJAK2 degradation is due to a defect

in F119S and L128F to engage SOCS1. As predicted, unlike VHL(WT), both F119S and L128F

mutants were severely compromised in binding SOCS1 (Fig. 4.3f), underscoring the potential

requirement of SOCS1, rather than ECV complex formation, in the degradation of pJAK2.

Notably, F119S and L128F mutants retained the ability to bind JAK2 (data not shown).

We asked whether the ability of SOCS1 to recruit the E3 ubiquitin ligase components was

required for VHL-dependent pJAK2 degradation. Analogous to the α domain of VHL, the

SOCS-box of SOCS1 facilitates the recruitment of the various ECS components including Cul5

91

or Cul2, Elongins BC and Rbx1 132,198,199. While both VHL(WT) and VHL(C162F; α domain

mutant that cannot form an ECV) mutant promoted pJAK2 degradation when co-expressed with

wild-type SOCS1, co-expression of SOCS1∆SOCS-box mutant abrogated pJAK2 degradation

(Fig. 4.3g). These results suggest that SOCS-box is required for enzymatic activity of the

VHL/SOCS1 heterodimer for the degradation of pJAK2.

92

93

Figure 4.3. VHL and SOCS1 cooperate to degrade pJAK2 in vivo.

94

(A) HEK293 cells transfected with the indicated plasmids were lysed, immunoprecipitated with

either anti-T7 (left panels) or anti-HA (right panels) antibody, resolved by SDS-PAGE, and

immunoblotted with the indicated antibodies. (B) BaF3 cells stably expressing EPOR were

stimulated with EPO in the presence (+) or absence (-) of MG132. Cells were lysed and

immunoprecipitated with anti-VHL or isotype-matched control antibody and immunoblotted

with the indicated antibodies. (C) 786-O subclones stably expressing the indicated HA-VHL

were lysed, immunoprecipitated with anti-HA antibody and immunoblotted with indicated

antibodies. (D) 786-O subclones stably expressing the indicated HA-VHL were lysed and

immunoblotted with indicated antibodies. (E) Equal amounts of whole cell extracts prepared

from HEK293 cells transfected with the indicated plasmids were resolved by SDS-PAGE and

immunoblotted with the indicated antibodies. (F) HEK293 cells transfected with the indicated

plasmids in combination with a plasmid encoding HA-SOCS1 were lysed, immunoprecipitated

with anti-VHL antibody and immunoblotted with anti-HA antibody. (G) Equal amounts of

whole cell extracts prepared from HEK293 cells transfected with the indicated plasmids and

stimulated with EPO were resolved by SDS-PAGE and immunoblotted with the indicated

antibodies. IP: immunoprecipitation; IB: immunoblot; WCE: whole cell extract. Asterisk

denotes non-specific protein bands.

95

4.3.5 CP-VHL/SOCS1 association inhibits pJAK2 binding and

degradation

We asked whether the observed defect in pJAK2 degradation via CP-VHL was due to a failure in

binding SOCS1. Unexpectedly, both VHL(R200W) and VHL(H191D) mutants showed a

dramatic increase in SOCS1 binding in comparison to their wild-type VHL counterpart (Fig.

4.4a), which suggests that CP-causing mutations confer significantly higher affinity for SOCS1.

We next asked whether this altered affinity of CP-VHL for SOCS1 affected pJAK2 recruitment.

HEK293 cells transfected with plasmids encoding EPOR, T7-JAK2 and HA-SOCS1 in

combination with plasmids encoding HA-VHL(WT or R200W or H191D) were stimulated with

EPO in the presence of MG132 to minimize the degradation of pJAK2. pJAK2 co-precipitated

significantly lower levels of CP-VHL mutants in comparison to VHL(WT), suggesting that the

abnormal association between CP-VHL and SOCS1 hinders pJAK2 substrate binding (Fig. 4.4b).

We next directly compared the efficiency of VHL(WT)/SOCS1 against CP-VHL/SOCS1 in

promoting pJAK2 degradation. T7-pJAK2 was first generated by ectopic expression of EPOR

and T7-JAK2 in HEK293 cells followed by EPO stimulation. Cells were lysed and

immunoprecipitated with an anti-T7 antibody. T7-pJAK2 enriched on beads were washed and

equally distributed into 4 reaction tubes, as confirmed by comparable levels of IgGL (Fig. 4.4c,

bottom panel), and mixed with HEK293 cell lysates expressing empty plasmid (MOCK), HA-

VHL(WT), HA-VHL(R200W) or HA-VHL(H191D) in combination with HA-SOCS1.

VHL(WT)/SOCS1 containing lysate markedly reduced the level of pJAK2 in comparison to CP-

VHL/SOCS1 or SOCS1 only containing lysates (Fig. 4.4c). These results collectively suggest

that the CP-VHL/SOCS1 heterocomplex is defective in promoting pJAK2 degradation.

96

97

Figure 4.4. CP-VHL mutants are defective in pJAK2 degradation and R200W/R200W CP

mice exhibit elevated pJAK2 and pSTAT5 levels.

(A) HEK293 cells transfected with the indicated plasmids in combination with HA-SOCS1 were

lysed, immunoprecipitated with anti-VHL antibody and blotted with anti-HA antibody. (B)

HEK293 cells transfected with the indicated plasmids were treated with EPO and MG132, lysed,

immunoprecipitated with anti-T7 antibody, and immunoblotted with the indicated antibodies.

(C) T7-pJAK2 was first generated by ectopic expression of EPOR and T7-JAK2 in HEK293

cells followed by EPO stimulation. Cells were lysed and immunoprecipitated with an anti-T7

antibody. T7-pJAK2 enriched on beads were washed and equally distributed into 4 reaction

tubes, as confirmed by comparable levels of IgGL (bottom panel), and mixed with HEK293 cell

lysates expressing empty plasmid (MOCK), HA-VHL(WT), HA-VHL(R200W) or HA-

VHL(H191D) in combination with HA-SOCS1. (D) Single cell suspensions enriched with

erythroid progenitors generated from spleens of phenylhydrazine-treated R200W/R200W or WT

mice were washed in cytokine-free media to remove any residual cytokines. Cells were cytokine

starved for additional 4 h to purge any pre-existing stimulation of JAK2-STAT5 pathway and

subsequently treated with increasing concentrations of exogenous EPO for 15 min. Equal

amounts of cell lysates were resolved on SDS-PAGE and immunoblotted with the indicated

antibodies. IP: immunoprecipitation; IB: immunoblot; WCE: whole cell extract. Asterisk

denotes non-specific protein bands.

98

4.3.6 pJAK2 and pSTAT5 are elevated in CP-mice

Erythroid progenitors from PV patients are hypersensitive to EPO due to JAK2 activating

mutations associated with increased levels of phosphorylated JAK2 and STAT5 139. Erythroid

progenitors from CP patients or R200W/R200W mice have likewise been shown to be

hypersensitive to EPO 40,142. Single cell suspensions enriched with erythroid progenitors were

generated from spleens of phenylhydrazine (PHZ)-treated R200W/R200W or WT mice and

residual cytokines were removed by washes in cytokine-free media. Cells were cytokine starved

for additional 4 h to purge any pre-existing stimulation of the JAK2-STAT5 pathway and

subsequently treated with increasing concentrations of exogenous EPO for 15 min. Expression

levels of pJAK2 and pSTAT5 were noticeably higher in R200W/R200W compared to the WT

erythroid progenitor-enriched cell lysates (Fig. 4.4d, compare lanes 3 and 4 against 7 and 8).

Densitometry performed on unsaturated exposures of the immunoblots validated the observed

trend (data not shown). These results demonstrate that homozygous inheritance of CP-causing

R200W mutation increases JAK2-STAT5 signalling pathway in vivo. Further experiments will

be conducted in order to determine JAK2-STAT5 sensitivity using both cellular methods (colony

forming assays) and additional biochemical stimulation assays.

