pi3k regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

221
PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary epithelial cells The Harvard community has made this article openly available. Please share how this access benefits you. Your story matters Citation Thorpe, Lauren Marie. 2015. PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary epithelial cells. Doctoral dissertation, Harvard University, Graduate School of Arts & Sciences. Citable link http://nrs.harvard.edu/urn-3:HUL.InstRepos:14226045 Terms of Use This article was downloaded from Harvard University’s DASH repository, and is made available under the terms and conditions applicable to Other Posted Material, as set forth at http:// nrs.harvard.edu/urn-3:HUL.InstRepos:dash.current.terms-of- use#LAA

Upload: others

Post on 11-Sep-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

PI3K regulatory subunit p85alphaplays a tumor suppressiverole in the transformation

of mammary epithelial cellsThe Harvard community has made this

article openly available. Please share howthis access benefits you. Your story matters

Citation Thorpe, Lauren Marie. 2015. PI3K regulatory subunit p85alphaplays a tumor suppressive role in the transformation of mammaryepithelial cells. Doctoral dissertation, Harvard University, GraduateSchool of Arts & Sciences.

Citable link http://nrs.harvard.edu/urn-3:HUL.InstRepos:14226045

Terms of Use This article was downloaded from Harvard University’s DASHrepository, and is made available under the terms and conditionsapplicable to Other Posted Material, as set forth at http://nrs.harvard.edu/urn-3:HUL.InstRepos:dash.current.terms-of-use#LAA

Page 2: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

PI3K regulatory subunit p85alpha plays a tumor suppressive role

in the transformation of mammary epithelial cells

A dissertation presented

by

Lauren Marie Thorpe

to

The Division of Medical Sciences

in partial fulfillment of the requirements

for the degree of

Doctor of Philosophy

in the subject of

Virology

Harvard University

Cambridge, Massachusetts

December 2014

Page 3: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

© 2014 Lauren Marie Thorpe

All rights reserved.

Page 4: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

iii

Dissertation Advisor: Dr. Jean J. Zhao Lauren Marie Thorpe

PI3K regulatory subunit p85alpha plays a tumor suppressive role

in the transformation of mammary epithelial cells

Abstract

Hyperactivation of the phosphatidylinositol 3-kinase (PI3K) pathway is one of the most

common events in human cancers. Class IA PI3Ks are heterodimers of a p110 catalytic

and a p85 regulatory subunit that coordinate the cellular response to extracellular stimuli.

Activating mutations in class IA PIγK catalytic isoform p110g are well established as

causative in a number of cancer types. More recently, mutation or loss of the class IA

regulatory isoform p85g (encoded by PIK3R1) has emerged as contributing to

oncogenesis. In this dissertation, we use both in vitro and in vivo approaches to examine

the role of p85g as a tumor suppressor in the transformation of mammary epithelial cells.

Using publically available online databases, we find heterozygous deletion of PIK3R1

occurs in 19-26% of breast tumors. Moreover, PIK3R1 expression is significantly

decreased in breast tumors compared to normal breast tissue. In human mammary

epithelial cells expressing dominant negative p53 (DDp53-HMECs), RNAi-mediated

knockdown of PIK3R1 increases PI3K/AKT activation in response to growth factor

stimulation and leads to transformation as assessed by anchorage-independent growth.

PIK3R1 knockdown also augments transformation of DDp53-HMECs by oncogenes,

including activated HER2/neu. In a mouse model of HER2/neu-driven breast cancer,

genetic ablation of Pik3r1 accelerates mammary tumor development. Transformation

driven by p85g loss is largely mediated by signaling through catalytic isoform p110g, as

Page 5: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

iv

selective pharmacological inhibition of p110g but not p110く effectively blocks colony

formation of PIK3R1 knockdown DDp53-HMECs and growth of Pik3r1 knockout tumors.

Mechanistically, we find that partial reduction of p85g increases the amount of p85-

p110g bound to activated receptors, augmenting PI3K signaling and oncogenic

transformation.

Together the work presented in this dissertation suggests that p85g depletion selectively

targets a free negative regulator pool of this regulatory subunit that modulates PI3K

activation under normal conditions, and transforms cells when lost. Furthermore, our

work indicates that p85g plays a tumor suppressive role in the pathogenesis of breast

tumors. Isoform-selective PI3K inhibitors are currently emerging in the clinic, and may

offer improved specificity and reduced toxicity over first-generation pan-PI3K inhibitors.

Our findings suggest p110g-selective therapies may be an effective treatment for breast

cancers with reduced p85g expression.

Page 6: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

v

Table of Contents Abstract iii Acknowledgements vi Index of Figures vii Index of Tables x List of Abbreviations xi Glossary of Terms xvi Chapter 1: Introduction 1 Chapter 2: PI3K regulatory subunit p85alpha plays a tumor suppressive role in human mammary epithelial cells 46 Chapter 3: PI3K regulatory subunit p85alpha plays a tumor suppressive role in a genetically engineered mouse model of mammary tumorigenesis 94 Chapter 4: Summary, discussion, and future directions 132 Materials and Methods 150 References 165 Appendix A: Supplemental table of class I PI3K alterations in cancer, with complete references 192 Appendix B: Supplemental table of genetically engineered mouse models of PI3K isoforms in cancer, with complete references 200

Page 7: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

vi

Acknowledgements

I am deeply grateful to the many people who have supported me both professionally and

personally throughout my years of study. Thank you to my undergraduate mentor, Dr.

Brooke McCartney, who gave me my first true research experience at Carnegie Mellon

University, and encouraged me to pursue graduate school. Thank you to my graduate

mentor, Dr. Jean Zhao, an insightful scientist and caring mentor, who has afforded me

many great opportunities during my time at Harvard University. Thank you to the past

and current members of the Zhao lab for their guidance, in particular Dr. Hailing Cheng,

Thanh Von, Stephanie Santiago, Dr. Linda Clayton, Carolynn Ohlson, and Dr. Haluk

Yuzugullu. Thank you to my Dissertation Advisory Committee, Drs. Karl Münger, Lewis

Cantley, and Myles Brown, for their time and scientific expertise over the years; I am

especially grateful to Karl for his mentoring and advice. Thank you to the Virology

program, in particular Dr. David Knipe, and to my cohort of twelve, the Vironauts. Thank

you to all of my friends and teammates, who have helped remind me that there is life

outside the lab, and that it is good. Thank you to Hyun Kim, who has supported me

every day and in every possible way. Thank you to my family: my parents Tom and Deb,

my sister Jessica, and my brother David. You’ve been there for me through it all. This

dissertation is dedicated to you.

Page 8: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

vii

Index of Figures Chapter 1

Figure 1.1 The PI3K family comprises multiple classes and isoforms 7 Figure 1.2 Signaling by class I PI3K isoforms 8 Figure 1.3 Signaling by class II PI3K isoforms 10 Figure 1.4 Signaling by class III PI3K isoforms 11 Figure 1.5 Divergent roles of class I PI3K catalytic isoforms in the context

of RTK, GPCR, and small GTPase inputs 19 Figure 1.6 Competition model for p110g and p110く regulation of RTK-

mediated PI3K signaling 20 Figure 1.7 Molecular contexts dictating applications for isoform-selective

PI3K inhibitors 21 Figure 1.8 Rational combination of PI3K inhibitors and other targeted

therapeutics 32 Chapter 2

Figure 2.1 PIK3R1 expression is significantly reduced in breast cancers 54 Figure 2.2 Generation of DDp53-HMECs with stable RNAi-mediated

PIK3R1 knockdown 57 Figure 2.3 PIK3R1 knockdown transforms DDp53-HMECs and increases

growth factor-stimulated PI3K/AKT activation 59 Figure 2.4 Augmented PI3K/AKT activation in PIK3R1 knockdown

DDp53-HMECs is rescued by ectopic expression of PIK3R1 61 Figure 2.5 Generation of DDp53-HMECs with activated HER2/neu

and RNAi-mediated PIK3R1 knockdown 63 Figure 2.6 PIK3R1 knockdown increases transformation and PI3K/AKT

signaling driven by activated HER2/neu in DDp53-HMECs 64 Figure 2.7 PIK3R1 knockdown increases transformation and PI3K/AKT

signaling driven by p110g-H1047R in DDp53-HMECs 66 Figure 2.8 Transformation of PIK3R1 knockdown DDp53-HMECs is

blocked by p110g-selective pharmacological inhibition 68

Page 9: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

viii

Figure 2.9 Transformation of PIK3R1 knockdown DDp53-HMECs with activated HER2/neu is blocked by p110g-selective inhibition 69

Figure 2.10 PI3K/AKT signaling in PIK3R1 knockdown DDp53-HMECs

with activated HER2/neu is blocked by p110g-selective inhibition 71 Figure 2.11 Endogenous p85g and PTEN do not appear to interact in

DDp53-HMECs 72 Figure 2.12 PTEN and p85g do not appear to interact in a variety of cell

types 74 Figure 2.13 PIK3R1 knockdown does not affect PTEN mRNA levels or lipid

phosphatase activity in DDp53-HMECs 77 Figure 2.14 PIK3R1 knockdown does not increase growth factor-stimulated

RTK phosphorylation in DDp53-HMECs 80 Figure 2.15 PIK3R1 knockdown does not affect growth factor-stimulated

RTK trafficking in DDp53-HMECs 82 Figure 2.16 PIK3R1 knockdown increases transformation of DDp53-HMECs

expressing activated ErbB3 84 Figure 2.17 PIK3R1 knockdown increases the amount of p85-p110g bound

to activated RTKs in DDp53-HMECs 86 Figure 2.18 Model: partial p85g loss leads to increased PI3K/AKT signaling

and transformation 87 Chapter 3

Figure 3.1 Schematic of mammary gland development in the mouse 97 Figure 3.2 Schematic of Pik3r1 conditional knockout allele and breeding

scheme for mammary-specific Pik3r1 ablation 101 Figure 3.3 Transgenic MMTV-Cre ablates Pik3r1 expression in mouse

mammary epithelial cells 103 Figure 3.4 Pik3r1 expression is not required for mouse mammary gland

development 104 Figure 3.5 PI3K/AKT pathway activation in spontaneous mammary tumors

from Pik3r1 knockout mice 109 Figure 3.6 Pathology of primary spontaneous mammary tumors and lung

metastases from Pik3r1 knockout mice 110 Figure 3.7 Adjacent mammary glands from Pik3r1 knockout mice with

spontaneous mammary tumors have a hypermorphic phenotype 111

Page 10: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

ix

Figure 3.8 Schematic of the transgenic NIC allele and breeding scheme for mammary-specific HER2/neu expression and Pik3r1 ablation 113

Figure 3.9 Pik3r1 ablation reduces the latency of HER2/neu-driven

mammary tumor development 114 Figure 3.10 Effect of Pik3r1 ablation on PI3K/AKT pathway activation in

HER2/neu-driven mammary tumors 116 Figure 3.11 Quantification of the effect of Pik3r1 ablation on PI3K/AKT

pathway activation in HER2/neu-driven mammary tumors 117

Figure 3.12 Effect of Pik3r1 ablation on tumor pathology and proliferation of HER2/neu-driven mammary tumors 119

Figure 3.13 Pan-PIγK or p110g-selective inhibitors block the growth of transplanted HER2/neu tumors with Pik3r1 ablation 121 Figure 3.14 Pan-PIγK or p110g-selective inhibitors suppress PI3K/AKT

activation in transplanted HER2/neu tumors with Pik3r1 ablation 123 Figure 3.15 Pan-PIγK or p110g-selective inhibitors suppress proliferation

and induce apoptosis in transplanted HER2/neu tumors with Pik3r1 ablation 124

Figure 3.16 Model: heterozygous or homozygous Pik3r1 ablation has a

similar effect on HER2/neu-driven tumorigenesis 130 Chapter 4

Figure 4.1 Model: modulation of p85g levels might produce a range of RTK-mediated PI3K activity 140

Page 11: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

x

Index of Tables

Chapter 1

Table 1.1 Class I PI3K isoform alterations in cancer 4 Table 1.2 Pan-PI3K inhibitors and their clinical applications 29 Table 1.3 Dual pan-PI3K/mTOR inhibitors and their clinical applications 30 Table 1.4 Isoform-selective PI3K inhibitors and their clinical applications 31 Table 1.5 Combination of PI3K inhibitors with other targeted therapies

in the clinic 33 Chapter 2

Table 2.1 PIK3R1 expression is significantly reduced in breast cancers across multiple microarray datasets 55

Chapter 3

Table 3.1 Nulliparous female mice with mammary-specific Pik3r1 ablation develop spontaneous mammary tumors 106