4.3.7 Discussion

Mapping of VHL disease-causing mutations on VHL/Elongin B/Elongin C (VBC) crystal

structure engaged with HIF1α peptide has revealed two major domains α and β required for

Elongin C and HIF1α binding, respectively 48,184,200. VHL mutations that disrupt (F119S and

L128F) or enhance (R200W and H191D) SOCS1 binding interestingly clustered to a unique

region of VHL, revealing a likely interface or ‘SOCS groove’ required for the engagement of

SOCS1 (Fig. 4.5a). Notably, the SOCS groove does not overlap with Elongin C or HIF1α

binding interface. This is consistent with the observed autonomy of HIF- and JAK2-associated

functions of VHL clearly revealed by specific mutants F119S and L128F, which retain the ability

to degrade HIFα but fail to degrade pJAK2 despite their ability to form ECV. Conversely,

C162F retains the ability to degrade pJAK2 despite its inability to form ECV or degrade HIFα.

99

Thus, mutations within the groove may alter VHL’s affinity for SOCS1 positively or negatively

via steric conformational change.

We propose the following model of CP. In normal individuals, VHL forms a proper ECV

complex and negatively regulate HIFα via the ubiquitin pathway. In contrast, CP-associated

mutations (e.g., R200W) attenuate HIFα binding and ECV complex formation, causing the

reported mild stabilization of HIFα, which leads to the overproduction of HIF-target EPO in the

kidney and secondary polycythemia (Fig. 4.5b). In normal individuals, VHL also binds SOCS1

through its SOCS groove and together recognize pJAK2 for ubiquitin-mediated degradation, and

thus negatively regulate the JAK2-STAT pathway. The R200W mutation in CP patients causes

conformational change within the SOCS groove, leading to an inordinately tight CP-

VHL/SOCS1 association and thereby blocking pJAK2 recruitment and degradation. Resulting

pJAK2 stabilization promotes hyperactivation of the JAK2-STAT pathway in erythroid

progenitors, causing hypersensitivity to EPO and primary polycythemia.

100

Figure 4.5. The ‘SOCS groove’ and the revised molecular model of CP.

(A) Mutations (red) that influence SOCS1 binding are indicated on the VHL/Elongin B/Elongin

C (VBC) crystal structure bound to HIF1α peptide and cluster within the ‘SOCS-groove’.

Analyzed using DeepView/Swiss-PdbViewer v4.0. (B) Molecular model of CP. See text for

details.

101

The present findings also provide molecular explanations to several mysteries and paradoxes in

VHL and CP fields. For example, it has been unclear why polycythemia is rarely observed in

VHL patients despite the fact that mutations that promote HIFα stabilization were common

among VHL patients 38,69,142. We show here that most tumour-causing VHL mutants, including

those that have lost the ability to degrade HIFα, retain the ability to negatively regulate pJAK2

downstream of EPO signalling, which likely explains the rarity of polycythemia among VHL

patients (see Fig. 4.5b). Furthermore, R200W/WT heterozygous mice, which do not show

detectable HIFα accumulation or EPO overproduction, have BFU-E cells that are modestly

hypersensitive to EPO ex vivo 142. Consistent with this observation is a report describing a

woman with Y175C/WT VHL genotype who has polycythemia without an elevated level of

serum EPO 201. These findings support the notion that HIF is unlikely involved in

hypersensitivity of erythroid progenitors to EPO, a hallmark feature of primary polycythemia.

PV-associated JAK2(V617F) mutation causes uncontrolled expansion of RBCs, but also gives

rise to pleomorphic and clustered megakaryocytes hypersensitive to thrombopoietin, which,

similar to EPO, signals through JAK2 134. Abnormal megakaryocyte function is thought to be

critical in thrombotic complications frequently observed in PV patients 192. Strikingly,

R200W/R200W mice exhibit increased number of megakaryocytes that cluster and CP patients,

like PV patients, often present with thrombotic complications 40,142. In contrast, secondary

polycythemia associated with elevated EPO does not give rise to megakaryocytic defects; an

observation supported in mice with constitutive overexpression of EPO that do not develop

thrombotic complications despite inordinately high RBC count 202. These observations suggest

that the hyperactive JAK2-STAT signalling, but not the increased EPO production due to a mild

defect in HIF regulation, is the principal mechanism underlying thrombotic complications

observed in CP patients.

The discovery of JAK2 mutations in PV patients has certainly expedited the clinical trials of

JAK2 inhibitors in the management of PV. However, despite clinical features shared between

PV and CP, including hypersensitivity to erythropoietin and megakaryocytic defects associated

with thrombotic complications, JAK2 inhibitors have not been considered for CP. Thus, the

present findings linking VHL to JAK2-STAT5 pathway provide a biochemical rationale for

JAK2-targeted therapies in CP.

102

Chapter 5 Conclusions and future directions

5.1 E-cadherin loss in RCC

In addition to our publication of the HIFα-mediated repression of E-cadherin, two independent

reports were published showing a very similar mechanism responsible for E-cadherin loss in

RCC35,203,204. In all three publications E-cadherin expression was recovered by reconstitution of

RCC cell lines with wild type-VHL. However, some differences were noted. Krishnamachary et

al. argue that the regulation of E-cadherin is strictly mediated through HIF1α since restoration of

VHL in HIF1α-deficient 786-O RCC cell line did not rescue E-cadherin expression203. In

contrast, Estaban et al. demonstrated both HIF1α and HIF2α were capable of repressing E-

cadherin expression35. Intriguingly, we were able to observe restoration of E-cadherin

expression upon reintroduction of wild-type VHL in 786-O cells, which do not express HIF1α,

suggesting that downregulation of HIF2α is sufficient for promotion of E-cadherin

expression204,205. It is possible that after extended passaging some immortalized cell lines may

lose the ability to recover E-cadherin expression, explaining the differences in these

observations. Similar to our findings that HIF2α drives the expression of E2 box-specific

transcriptional repressors, Krishnamachary et al. found elevation of SIP1 as well as modest

induction of two additional repressors TCF-3 and δEF1 (ZFHX1A). Interestingly, in our

luciferase assay we saw an incomplete repression of the E-cadherin promoter with SIP1 alone,

indicating that additional repressors may play a role in HIF-mediated repression of E-cadherin.

Taken together, these reports highlight a previously unknown link between the major cause of

RCC development (i.e., VHL inactivation) and the loss of a critical invasion suppressor, E-

cadherin. Early loss of both VHL and E-cadherin observed in the pre-malignant foci of VHL

patients undoubtedly sets the stage for disease progression.

The VHL-mediated regulation of E-cadherin is dependent on HIF, which also governs the

transcription of more than 60 genes, including VEGF, PDGFβ, GLUT1, transferrin and its

receptor, EPO, TGFα/β3, and CXCR4 and its ligand SDF1α55,173. In addition, VHL also

regulates the expression of genes independent of HIF, such as MMPs, TIMPs and FN55. Many

103

of these genes, both HIF-dependent and independent, have been shown to affect the metastatic

potential of RCC; however, the biological contribution of an individual gene product upon VHL

loss during RCC progression is at present unclear and difficult to discern. Interestingly, shRNA-

targeted reduction of E-cadherin in RCC cells with a restored VHL-HIF pathway markedly

increased the invasive potential, suggesting that the loss of E-cadherin plays a critical role in

promoting the malignant behavior of RCC. In keeping with this hypothesis, it was recently

shown that a lack of E-cadherin and VHL staining by IHC correlated with high grade tumours

and poor prognosis206.

Loss of E-cadherin, in addition to promoting invasiveness and EMT, may also influence

the Wnt signalling pathway (see Fig. 5.1)143. For example, phosphorylation of β-catenin via

receptor tyrosine kinases (e.g., c-Met, fyn, fer) or c-src can disassociate β-catenin from the

cytoplasmic tail of E-cadherin and translocate to the nucleus where it binds TCF to drive the

transcription of genes responsible for proliferation and differentiation108,143. Peruzzi et al.

showed that the loss of VHL in RCC triggers HGF-driven β-catenin signalling that induced

branched morphogenesis. Reintroduction of VHL repressed the accumulation of the active

cytoplasmic β-catenin and the disruption of adherens junction106. Nakaigawa et al. showed that

c-Met is phosphorylated upon VHL loss in the absence of HGF ligand stimulation. The

inhibition of c-Met using a c-Met inhibitor attenuated the growth of VHL-null RCC tumors in

nude mice, implicating the importance of c-Met signaling in renal epithelial oncogenesis207. It is

tempting to speculate that a loss of VHL may not only decrease the adhesive potential (and

therefore, increase invasive potential), but also simultaneously increase the availability of active

β-catenin (and β-catenin-mediated gene transcription, such as Cyclin D1) by either constitutive

c-Met phosphorylation and/or inhibition of β-catenin degradation. Thus, the inactivation of a

single gene VHL may cause a pronounced shift from a normal epithelial homeostasis towards an

invasive de-differentiated cellular state with enhanced proliferative capacity. These recent

studies have begun to unravel the molecular pathways regulating the development of aggressive

RCC upon VHL inactivation, involving both HIF-dependent and -independent mechanisms.

Determining the pathophysiologic relevance, as well as the relative contribution, of these distinct

pathways will undoubtedly shed important insight into the understanding the molecular basis of

EMT in kidney cancer.

104

Figure 5.1. Role of VHL in the regulation of E-cadherin and β-catenin. Loss of VHL leads to the stabilization of HIFα (1), which promotes the transactivation of E-cadherin-specific repressors, curtailing E-cadherin expression. Loss of VHL also causes constitutive phosphorylation of c-Met (2) and subsequent release and activation of β-catenin, as well as stabilization of active β-catenin (3). See text for details.

105

5.2 Uncovering the mechanism of VHL mediated FN assembly

The dual role of VHL in the promotion of ECM and the negative regulation of the hypoxic

response is well established. While the mechanisms that underlie HIFα destruction are well

understood, the mechanism by which VHL is able to promote an ECM is significantly less clear.

We have demonstrated that the neddylation of VHL promotes direct association with FN by

displacing Cul2; however, the processes that regulate NEDD8 conjugation to VHL remain

unclear. The specificity and regulation of NEDD8 conjugation remains at the level of the E3-

ligase. Therefore, discovery and characterization of the E3 for VHL neddylation will likely yield

clues as to what physiologic conditions promote VHL-FN interaction and subsequent

extracellular fibril formation.

VHL binds FN via its beta-domain, while VHL itself has no known enzymatic function, it likely

acts as an adaptor bringing in additional proteins through its association with elongins B and C.

As the discovery of VHL associated proteins yielded significant clues to the function of VHL in

the ECV, it stands to reason that identification of additional proteins in the VHL-FN complex

will shed light on the function of the VHL-FN complex. Intriguingly, the mechanisms that have

been proposed for FN and ColIV interaction with VHL differ greatly. VHL has been proposed

to bind hydroxylated collagen and the competition for VHL by HIF and ColIV determines the

ability of VHL to bind ColIV. We have not seen a difference in VHL-FN association under

hypoxia, which is in keeping with the observation that overexpression of HIFα does not affect

FN deposition. The differences in proposed mechanisms are in contrast to the striking similarity

that is seen in the defects and production of FN and ColIV matrices. While it is certainly

possible that the rules that govern VHL interaction with the two matrix proteins are entirely

divergent, there is also the possibility that a greater understanding of VHL matrix function will

produce a more unified theory of VHL matrix function. It is interesting to note that VHL-ColIV

interaction was not tested under limiting oxygen tension in vivo, instead a potent iron chelator,

DFO was used as a mimetic for hypoxia. Treatment of DFO resulted in a shift of approximately

30KDa in the ColIV protein. This shift is likely due to a inhibition of glycosylation, which

requires hydroxylated prolines as a substrate for the addition of glycosyl groups in the ER208. It

106

is interesting to note that a Far Western of DFO-treated lysates revealed that VHL only

associates with the upper-glycosylated ColIV band and this association does not appear to be

lessened by DFO treatment, which would be expected if DFO inhibited the hydroxylation of

ColIV. An alternate explanation is that DFO reduces glycosylation and that VHL specifically

recognizes the glycosylated form of ColIV. Interestingly, we have seen VHL colocalize with FN

specifically in the Golgi. Given the role of the Golgi in the refinement of glycosylation and the

fact that both ColIV and FN are heavily glycosylated proteins, it is tempting to hypothesize that

future work may uncover a role of VHL in promoting matrix deposition via refinement of

glycosylation of target proteins in the Golgi.

5.3 Characterization of VHL mutation in additional haematopoietic malignancies

Mutations in JAK2 have been identified in a majority of patients with PV, essential

thrombocythemia (ET) and myelofibrosis with myeloid metaplasia (MMM)127. Increases in

signalling downstream of JAK2 have been observed in mediastinal B-cell lymphoma (PMBL),

classical Hodgkin lymphoma (cHL), chronic lymphocytic leukemia (CLL), and large granular

lymphocyte (LGL) leukemia209. Constitutive signalling downstream of JAK2 in leukemia has

been shown to prevent apoptosis of leukemic cells to stimulus, such as Fas ligand, via the

upregulation of anti-apoptotic proteins including Bcl-xL210,211.

Interestingly, increased signalling of tyrosine kinases including JAK2 has recently been

described in Hodgkin’s lymphoma in the absence of activating mutations of JAK2212. Although

JAK2 is often amplified due to inherent genomic instability, the activation status of JAK2 does

not correlate with copy number213. This paradox was understood upon the identification of

SOCS1 mutation in nearly 50% of cHL samples214. However, JAK2 expression is elevated in

over 85% of cHL allowing for the possibility that additional mutations contribute to the increase

of tyrosine kinase activity in this lymphoma. Given the importance of proper VHL-SOCS1

function in polycythemia, it stands to reason that VHL mutations could affect a similar

hyperactivation of JAK2 in the background of cHL.

Our understanding of the role of JAK2 deregulation in haematopoietic malignancies has rapidly

expanded in the past four years. The negative regulators of JAK2, VHL and SOCS1 represent an

107

important impediment for haematopoietic disease development. Our identification of the role of

VHL in a myeloproliferative disorder brings CP in line with our understanding of MPDs.

108

References

1. Collins, E.T. Intra-ocular growths (two cases, brother and sister, with peculiar vascular new growth, probably retinal, affecting both eyes). Trans. Ophthal. Soc. U.K. 14, 141-149 (1894).

2. Neumann, H.P.H. & Wiestler, O.D. Clustering of features of von Hippel-Lindau disease: evidence of a complex genetic locus. Lancet 337, 1052-1054 (1991).

3. Maddock, I., et al. A genetic register for von Hippel-Lindau disease. J Med Genet 33 33, 120-127 (1996).

4. von Hippel, E. Ueber eine sehr seltene Erkrankung der Nethaut. Graefe Arch Ophthal 59, 83-106 (1904).

5. Lindau, A. Zur Frage der Angiomatosis Retinae und ihrer Hirnkomplikation. Acta Opthal 4, 193-226 (1927).

6. Richard, S., Campello, C., Taillandier, L., Parker, F. & Resche, F. Haemangioblastoma of the central nervous system in von Hippel-Lindau disease. French VHL Study Group. J

Intern Med 243, 547-553 (1998).

7. Frew, I.J., et al. Combined VHLH and PTEN mutation causes genital tract cystadenoma and squamous metaplasia. Mol Cell Biol 28, 4536-4548 (2008).

8. Hough, D.M., Stephens, D.H., Johnson, C.D. & Binkovitz, L.A. Pancreatic lesions in von Hippel-Lindau disease: prevalence, clinical significance, and CT findings. AJR Am J

Roentgenol 162, 1091-1094 (1994).

9. Neumann, H.P.H., et al. Pheochromocytomas, multiple endocrine neoplasia type 2, and von Hippel-Landau Disease. N. Engl. J. Med. 329, 1531-1538 (1993).

10. Manski, T., et al. Endolymphatic sac tumors - A source of morbid hearing loss in von Hippel-Lindau disease. JAMA 277, 1461-1466 (1997).

11. Sung, D.I., Chang, C.H. & Harisiadis, L. Cerebellar hemangioblastomas. Cancer 49, 553-555 (1982).

12. Dan, N.G. & Smith, D.E. Pituitary hemangioblastoma in a patient with von Hippel-Lindau disease. Case report. J Neurosurg 42, 232-235 (1975).