Table 3.2 Comparison of mammary tumor development in Pik3r1 knockout

mice to other established GEMMs of breast cancer 107

Page 12: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

xi

List of Abbreviations

4G10 antibody clone specific to phosphotyrosines

ALL acute lymphoblastic leukemia

AML acute myeloid leukemia

ANOVA analysis of variance

BCL2 B cell lymphoma 2

BCL-XL B cell lymphoma-extra large

BCR B cell receptor

BET bromodomain and extraterminal domain

BH breakpoint cluster homology domain

BSA bovine serum albumin

C/EBPく CCAAT/enhancer binding protein beta

Ca2+ calcium ion

CDK4 cyclin-dependent kinase 4

CDK6 cyclin-dependent kinase 6

CEF chicken embryo fibroblast

CLL chronic lymphocytic leukemia

CML chronic myelogenous leukemia

CRPC castration-resistant prostate cancer

CSF1 colony stimulating factor 1

DDp53 dominant negative p53 mutant

DM1 mertansine, a cytotoxic agent

DMEM Dulbecco’s modified eagle medium

DMEM/F12 Dulbecco’s modified eagle medium, nutrient mixture F1β

DTT dithiothreitol

EDTA ethylenediaminetetraacetic acid

Page 13: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

xii

EGF epidermal growth factor

EGFR epidermal growth factor receptor

ERK extracellular signal-regulated kinase

ESCC esophageal squamous cell carcinoma

FBS fetal bovine serum

FOXO forkhead box O transcription factors

GAP GTPase activating protein

GBM glioblastoma multiforme

GEF guanine nucleotide exchange factor

GEMM genetically engineered mouse model

GIST gastrointestinal stromal tumor

GPCR G-protein coupled receptor

H&E hematoxylin and eosin

HCC hepatocellular carcinoma

HMEC human mammary epithelial cell

hTERT human telomerase reverse transcriptase

ID2 inhibitor of DNA binding 2

IGF1 insulin-like growth factor 1

IHC immunohistochemistry

INHL indolent non-Hodgkin lymphoma

INPP4B type II inositol 3,4-bisphosphate 4-phosphatase

IP immunoprecipitation

IR insulin receptor

IRES internal ribosomal entry site

IRS1 insulin receptor substrate 1

iSH2 inter-SH2 domain

Page 14: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

xiii

JAK2 Janus kinase 2

KI gene knock-in

KO gene knockout

LPA lysophosphatidic acid

LTR long terminal repeat

MAPK mitogen-activated protein kinase

MCL mantle cell lymphoma

MCL1 myeloid cell leukemia sequence 1

MEF mouse embryo fibroblast

MEK MAPK/ERK kinase

miRNA micro RNA

MM multiple myeloma

MMEC mouse mammary epithelial cell

MMTV mouse mammary tumor virus

mRNA messenger RNA

MTM myotubularin family phosphatases

mTOR mammalian target of rapamycin

mTORC1 mTOR complex 1

NcrNu Ncr nude athymic mouse strain

NRG1 neuregulin 1

NSCLC non-small cell lung carcinoma

P proline-rich domain

p16INK4A cyclin-dependent kinase 4 inhibitor A

p85 BD p85-binding domain

p90RSK p90 ribosomal S6 kinase

PAGE polyacrylamide gel electrophoresis

Page 15: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

xiv

PARP poly-(ADP-ribose) polymerase

PBS phosphate buffered saline

PCR polymerase chain reaction

PHTS PTEN hamartoma tumor syndrome

PI phosphatidylinositide; see also PtdIns

PI3K phosphatidylinositol 3-kinase

PIN prostatic intraepithelial neoplasia

PIP phosphatidylinositol 3-phosphate; see also PtdIns(3)P

PIP2 PtdIns 4,5-bisphosphate; see also PtdIns(4,5)P2

PIP3 PtdIns 3,4,5-trisphosphate; see also PtdIns(3,4,5)P3

PPARけ peroxisome proliferator-activated receptor gamma

PPRE peroxisome proliferator response element

PR prolactin receptor

PtdIns phosphatidylinositide; see also PI

PtdIns(3)P phosphatidylinositol 3-phosphate; see also PIP

PtdIns(3,4)P2 PtdIns 3,4-bisphosphate

PtdIns(3,4,5)P3 PtdIns 3,4,5-trisphosphate; see also PIP3

PtdIns(4,5)P2 phosphatidylinositol 4,5-bisphosphate; see also PIP2

PTEN phosphatase and tensin homolog phosphatase

PyMT polyoma middle T

qPCR quantitative polymerase chain reaction

RANKL receptor activator of nuclear factor kappa-B ligand

RBD RAS-binding domain

RNAi RNA interference

RTK receptor tyrosine kinase

S6K p70 ribosomal protein S6 kinase

Page 16: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

xv

SABCS San Antonio Breast Cancer Symposium

SCCHN squamous cell carcinoma of the head and neck

SD standard deviation

SDS sodium dodecyl sulfate

SEM standard error of the mean

SHH sonic hedgehog

shRNA short hairpin RNA

SLL small lymphocytic leukemia

SMO smoothened

sqNSCLC squamous non-small cell lung cancer

STAT3 signal transducer and activator of transcription 3

STAT5 signal transducer and activator of transcription 5

T-ALL T cell acute lymphoblastic leukemia

TBS tris-buffered saline

TBST tris-buffered saline with tween

TCC transitional cell carcinoma

TCGA The Cancer Genome Atlas

TCR T cell receptor

T-DM1 conjugate of DM1 to the monoclonal antibody Trastuzumab

TEB terminal end bud

TNBC triple-negative breast cancer

TUNEL terminal deoxynucleotidyl transferase dUTP nick end labeling

UTR untranslated region

WCL whole cell lysate

Page 17: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

xvi

Glossary of Terms

Angiogenesis: the formation of new blood vessels from pre-existing vessels. Physiological angiogenesis is critical for normal growth and development, while pathophysiological angiogenesis is important for tumor growth. Myristoylated: irreversible co-translational modification of proteins in which a myristoyl group is covalently attached to an N-terminal amino acid of a nascent polypeptide, promoting membrane localization of the modified protein. Congenital mosaic overgrowth syndromes: a clinically heterogeneous group of genetic disorders characterized by abnormal progressive localized growth. They are caused by diverse somatic mutations and associated with increased cancer risk. Inter-SH2 (iSH2) domain: the domain of p85 regulatory PI3K isoforms that is located between the C- and N-terminal SH2 domains and directly interacts with class IA p110 catalytic isoforms. Megalencephaly syndromes: a collection of sporadic overgrowth disorders characterized by enlarged brain size and other distinct features. SH2 domain: SRC homology 2 domain; a structurally conserved protein–protein interaction domain that facilitates interaction with phosphorylated tyrosine residues on other proteins. RAS superfamily proteins: small monomeric membrane-associated GTPases, which are divided into the RAS, RHO, RAB, ARF, and RAN subfamilies based on structure and function. RAS GTPases: subfamily of RAS superfamily GTPases that plays critical roles in signal transduction. In mammals, the three major RAS subfamily members are HRAS, KRAS, and NRAS. RHO GTPases: subfamily of RAS superfamily proteins that shares similar roles in signal transduction to RAS GTPases and is best characterized for the regulation of cell shape, movement, and polarity. Xenograft: transplantation of living cells, tissues, or organs from one species to another. Human cell lines are often xenografted into mice to study factors affecting tumor growth.

Page 18: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

Chapter 1: Introduction

Page 19: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

2

Acknowledgements

A manuscript based on the work in this chapter was accepted for publication in Nature

Reviews Cancer and is currently in press with the following authors: Lauren Thorpe,

Haluk Yuzugullu, and Jean Zhao. This content has been used in accordance with Nature

Publishing Group policy, which states that authors retain the right to reproduce their

contribution in whole or in part in any printed volume of which they are an author. It has

been reformatted to adhere to dissertation formatting guidelines.

Haluk Yuzugullu and Jean Zhao wrote the section on therapeutic targeting of PI3K

isoforms in cancer, and Lauren Thorpe wrote all other sections. Lauren Thorpe, Haluk

Yuzugullu, and Jean Zhao edited the manuscript, and Lauren Thorpe and Haluk

Yuzugullu prepared the tables and figures.

We would additionally like to thank Tom Roberts for critical reading of the manuscript,

and Tom Roberts and Lewis Cantley for helpful discussions.

Page 20: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

3

Preface

Phosphatidylinositol 3-Kinases (PI3Ks) are critical coordinators of intracellular signaling

in response to extracellular stimuli. Hyperactivation of PI3K signaling cascades is one of

the most common events in human cancers. In this chapter, we discuss recent advances

in our knowledge of the roles of distinct PI3K isoforms in normal and oncogenic

signaling, the different ways in which PI3K can be upregulated, and the current state and

future potential of targeting this pathway in the clinic.

Introduction

Phosphatidylinositol 3-Kinases (PI3Ks) are a family of lipid kinases that integrate signals

from growth factors, cytokines, and other environmental cues, translating them into

intracellular signals that regulate multiple signaling pathways. These pathways control

many physiological functions and cellular processes, including cell proliferation, growth,

survival, motility, and metabolism (Engelman et al., 2006; Liu et al., 2009;

Vanhaesebroeck et al., 2010). Activating alterations in PI3K are frequent in a variety of

cancers (Table 1.1; for a fully referenced version see Appendix A), making this class of

enzymes a prime drug target (Engelman, 2009; Liu et al., 2009). Tremendous efforts

have been devoted to the development of effective PI3K inhibitors for cancer therapy.

Initial PI3K-directed drugs in clinical trials, consisting largely of non-isoform-selective

pan-PI3K inhibitors, have not yielded exciting results. However, recent preclinical studies

have demonstrated that different PI3K isoforms play divergent roles in cellular signaling

and cancer, suggesting that inhibitors targeting individual isoforms may be able to

achieve greater therapeutic efficacy. Isoform-selective inhibitors are now emerging in the

clinic, and have had promising success. In this chapter, we provide an update on what

has been learned in recent years about PI3K isoform-specific functions, differences in

the modes of PI3K isoform activation, and the progress of isoform-selective inhibitors in

Page 21: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

4

Table 1.1: Class I PI3K isoform alterations in cancer

Alteration Type Cancer Type Frequency of Alteration

Sample Size Range

Class IA

PIK3CA (p110g) Mutation Endometrial 10.3-53.0% 29-232 Breast 7.1-35.5% 65-507 Ovarian 33.0% 97 Colorectal 16.9

†-30.6%

72-195

Bladder 5.0-20.0% 20-130 Lung 0.6-20.0% 5-183 Cervical 13.6% 22 Glioblastoma 4.3-11.0% 91-291 Head and neck 8.1-9.4%

32-74

Esophageal 5.5%

145 Melanoma 5.0%

121

Prostate 1.3-3.6%

55-156 Sarcoma 2.9%

207

Renal 1.0-2.9%

98-417 Liver 1.6% 125 Megalencephaly

‡ 48.0% 50

Copy number Head and neck 9.1-100% 11-117 gain/amplification Cervical 9.1-76.4% 22-55 Lung 9.5-69.6% 3-92 Lymphoma 16.7-68.2% 22-60 Ovarian 13.3-39.8% 60-93 Gastric 36.4% 55 Thyroid 30.0% 110 Prostate 28.1% 32 Breast 8.7-13.4% 92-209 Glioblastoma 1.9-12.2% 139-206 Endometrial 10.3% 29 Thyroid 9.4% 128 Esophageal 5.7% 87 Leukemia 5.6% 161

Increased expression Prostate 40.0% 25

PIK3CB (p110く) Mutation Breast 0.5% 183

Copy number Lung 56.5% 46 gain/amplification Thyroid 42.3% 97 Ovarian 5-26.9% NA-93

Lymphoma 20.0% 60

Glioblastoma 5.8% 103 Breast 4.9-5% NA-81

Increased expression Prostate 46.7% 30 Glioblastoma 3.9% 103

PIK3CD (p110h) Copy number gain Glioblastoma 40.0% 10

Increased expression Neuroblastoma 52.6% 19 Glioblastoma 5.8% 103

Page 22: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

5

Table 1.1: Class I PI3K isoform alterations in cancer (continued)

Alteration Type Cancer Type Frequency of Alteration

Sample Size Range

Class IA

PIK3R1 (p85g, p55g, p50g) Mutation Endometrial 19.8-32.8% 108-243 Pancreatic 16.7% 6 Glioblastoma 7.6-11.3% 91-291 Colorectal 4.6

†-8.3% 108-195

Melanoma 4.4% 68 Ovarian 3.8% 80 Esophageal 3.4%

145

Breast 1.1-2.8% 62-507 Colon 1.7% 60

Decreased expression Breast 61.8% 458 Prostate 17-75%* NA Lung 19-46%* NA Ovarian 22%* NA Breast 18%* NA Bladder 18%* NA

Copy number loss Ovarian 21.5% 93

PIK3R2 (p85く) Mutation Endometrial 4.9% 243 Colorectal 0.9% 108 Megalencephaly

‡ 22.0% 50

Amplification Lymphoma 23.3% 60

Increased expression Colon 55.0% 20 Breast 45.7% 35

PIK3R3 (p55け) Copy number gain Ovarian 15.0% 93

Class IB

PIK3CG (p110け) Copy number gain Ovarian 19.3% 93

Increased expression Breast 77.5% 40

Prostate 72.4% 29

Medulloblastoma 52.9% 17

PIK3R5 (p101)

Mutation Melanoma 38.2% 68 Gastric 2.7% 37

For further detail and references, see the expanded version of this table in Appendix A. ‡ Megalencephaly syndromes are a collection of sporadic overgrowth disorders characterized by enlarged brain size and other distinct features. † Combined number of hypermutated and non-hypermutated colon and colorectal patient samples with mutations in the indicated gene. * Represents the percent reduction in gene expression. NA Sample size not available for this study.

Page 23: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

6

preclinical and early clinical studies.

Multiple PI3K classes and isoforms

PIγKs phosphorylate the γ’-hydroxyl group of phosphatidylinositides (PtdIns). They are

divided into three classes based on their structures and substrate specificities (Figure

1.1). In mammals, class I PI3Ks are further divided into subclasses IA and IB based on

their modes of regulation. Class IA PI3Ks are heterodimers of a p110 catalytic subunit

and a p85 regulatory subunit. The genes PIK3CA, PIK3CB, and PIK3CD respectively

encode three highly homologous class IA catalytic isoforms, p110g, p110く, and p110h.

These isoforms associate with any of five regulatory isoforms, p85g (and its splicing

variants p55g and p50g, encoded by PIK3R1), p85く (PIK3R2), and p55け (PIK3R3),

collectively called p85 type regulatory subunits (reviewed in (Engelman et al., 2006;

Mellor et al., 2012)). Class IB PIγKs are heterodimers of a p110け catalytic subunit

(encoded by PIK3CG) coupled with regulatory isoforms p101 (PIK3R5) or p87 (p84 or

p87PIKAP, encoded by PIK3R6). While p110g and p110く are ubiquitously expressed,

p110h and p110け expression is largely restricted to leukocytes (Okkenhaug and

Vanhaesebroeck, 2003).