13. Ginzburg, B.M., et al. Diagnosis of von Hippel-Lindau disease in a patient with blindness resulting from bilateral optic nerve hemangioblastomas. AJR Am J Roentgenol 159, 403-405 (1992).

14. Benjamin, L.E. & Keshet, E. Conditional switching of vascular endothelial growth factor (VEGF) expression in tumors: induction of endothelial cell shedding and regression of

109

hemangioblastoma-like vessels by VEGF withdrawal. Proc Natl Acad Sci (USA) 94, 8761-8766 (1997).

15. Vortmeyer, A.O., et al. Developmental arrest of angioblastic lineage initiates tumorigenesis in von Hippel-Lindau disease. Cancer Res 63, 7051-7055 (2003).

16. Reifenberger, G., Reifenberger, J., Bilzer, T., Wechsler, W. & Collins, V. Coexpression of transforming growth factor-alpha and epidermal growth factor receptor in capillary hemangioblastomas of the central nervous system. Am J Pathol 147, 245 (1995).

17. Wizigmann-Voos, S., Breier, G., Risau, W. & Plate, K. Up-regulation of vascular endothelial growth factor and its receptors in von Hippel-Lindau disease-associated and sporadic hemangioblastomas. Cancer Res 55, 1358-1364 (1995).

18. Bohling, T., et al. Expression of growth factors and growth factor receptors in capillary hemangioblastoma. J Neuropathol Exp Neurol 55, 522-527 (1996).

19. Krieg, M., Marti, H. & KH, P. Coexpression of erythropoietin and vascular endothelial growth factor in nervous system tumors associated with von hippel-lindau tumor suppressor gene loss of function. Blood 92, 3388-3393 (1998).

20. Glenn, F. Pheochromocytoma, a chromaffin tumor of the adrenal medulla. Harlem Hosp

Bull 10, 1-20 (1957).

21. Reisch, N., Peczkowska, M., Januszewicz, A. & Neumann, H.P. Pheochromocytoma: presentation, diagnosis and treatment. J Hypertens 24, 2331-2339 (2006).

22. Zelinka, T., Eisenhofer, G. & Pacak, K. Pheochromocytoma as a catecholamine producing tumor: implications for clinical practice. Stress 10, 195-203 (2007).

23. Gimm, O. Pheochromocytoma-associated syndromes: genes, proteins and functions of RET, VHL and SDHx. Fam Cancer 4, 17-23 (2005).

24. Lee, S., et al. Neuronal apoptosis linked to EglN3 prolyl hydroxylase and familial pheochromocytoma genes: developmental culling and cancer. Cancer Cell 8, 155-167 (2005).

25. Cohen, H.T. & McGovern, F.J. Renal-cell carcinoma. N Engl J Med 353, 2477-2490 (2005).

26. Brugarolas, J. Renal-cell carcinoma--molecular pathways and therapies. N Engl J Med 356, 185-187 (2007).

27. Gunaratnam, L., et al. Hypoxia inducible factor activates the transforming growth factor-alpha/epidermal growth factor receptor growth stimulatory pathway in VHL(-/-) renal cell carcinoma cells. J Biol Chem 278, 44966-44974 (2003).

110

28. Wiesener, M.S., et al. Paraneoplastic erythrocytosis associated with an inactivating point mutation of the von Hippel-Lindau gene in a renal cell carcinoma. Blood 99, 3562-3565 (2002).

29. Lager, D.J., Slagel, D.D. & Palechek, P.L. The expression of epidermal growth factor receptor and transforming growth factor alpha in renal cell carcinoma. Mod Pathol 7, 544-548 (1994).

30. Choyke, P.L., et al. The natural history of renal lesions in von Hippel-Lindau disease: a serial CT study in 28 patients. AJR Am J Roentgenol 159, 1229-1234 (1992).

31. Paraf, F., et al. Renal lesions in von Hippel-Lindau disease: immunohistochemical expression of nephron differentiation molecules, adhesion molecules and apoptosis proteins. Histopathology 36, 457-465 (2000).

32. Mandriota, S.J., et al. HIF activation identifies early lesions in VHL kidneys: evidence for site-specific tumor suppressor function in the nephron. Cancer Cell 1, 459-468 (2002).

33. Lee, Y.S., et al. Coexpression of erythropoietin and erythropoietin receptor in von Hippel-Lindau disease-associated renal cysts and renal cell carcinoma. Clin Cancer Res 11, 1059-1064 (2005).

34. Nogueira, E., Klimek, F., Weber, E. & Bannasch, P. Collecting duct origin of rat renal clear cell tumors. Virchows Arch B Cell Pathol Incl Mol Pathol 57, 275-283 (1989).

35. Esteban, M.A., et al. Regulation of E-cadherin expression by VHL and hypoxia-inducible factor. Cancer Res 66, 3567-3575 (2006).

36. Young, A.P., et al. VHL loss actuates a HIF-independent senescence programme mediated by Rb and p400. Nat Cell Biol 10, 361-369 (2008).

37. Neumann, H.P. & Wiestler, O.D. Clustering of features of von Hippel-Lindau syndrome: evidence for a complex genetic locus. Lancet 337, 1052-1054 (1991).

38. Hoffman, M.A., et al. von Hippel-Lindau protein mutants linked to type 2C VHL disease preserve the ability to downregulate HIF. Hum Mol Genet 10, 1019-1027 (2001).

39. Brauch, H., et al. von Hippel-Lindau disease with pheochromocytoma in the Black Forest region in Germany: evidence for a founder effect. Hum Genet 95, 551-556 (1995).

40. Ang, S.O., et al. Disruption of oxygen homeostasis underlies congenital Chuvash polycythemia. Nat Genet 32, 614-621 (2002).

41. Seizinger, B.R., et al. Von-hippel lindau disease maps to the region of chromosome 3 associated with renal cell carcinoma. Nature 332, 268-269 (1988).

42. Latif, F., et al. Identification of the von Hippel-Lindau Disease Tumor Suppressor Gene. Science 260, 1317-1320 (1993).

111

43. Gnarra, J., et al. Defective placental vasculogenesis causes embryonic lethality in VHL-deficient mice. Proc Natl Acad Sci U S A 94, 9102-9107 (1997).

44. Iliopoulos, O., Ohh, M. & Kaelin, W. pVHL19 is a biologically active product of the von Hippel-Lindau gene arising from internal translation initiation. Proc Natl Acad Sci U S A 95, 11661-11666 (1998).

45. Schoenfeld, A., Davidowitz, E.J. & Burk, R.D. A second major native von Hippel-Lindau gene product, initiated from an internal translation start site, functions as a tumor suppressor. Proc Natl Acad Sci U S A 95, 8817-8822. (1998).

46. Lee, S., et al. Nuclear/cytoplasmic localization of the von Hippel-Lindau tumor suppressor gene product is determined by cell density. Proc Natl Acad Sci 93, 1770-1775 (1996).

47. Mekhail, K., Gunaratnam, L., Bonicalzi, M.E. & Lee, S. HIF activation by pH-dependent nucleolar sequestration of VHL. Nat Cell Biol 6, 642-647 (2004).

48. Stebbins, C.E., Kaelin, W.G. & Pavletich, N.P. Structure of the VHL-ElonginC-elonginB complex: implications for VHL tumor suppressor function. Science 284, 455-461 (1999).

49. Li, L., et al. Hypoxia-inducible factor linked to differential kidney cancer risk seen with type 2A and type 2B VHL mutations. Mol Cell Biol 27, 5381-5392 (2007).

50. Pastore, Y.D., et al. Mutations in the VHL gene in sporadic apparently congenital polycythemia. Blood 101, 1591-1595 (2003).

51. Pastore, Y., et al. Mutations of von Hippel-Lindau tumor-suppressor gene and congenital polycythemia. Am J Hum Genet 73, 412-419 (2003).

52. Ohh, M. & Kaelin, W.G., Jr. VHL and kidney cancer. Methods Mol Biol 222, 167-183 (2003).

53. Jariel-Encontre, I., Bossis, G. & Piechaczyk, M. Ubiquitin-independent degradation of proteins by the proteasome. Biochim Biophys Acta 1786, 153-177 (2008).