In the absence of activating signals, p85 interacts with p110, inhibiting p110 kinase

activity. Upon receptor tyrosine kinase (RTK) or G-protein coupled receptor (GPCR)

activation, class I PI3Ks are recruited to the plasma membrane, where p85 inhibition of

p110 is relieved and p110 phosphorylates PtdIns 4,5-bisphosphate (PtdIns(4,5)P2) to

generate PtdIns(3,4,5)P3 (Figure 1.2). This lipid product acts as a second messenger,

activating AKT-dependent and –independent downstream signaling pathways (reviewed

in (Engelman et al., 2006; Liu et al., 2009; Vanhaesebroeck et al., 2010)). The

phosphatase and tensin homolog (PTEN) lipid phosphatase removes the γ’ phosphate

Page 24: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

7

Figure 1.1: The PI3K family comprises multiple classes and isoforms. PI3Ks are classified based on their substrate specificities and structures. In vivo, class IA and IB PI3Ks phosphorylate PtdIns(4,5)P2, while class III PI3Ks phosphorylate PtdIns. Some evidence suggests that class II PI3Ks may also preferentially phosphorylate PtdIns in vivo (Falasca et al., 2007; Maffucci et al., 2005; Yoshioka et al., 2012). Class IA PI3Ks are heterodimers of a p110 catalytic subunit and a p85 regulatory subunit. Class IA catalytic isoforms (p110g, p110く, and p110h) possess a p85-binding domain (p85 BD), RAS-binding domain (RBD), helical domain, and catalytic domain. Class IA p85 regulatory isoforms (p85g, p85く, p55g, p55け, and p50g) possess an inter-SH2 (iSH2) domain that binds class IA catalytic subunits, flanked by SH2 domains that bind phosphorylated YXXM motifs. The longer isoforms, p85g and p85く, additionally possess N-terminal SH3 and breakpoint cluster homology (BH) domains. Class IB PI3Ks are heterodimers of a p110け catalytic subunit and a p101 or p87 regulatory subunit. p110け possesses an RBD, helical domain, and catalytic domain. The domain structures of p101 and p87 are not fully known, but a C-terminal region of p101 has been identified as binding Gくけ subunits (Vadas et al., 2013). The monomeric class II isoforms (PI3K-Cβg, PI3K-Cβく, and PIγK-Cβけ) possess an RBD, helical domain, and catalytic domain. VPS34, the only class III PI3K, possesses helical and catalytic domains. VPS34 forms a constitutive heterodimer with the myristoylated, membrane-associated VPS15 protein. Other indicated domains include proline-rich (P) domains and membrane-interacting C2 domains. Modified with permission from (Liu et al., 2009).

Page 25: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

8

Figure 1.2: Signaling by class I PI3K isoforms. Upon receptor tyrosine kinase (RTK) or G-protein coupled receptor (GPCR) activation, class I PI3Ks are recruited to the plasma membrane by interaction with phosphorylated YXXM motifs on RTKs or their adaptors, or with GPCR-associated Gくけ subunits. There they phosphorylate PtdIns(4,5)P2 (PIP2) to generate PtdIns(3,4,5)P3 (PIP3), a second messenger which activates a number of AKT-dependent and –independent downstream signaling pathways regulating diverse cellular functions including growth, metabolism, motility, survival, and transformation. The phosphatase and tensin homolog (PTEN) lipid phosphatase removes the γ’ phosphate from PtdIns(γ,4,5)P3 to inactivate class I PI3K signaling. Modified with permission from (Liu et al., 2009).

Page 26: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

9

from PtdIns(3,4,5)P3 to inactivate PI3K signaling.

Relatively little is known about class II PI3Ks. There are three class II isoforms, PI3K-

Cβg, PIγK-Cβく, and PIγK-Cβけ, respectively encoded by PIK3C2A, PIK3C2B, and

PIK3C2G. These monomeric lipid kinases do not possess a regulatory subunit. PI3K-

Cβg and PIγK-Cβく are broadly expressed, while PIγK-Cβけ expression is limited to the

liver, prostate, and breast (Falasca and Maffucci, 2012). Although early experiments

indicated that PI3K-Cβg and PIγK-Cβく could phosphorylate both PtdIns and PtdIns(4)P,

in vivo PtdIns may be the preferred substrate, generating PtdIns(3)P (Falasca et al.,

2007; Maffucci et al., 2005; Yoshioka et al., 2012). The physiological roles of class II

PI3Ks are not fully understood, but recent studies suggest that PI3K-Cβg is important in

angiogenesis (Yoshioka et al., 2012) and primary cilium function (Franco et al., 2014). In

addition, PI3K-Cβg and PIγK-Cβく have been reported to regulate cellular functions

including growth and survival (reviewed in (Falasca and Maffucci, 2012;

Vanhaesebroeck et al., 2010)) (Figure 1.3).

The single class III PI3K, VPS34, is encoded by PIK3C3. VPS34 forms a constitutive

heterodimer with the myristoylated, membrane-associated VPS15 (encoded by PIK3R4),

and phosphorylates PtdIns to produce PtdIns(3)P (Schu et al., 1993; Volinia et al.,

1995). In mammals, VPS34 is ubiquitously expressed (Volinia et al., 1995). The VPS34-

VPS15 dimer is found in distinct multiprotein complexes, which have critical roles in

intracellular trafficking and autophagy (reviewed in (Backer, 2008; Vanhaesebroeck et

al., 2010)) (Figure 1.4). The myotubularin (MTM) family phosphatases MTM1 and

MTMRβ remove the γ’ phosphate from PtdIns(γ)P, regulating the lipid products of class

II and III PI3Ks (Blondeau et al., 2000; Cao et al., 2008; Lu et al., 2012; Velichkova et al.,

2010).

Page 27: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

10

Figure 1.3: Signaling by class II PI3K isoforms. Class II PI3Ks are not well understood, but may be activated by a number of different stimuli, including hormones, growth factors, chemokines, cytokines, phospholipids, and calcium (Ca2+). Although in vitro class II PI3Ks can phosphorylate both PtdIns and PtdIns(4)P, in vivo this class may preferentially phosphorylate PtdIns (PI) to generate PtdIns(3)P (PIP) (Falasca et al., 2007; Maffucci et al., 2005; Yoshioka et al., 2012). Class II PI3Ks regulate cellular functions including glucose transport, endocytosis, cell migration, and survival. Myotubularin (MTM) family phosphatases remove the γ’ phosphate from PtdIns(γ)P to inactivate class II PI3K signaling.

Page 28: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

11

Figure 1.4: Signaling by class III PI3K isoforms. The class III VPS34-VPS15 heterodimer is found in distinct multiprotein complexes, which perform specific cellular functions. VPS34 may be activated by stimuli including amino acids, glucose, and other nutrients, and phosphorylates PtdIns (PI) to generate PtdIns(3)P (PIP). It plays critical roles in autophagy, endosomal trafficking, and phagocytosis. MTM family phosphatases remove the γ’ phosphate from PtdIns(γ)P to inactivate class III PIγK signaling.

Page 29: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

12

Alterations of PI3K isoforms in cancer

Overactivation of the PI3K pathway is one of the most frequent events in human

cancers. The most common mechanism leading to aberrant PI3K signaling is somatic

loss of PTEN via genetic or epigenetic alterations (reviewed in (Parsons, 2004; Song et

al., 2012)). The PI3K pathway can also be upregulated by activation of RTKs, or

alterations in isoforms of PI3K itself (Table 1.1).

Class I PI3K catalytic isoform alterations

The transforming potential of class I PI3K catalytic isoforms was first demonstrated by

studies in the late 1990s and early β000s, which showed that fusion of p110g to viral

sequences (Chang et al., 1997) or the SRC myristoylation sequence (Klippel et al.,

1996; Zhao et al., 2003; Zhao et al., 2005) was activating and highly oncogenic. The

2004 discovery of frequent PIK3CA mutations in human cancers (Samuels et al., 2004)

brought PI3K to the forefront as a major cancer driver and potential drug target. PIK3CA

mutation has since been firmly established as causative in many cancer types (Table

1.1). Missense mutations occur in all domains of p110g, but the majority cluster in two

hotspots, the most common being E542K and E545K in the helical domain and H1047R

in the kinase domain. Cell-based analyses confirmed that these hotspot mutations

confer transformation via constitutive activation of p110g (Isakoff et al., 2005; Kang et

al., 2005; Zhao et al., 2005). Subsequently, several studies using genetically engineered

mouse models (GEMMs) demonstrated roles for mutant PIK3CA in tumor initiation,

progression, and maintenance (Engelman et al., 2008; Kinross et al., 2012; Liu et al.,

2011; Wu et al., 2013; Yuan et al., 2013) (Appendix B). Helical domain mutations

reduce inhibition of p110g by p85 (Burke et al., 2012; Huang et al., 2007; Miled et al.,

2007; Zhao and Vogt, 2010) or facilitate direct interaction of p110g with insulin receptor

substrate 1 (IRS1) (Hao et al., 2013), while kinase domain mutations increase interaction

Page 30: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

13

of p110g with lipid membranes (Burke et al., 2012; Huang et al., 2007; Mandelker et al.,

2009). Other PIK3CA mutations mimic distinct structural conformation changes that

occur during activation of PI3K (Burke et al., 2012). Interestingly, some of these

mutations in PIK3CA have also been reported in congenital mosaic overgrowth

syndromes (Kurek et al., 2012; Orloff et al., 2013; Rios et al., 2013; Riviere et al., 2012).

In contrast, mutations in other class I catalytic isoforms are rare. While activating

PIK3CD mutations have been described in immune deficiencies (Angulo et al., 2013;

Lucas et al., 2014), they have not been linked to cancer. One PIK3CB mutation was

detected in a single case of breast cancer (Kan et al., 2010); this helical domain

substitution enhances basal PIγK activation, potentially by increasing p110く association

with membranes (Dbouk et al., 2013). Recent structural studies have indicated that

p110く may be less inhibited by p85 (Dbouk et al., 2010; Vogt, 2011; Zhang et al., 2011)

and thus has higher basal transforming potential. Interestingly, p110h expression has

been detected in some human solid cancer cell lines (Sawyer et al., 2003), and

overexpression of wildtype p110く, p110h, or p110け, but not p110g, transforms cells in

vitro (Kang et al., 2006). This is consistent with the fact that PIK3CB, PIK3CD, and

PIK3CG are generally amplified or overexpressed, but not mutated, in cancers (Table

1.1).

Class I PI3K regulatory isoform alterations

Recent studies have converged to implicate the p85 regulatory isoforms in

tumorigenesis. Since the initial discovery of PIK3R1 mutations in human cancer cell lines

and primary tumors (Philp et al., 2001), somatic mutations in PIK3R1 have been

identified in a number of different cancers (Cancer Genome Atlas Research, 2008;

Cheung et al., 2011; Cizkova et al., 2013; Jaiswal et al., 2009; Urick et al., 2011) (Table

Page 31: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

14

1.1). The majority are substitutions or in-frame insertions or deletions in the inter-SH2

(iSHβ) domain of p85g (Cancer Genome Atlas Research, 2008; Cheung et al., 2011;

Jaiswal et al., 2009; Urick et al., 2011), the region of the protein that makes contact with

p110 (Huang et al., 2007), indicating this domain as a mutation hotspot (Cheung et al.,

2011). A number of these iSH2 domain mutants retain the ability to bind and stabilize

p110 isoforms, but promote enhanced PI3K activity and transformation due to reduced

ability to inhibit p110 (Cheung et al., 2011; Jaiswal et al., 2009; Sun et al., 2010; Urick et

al., 2011; Wu et al., 2009).

In addition, reduced expression of PIK3R1 has been reported in some cancers (Cizkova

et al., 2013; Taniguchi et al., 2010) (Table 1.1). PIK3R1 mRNA levels inversely

correlated with malignancy grade and incidence of metastasis in both breast and liver

cancers (Cizkova et al., 2013; Taniguchi et al., 2010). In mice, Pik3r1 ablation increased

epithelial neoplasia driven by Pten loss (Luo et al., 2005c) and led to spontaneous

development of aggressive liver tumors (Taniguchi et al., 2010). This work indicates that

p85g can negatively regulate PIγK signaling in cancer, and suggests that p85g has

tumor suppressive functions in certain tissues (Luo and Cantley, 2005).

Alterations in genes encoding other regulatory isoforms have also been detected, albeit

at a lower frequency. Increased PIK3R2 expression has been reported in breast and

colon cancers (Cortes et al., 2012) (Table 1.1). Consistent with this, overexpression of

wildtype p85く increased PIγK pathway activation in cells and tumor formation in mice

(Cortes et al., 2012). Somatic PIK3R2 mutations have been found in endometrial and

colorectal cancers (Cheung et al., 2011; Jaiswal et al., 2009), and causative germline

PIK3R2 mutations have been reported in megalencephaly syndromes (Riviere et al.,

2012). All PIK3R2 mutations described to date are substitutions with no apparent

Page 32: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

15

hotspot region, and similar to some p85g mutants, mutations in p85く increase PIγK

activation without affecting p110 binding (Cheung et al., 2011). Together these studies

indicate that PI3K regulatory isoforms may contribute to tumorigenesis by multiple

mechanisms.

Class II PI3K isoform alterations

Although class II PI3Ks are not well understood, PIK3C2A or PIK3C2B expression has

been implicated in physiological functions important to tumorigenesis (Biswas et al.,

2013; Diouf et al., 2011; Elis et al., 2008; Katso et al., 2006; Maffucci et al., 2005).

PIK3C2B amplification has been reported in glioblastoma (Knobbe and Reifenberger,

2003; Nobusawa et al., 2010; Rao et al., 2010), and somatic PIK3C2B mutations were

detected in non-small cell lung cancer (Liu et al., 2012), but the functional consequence

of these mutations is unknown. Perhaps the most convincing evidence towards a role for

class II PI3Ks in tumorigenesis comes from a recent study demonstrating that mice with

Pik3c2a ablation had compromised angiogenesis and vascular barrier integrity, and

significant reduction in the size and microvessel density of implanted tumors (Yoshioka

et al., 2012). Since mice with embryonic Pik3c2a or Pik3c2b knockout (KO) are viable

(Harada et al., 2005; Harris et al., 2011), a class II-selective PI3K inhibitor might target

tumor angiogenesis with tolerable side effects, although toxicity due to the critical role of

PI3K-Cβg in maintaining normal renal homeostasis (Harris et al., 2011) would need to be

considered.

Type II inositol 3,4-bisphosphate 4-phosphatase (INPP4B), the phosphatase responsible

for dephosphorylation of PtdIns(3,4)P2 to PtdIns(3)P (Gewinner et al., 2009; Norris et al.,

1997), has also been implicated in cancer. In human mammary cell lines, INPP4B

knockdown increased AKT activation and transformation (Fedele et al., 2010; Gewinner

Page 33: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

16

et al., 2009). INPP4B loss-of-heterozygosity has been detected in cancers (Gewinner et

al., 2009; Stjernstrom et al., 2014), and reduced INPP4B expression has been correlated

with high tumor grade, earlier recurrence, and decreased survival (Fedele et al., 2010;

Gewinner et al., 2009; Hodgson et al., 2011). Identification of INPP4B as a tumor

suppressor suggests that deregulation of the class II PI3K lipid products may contribute

to tumorigenesis.

Class III PI3K isoform alterations

There is currently little evidence indicating an oncogenic role for VPS34. One recent

study suggested that VPS34 is tyrosine-phosphorylated and activated downstream of

SRC, and its lipid kinase activity is required for SRC-mediated transformation (Hirsch et

al., 2010). However, overexpression of wildtype or myristoylated VPS34 was not

sufficient to induce cellular transformation (Denley et al., 2009). Another study indicated

that VPS34 activity might be decreased in the context of activated epidermal growth

factor receptor (EGFR) (Wei et al., 2013). Further investigation is needed to determine

whether VPS34 plays a role in transformation.

Divergent roles of class I PI3K catalytic isoforms

Class I PI3K catalytic isoforms share a conserved domain structure. They utilize the

same lipid substrates and generate the same lipid products. Despite their similarities,

accumulating evidence indicates these isoforms have distinct roles in mediating PI3K

signaling in physiological and oncogenic contexts.

GEMMs have been used to elucidate the roles of individual class I PI3K isoforms. Mice

with germline KO of Pik3ca or knock-in (KI) of a kinase-dead Pik3ca allele die at day

E10.5 (Bi et al., 2002; Bi et al., 1999). Interestingly, Pik3cb KO mice die much earlier at

Page 34: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

17

day E3.5 (Bi et al., 2002), while kinase-dead Pik3cb KI mice develop to maturity with

minor defects in size and glucose metabolism, and major defects in male fertility (Ciraolo

et al., 2008; Ciraolo et al., 2010). These differences suggest an important kinase-

independent scaffolding role for p110く (Ciraolo et al., 2008). Germline inactivation of

Pik3cd or Pik3cg by KO or KI of kinase-dead alleles yields viable mice that grow to

adulthood; however, loss of p110h results in functional defects in lymphocytes,

neutrophils, and mast cells (Ali et al., 2004; Clayton et al., 2002; Jou et al., 2002;

Okkenhaug et al., 2002), while loss of p110け impairs thymocyte development, T cell

activation, and neutrophil migration (Martin et al., 2008; Sasaki et al., 2000; Yum et al.,

2001). These studies indicate non-redundant roles in mouse embryonic development for

p110g and p110く, the two ubiquitously expressed class I PIγK isoforms, and distinct

roles in the immune system and inflammatory response for p110h and p110け, the two

leukocyte-restricted isoforms.

Technological developments have facilitated further insight into the individual roles of

PI3K enzymes. The generation of conditional KO animals using the Cre/loxP

recombination system has allowed the functions of each isoform to be studied in

different tissues, stages of development, and pathological settings (Appendix B).

Additional progress has come from studies using RNA interference (RNAi) and a new

generation of isoform-selective PI3K inhibitors. These have advanced our understanding

of the roles of class I catalytic isoforms in mediating signaling downstream of RTKs,

GPCRs, and small GTPases (Figure 1.5 and Figure 1.6), and in the context of PTEN

deficiency (Figure 1.7).

In mediating RTK signaling

Binding of growth factor ligands induces RTK dimerization, activation, and auto-

Page 35: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

18

phosphorylation of tyrosine-containing YXXM motifs on the receptors or their associated

adaptor proteins. Class IA p110-p85 heterodimers are then recruited to activated RTKs

through direct interaction of p85 SH2 domains with these phosphorylated YXXM motifs

(Rameh et al., 1995; Yu et al., 1998a; Yu et al., 1998b) (Figure 1.2). Accordingly p110g,

p110く, and p110h can complex with activated RTKs (Figure 1.5), and might be

expected to mediate growth factor signaling.

Studies using isoform-selective pharmacological inhibitors and genetic inactivation or

ablation indicated that loss of p110g activity was sufficient to largely block PI3K signaling

in response to a number of growth factors (Foukas et al., 2006; Graupera et al., 2008;

Knight et al., 2006; Sopasakis et al., 2010; Utermark et al., 2012; Zhao et al., 2006).

Notably, genetic ablation or inactivation of p110く had only a modest effect on PIγK

signaling following acute RTK activation (Ciraolo et al., 2008; Guillermet-Guibert et al.,

2008; Jia et al., 2008). It was suggested that the relative abundance of catalytic isoforms

in a particular tissue might dictate which isoforms are dominant in mediating RTK

signaling (Chaussade et al., 2007). This may explain the role of p110h, which is mainly

expressed in leukocytes and is the primary isoform regulating PI3K signaling

downstream of certain RTKs in mast cells and macrophages (Ali et al., 2004;

Papakonstanti et al., 2008; Vanhaesebroeck et al., 1999). However, differential

expression does not completely explain isoform dependence, as in many tissues p110く

levels are comparable to or even higher than levels of p110g (Geering et al., 2007).

The involvement of p110く in RTK signaling remained puzzling, until a recent study from

our group suggested a new model. In mice, while p110g ablation blocked normal

mammary development and mammary tumorigenesis driven by polyoma middle T

(PyMT) or HERβ (also known as ERBBβ), p110く ablation increased mammary gland

Page 36: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

19

Figure 1.5: Divergent roles of class I PI3K catalytic isoforms in the context of RTK, GPCR, and small GTPase inputs. Class I PI3Ks mediate signaling downstream of RTKs, GPCRs, and small GTPases. Left: p85 regulatory subunits bind phosphorylated YXXM motifs on activated RTKs. Because p110g, p110く, and p110h bind p85, these isoforms mediate signaling downstream of RTKs. Recent evidence also suggests that p87-p110け may be activated by certain RTKs (Schmid et al., 2011). Middle: Small GTPases synergize with RTK and GPCR signals to directly activate PI3Ks by interacting with their RAS-binding domains (RBDs). Isoforms p110g, p110h, and p110け bind RAS family GTPases, while p110く binds the RHO family GTPases RAC1 and CDC42 (Fritsch et al., 2013). Right: Gg and Gくけ proteins dissociate from activated GPCRs. Catalytic isoforms p110く and p110け, and regulatory isoform p101, directly bind and are activated by Gくけ. p110h may be activated downstream of GPCRs, but the mechanism is unknown (Durand et al., 2009; Reif et al., 2004; Saudemont et al., 2009). Gg proteins have been reported to directly bind and inhibit p110g (Ballou et al., 2006; Ballou et al., 2003; Yeung and Wong, 2010). Modified with permission from (Vanhaesebroeck et al., 2010).

Page 37: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

20

Figure 1.6: Competition model for p110g and p110く regulation of RTK-mediated PI3K signaling. Based on work presented in (Utermark et al., 2012). Both p85-p110g and p85-p110く compete for phosphorylated YXXM sites on activated RTKs. However, the maximal specific activity and enzymatic rate of p110g are higher than that of p110く (Beeton et al., 2000; Meier et al., 2004), and RTK-associated p110g may have higher lipid kinase activity than p110く (Utermark et al., 2012). By this model, loss or inactivation of p110g or p110く differentially modulates RTK signaling. Knockout of p110g allows all sites to be occupied by the less active p110く, decreasing RTK output. Conversely, knockout of p110く allows all sites to be bound by the more active p110g, increasing RTK output. Genetically or pharmacologically inactivated p110g or p110く can still bind RTKs but cannot signal, reducing RTK output.

Page 38: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

21

Figure 1.7: Molecular contexts dictating applications for isoform-selective PI3K inhibitors. Red box: Upregulation or mutation of receptor tyrosine kinases (RTKs), oncogenic RAS mutations, or activating p110g mutations all increase PtdIns(γ,4,5)P3

production through p110g, which can be amplified by mutation or loss of PTEN. In these contexts use of p110g-selective inhibitors is effective. Blue box: In the absence of other oncogenic alterations, PTEN loss or mutation increases PtdIns(3,4,5)P3 production through p110く, perhaps due to RAC1- or CDC42-mediated p110く activation, or the basal activity of this isoform. In this context use of p110く-selective inhibitors is effective. Green box: Upregulation or mutation of B cell receptors (BCRs), cytokine receptors, or other immune cell surface markers increases PtdIns(3,4,5)P3 production through p110h. In this context use of p110h-selective inhibitors is effective.

Page 39: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

22

outgrowth and accelerated tumor formation driven by these oncogenic RTKs (Utermark

et al., 2012). To explain this negative role of p110く, a competition model was proposed:

if p110g has higher RTK-associated lipid kinase activity than p110く, the less-active

p110く could compete with p110g for phosphorylated YXXM sites on receptors to

modulate PI3K signal strength downstream of RTKs (Utermark et al., 2012) (Figure 1.6).

Although direct comparison of RTK-associated p110g and p110く lipid kinase activity has

not been shown, the maximal specific activity and enzymatic rate of p110g are higher

than that of p110く (Beeton et al., 2000; Meier et al., 2004). Biochemical data were

consistent with this proposed model, demonstrating that in p110く KO cells, activated

RTKs had more bound p110g and higher associated lipid kinase activity (Utermark et al.,

2012). Furthermore, pharmacologically inactivated p110く could still compete with p110g

for binding sites on activated receptors, modestly reducing signaling and tumor growth

driven by PyMT or HER2 (Utermark et al., 2012). This model also explains moderately

decreased AKT activation, mild hyperglycemia, and delayed HER2-driven tumor

formation observed in mice with KI of kinase-dead p110く (Ciraolo et al., 2008), a

scenario mimicking p110く-selective kinase inhibition. These studies not only reveal a

novel p110く-based regulatory mechanism in RTK-mediated PI3K signaling, but also

identify p110g as an important target in cancers driven by oncogenic RTKs.

Initial studies suggested that class IA isoforms mediated signaling downstream of RTKs,

while the class IB isoform signaled downstream of GPCRs. Although p110け activation by

GPCRs is well established, a recent report suggested that this class IB isoform might

also function downstream of RTKs through regulatory isoform p87 in mouse myeloid

cells (Schmid et al., 2011) (Figure 1.5). Given that p87 and p101 may have distinct

tissue distribution (Bohnacker et al., 2009; Shymanets et al., 2013; Voigt et al., 2006)

and non-redundant functions (Bohnacker et al., 2009; Kurig et al., 2009; Schmid et al.,

Page 40: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

23

2011; Shymanets et al., 2013), this suggests that the two class IB regulatory isoforms

may mediate p110け activation in response to specific upstream signals.

In mediating GPCR signaling

GPCRs are a family of seven-transmembrane domain receptors that associate with

heterotrimeric G proteins composed of the Gg and Gくけ subunits. Ligand binding to

GPCRs results in allosteric activation and disassociation of bound G proteins into their

separate subunits, which can then act on intracellular targets.

The single class IB PIγK isoform, p110け, is activated by G proteins (Brock et al., 2003;

Maier et al., 1999; Stoyanov et al., 1995) (Figure 1.2). Although association of p110け

with either its p101 or p87 regulatory isoforms increased its activation in response to Gくけ

(Brock et al., 2003; Stephens et al., 1997; Suire et al., 2005), recent evidence indicated

that p101 is the main regulatory isoform involved in GPCR-mediated p110け signaling

(Kurig et al., 2009; Schmid et al., 2011) (Figure 1.5). Both p110け and p101 interact

directly with Gくけ heterodimers, and these contacts are critical for signaling and

transformation mediated by p110け (Brock et al., 2003; Vadas et al., 2013). Recent

studies have shown that in myeloid cells, p110け can be activated by GPCR and RTK

signals in a RAS- or RAP1A-dependent manner to mediate integrin g4く1 activity,

leading to tumor inflammation and progression (Schmid et al., 2011; Schmid et al.,

2013). Thus p110け-mediated signaling may contribute to tumorigenesis by controlling

both tumor cell characteristics and the tumor microenvironment.

Interestingly, in vitro experiments (Kubo et al., 2005; Kurosu et al., 1997; Maier et al.,

1999; Murga et al., 2000) and subsequent GEMM studies (Ciraolo et al., 2008;

Guillermet-Guibert et al., 2008; Jia et al., 2008) demonstrated a role for p110く in G

Page 41: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

24

protein-mediated PI3K signaling (Figure 1.2). Recently a region in the C2-helical domain

linker of p110く was shown to bind Gくけ subunits (Figure 1.5); this region is not present in

other class IA isoforms (Dbouk et al., 2012), and is similar to the region of p110け that

binds Gくけ (Vadas et al., 2013). Abrogation of p110く-Gくけ interaction blocked p110く-

mediated signaling and transformation downstream of GPCRs, and inhibited the

proliferation and invasiveness of cancer cells (Dbouk et al., 2012). Although p110h does

not directly interact with G proteins, a non-redundant role for this isoform in GPCR-

mediated leukocyte migration has been demonstrated in certain contexts (Durand et al.,

2009; Reif et al., 2004; Saudemont et al., 2009); however, the mechanism of p110h

activation downstream of GPCRs is unknown. It has also been reported that some Gg

proteins directly bind and inhibit p110g (Ballou et al., 2006; Ballou et al., 2003; Yeung

and Wong, 2010). Clearly, class I PI3K isoforms cooperate with GPCRs in a number of

different ways to regulate signaling and transformation.

Downstream of RAS and other small GTPases

RAS superfamily proteins are direct activators of the PI3K pathway. All class I PI3K

catalytic isoforms possess an N-terminal RAS-binding domain (RBD) (Figure 1.1)

allowing them to interact with RAS GTPases or other RAS superfamily members (Figure

1.5).

Activated or oncogenic mutant RAS proteins directly bind and increase the enzymatic

activity of both p110g (Rodriguez-Viciana et al., 1994; Rodriguez-Viciana et al., 1996)

and p110け (Pacold et al., 2000; Rubio et al., 1997; Suire et al., 2002). Cellular and

structural studies suggest that p110け association with RAS might both increase its

membrane translocation (Kurig et al., 2009; Pacold et al., 2000) and allosterically

increase p110け kinase activity (Pacold et al., 2000). Interestingly, RAS is required for

Page 42: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

25

activation of p110け bound to regulatory isoform p87, but not p101 (Kurig et al., 2009). In

vitro, the transforming capability of both helical domain p110g mutants (Zhao and Vogt,

2008, 2010) and of overexpressed wildtype p110け (Denley et al., 2008; Kang et al.,

2006) are dependent on their association with RAS. GEMM studies using KI of Pik3ca

with an RBD mutation or KO of endogenous Pik3ca revealed that the p110g-RAS

interaction is critical for both the initiation and maintenance of lung tumors (Castellano et

al., 2013; Gupta et al., 2007) and the development of myeloid leukemia (Gritsman et al.,

2014) driven by oncogenic KRAS. In mice, p110け-RAS binding is required for

inflammation-induced PtdIns(3,4,5)P3 accumulation (Suire et al., 2006) and

inflammation-associated tumor progression (Schmid et al., 2011; Schmid et al., 2013).

These studies highlight the importance of p110g or p110け interaction with RAS in both

normal PI3K signaling and transformation.