54. Sufan, R.I. & Ohh, M. Role of the NEDD8 modification of Cul2 in the sequential activation of ECV complex. Neoplasia 8, 956-963 (2006).

55. Maynard, M.A. & Ohh, M. von Hippel-Lindau tumor suppressor protein and hypoxia-inducible factor in kidney cancer. Am J Nephrol 24, 1-13 (2004).

56. Maynard, M.A. & Ohh, M. The role of hypoxia-inducible factors in cancer. Cell Mol Life

Sci 64, 2170-2180 (2007).

57. Maxwell, P., et al. The von Hippel-Lindau gene product is necessary for oxgyen-dependent proteolysis of hypoxia-inducible factor α subunits. Nature 399, 271-275 (1999).

112

58. Maxwell, P.H., et al. The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature 399, 271-275 (1999).

59. Masson, N., Willam, C., Maxwell, P.H., Pugh, C.W. & Ratcliffe, P.J. Independent function of two destruction domains in hypoxia-inducible factor-alpha chains activated by prolyl hydroxylation. EMBO J. 20, 5197-5206 (2001).

60. Ivan, M., et al. HIFalpha targeted for VHL-mediated destruction by proline hydroxylation: implications for O2 sensing. Science 292, 464-468. (2001).

61. Jaakkola, P., et al. Targeting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science 292, 468-472. (2001).

62. Epstein, A.C., et al. C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. [see comments]. Cell 107, 43-54 (2001).

63. D'Angelo, G., Duplan, E., Boyer, N., Vigne, P. & Frelin, C. Hypoxia up-regulates prolyl hydroxylase activity: a feedback mechanism that limits HIF-1 responses during reoxygenation. J Biol Chem 278, 38183-38187 (2003).

64. Maynard, M.A., et al. Multiple splice variants of the human HIF-3alpha locus are targets of the VHL E3 ubiquitin ligase complex. J Bio Chem 278, 11032-11040 (2003).

65. Lando, D., Peet, D.J., Whelan, D.A., Gorman, J.J. & Whitelaw, M.L. Asparagine hydroxylation of the HIF transactivation domain a hypoxic switch. Science 295, 858-861 (2002).

66. Lando, D., et al. FIH-1 is an asparaginyl hydroxylase enzyme that regulates the transcriptional activity of hypoxia-inducible factor. Genes Dev 16, 1466-1471 (2002).

67. Kondo, K., Klco, J., Nakamura, E., Lechpammer, M. & Kaelin Jr, W.G. Inhibition of HIF is necessary for tumor suppression by the von Hippel-Lindau protein. Cancer Cell 1, 237-246 (2002).

68. Kondo, K., Kim, W.Y., Lechpammer, M. & Kaelin, W.G., Jr. Inhibition of HIF2alpha is sufficient to suppress pVHL-defective tumor growth. PLoS Biol 1, E83 (2003).

69. Clifford, S.C., et al. Contrasting effects on HIF-1alpha regulation by disease-causing pVHL mutations correlate with patterns of tumourigenesis in von Hippel-Lindau disease. Hum Mol Genet 10, 1029-1038 (2001).

70. Hoffman, M.A., et al. von Hippel-Lindau protein mutants linked to type 2C VHL disease preserve the ability to downregulate HIF. Hum Mol Genet 10, 1019-1027. (2001).

71. Ohh, M., et al. The von Hippel-Lindau Tumor Suppressor Protein is Required for Proper Assembly of an Extracellular Fibronectin Matrix. Mol. Cell 1, 959-968 (1998).

113

72. Kurban, G., Hudon, V., Duplan, E., Ohh, M. & Pause, A. Characterization of a von Hippel Lindau pathway involved in extracellular matrix remodeling, cell invasion, and angiogenesis. Cancer Res 66, 1313-1319 (2006).

73. Lieubeau-Teillet, B., et al. von Hippel-Lindau gene-mediated growth suppression and induction of differentiation in renal cell carcinoma cells grown as multicellular tumor spheroids. Cancer Res 58, 4957-4962 (1998).

74. Bishop, T., et al. Genetic Analysis of Pathways Regulated by the von Hippel-Lindau Tumor Suppressor in Caenorhabditis elegans. PLoS Biol 2, E289 (2004).

75. Grosfeld, A., et al. Interaction of hydroxylated collagen IV with the von hippel-lindau tumor suppressor. J Biol Chem 282, 13264-13269 (2007).

76. Kurban, G., et al. Collagen matrix assembly is driven by the interaction of von Hippel-Lindau tumor suppressor protein with hydroxylated collagen IV alpha 2. Oncogene (2007).

77. Pan, Z.Q., Kentsis, A., Dias, D.C., Yamoah, K. & Wu, K. Nedd8 on cullin: building an expressway to protein destruction. Oncogene 23, 1985-1997 (2004).

78. Stickle, N.H., et al. pVHL modification by NEDD8 is required for fibronectin matrix assembly and suppression of tumor development. Mol Cell Biol 24, 3251-3261 (2004).

79. Xirodimas, D.P., Saville, M.K., Bourdon, J.C., Hay, R.T. & Lane, D.P. Mdm2-mediated NEDD8 conjugation of p53 inhibits its transcriptional activity. Cell 118, 83-97 (2004).

80. Watson, I.R., Blanch, A., Lin, D.C., Ohh, M. & Irwin, M.S. Mdm2-mediated NEDD8 modification of TAp73 regulates its transactivation function. J Biol Chem 281, 34096-34103 (2006).

81. Oved, S., et al. Conjugation to Nedd8 instigates ubiquitylation and down-regulation of activated receptor tyrosine kinases. J Biol Chem 281, 21640-21651 (2006).

82. Gao, F., Cheng, J., Shi, T. & Yeh, E.T. Neddylation of a breast cancer-associated protein recruits a class III histone deacetylase that represses NFkappaB-dependent transcription. Nat Cell Biol 8, 1171-1177 (2006).

83. Russell, R.C. & Ohh, M. NEDD8 acts as a 'molecular switch' defining the functional selectivity of VHL. EMBO Rep 9, 486-491 (2008).

84. Neumann, H.P.H. & Zbar, B. Renal cysts, renal cancer and von Hippel-Lindau disease. Kidney Intl. 51, 16-26 (1997).

85. Esteban, M.A., Harten, S.K., Tran, M.G. & Maxwell, P.H. Formation of primary cilia in the renal epithelium is regulated by the von Hippel-Lindau tumor suppressor protein. J

Am Soc Nephrol 17, 1801-1806 (2006).

114

86. Lutz, M.S. & Burk, R.D. Primary cilium formation requires von hippel-lindau gene function in renal-derived cells. Cancer Res 66, 6903-6907 (2006).

87. Schermer, B., et al. The von Hippel-Lindau tumor suppressor protein controls ciliogenesis by orienting microtubule growth. J Cell Biol 175, 547-554 (2006).

88. Thoma, C.R., et al. pVHL and GSK3beta are components of a primary cilium-maintenance signalling network. Nat Cell Biol 9, 588-595 (2007).

89. Kaelin, W.G., Jr. Molecular basis of the VHL hereditary cancer syndrome. Nat Rev

Cancer 2, 673-682 (2002).

90. Hofstra, R.M.W., et al. Extensive mutation scanning of RET in sporadic medullary thyroid carcinoma and of RET and VHL in sporadic pheochromocytoma reveals involvement of these genes in only a minority of cases. Journal of Clinical

Endocrinology and Metabolism 81, 2881-2884 (1996).

91. Estus, S., et al. Altered gene expression in neurons during programmed cell death: identification of c-jun as necessary for neuronal apoptosis. J Cell Biol 127, 1717-1727 (1994).

92. Schlisio, S., et al. The kinesin KIF1Bbeta acts downstream from EglN3 to induce apoptosis and is a potential 1p36 tumor suppressor. Genes Dev 22, 884-893 (2008).

93. Kaelin, W.G., Jr. The von Hippel-Lindau tumour suppressor protein: O2 sensing and cancer. Nat Rev Cancer 8, 865-873 (2008).

94. Vogel, K.S., Brannan, C.I., Jenkins, N.A., Copeland, N.G. & Parada, L.F. Loss of neurofibromin results in neurotrophin-independent survival of embryonic sensory and sympathetic neurons. Cell 82, 733-742 (1995).