Although p110h was shown to bind RAS in vitro (Fritsch et al., 2013; Vanhaesebroeck et

al., 1997), some studies indicated that p110h kinase activity was not stimulated by

HRAS, NRAS, or KRAS, but instead by RRAS and TC21 (also known as RRAS2)

(Murphy et al., 2002; Rodriguez-Viciana et al., 2004). Furthermore, B and T cells derived

from Tc21 KO mice displayed diminished PIγK activity and recruitment of p110h to T cell

receptors (TCRs) and B cell receptors (BCRs), suggesting that TC21 might function

upstream of p110h (Delgado et al., 2009). Thus PI3K signaling through p110h may be

regulated by additional RAS subfamily members.

It was initially anticipated that all p110 isoforms bearing a RBD might interact with RAS

GTPases. Surprisingly, in vitro studies determined that p110く kinase activity was not

stimulated by any RAS subfamily members (Rodriguez-Viciana et al., 2004). A recent

extensive biochemical study demonstrated that p110く is instead regulated by RAC1 and

Page 43: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

26

CDC42 of the RHO GTPase subfamily (Fritsch et al., 2013) (Figure 1.5). Direct

interaction between the p110く RBD and RAC1 is important for GPCR-mediated

activation of p110く (Fritsch et al., 2013), indicating cooperative Gくけ and RHO GTPase

signaling through p110く. Previous studies reported that an intact RBD was required for

signaling and oncogenic transformation by wildtype p110く in cultured cells (Denley et

al., 2008; Kang et al., 2006), suggesting a potential role for RHO GTPase interaction

with p110く in transformation. Notably, RAC1 and CDC4β can also be activated

downstream of PI3K by PtdIns(3,4,5)P3-dependent guanine nucleotide exchange factors

(GEFs) and GTPase activating proteins (GAPs) (Klarlund et al., 1997; Krugmann et al.,

2002; Welch et al., 2002). The finding of distinct p110く regulation by RAC1 and CDC4β

expands PI3K signaling input by GTPases beyond the RAS subfamily, and also supports

the notion that PI3K can act both upstream and downstream of GTPases, potentially

allowing for positive feedback loops in cancer settings.

In PTEN deficiency

The PTEN lipid phosphatase counteracts class I PI3K activity, making it an important

tumor suppressor. Somatic loss of PTEN in human cancers is common. Germline PTEN

mutations are also found in several genetic disorders characterized by multiple

hamartomas with overgrowth phenotypes, collectively termed PTEN hamartoma tumor

syndromes (PHTS) (Liaw et al., 1997).

Pten KO mouse models provided a tool to explore the molecular mechanisms underlying

diseases caused by PTEN loss. While embryonic Pten KO is lethal (Di Cristofano et al.,

1998; Stambolic et al., 1998), heterozygous or conditional Pten KO animals

recapitulated human disease phenotypes, including development of prostate cancer

(Alimonti et al., 2010; Kwabi-Addo et al., 2001; Wang et al., 2003). Surprisingly, ablation

Page 44: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

27

of p110く, but not p110g, blocked prostatic intraepithelial neoplasia (PIN) induced by

PTEN loss (Jia et al., 2008). Subsequent studies demonstrated a correlation between

PTEN deficiency and sensitivity to p110く knockdown or inhibition in human cancer cell

lines both in vitro and in mouse xenografts (Ni et al., 2012; Torbett et al., 2008; Wee et

al., 2008). However, the mechanism governing the specific importance of p110く in the

context of PTEN loss remains elusive. Perhaps the unique role for p110く as a

convergence point for GPCR and RAC1 or CDC42 signals (Figure 1.7) contributes to

transformation induced by PTEN deficiency. Structural studies have also suggested that

compared to p110g, p110く is less inhibited by p85, and may supply a basal level of

PtdIns(3,4,5)P3 (Dbouk et al., 2010; Vogt, 2011; Zhang et al., 2011). This may explain

why wildtype p110く can be oncogenic when it is overexpressed (Denley et al., 2008;

Kang et al., 2006) or when PTEN is lost.

Although p110く is the primary PIγK isoform involved in many cases of tumorigenesis

driven by PTEN loss, studies have shown that depending on the tissue type and

pathology both p110g and p110く may be involved (Berenjeno et al., 2012; Jia et al.,

2013; Schmit et al., 2014). Mice with Pten ablation in the basal epidermal compartment

require both p110g and p110く for the development of hyperproliferative epidermal

lesions closely resembling PHTS (Wang et al., 2013a; Wang et al., 2013b). In this

model, spatially distinct roles for these isoforms in epidermal compartments were

identified: p110g is responsible for RTK signaling in and survival of suprabasal cells,

whereas p110く is important for GPCR signaling in and proliferation of basal cells (Wang

et al., 2013a). In mice with thymocyte-specific Pten KO, not surprisingly both p110h and

p110け were required for the development of T cell acute lymphoblastic leukemia (T-ALL)

(Subramaniam et al., 2012). This suggests that in certain contexts, transformation driven

by PTEN loss may be governed by the PI3K isoforms that are dominant in that tissue or

Page 45: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

28

compartment.

Since PTEN loss removes one mechanism of PI3K pathway negative regulation, the

specific roles of p110 isoforms in this pathogenic context can be influenced by other

activating inputs. These can be cues from the tissue microenvironment, or other co-

existing genetic events. A recent GEMM study demonstrated that concomitant activation

of oncogenic KRAS in ovarian endometrioid adenocarcinoma driven by Pten ablation

shifted the PIγK isoform reliance from p110く to p110g (Schmit et al., 2014) (Figure 1.7).

Consistent with this, a subset of PTEN-mutant human endometrioid endometrial cancer

cell lines harboring other PI3K-activating mutations were found to be resistant to p110く

inhibition (Weigelt et al., 2013). It is also possible that other genetic events downstream

of PI3K or in PI3K-independent pathways may render PTEN-null tumors less reliant on

PI3K. Thus determination of isoform dependency in PTEN-deficient tumors remains a

challenge.

Therapeutic targeting of PI3K isoforms in cancer

The central role of PI3K in cancer makes it an attractive therapeutic target. Enormous

efforts have focused on the development of drugs targeting PI3K, many of which are

undergoing clinical evaluation (Table 1.2, Table 1.3, and Table 1.4). Unlike drugs

targeting other oncogenic kinases, such as EGFR, BRAF, and ALK, PI3K inhibitors have

shown limited efficacy as mono-therapies in early trials on patients with tumors harboring

PI3K pathway activation (Rodon et al., 2013). The effectiveness of these early PI3K

inhibitors may have been limited by their lack of specificity, and by compensatory

signaling feedback loops and co-existing genetic and epigenetic alterations. The

development of novel isoform-selective PI3K inhibitors (Figure 1.7) and their rational

combination with other therapeutics (Figure 1.8 and Table 1.5) may substantially

Page 46: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

29

Table 1.2: Pan-PI3K inhibitors and their clinical applications

Agent Company Target Trial stage* Tumor types*

BKM120 Novartis Class I PI3Ks I, II, and III NSCLC Endometrial Thyroid CRPC Breast Colorectal Head and neck GBM Renal cell B cell lymphoma GIST Melanoma Ovarian Prostate Pancreatic Leukemia Esophageal Cervical Non-Hodgkin lymphoma Squamous NSCLC Adv. solid tumors

GDC0941 Genentech Class I PI3Ks I and II Breast NSCLC Non-Hodgkin lymphoma Adv. solid tumors

BAY80-6946 Bayer Class I PI3Ks I and II Non-Hodgkin lymphoma Adv. solid tumors

ZSTK474 Zenyaku Kogyo Co. Class I PI3Ks I and II Adv. solid tumors

PX866 Oncothyreon Class I PI3Ks I and II Colorectal SCCHN Melanoma NSCLC Prostate GBM Adv. solid tumors

XL147 Exelixis/Sanofi-Aventis

Class I PI3Ks I and II Breast Endometrial Ovarian Lymphoma GBM NSCLC Adv. solid tumors

CH5132799 Chugai Pharma Europe

Class I PI3Ks I Adv. solid tumors

NSCLC, non-small cell lung carcinoma; CRPC, castration-resistant prostate cancer; GIST, gastrointestinal stromal tumor; SCCHN, squamous cell carcinoma of the head and neck; GBM, glioblastoma multiforme. * Data taken from an April 2014 search of http://www.clinicaltrials.gov.

Page 47: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

30

Table 1.3: Dual pan-PI3K/mTOR inhibitors and their clinical applications

Agent Company Target Trial stage* Tumor types*

GDC0980 Genentech PI3K/mTOR I and II Prostate Breast Endometrial Renal cell Non-Hodgkin

lymphoma Adv. solid tumors

PF04691502 Pfizer PI3K/mTOR I Adv. solid tumors

BGT226 Novartis PI3K/mTOR I and II Adv. solid tumors

BEZ235 Novartis PI3K/mTOR I and II Breast Renal cell Prostate TCC ALL AML CML Pancreatic

neuroendocrine Adv. solid tumors

XL765 Sanofi PI3K/mTOR I GBM Adv. solid tumors

GSK2126458 GlaxoSmithKline PI3K/mTOR I Solid tumors Lymphoma

DS7423 Daiichi Sankyo PI3K/mTOR I Solid tumors

PWT33597 Pathway Therapeutics PI3K/mTOR I Adv. solid tumors

SF1126 Semafore Pharmaceuticals

PI3K/mTOR I Adv. solid tumors

PF05212384 Pfizer PI3K/mTOR I and II Adv. solid tumors

TCC, transitional cell carcinoma; ALL, acute lymphoblastic leukemia; AML, acute myeloid leukemia; CML, chronic myelogenous leukemia; GBM, glioblastoma multiforme. * Data taken from an April 2014 search of http://www.clinicaltrials.gov.

Page 48: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

31

Table 1.4: Isoform-selective PI3K inhibitors and their clinical applications

Agent Company Target Trial stage* Tumor types*

BYL719 Novartis p110g I and II SCCHN ESCC Colorectal Breast GIST Kidney Pancreas Gastric Adv. solid tumors

GDC0032 Genentech p110g I Breast Adv. solid tumors

INK1117 Intellikine/Millenium p110g I Adv. solid tumors

AZD8186 Astra-Zeneca p110く I CRPC sqNSCLC TNBC Adv. solid tumors with

PTEN deficiency

GSK2636771 GlaxoSmithKline p110く I and II Adv. solid tumors with PTEN deficiency

SAR260301 Sanofi p110く I Adv. solid tumors Lymphoma

IPI145 Infinity p110h and p110け

I, II, and III CLL SLL ALL INHL Hematologic malignancies

AMG319 Amgen p110h I Lymphoid malignancies

CAL101 (GS101)

Gilead Sciences p110h I, II, and III INHL CLL MCL SLL Hodgkin lymphoma Non-Hodgkin lymphoma Other lymphomas AML MM Hematologic malignancies

GS9820 Gilead Sciences p110く and p110h

I Lymphoid malignancies

SCCHN, squamous cell carcinoma of the head and neck; ESCC, esophageal squamous cell carcinoma; GIST, gastrointestinal stromal tumor; CRPC, castration-resistant prostate cancer; sqNSCLC, squamous non-small cell lung cancer; TNBC, triple-negative breast cancer; CLL, chronic lymphocytic leukemia; SLL, small lymphocytic leukemia; ALL, acute lymphoblastic leukemia; INHL, indolent non-Hodgkin lymphoma; MCL, mantle cell lymphoma; AML, acute myeloid leukemia; MM, multiple myeloma. * Data taken from an April 2014 search of http://www.clinicaltrials.gov.

Page 49: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

32

Figure 1.8: Rational combination of PI3K inhibitors and other targeted therapeutics. Pan-PI3K and dual pan-PI3K/mTOR inhibitors are currently being tested in clinical trials (grey box). These agents are being combined with mTOR-selective inhibitors (also in grey box), RAS-RAF-MEK-ERK pathway inhibitors (yellow box), RTK or other membrane-associated protein inhibitors (blue box), hormone signaling inhibitors (purple box), and other agents inhibiting the cell cycle, apoptosis machinery, or other signaling pathways (red box). Colored number symbols indicate targeted therapeutics currently in clinical trials for combination with the designated PI3K inhibitor. Asterisks denote targeted therapeutics expected to cooperate with PI3K therapies based on preclinical studies. For further detail, see Table 1.5.

Page 50: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

33

Ta

ble

1.5

: C

om

bin

ati

on

of

PI3

K in

hib

ito

rs w

ith

oth

er

targ

ete

d t

hera

pie

s in

th

e c

lin

ic

Ag

en

t C

om

pa

ny

Ta

rget

Co

mb

inati

on

th

era

py t

ria

ls

Ag

en

t T

arg

et

Tu

mo

r ty

pe

s*

Cli

nic

al tr

ial*

Cla

ss

I p

an

-PI3

K in

hib

ito

rs

BK

M1

20

No

va

rtis

C

lass I

PI3

Ks

L

ap

atin

ib

F

ulv

estr

an

t

Tra

stu

zu

ma

b

L

etr

ozo

le

E

GF

R/H

ER

2

E

R

H

ER

2

A

rom

ata

se

B

reast

NC

T0

158

9861

N

CT

01

33

9442

N

CT

01

13

2664

N

CT

01

24

8494

G

efitin

ib

E

rlo

tin

ib

E

GF

R

E

GF

R

N

SC

LC

N

CT

01

57

0296

N

CT

01

48

7265

P

an

itu

mu

mab

EG

FR

Co

lore

cta

l N

CT

01

59

1421

C

etu

xim

ab

E

GF

R

H

ead

an

d n

eck

NC

T0

181

6984

B

eva

ciz

um

ab

V

EG

FR

GB

M

R

ena

l ce

ll N

CT

01

34

9660

N

CT

01

28

3048

IN

C2

80

c-M

ET

GB

M

NC

T0

187

0726

R

itu

xim

ab

C

D2

0

B

ce

ll ly

mp

hom

a

NC

T0

204

9541

Im

atin

ib

B

CR

-AB

L

G

IST

N

CT

01

46

8688

V

em

ura

fen

ib

E

ncora

fen

ib

B

RA

F

B

RA

F

M

ela

nom

a

NC

T0

151

2251

N

CT

01

82

0364

O

lap

arib

P

AR

P

T

NB

C

O

va

ria

n

NC

T0

162

3349

A

bira

tero

ne

a

ce

tate

CY

P1

7

P

rosta

te

NC

T0

163

4061

E

rism

od

eg

ib

T

ram

etin

ib

M

EK

16

2

E

ve

rolim

us

S

mo

oth

ene

d

M

EK

1/2

ME

K1

/2

m

TO

R

A

dv. so

lid tu

mo

rs

NC

T0

157

6666

N

CT

01

15

5453

N

CT

01

36

3232

N

CT

01

47

0209

GD

C0

94

1

Ge

ne

nte

ch

Cla

ss I

PI3

Ks

F

ulv

estr

an

t

Tra

stu

zu

ma

b

E

R

H

ER

2

B

reast

NC

T0

143

7566

N

CT

00

92

8330

B

eva

ciz

um

ab

V

EG

FR

Bre

ast

N

SC

LC

N

CT

00

96

0960

N

CT

00

97

4584

E

rlo

tin

ib

C

ob

ime

tin

ib

E

GF

R

M

EK

1

A

dv. so

lid tu

mo

rs

NC

T0

097

5182

N

CT

00

99

6892

Page 51: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

34

Ta

ble

1.5

: C

om

bin

ati

on

of

PI3

K in

hib

ito

rs w

ith

oth

er

targ

ete

d t

hera

pie

s i

n t

he

clin

ic (

co

nti

nu

ed

)

Ag

en

t C

om

pa

ny

Ta

rget

Co

mb

inati

on

th

era

py t

ria

ls

Ag

en

t T

arg

et

Tu

mo

r ty

pe

s*

Cli

nic

al tr

ial*

Cla

ss

I p

an

-PI3

K in

hib

ito

rs

PX

86

6

On

co

thyre

on

Cla

ss I

PI3

Ks

C

etu

xim

ab

E

GF

R

C

olo

recta

l

SC

CH

N

NC

T0

125

2628

N

CT

01

25

2628

V

em

ura

fen

ib

B

RA

F

M

ela

nom

a

NC

T0

161

6199

XL

14

7

Exe

lixis

/ S

an

ofi-A

ve

ntis

Cla

ss I

PI3

Ks

T

rastu

zu

ma

b

L

etr

ozo

le

H

ER

2

A

rom

ata

se

B

reast

NC

T0

104

2925

N

CT

01

08

2068

E

rlo

tin

ib

M

M1

21

X

L6

47

E

GF

R

H

ER

3

R

TK

s

A

dv. so

lid tu

mo

rs

NC

T0

069

2640

N

CT

00

70

4392

N

CT

01

43

6565

Iso

form

-se

lec

tive

PI3

K i

nh

ibit

ors

BY

L7

19

No

va

rtis

p

11

0g

C

etu

xim

ab

E

GF

R

S

CC

HN

N

CT

01

60

2315

L

JM

71

6

H

ER

3

E

SC

C

NC

T0

182

2613

E

ncora

fen

ib

C

etu

xim

ab

B

RA

F

E

GF

R

C

olo

recta

l N

CT

01

71

9380

F

ulv

estr

an

t

ER

Bre

ast

A

dv. so

lid tu

mo

rs

NC

T0

208

8684

N

CT

01

21

9699

Im

atin

ib

B

CR

-AB

L

G

IST

N

CT

01

73

5968

L

etr

ozo

le

E

xe

me

sta

ne

T

DM

-1

L

EE

01

1

A

rom

ata

se

A

rom

ata

se

H

ER

2‡

C

DK

4/6

B

reast

NC

T0

187

0505

N

CT

01

87

0505

N

CT

02

03

8010

N

CT

02

08

8684

E

ve

rolim

us

E

xe

me

sta

ne

m

TO

R

A

rom

ata

se

B

reast

K

idn

ey

P

an

cre

as

NC

T0

207

7933

A

UY

92

2

H

SP

90

G

astr

ic

NC

T0

161

3950

G

an

itu

mab

B

GJ3

98

M

EK

16

2

IG

F1

R

F

GF

R

M

EK

1/2

A

dv. so

lid tu

mo

rs

NC

T0

170

8161

N

CT

01

92

8459

N

CT

01

44

9058

GD

C0

03

2

Ge

ne

nte

ch

p1

10g

L

etr

ozo

le

F

ulv

estr

an

t

Aro

ma

tase

E

R

B

reast

NC

T0

129

6555

Page 52: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

35

Ta

ble

1.5

: C

om

bin

ati

on

of

PI3

K in

hib

ito

rs w

ith

oth

er

targ

ete

d t

hera

pie

s i

n t

he

clin

ic (

co

nti

nu

ed

)

Ag

en

t C

om

pa

ny

Ta

rget

Co

mb

inati

on

th

era

py t

ria

ls

Ag

en

t T

arg

et

Tu

mo

r ty

pe

s*

Cli

nic

al tr

ial*

Iso

form

-se

lec

tive

PI3

K i

nh

ibit

ors

INK

11

17

Inte

llikin

e/

Mill

en

ium

p

11

0g

M

LN

01

28

m

TO

RC

1/2

Ad

v. n

on

-h

em

ato

log

ical

ma

lign

an

cie

s

NC

T0

189

9053

SA

R2

60

30

1

Sa

no

fi

p1

10

Ve

mu

rafe

nib

BR

AF

Me

lan

om

a

NC

T0

167

3737

IPI1

45

In

fin

ity

p1

10

h an

d

p1

10け

O

fatu

mu

mab

C

D2

0

C

LL

S

LL

NC

T0

204

9515

R

itu

xim

ab

C

D2

0

H

em

ato

log

ic

ma

lign

an

cie

s

NC

T0

187

1675

AM

G3

19

Am

ge

n

p1

10

h

NC

T0

130

0026

CA

L1

01

(G

S1

01

) G

ilea

d S

cie

nce

s

p1

10

h

Ritu

xim

ab

Ofa

tum

um

ab

E

ve

rolim

us

B

ort

ezo

mib

C

D2

0

C

D2

0

m

TO

R

N

FせB

IN

HL

C

LL

M

CL

NC

T0

108

8048

G

S9

97

3

S

YK

He

ma

tolo

gic

m

alig

nan

cie

s

NC

T0

179

6470

E

ve

rolim

us

m

TO

R

M

CL

NC

T0

108

8048

Du

al p

an

-PI3

K/m

TO

R in

hib

ito

rs

GD

C0

98

0

Ge

ne

nte

ch

PI3

K/m

TO

R

F

ulv

estr

an

t

ER

Bre

ast

NC

T0

143

7566

A

bira

tero

ne

a

ce

tate

CY

P1

7

P

rosta

te

NC

T0

148

5861

B

eva

ciz

um

ab

V

EG

FR

Bre

ast

A

dv. so

lid tu

mo

rs

NC

T0

125

4526

N

CT

01

33

2604

PF

04

691

502

Pfize

r P

I3K

/mT

OR

PD

03

25

901

M

EK

Ad

v. so

lid tu

mo

rs

NC

T0

134

7866

Page 53: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

36

Ta

ble

1.5

: C

om

bin

ati

on

of

PI3

K in

hib

ito

rs w

ith

oth

er

targ

ete

d t

hera

pie

s i

n t

he

clin

ic (

co

nti

nu

ed

)

Ag

en

t C

om

pa

ny

Ta

rget

Co

mb

inati

on

th

era

py t

ria

ls

Ag

en

t T

arg

et

Tu

mo

r ty

pe

s*

Cli

nic

al tr

ial*

Du

al p

an

-PI3

K/m

TO

R in

hib

ito

rs

BE

Z2

35

No

va

rtis

P

I3K

/mT

OR

Tra

stu

zu

ma

b

H

ER

2

B

reast

A

dv. so

lid tu

mo

rs

NC

T0

147

1847

N

CT

01

28

5466

E

ve

rolim

us

m

TO

R

B

reast

R

ena

l ce

ll

Ad

v. so

lid tu

mo

rs

NC

T0

148

2156

N

CT

01

50

8104

A

bira

tero

ne

a

ce

tate

CY

P1

7

P

rosta

te

NC

T0

171

7898

L

etr

ozo

le

A

rom

ata

se

B

reast

NC

T0

124

8494

E

ve

rolim

us

M

EK

16

2

m

TO

R

M

EK

Ad

v. so

lid tu

mo

rs

NC

T0

148

2156

N

CT

01

33

7765

XL

76

5

Sa

no

fi

PI3

K/m

TO

R

L

etr

ozo

le

A

rom

ata

se

B

reast

NC

T0

108

2068

E

rlo

tin

ib

E

GF

R

A

dv. so

lid tu

mo

rs

NC

T0

077

7699

PF

05

212

384

Pfize

r P

I3K

/mT

OR

PD

03

25

901

M

EK

Ad

v. so

lid tu

mo

rs

NC

T0

134

7866

C

etu

xim

ab

E

GF

R

C

olo

recta

l ca

ncer

NC

T0

192

5274

B

eva

ciz

um

ab

V

EF

GR

Co

lore

cta

l ca

ncer

NC

T0

193

7715

T

NB

C, tr

iple

-neg

ative b

rea

st

cancer;

GIS

T, ga

str

oin

testin

al str

om

al tu

mor;

NS

CLC

, n

on

-sm

all

cell

lung c

arc

inom

a; E

SC

C,

esoph

age

al sq

uam

ou

s c

ell

ca

rcin

om

a; S

CC

HN

, sq

uam

ou

s c

ell

ca

rcin

om

a o

f th

e h

ead a

nd

ne

ck; sq

NS

CLC

, squa

mous n

on

-sm

all

cell

lung c

ancer;

TC

C,

tran

sitio

nal cell

carc

inom

a; IN

HL,

indole

nt non

-Hodg

kin

lym

pho

ma

; M

CL,

mantle c

ell

lym

ph

om

a; C

LL

, chro

nic

ly

mpho

cytic le

ukem

ia;

SLL

, sm

all

lym

phocytic le

ukem

ia;

ALL

, acute

lym

phob

lastic le

uke

mia

; A

ML,

acute

myelo

id leukem

ia; C

ML,

ch

ron

ic m

yelo

ge

nou

s leuke

mia

; M

M,

multip

le m

yelo

ma; C

RP

C, castr

ation

-resis

tan

t pro

sta

te c

ancer;

GB

M, glio

bla

sto

ma m

ultiform

e.

* D

ata

taken f

rom

an

Apri

l 20

14 s

earc

h o

f http

://w

ww

.clin

icaltrials

.gov.

† Bevaciz

um

ab is a

mono

clo

na

l a

ntibod

y targ

etin

g V

EG

F th

at p

revents

sig

nalin

g thro

ug

h V

EF

GR

. ‡ T

-DM

1 is a

conju

ga

te o

f th

e c

yto

toxic

ag

ent

mert

an

sin

e (

DM

1)

to t

he m

onoclo

nal an

tib

ody T

rastu

zum

ab t

arg

eting H

ER

2.

Page 54: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

37

improve therapeutic outcomes.

Emerging isoform-selective PI3K inhibitors

Most PI3K inhibitors in early clinical trials are ATP-competitive agents that target all

class I isoforms with similar potencies. These include pan-PI3K inhibitors (Table 1.2)

such as GDC0941 (Raynaud et al., 2009) and dual pan-PI3K/mTOR inhibitors (Table

1.3) such as BEZ235 (Maira et al., 2008). Though these drugs display potent preclinical

anti-tumor activity, their success in clinical trials as single agents has been modest

(Rodon et al., 2013). The therapeutic window and efficacy of pan-PI3K inhibitors are

limited in some cases by adverse effects arising from a broader spectrum of off-target

effects (Fruman and Rommel, 2014). Furthermore, while both pan-PI3K and isoform-

selective inhibitors have on-target effects from suppression of essential PI3K functions,

for example glucose homeostasis, pan-PI3K inhibitors likely have additional on-target

effects from inhibiting isoforms that are not contributing to tumorigenesis. Isoform-

selective inhibitors may achieve greater efficacy with fewer toxic effects, and are

emerging in the clinic (Table 1.4).

The most effective single agent PI3K-based therapy to date is idelalisib (CAL101 or

GS1101), a p110h-selective inhibitor. Idelalisib has achieved notable success in early

trials for patients with chronic lymphocytic leukemia or indolent lymphoma, and is

currently in phase III clinical trials (Furman et al., 2014; Gopal et al., 2014). Interestingly,

this dramatic response is not due to genetic activation of the PI3K pathway, as neither

PI3K mutation nor PTEN loss is common in these malignancies. Given the important role

of p110h in signaling downstream of BCRs (Clayton et al., 2002; Jou et al., 2002;

Okkenhaug et al., 2002) and the fact that leukemic B cells have been shown to be

dependent on BCR signaling, it is likely that idelalisib functions by blocking essential

Page 55: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

38

BCR signals. Two recent articles provide great insight into the success of idelalisib trials

(see (Fruman and Cantley, 2014) and (Vanhaesebroeck and Khwaja, 2014)).

In addition to the role of p110h in B cell malignancies, a recent preclinical study showed

that this isoform also contributes to PTEN-null T-ALL (Subramaniam et al., 2012).

However, p110h-selective inhibition in this study was insufficient to suppress

tumorigenesis; combined inhibition of both p110h and p110け was required for effective

anti-PI3K therapy (Subramaniam et al., 2012). The involvement of p110h and p110け in

leukocyte signaling and hematological malignancies has drawn great attention, and new

inhibitors that target both isoforms simultaneously are in clinical trials for B and T cell

lymphomas (Table 1.4). These isoforms may also mediate immune responses that

support the growth of solid tumors. In a mouse model, p110け inhibition blocked myeloid

cell recruitment to tumors, thus suppressing malignancy by targeting the tumor

microenvironment (Schmid et al., 2011). Another study indicated that p110h inhibition

impaired tumor growth by disrupting regulatory T cell-mediated immune tolerance (Ali et

al., 2014). These findings indicate potential new applications for p110h- or p110け-

selective therapies in cancer.

The frequency of PIK3CA mutations in solid tumors has generated great interest in the

potential for p110g-selective inhibitors in targeting these cancers. Data presented at the

2013 San Antonio Breast Cancer Symposium (SABCS) indicated promising early clinical

activity of p110g-selective inhibitors BYL719 or GDC0032 as single agents in patients

with PIK3CA-mutant advanced breast tumors (Juric et al., 2013b). Recent preclinical

findings that HER2- or KRAS-driven tumors rely on p110g (Castellano et al., 2013;

Gritsman et al., 2014; Gupta et al., 2007; Schmit et al., 2014; Utermark et al., 2012)

underscore the need for clinical evaluation of p110g-selective drugs in these disease

Page 56: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

39

settings. In these studies, growth of HER2- or KRAS-driven solid tumors is inhibited

similarly by pan- and p110g-selective inhibitors (Castellano et al., 2013; Utermark et al.,

2012), but only modestly by p110く-selective inhibition (Schmit et al., 2014; Utermark et

al., 2012). However, further study is needed to determine the contexts in which

simultaneous inhibition of p110g and p110く can improve outcomes of KRAS- or HER2-

driven disease.