95. Smith, K., et al. Silencing of epidermal growth factor receptor suppresses hypoxia-inducible factor-2-driven VHL-/- renal cancer. Cancer Res 65, 5221-5230 (2005).

96. Oh, R.R., et al. Expression of HGF/SF and Met protein is associated with genetic alterations of VHL gene in primary renal cell carcinomas. APMIS 110, 229-238 (2002).

97. Mukhopadhyay, D. & Riezman, H. Proteasome-independent functions of ubiquitin in endocytosis and signaling. Science 315, 201-205 (2007).

98. Wiley, H.S. & Burke, P.M. Regulation of receptor tyrosine kinase signaling by endocytic trafficking. Traffic 2, 12-18 (2001).

99. Wang, Y., et al. Regulation of endocytosis via the oxygen-sensing pathway. Nat Med 15, 319-324 (2009).

100. Bucci, C., et al. The small GTPase rab5 functions as a regulatory factor in the early endocytic pathway. Cell 70, 715-728 (1992).

115

101. Ceresa, B.P. Regulation of EGFR endocytic trafficking by rab proteins. Histol

Histopathol 21, 987-993 (2006).

102. Davidowitz, E.J., Schoenfeld, A.R. & Burk, R.D. VHL induces renal cell differentiation and growth arrest through integration of cell-cell and cell-extracellular matrix signaling. Mol Cell Biol 21, 865-874. (2001).

103. Lieubeau-Teillet, B., et al. von Hippel-Lindau gene-mediated growth suppression and induction of differentiation in renal cell carcinoma cells grown as multicellular tumor spheroids. Cancer Res. 58, 4957-4962 (1998).

104. Harten, S.K., et al. Regulation of renal epithelial tight junctions by the von Hippel-Lindau tumor suppressor gene involves occludin and claudin 1 and is independent of E-cadherin. Mol Biol Cell 20, 1089-1101 (2009).

105. Calzada, M.J., et al. von Hippel-Lindau tumor suppressor protein regulates the assembly of intercellular junctions in renal cancer cells through hypoxia-inducible factor-independent mechanisms. Cancer Res 66, 1553-1560 (2006).

106. Peruzzi, B., Athauda, G. & Bottaro, D.P. The von Hippel-Lindau tumor suppressor gene product represses oncogenic beta-catenin signaling in renal carcinoma cells. Proc Natl

Acad Sci U S A 103, 14531-14536 (2006).

107. Jin, T., George Fantus, I. & Sun, J. Wnt and beyond Wnt: multiple mechanisms control the transcriptional property of beta-catenin. Cell Signal 20, 1697-1704 (2008).

108. Brembeck, F.H., Rosario, M. & Birchmeier, W. Balancing cell adhesion and Wnt signaling, the key role of beta-catenin. Curr Opin Genet Dev 16, 51-59 (2006).

109. Barasch, J. Genes and proteins involved in mesenchymal to epithelial transition. Curr

Opin Nephrol Hypertens 10, 429-436 (2001).

110. Nollet, F., Berx, G. & van Roy, F. The role of the E-cadherin/catenin adhesion complex in the development and progression of cancer. Mol Cell Biol Res Commun 2, 77-85 (1999).

111. Cano, A., et al. The transcription factor snail controls epithelial-mesenchymal transitions by repressing E-cadherin expression. Nat Cell Biol 2, 76-83 (2000).

112. Takeichi, M. Morphogenetic roles of classic cadherins. Curr Opin Cell Biol 7, 619-627 (1995).

113. Yoshiura, K., et al. Silencing of the E-cadherin invasion-suppressor gene by CpG methylation in human carcinomas. Proc Natl Acad Sci U S A 92, 7416-7419 (1995).

114. Strathdee, G. Epigenetic versus genetic alterations in the inactivation of E-cadherin. Semin Cancer Biol 12, 373-379 (2002).

116

115. Moreno-Bueno, G., et al. Genetic profiling of epithelial cells expressing e-cadherin repressors reveals a distinct role for snail, slug, and e47 factors in epithelial-mesenchymal transition. Cancer Res 66, 9543-9556 (2006).

116. Imai, T., et al. Hypoxia attenuates the expression of E-cadherin via up-regulation of SNAIL in ovarian carcinoma cells. Am J Pathol 163, 1437-1447 (2003).

117. Come, C., et al. Snail and slug play distinct roles during breast carcinoma progression. Clin Cancer Res 12, 5395-5402 (2006).

118. Rosivatz, E., et al. Differential expression of the epithelial-mesenchymal transition regulators snail, SIP1, and twist in gastric cancer. Am J Pathol 161, 1881-1891 (2002).

119. Messinezy, M. & Pearson, T.C. The classification and diagnostic criteria of the erythrocytoses (polycythaemias). Clin Lab Haematol 21, 309-316 (1999).

120. Percy, M.J., et al. A novel erythrocytosis-associated PHD2 mutation suggests the location of a HIF binding groove. Blood 110, 2193-2196 (2007).

121. Percy, M.J., et al. A gain-of-function mutation in the HIF2A gene in familial erythrocytosis. N Engl J Med 358, 162-168 (2008).

122. Percy, M.J., et al. A family with erythrocytosis establishes a role for prolyl hydroxylase domain protein 2 in oxygen homeostasis. Proc Natl Acad Sci U S A 103, 654-659 (2006).

123. Kim, W.Y., et al. Failure to prolyl hydroxylate hypoxia-inducible factor alpha phenocopies VHL inactivation in vivo. EMBO J 25, 4650-4662 (2006).

124. Baxter, E.J., et al. Acquired mutation of the tyrosine kinase JAK2 in human myeloproliferative disorders. Lancet 365, 1054-1061 (2005).

125. James, C., et al. A unique clonal JAK2 mutation leading to constitutive signalling causes polycythaemia vera. Nature 434, 1144-1148 (2005).

126. Kralovics, R., et al. A gain-of-function mutation of JAK2 in myeloproliferative disorders. N Engl J Med 352, 1779-1790 (2005).

127. Levine, R.L., et al. Activating mutation in the tyrosine kinase JAK2 in polycythemia vera, essential thrombocythemia, and myeloid metaplasia with myelofibrosis. Cancer

Cell 7, 387-397 (2005).

128. Zhao, R., et al. Identification of an acquired JAK2 mutation in polycythemia vera. J Biol

Chem 280, 22788-22792 (2005).

129. Watowich, S.S., et al. Cytokine receptor signal transduction and the control of hematopoietic cell development. Annu Rev Cell Dev Biol 12, 91-128 (1996).

130. Constantinescu, S.N., Girardot, M. & Pecquet, C. Mining for JAK-STAT mutations in cancer. Trends Biochem Sci 33, 122-131 (2008).

117

131. Ungureanu, D., Saharinen, P., Junttila, I., Hilton, D.J. & Silvennoinen, O. Regulation of Jak2 through the ubiquitin-proteasome pathway involves phosphorylation of Jak2 on Y1007 and interaction with SOCS-1. Mol Cell Biol 22, 3316-3326 (2002).

132. Kamizono, S., et al. The SOCS box of SOCS-1 accelerates ubiquitin-dependent proteolysis of TEL-JAK2. Journal of Biological Chemistry 276, 12530-12538 (2001).

133. Sarna, M.K., et al. Differential regulation of SOCS genes in normal and transformed erythroid cells. Oncogene 22, 3221-3230 (2003).

134. Wernig, G., et al. Expression of Jak2V617F causes a polycythemia vera-like disease with associated myelofibrosis in a murine bone marrow transplant model. Blood 107, 4274-4281 (2006).

135. Lacout, C., et al. JAK2V617F expression in murine hematopoietic cells leads to MPD mimicking human PV with secondary myelofibrosis. Blood 108, 1652-1660 (2006).

136. Bumm, T.G., et al. Characterization of murine JAK2V617F-positive myeloproliferative disease. Cancer Res 66, 11156-11165 (2006).

137. Tiedt, R., et al. Ratio of mutant JAK2-V617F to wild-type Jak2 determines the MPD phenotypes in transgenic mice. Blood 111, 3931-3940 (2008).