One drawback of p110g-selective inhibitors is their inevitable on-target adverse effects

on insulin signaling and glucose metabolism, since p110g is the major isoform mediating

these functions (Knight et al., 2006; Sopasakis et al., 2010). In the clinic, the effect of

p110g-selective inhibitors on glucose homeostasis must be carefully managed (Busaidy

et al., 2012), and is in some cases limiting (Rodon et al., 2013). To circumvent this,

inhibitors are being developed that specifically target p110g harboring hotspot mutations.

Such agents might be used at high doses with low toxicity, similar to mutant-selective

BRAF inhibitors that have had great clinical success (Bollag et al., 2010; Chapman et al.,

2011). A major obstacle to this approach is the heterogeneity of oncogenic PIK3CA

mutations. Some progress has been made with the discovery of GDC0032, which was

reported at the 2013 SABCS to have enhanced potency in PIK3CA mutant breast cancer

models (Sampath et al., 2013); one preclinical study also reported success using stapled

peptides to specifically disrupt the interaction of p110g-E545K with IRS1 (Hao et al.,

2013). However, devising strategies to selectively interrupt mutant-specific function

remains challenging. If developed, this class of inhibitor will likely be most effective in

early stage tumors with PIK3CA mutations, as advanced PIK3CA-mutant tumors may

have escaped their dependency on oncogenic p110g (Liu et al., 2011). Such drugs

would also be ideal for treating congenital overgrowth syndromes caused by PIK3CA

mutations occurring during early embryonic development (Kurek et al., 2012; Orloff et

Page 57: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

40

al., 2013; Rios et al., 2013; Riviere et al., 2012). In these contexts, p110g mutant-

selective inhibitors may yield improved therapeutic index.

Several preclinical studies have documented that certain PTEN-deficient tumors depend

on p110く (Jia et al., 2008; Ni et al., 2012; Wee et al., 2008), prompting a new clinical

trial with the p110く-selective inhibitor GSK2636771 in patients with PTEN-deficient

advanced solid tumors (NCT01458067). However, since PTEN is a negative regulator of

PI3K, isoform-dependency of PTEN-deficient tumors can be complicated as it can be

affected by tissue type, co-existing genetic events, and microenvironmental cues that

fuel cancer cells. In model systems where PTEN-deficient tumors are found to be

dependent on p110く, addition of oncogenic RTKs, RAS, or mutant PIK3CA can shift

dependency partially or totally to p110g (Figure 1.7). Recent studies also show that

prolonged treatment of PTEN-deficient tumor cells with p110く-selective inhibitors can

shift isoform dependency from p110く to p110g (N. Rosen, unpublished observations).

Therefore in most PTEN-deficient solid tumors, both p110g and p110く should be

targeted.

Although development of dual p110g/p110く-selective inhibitors has proven difficult

(Knight et al., 2006), combination of individual p110g- and p110く-selective inhibitors

might offer flexibility in the dosing of each isoform-selective inhibitor to further reduce

toxicity and increase the therapeutic window. One approach could involve continuous

inhibition of p110く to suppress elevated basal PIγK activity due to PTEN loss, combined

with pulsatile inhibition of p110g to avoid toxicity due to glucose elevation. Such a

strategy might also avoid the reported shift in isoform dependency of tumors from p110g

to p110く after prolonged treatment with the p110g-selective inhibitor BYL719 (J.A.

Engelman, unpublished observations). Ultimately, the success of targeting PI3K in

Page 58: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

41

cancer will likely require better understanding of which PI3K isoforms to target in a given

disease setting, improved inhibitors, and more careful dosing strategies.

Resistance mechanisms and combination therapeutic strategies

PI3K-based therapeutic approaches have encountered a number of roadblocks in the

form of intrinsic and acquired resistance mechanisms. A large body of work has

identified multiple signaling feedback loops, compensatory parallel signaling pathways,

and modes of downstream pathway activation that may result in clinical resistance to

PI3K inhibitors. Consequently, combination therapies are being developed and

evaluated in both preclinical and clinical settings (Figure 1.8 and Table 1.5), and will be

necessary to maximize clinical efficacy of PI3K inhibitors.

The first indication of feedback loops in the PI3K pathway came from experiments with

mTOR inhibitors. In early studies mTOR inhibition led to p70 ribosomal protein S6 kinase

(S6K) suppression, IRS1 upregulation, and PI3K-AKT activation (O'Reilly et al., 2006).

This prompted the development of dual pan-PI3K/mTOR inhibitors that are currently in

clinical trials (Table 1.3). Interestingly, feedback loops can also arise from dual pan-

PI3K/mTOR inhibition. A recent preclinical report suggested that PI3K and mTOR

blockade activated the Janus kinase 2 (JAK2)-signal transducer and activator of

transcription 5 (STAT5) signaling axis via IRS1, generating resistance to dual

PI3K/mTOR inhibition, which could be overcome by targeting JAK2 (Britschgi et al.,

2012). Similarly, in another preclinical study treatment with BEZ235 increased

phosphorylation of multiple signaling molecules, including STAT3, STAT5, JUN, and p90

ribosomal S6 kinase (p90RSK) (Muranen et al., 2012). Isoform-selective PI3K inhibitors

can also generate feedback loops: in a recent study of PIK3CA mutant breast tumors,

mTOR complex 1 (mTORC1) reactivation by insulin-like growth factor 1 (IGF1) and

Page 59: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

42

neuregulin 1 (NRG1) was associated with tumor resistance to the p110g-selective agent

BYL719, necessitating concurrent mTORC1 inhibition using RAD001 (Elkabets et al.,

2013). Inhibiting both PI3K and mTOR, possibly in conjunction with additional signaling

pathways, may be required to achieve effective anti-tumor activity.

Another important resistance mechanism to PI3K pathway inhibition is increased

expression of RTKs, such as HER3, IGF1R, insulin receptor (IR), and EGFR, via

forkhead box O (FOXO)-mediated transcriptional upregulation (Chandarlapaty et al.,

2011). Robust HER3 induction in response to PI3K inhibition has been reported in

several tumor types (Garrett et al., 2011; Muranen et al., 2012; Sergina et al., 2007).

While HER3 itself does not possess strong tyrosine kinase activity, it dimerizes with

EGFR, HER2, or HER4, hyperactivating the PI3K pathway and dampening the efficacy

of PI3K drugs. A preclinical study demonstrated that combination of the HER3-

neutralizing antibody LJM716 and the p110g-selective inhibitor BYL719 potently blocked

PI3K signaling and growth of HER2-positive breast tumor xenografts, even without a

direct HER2 antagonist (Garrett et al., 2013). Similarly, combination of the dual

EGFR/HER3 inhibitor MEHD7945A with a PI3K inhibitor (GDC0941) or AKT inhibitor

(GDC0068) effectively blocked the growth of triple-negative breast cancer cells in vitro

and in xenografts in a preclinical study (Tao et al., 2014). Blockade of PI3K along with

upstream RTKs may therefore circumvent certain PI3K therapy resistance mechanisms

(Figure 1.8).

Activation of convergent signaling pathways, for example the RAS-RAF-MEK-ERK

pathway, can also lead to PI3K pathway inhibition resistance. Mutant RAS can activate

both the RAF-ERK and PI3K-AKT-mTOR pathways in cancer cells; blocking the PI3K

pathway in such cells leads to upregulation of the ERK pathway (Serra et al., 2011).

Page 60: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

43

Inhibition of both PI3K and ERK pathways successfully suppressed the growth of cancer

cells in mouse models (Castellano et al., 2013; Engelman et al., 2008; Will et al., 2014),

and combinations of MEK inhibitors and pan- or isoform-selective PI3K agents are being

evaluated in clinical trials. However, there is preclinical evidence that some of these

combinations may be limited due to synergistic toxicity (Castellano et al., 2013).

Preclinical studies indicate that pulsatile inhibition of both PI3K and ERK pathways may

provide more effective anti-tumor activity while limiting toxic effects (Will et al., 2014),

suggesting that optimization of such combinations in the clinic will require careful dosing

strategies.

Another mode of resistance to PI3K-directed therapies arises from the activation of

transcription downstream or outside of the PI3K pathway. Several reports have indicated

MYC amplification or overexpression (Ilic et al., 2011; Liu et al., 2011) or activation of the

Notch and WNT/く-catenin pathways (Muellner et al., 2011; Tenbaum et al., 2012) as

mechanisms of resistance to PI3K inhibition. Recently, the bromodomain and

extraterminal (BET) inhibitor JQ1 has been shown to downregulate transcription of MYC,

among other targets (Delmore et al., 2011). XAV939 has also been identified as an

inhibitor of WNT/く-catenin-mediated transcription (Huang et al., 2009). Combination of

PI3K inhibition with these agents is being actively pursued in preclinical settings.

Other combination therapies have been suggested by assessing pathways that may

synergize with PI3K (Figure 1.8). As presented at the 2012 and 2013 SABCS, anti-

estrogen therapies are being tested in combination with PI3K inhibitors in clinical trials

for breast cancer patients (Juric et al., 2012; Juric et al., 2013a; Juric et al., 2013b). In a

brain tumor study, coordinate activation of sonic hedgehog (SHH) and PI3K signaling

was found in PTEN-deficient glioblastoma; combination of BKM120, a pan-PI3K

Page 61: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

44

inhibitor, and LED225, a smoothened (SMO) inhibitor that blocks SHH signaling,

resulted in synergistic anti-tumor effects (Gruber Filbin et al., 2013). Poly-(ADP-ribose)

polymerase (PARP) and PI3K inhibitors have been found to cooperate in prostate and

triple-negative breast cancers (Gonzalez-Billalabeitia et al., 2014; Ibrahim et al., 2012;

Juvekar et al., 2012). It appears that PI3K inhibition downregulates BRCA1 and BRCA2,

impairing homologous recombination and sensitizing BRCA-wildtype cancer cells to

PARP inhibition. Another attractive approach is combination of PI3K-targeted agents

with drugs that suppress anti-apoptotic factors. B cell lymphoma 2 (BCL2), myeloid cell

leukemia sequence 1 (MCL1), and other pro-survival proteins are frequently upregulated

in cancer, and may explain why PI3K inhibition is often cytostatic in tumor cells. BCL2 or

MCL1 suppression may induce cytotoxicity in response to PI3K inhibition (Rahmani et

al., 2013). Finally, an emerging approach is to combine PI3K inhibitors with agents that

disrupt cell cycle machinery (Vora et al., 2014). The p16-Cyclin D-cyclin-dependent

kinase 4 (CDK4)-CDK6 pathway is frequently dysregulated in cancer. A number of

CDK4/CDK6 inhibitors, including LEE011 and palbociclib (PD0332991), are entering

clinical trials for combination with pan- or p110g-selective inhibitors. Such rational

combination therapies will be required to increase the success of PI3K inhibitors.

Conclusions and perspective

Targeting the PI3K pathway remains both an opportunity and a challenge for cancer

therapy. Recent advances have provided the framework and rationale for inhibiting

select class I PI3K catalytic isoforms. We have learned a great deal about the divergent

roles of these isoforms in different signaling contexts, and are beginning to understand

the importance of each isoform in various tissues, compartments, and cancer types.

These findings have informed preclinical and clinical studies with isoform-selective PI3K

agents, which offer improved specificity and reduced toxicity over first-generation pan-

Page 62: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

45

PI3K drugs. Isoform-selective PI3K inhibitors have seen promising success in early- and

late-stage clinical trials for solid and hematological malignancies, highlighting the

potential for isoform-selective PI3K therapeutics.

Although we have made substantial progress, further efforts are needed. We have only

recently begun to appreciate the importance of class I regulatory isoforms in

tumorigenesis. The different ways in which p85 subunits contribute to cancer, and the

effective means to pharmacologically inhibit these mechanisms, are still not fully

understood. Similarly, while a recent study indicates that class II isoform PI3K-Cβg is

important for pathophysiological angiogenesis, the roles of class II and III PI3Ks in

cancer remain unclear.

For the class I catalytic isoforms, we must continue to precisely define the disease

settings in which different PI3K isoforms will need to be targeted. To better inform

isoform-selective therapeutic strategies, a set of biomarkers to predict the active p110

isoforms in a given tumor would be ideal, but development of this will require systematic

studies. Continued work to understand the underlying cellular programs that protect

tumors with aberrant PI3K activation from PI3K-targeted therapy will also be important.

This will allow for better rational design of combination therapies, which will be

necessary to overcome compensatory pathway activation and acquired resistance

mechanisms and maximize the anti-tumor activity of PI3K inhibitors. Dosing strategies

will also need to be carefully considered, as recent studies suggest that in some cases

pulsatile inhibition may reduce toxicity without sacrificing efficacy. Progress in these

areas should increase the effectiveness of PI3K-directed therapies in the clinic.

Page 63: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

Chapter 2: PI3K regulatory subunit p85alpha plays a tumor

suppressive role in human mammary epithelial cells

Page 64: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

47

Acknowledgements

Hailing Cheng and Jean Zhao conceived the initial project idea. Lauren Thorpe

performed all experiments and data analysis.

I would additionally like to thank Hailing Cheng for providing HMECs and plasmids, for

teaching initial techniques, and for assistance in troubleshooting experiments. Thank you

to Haluk Yuzugullu for assistance in subcloning the neuT construct. Thank you to

Jennifer Spangle for providing detailed protocols for receptor internalization and

degradation assays. I would also like to thank Lewis Cantley for helpful discussions

conceptualizing the mechanism portion of this work, and Lewis Cantley, Karl Münger,

and Myles Brown for helpful discussions on experimental confirmation of this

mechanism.

Page 65: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

48

Preface

In this chapter, we use human mammary epithelial cells (HMECs), a normal, non-

transformed cell culture system, to examine the in vitro effects of partial p85g loss on

PI3K signaling and cell transformation. We find that RNAi-mediated PIK3R1 knockdown

increases growth factor-dependent PIγK signaling through catalytic isoform p110g,

facilitating cellular transformation. We furthermore demonstrate that knockdown of

PIK3R1 augments PI3K signaling and transformation mediated by oncogenes, including

activated HERβ/neu. Finally, we show that partial reduction of p85g leads to an

increased amount of p110g bound to activated receptor tyrosine kinases (RTKs).