138. Feng, J., et al. Activation of Jak2 catalytic activity requires phosphorylation of Y1007 in the kinase activation loop. Mol Cell Biol 17, 2497-2501 (1997).

139. Teofili, L., et al. Different STAT-3 and STAT-5 phosphorylation discriminates among Ph-negative chronic myeloproliferative diseases and is independent of the V617F JAK-2 mutation. Blood 110, 354-359 (2007).

140. Sergeyeva, A., et al. Congenital polycythemia in Chuvashia. Blood 89, 2148-2154 (1997).

141. Perrotta, S., et al. Von Hippel-Lindau-dependent polycythemia is endemic on the island of Ischia: identification of a novel cluster. Blood 107, 514-519 (2006).

142. Hickey, M.M., Lam, J.C., Bezman, N.A., Rathmell, W.K. & Simon, M.C. von Hippel-Lindau mutation in mice recapitulates Chuvash polycythemia via hypoxia-inducible factor-2alpha signaling and splenic erythropoiesis. J Clin Invest 117, 3879-3889 (2007).

143. Nelson, W.J. & Nusse, R. Convergence of Wnt, beta-catenin, and cadherin pathways. Science 303, 1483-1487 (2004).

144. Lonergan, K.M., et al. Regulation of Hypoxia-Inducible mRNAs by the von Hippel-LIndau Protein Requires Binding to Complexes containing Elongins B/C and Cul2. Mol.

Cell. Biol. 18, 732-741 (1998).

145. Kibel, A., Iliopoulos, O., DeCaprio, J.D. & Kaelin, W.G. Binding of the von Hippel-Lindau tumor suppressor protein to elongin B and C. Science 269, 1444-1446 (1995).

118

146. Comijn, J., et al. The two-handed E box binding zinc finger protein SIP1 downregulates E-cadherin and induces invasion. Mol Cell 7, 1267-1278 (2001).

147. Eble, J.N., Sauter, G., Epstein, J.I. & Sesterhenn, I.A. World Health Organization Classification of Tumors: Pathology and Genetics of Tumors of the Urinary System and Male Genital Organs. IRAC press, Lyon (2004).

148. Furge, K.A., et al. Robust classification of renal cell carcinoma based on gene expression data and predicted cytogenetic profiles. Cancer Res 64, 4117-4121 (2004).

149. Takahashi, M., et al. Gene expression profiling of clear cell renal cell carcinoma: gene identification and prognostic classification. Proc Natl Acad Sci U S A 98, 9754-9759 (2001).

150. Fish, J.E., et al. The expression of endothelial nitric-oxide synthase is controlled by a cell-specific histone code. J Biol Chem 280, 24824-24838 (2005).

151. Shimazui, T., Giroldi, L.A., Bringuier, P.P., Oosterwijk, E. & Schalken, J.A. Complex cadherin expression in renal cell carcinoma. Cancer Res 56, 3234-3237 (1996).

152. Morell-Quadreny, L., et al. Disruption of basement membrane, extracellular matrix metalloproteinases and E-cadherin in renal-cell carcinoma. Anticancer Res 23, 5005-5010 (2003).

153. Baba, M., et al. Loss of von Hippel-Lindau protein causes cell density dependent deregulation of CyclinD1 expression through hypoxia-inducible factor. Oncogene 22, 2728-2738 (2003).

154. Berx, G., et al. E-cadherin is a tumour/invasion suppressor gene mutated in human lobular breast cancers. Embo J 14, 6107-6115 (1995).

155. Berx, G., et al. E-cadherin is inactivated in a majority of invasive human lobular breast cancers by truncation mutations throughout its extracellular domain. Oncogene 13, 1919-1925 (1996).

156. Behrens, J., Lowrick, O., Klein-Hitpass, L. & Birchmeier, W. The E-cadherin promoter: functional analysis of a G.C-rich region and an epithelial cell-specific palindromic regulatory element. Proc Natl Acad Sci U S A 88, 11495-11499 (1991).

157. Graff, J.R., et al. E-cadherin expression is silenced by DNA hypermethylation in human breast and prostate carcinomas. Cancer Res 55, 5195-5199 (1995).

158. Ji, X., Woodard, A.S., Rimm, D.L. & Fearon, E.R. Transcriptional defects underlie loss of E-cadherin expression in breast cancer. Cell Growth Differ 8, 773-778 (1997).

159. Batlle, E., et al. The transcription factor snail is a repressor of E-cadherin gene expression in epithelial tumour cells. Nat Cell Biol 2, 84-89 (2000).

119

160. Peinado, H., Ballestar, E., Esteller, M. & Cano, A. Snail mediates E-cadherin repression by the recruitment of the Sin3A/histone deacetylase 1 (HDAC1)/HDAC2 complex. Mol

Cell Biol 24, 306-319 (2004).

161. Luecke, H.F. & Yamamoto, K.R. The glucocorticoid receptor blocks P-TEFb recruitment by NFkappaB to effect promoter-specific transcriptional repression. Genes Dev 19, 1116-1127 (2005).

162. Baba, M., et al. Tumor suppressor protein VHL is induced at high cell density and mediates contact inhibition of cell growth. Oncogene 20, 2727-2736 (2001).

163. Mohan, S. & Burk, R.D. von Hippel-Lindau protein complex is regulated by cell density. Oncogene 22, 5270-5280 (2003).

164. Groulx, I. & Lee, S. Oxygen-dependent ubiquitination and degradation of hypoxia-inducible factor requires nuclear-cytoplasmic trafficking of the von Hippel-Lindau tumor suppressor protein. Mol Cell Biol 22, 5319-5336 (2002).

165. Vandewalle, C., et al. SIP1/ZEB2 induces EMT by repressing genes of different epithelial cell-cell junctions. Nucleic Acids Res 33, 6566-6578 (2005).

166. Spagnoli, F.M., Cicchini, C., Tripodi, M. & Weiss, M.C. Inhibition of MMH (Met murine hepatocyte) cell differentiation by TGF(beta) is abrogated by pre-treatment with the heritable differentiation effector FGF1. J Cell Sci 113 ( Pt 20), 3639-3647 (2000).

167. Jamora, C., et al. A signaling pathway involving TGF-beta2 and snail in hair follicle morphogenesis. PLoS Biol 3, e11 (2005).

168. Zhang, H., et al. Cellular response to hypoxia involves signaling via Smad proteins. Blood 101, 2253-2260 (2003).

169. Scheid, A., et al. Physiologically low oxygen concentrations in fetal skin regulate hypoxia-inducible factor 1 and transforming growth factor-beta3. Faseb J 16, 411-413 (2002).

170. Nishi, H., et al. Hypoxia-inducible factor-1 transactivates transforming growth factor-beta3 in trophoblast. Endocrinology 145, 4113-4118 (2004).

171. Ananth, S., et al. Transforming growth factor beta1 is a target for the von Hippel-Lindau tumor suppressor and a critical growth factor for clear cell renal carcinoma. Cancer Res 59, 2210-2216 (1999).

172. Staller, P., et al. Chemokine receptor CXCR4 downregulated by von Hippel-Lindau tumour suppressor pVHL. Nature 425, 307-311 (2003).

173. Zagzag, D., et al. Stromal cell-derived factor-1alpha and CXCR4 expression in hemangioblastoma and clear cell-renal cell carcinoma: von Hippel-Lindau loss-of-function induces expression of a ligand and its receptor. Cancer Res 65, 6178-6188 (2005).

120

174. Brugarolas, J.B., Vazquez, F., Reddy, A., Sellers, W.R. & Kaelin, W.G., Jr. TSC2 regulates VEGF through mTOR-dependent and -independent pathways. Cancer Cell 4, 147-158 (2003).

175. Zundel, W., et al. Loss of PTEN facilitates HIF-1-mediated gene expression. Genes Dev. 14, 391-396 (2000).

176. Ravi, R., et al. Regulation of tumor angiogenesis by p53-induced degradation of hypoxia-inducible factor 1alpha. Genes Dev 14, 34-44 (2000).

177. Laughner, E., Taghavi, P., Chiles, K., Mahon, P.C. & Semenza, G.L. HER2 (neu) signaling increases the rate of hypoxia-inducible factor 1alpha (HIF-1alpha) synthesis: novel mechanism for HIF-1-mediated vascular endothelial growth factor expression. Molecular & Cellular Biology 21, 3995-4004 (2001).

178. Chen, Y., McPhie, D., Hirschberg, J. & Neve, R. The amyloid precursor protein-binding protein APP-BP1 drives the cell cycle through the S-M checkpoint and causes apoptosis in neurons. J Biol Chem 275, 8929-8935 (2000).

179. Ohh, M., et al. Ubiquitination of HIF requires direct binding to the von Hippel-Lindau protein beta domain. Nature Cell Biology 2, 423-427 (2000).

180. Ohh, M., et al. An intact NEDD8 pathway is required for Cullin-dependent ubiquitylation in mammalian cells. EMBO Reports 3, 177-182 (2002).

181. Vuento, M. & Vaheri, A. Dissociation of fibronectin from gelatin-agarose by amino compounds. Biochem J 175, 333-336 (1978).

182. Ohh, M., et al. Synthetic peptides define critical contacts between elongin C, elongin B, and the von hippel-lindau protein. J Clin Invest 104, 1583-1591 (1999).

183. Zheng, N., et al. Structure of the Cul1-Rbx1-Skp1-F boxSkp2 SCF ubiquitin ligase complex. Nature 416, 703-709 (2002).

184. Min, J.H., et al. Structure of an HIF-1alpha -pVHL complex: hydroxyproline recognition in signaling. Science 296, 1886-1889 (2002).

185. Feldman, D.E., Spiess, C., Howard, D.E. & Frydman, J. Tumorigenic mutations in VHL disrupt folding in vivo by interfering with chaperonin binding. Mol Cell 12, 1213-1224 (2003).

186. Lonergan, K.M., et al. Regulation of hypoxia-inducible mRNAs by the von Hippel-Lindau tumor suppressor protein requires binding to complexes containing elongins B/C and Cul2. Mol Cell Biol 18, 732-741 (1998).

187. Rathmell, W.K., et al. In vitro and in vivo models analyzing von Hippel-Lindau disease-specific mutations. Cancer Res 64, 8595-8603 (2004).

121

188. Ohh, M., et al. The von Hippel-Lindau tumor suppressor protein is required for proper assembly of an extracellular fibronectin matrix. Mol Cell 1, 959-968 (1998).

189. Ohh, M., et al. Ubiquitination of hypoxia-inducible factor requires direct binding to the beta-domain of the von Hippel-Lindau protein. Nat Cell Biol 2, 423-427 (2000).

190. Ohh, M., et al. Synthetic peptides define critical contacts between elongin C, elongin B, and the von Hippel-Lindau protein. J Clin Invest 104, 1583-1591 (1999).

191. Prchal, J.F. & Axelrad, A.A. Letter: Bone-marrow responses in polycythemia vera. N

Engl J Med 290, 1382 (1974).

192. Bellucci, S. & Michiels, J.J. The role of JAK2 V617F mutation, spontaneous erythropoiesis and megakaryocytopoiesis, hypersensitive platelets, activated leukocytes, and endothelial cells in the etiology of thrombotic manifestations in polycythemia vera and essential thrombocythemia. Semin Thromb Hemost 32, 381-398 (2006).

193. Chung, J., Roberts, A.M., Chow, J., Coady-Osberg, N. & Ohh, M. Homotypic association between tumour-associated VHL proteins leads to the restoration of HIF pathway. Oncogene 25, 3079-3083 (2006).

194. Dixon, C., et al. Overproduction of polypeptides corresponding to the amino terminus of the F-box proteins Cdc4p and Met30p inhibits ubiquitin ligase activities of their SCF complexes. Eukaryot Cell 2, 123-133 (2003).

195. Kominami, K., Ochotorena, I. & Toda, T. Two F-box/WD-repeat proteins Pop1 and Pop2 form hetero- and homo-complexes together with cullin-1 in the fission yeast SCF (Skp1-Cullin-1-F-box) ubiquitin ligase. Genes Cells 3, 721-735 (1998).

196. Suzuki, H., et al. Homodimer of two F-box proteins betaTrCP1 or betaTrCP2 binds to IkappaBalpha for signal-dependent ubiquitination. J Biol Chem 275, 2877-2884 (2000).

197. Tang, X., et al. Suprafacial orientation of the SCFCdc4 dimer accommodates multiple geometries for substrate ubiquitination. Cell 129, 1165-1176 (2007).

198. Kamura, T., et al. Muf1, a novel Elongin BC-interacting leucine-rich repeat protein that can assemble with Cul5 and Rbx1 to reconstitute a ubiquitin ligase. J Biol Chem 276, 29748-29753 (2001).

199. Rui, L., Yuan, M., Frantz, D., Shoelson, S. & White, M.F. SOCS-1 and SOCS-3 block insulin signaling by ubiquitin-mediated degradation of IRS1 and IRS2. J Biol Chem 277, 42394-42398 (2002).

200. Hon, W.C., et al. Structural basis for the recognition of hydroxyproline in HIF-1 alpha by pVHL. Nature 417, 975-978 (2002).

201. Bento, M.C., et al. Congenital polycythemia with homozygous and heterozygous mutations of von Hippel-Lindau gene: five new Caucasian patients. Haematologica 90, 128-129 (2005).

122

202. Shibata, J., et al. Hemostasis and coagulation at a hematocrit level of 0.85: functional consequences of erythrocytosis. Blood 101, 4416-4422 (2003).

203. Krishnamachary, B., et al. Hypoxia-inducible factor-1-dependent repression of E-cadherin in von Hippel-Lindau tumor suppressor-null renal cell carcinoma mediated by TCF3, ZFHX1A, and ZFHX1B. Cancer Res 66, 2725-2731 (2006).

204. Evans, A.J., et al. VHL promotes E2 box-dependent E-cadherin transcription by HIF-mediated regulation of SIP1 and snail. Mol Cell Biol 27, 157-169 (2007).

205. Evans, A.J., et al. VHL Promotes E2 Box-dependent E-cadherin Transcription by HIF-mediated Regulation of SIP1 and Snail. Mol Cell Biol (2006).

206. Gervais, M.L., et al. Nuclear E-cadherin and VHL immunoreactivity are prognostic indicators of clear-cell renal cell carcinoma. Lab Invest 87, 1252-1264 (2007).

207. Nakaigawa, N., et al. Inactivation of von Hippel-Lindau gene induces constitutive phosphorylation of MET protein in clear cell renal carcinoma. Cancer Res 66, 3699-3705 (2006).

208. Prockop, D.J. & Kivirikko, K.I. Collagens: molecular biology, diseases, and potentials for therapy. Annu Rev Biochem 64, 403-434 (1995).

209. Ferrajoli, A., Faderl, S., Ravandi, F. & Estrov, Z. The JAK-STAT pathway: a therapeutic target in hematological malignancies. Curr Cancer Drug Targets 6, 671-679 (2006).

210. Mullighan, C.G., et al. JAK mutations in high-risk childhood acute lymphoblastic leukemia. Proc Natl Acad Sci U S A 106, 9414-9418 (2009).

211. Epling-Burnette, P.K., et al. Inhibition of STAT3 signaling leads to apoptosis of leukemic large granular lymphocytes and decreased Mcl-1 expression. J Clin Invest 107, 351-362 (2001).

212. Melzner, I., Weniger, M.A., Menz, C.K. & Moller, P. Absence of the JAK2 V617F activating mutation in classical Hodgkin lymphoma and primary mediastinal B-cell lymphoma. Leukemia 20, 157-158 (2006).

213. Renne, C., et al. High expression of several tyrosine kinases and activation of the PI3K/AKT pathway in mediastinal large B cell lymphoma reveals further similarities to Hodgkin lymphoma. Leukemia 21, 780-787 (2007).

214. Mottok, A., Renne, C., Willenbrock, K., Hansmann, M.L. & Brauninger, A. Somatic hypermutation of SOCS1 in lymphocyte-predominant Hodgkin lymphoma is accompanied by high JAK2 expression and activation of STAT6. Blood 110, 3387-3390 (2007).