Together these findings suggest that p85g depletion selectively targets a free negative

regulator pool of the PI3K regulatory subunit that fine tunes RTK-mediated PI3K

activation under normal conditions, and transforms cells when lost.

Introduction

Class IA phosphatidylinositol 3-kinases (PI3Ks) are critical coordinators of the cellular

response to extracellular signals. They are heterodimers comprised of a p110 catalytic

subunit (p110g, p110く, or p110h) and a p85 regulatory subunit (p85g, p55g, p50g,

p85く, or p55け, collectively referred to as p85) (Figure 1.1). In quiescent cells, the iSH2

domain of the p85 regulatory subunit binds and stabilizes the p110 catalytic subunit

while maintaining p110 in a low-activity state (Yu et al., 1998b). Upon growth factor

stimulation, the SH2 domains of p85 bind tyrosine-phosphorylated YXXM motifs on

activated RTKs or their adaptors, recruiting p110 to the plasma membrane and

simultaneously relieving its inhibition (Yu et al., 1998a) (Figure 1.2). The activated p110

catalytic subunit phosphorylates the γ’-hydroxyl group of phosphatidylinositol 4,5-

bisphosphate (PtdIns(4,5)P2) to produce phosphatidylinositol 3,4,5-trisphosphate

(PtdIns(3,4,5)P3), a cellular second messenger which goes on to activate multiple AKT-

Page 66: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

49

dependent and –independent downstream signaling pathways. PI3K signaling controls

diverse cellular activities, including cell growth, proliferation, survival, and transformation

(Liu et al., 2009; Wong et al., 2010). The phosphatase and tensin homologue (PTEN)

lipid phosphatase removes the γ’ phosphate from PtdIns(γ,4,5)P3 to inactivate PI3K

signals.

Several reports have linked p85g to activity or stability of the PTEN lipid phosphatase. In

liver lysates from mice with liver-specific Pik3r1 knockout, PtdIns(3,4,5)P3 production

was more sustained and AKT activation was increased, while the lipid phosphatase

activity of PTEN was found to be reduced, suggesting that p85g may enhance PTEN

activity (Taniguchi et al., 2006). This finding was corroborated by in vitro assays, where

addition of increasing amounts of purified p85g augmented dephosphorylation of

PtdIns(3,4,5)P3 by PTEN (Chagpar et al., 2010). Furthermore, co-immunoprecipitation

experiments have suggested an interaction between p85g and PTEN (Chagpar et al.,

2010; Cheung et al., 2011; Rabinovsky et al., 2009); this interaction was shown to be

direct and to require the N-terminal SHγ and BH domains of p85g (Chagpar et al., 2010).

In addition to effects on PTEN lipid phosphatase activity, p85g has been linked to PTEN

expression and protein stability. Although liver-specific ablation of Pik3r1 had no effect

on the levels of PTEN protein in mouse liver lysates (Taniguchi et al., 2006; Taniguchi et

al., 2010), these mice developed liver tumors within 14-20 months, and lysates from

these tumors had significantly reduced PTEN protein and mRNA levels (Taniguchi et al.,

2010). In a separate study, ectopic expression of wildtype p85g increased PTEN protein

levels, while expression of the endometrial cancer-associated p85g-E160* mutant

enhanced ubiquitination and proteasomal degradation of PTEN; in these contexts, PTEN

mRNA levels were unchanged (Cheung et al., 2011). Together these findings suggest

that p85g may affect PTEN expression, stability, or activity, potentially contributing to

Page 67: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

50

negative regulation of the PI3K pathway.

In addition to its reported effects on PTEN, p85g has also been shown to have a

GTPase activating protein (GAP) function towards select Rab proteins. Rabs are a large

family of small GTPases critical for the regulation of intracellular trafficking (Jean and

Kiger, 2012; Stenmark, 2009). One mechanism controlling receptor-mediated signaling

is the endocytosis of receptors following their activation; receptors are targeted to the

early endosome, then either transported to the lysosome for degradation, terminating

signaling, or recycled back to the plasma membrane, allowing for potentially sustained

signaling (Miaczynska, 2013). Of the many Rab proteins, in particular Rab4 and Rab5

are important for transport from the plasma membrane to early endosomes, and Rab11

is important for recycling of endosomes back to the plasma membrane (Jean and Kiger,

2012; Stenmark, 2009). The N-terminal breakpoint cluster homology (BH) domain of

p85g has been reported to directly bind and have GAP activity towards Rab5; in

addition, p85g showed GAP activity toward Rab4, but not Rab11 (Chamberlain et al.,

2004). Ectopic expression of a p85g mutant with disrupted Rab-GAP activity led to

increased growth factor-stimulated PI3K/AKT activation and cellular transformation

(Chamberlain et al., 2004; Chamberlain et al., 2008), apparently due to more rapid and

sustained internalization and reduced degradation of RTKs (Chamberlain et al., 2008;

Chamberlain et al., 2010). Similarly, a recent report demonstrated that RNAi-mediated

p85g downregulation lead to an increase in active GTP-bound Rab5 and PI3K/AKT

pathway activation (Dou et al., 2013). Together this suggests that p85g might also

contribute to regulation of PI3K signaling by effects on activation of Rab proteins

controlling receptor trafficking.

Hyperactivation of the PI3K pathway is one of the most common events in human

Page 68: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

51

cancers (Table 1.1 and Appendix A). This can be a result of lesions along the pathway,

such as loss of the PTEN phosphatase opposing PI3K (Cully et al., 2006), or of

mutations in PI3K itself. Oncogenic mutations frequently occur in the PIK3CA gene

encoding catalytic isoform p110g (Cancer Genome Atlas Research, 2008; Parsons et

al., 2008; Samuels et al., 2004; Thomas et al., 2007), and confer constitutive p110g

activation leading to transformation (Gymnopoulos et al., 2007; Hon et al., 2012; Isakoff

et al., 2005; Kang et al., 2005; Mandelker et al., 2009; Samuels et al., 2005; Zhao et al.,

2005). More recently, somatic mutations in PIK3R1 (encoding p85g and its splicing

variants p55g and p50g) and PIK3R2 (encoding p85く) have been reported in

glioblastoma (Cancer Genome Atlas Research, 2008; Parsons et al., 2008) and in

endometrial cancer (Cheung et al., 2011; Urick et al., 2011). A majority of PIK3R1

mutations occur in iSH2 domain hotspots (Cancer Genome Atlas Research, 2008;

Cheung et al., 2011; Jaiswal et al., 2009; Urick et al., 2011); subsequent studies

demonstrated that many of these iSH2 mutants retain the ability to bind and stabilize

p110, but are less able to inhibit p110 catalytic activity, resulting in increased PI3K

activation and transformation (Cheung et al., 2011; Jaiswal et al., 2009; Urick et al.,

2011; Wu et al., 2009). A smaller fraction of PIK3R1 mutations found in cancers occur in

other domains of the protein; interestingly, some of these mutants lack the iSH2 domain

and are unable to bind p110 (Cheung et al., 2011; Jaiswal et al., 2009; Urick et al.,

2011). Ectopic expression of the truncation mutant p85g-E160* has been reported to

contribute to destabilization of PTEN protein (Cheung et al., 2011), but the effects of

other truncation mutations has not yet been determined. The discovery and

characterization of mutations in PIK3R1 has established a previously unreported role for

the p85 regulatory subunit in cancer, potentially by multiple distinct mechanisms.

In addition to oncogenic mutations in PIK3R1, accumulating evidence suggests that

Page 69: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

52

changes in the levels of p85g can have an important effect on PI3K activation. In mice,

although complete Pik3r1 ablation is perinatally lethal (Fruman et al., 2000), partial loss

of different p85 isoforms improves insulin sensitivity and insulin-stimulated AKT

activation (Chen et al., 2004; Mauvais-Jarvis et al., 2002; Terauchi et al., 1999; Ueki et

al., 2002a; Ueki et al., 2002b). It has been suggested that in some tissues, the p85

regulatory subunit is present in excess of p110, and that monomeric free p85 is capable

of acting as a negative regulator of PI3K signaling (Luo and Cantley, 2005; Mauvais-

Jarvis et al., 2002; Ueki et al., 2002a; Ueki et al., 2003). In support of this model, p85g

overexpression in L6 myotubes significantly reduced insulin-stimulated PI3K/AKT

activation (Ueki et al., 2000), and mice with p85g overexpression had decreased skeletal

muscle insulin signaling (Barbour et al., 2005). Furthermore, monomeric p85g has been

shown to downregulate insulin-stimulated PI3K activity by forming a sequestration

complex with non-signaling IRS1 adaptors (Luo et al., 2005a). Loss of p85g has also

recently been proposed to play a role in cancer. Significant PIK3R1 underexpression

was recently detected in breast cancers, and correlated with poorer metastasis-free

survival (Cizkova et al., 2013). In mice, liver-specific Pik3r1 deletion eventually led to

development of aggressive high-grade hepatocellular carcinoma (HCC)-like tumors with

upregulated AKT activation (Taniguchi et al., 2010). Together these studies indicate that

partial reduction of p85 can upregulate PI3K in response to insulin, and that furthermore

loss of p85 may contribute to tumorigenesis in the breast and liver.

In this chapter, we examine the importance of p85g levels in regulating PIγK/AKT

activation in mammary epithelial cells. We use RNAi techniques to downregulate

PIK3R1 expression in human mammary epithelial cells, and explore the effects on

PI3K/AKT signaling and cellular transformation. We then use isoform-selective

pharmacological inhibitors to identify the PI3K catalytic isoforms that contribute to

Page 70: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

53

signaling in the context of reduced p85g. Finally, we use immunoprecipitation

experiments to identify a potential mechanism linking the levels of p85g to the magnitude

of RTK-mediated PI3K/AKT signaling activation. Our findings support previous reports

that p85g levels modulate PIγK output, and importantly demonstrate that a reduction in

p85g is sufficient to transform mammary epithelial cells in vitro.

Results

PIK3R1 expression is significantly reduced in breast cancer

To determine whether p85g may play a tumor suppressive role in breast cancer, we

analyzed expression levels of the PIK3R1 gene encoding p85g in different publically

available datasets from breast cancer patients. We used the cBioPortal for Cancer

Genomics (http://www.cbioportal.org, (Cerami et al., 2012; Gao et al., 2013)) to query

data from a 2012 comprehensive study of human breast tumors by The Cancer Genome

Atlas Network (TCGA) (Cancer Genome Atlas, 2012) and an additional provisional

TCGA data set for copy number loss (as determined by GISTIC) or mutation of PIK3R1.

Such alterations in PIK3R1 occurred in 23% and 28% of breast cancer cases in these

studies respectively, with the vast majority of these being heterozygous loss of the gene

(Figure 2.1 A). We also used Oncomine, an online database compiling expression data

from thousands of microarray experiments (https://www.oncomine.org, (Rhodes et al.,

2007; Rhodes et al., 2004)), to analyze PIK3R1 expression in other breast cancer

datasets. Across multiple studies, PIK3R1 mRNA expression was significantly reduced

by 50-77% in breast cancer samples when compared to normal breast tissue (Figure

2.1 B and Table 2.1). Together these results indicate that in breast cancers, PIK3R1 is

consistently decreased at both the genomic and mRNA expression levels, suggesting

that reduced levels of p85g may play a functional role in tumorigenesis of this tissue.

Page 71: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

54

Figure 2.1: PIK3R1 expression is significantly reduced in breast cancers. Publically available datasets were queried for changes in PIK3R1 associated with breast cancer. (A) The cBio Portal for Cancer Genomics (http://www.cbioportal.org) was used to determine the incidence of PIK3R1 copy number loss or mutation in both a comprehensive 2012 study and an additional provisional study of breast cancer from The Cancer Genome Atlas Network (TCGA). In the 2012 study, such alterations occurred in 23% (111/482) of cases; in the provisional study, these alterations occurred in 28% (271/962) of cases. (B) The Oncomine database (https://www.oncomine.org/) was used to determine changes in PIK3R1 expression in breast cancer as compared to normal breast tissue. Microarray data from representative studies was converted to raw expression levels by taking the inverse log2, and then normalized to the mean raw expression level of normal breast tissue in that specific study. Means ± SEM are shown. Statistical significance was determined using unpaired t-test. For more detail on individual studies, see Table 2.1. *, P < 0.05; **, P < 0.01; ****, P < 0.0001.

Page 72: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

55

Ta

ble

2.1

: P

IK3R

1 e

xp

ress

ion

is s

ign

ific

an

tly r

ed

uc

ed

in

bre

as

t ca

nce

rs a

cro

ss m

ult

iple

mic

roarr

ay d

ata

sets

Stu

dy n

am

e

Pu

blic

ati

on

C

on

tro

l ti

ssu

e

Co

ntr

ol N

C

an

ce

r ti

ssu

e

Ca

nce

r N

%

re

du

cti

on

in

PIK

3R

1

p v

alu

e

So

rlie

Bre

ast

(So

rlie

et a

l., 2

00

1)

No

rma

l b

reast

4

Du

cta

l b

rea

st

ca

rcin

om

a

65

50

.80

%

0.0

16

0

So

rlie

Bre

ast 2

(S

orlie

et a

l., 2

00

3)

No

rma

l b

reast

4

Du

cta

l b

rea

st

ca

rcin

om

a

81

50

.06

%

0.0

07

9

Ric

ha

rdso

n B

reast

2

(Ric

hard

son

et a

l.,

20

06)

No

rma

l b

reast

7

Du

cta

l b

rea

st

ca

rcin

om

a

40

51

.77

%

0.0

18

6

Ma

Bre

ast

4

(Ma

et a

l., 2

00

9)

No

rma

l b

reast

28

Inva

siv

e d

ucta

l b

reast

carc

ino

ma

18

77

.34

%

0.0

03

4

TC

GA

Bre

ast

TC

GA

Ne

twork

2

01

1, n

o a

sso

cia

ted

p

ub

lica

tio

n

No

rma

l b

reast

61

Inva

siv

e d

ucta

l b

reast

carc

ino

ma

38

9

66

.71

%

< 0

.00

01

T

he

Onco

min

e d

ata

ba

se v

ers

ion 4

.4.4

.3 (

http

s:/

/ww

w.o

ncom

ine.o

rg)

was q

ueri

ed f

or

expre

ssio

n o

f P

IK3R

1 in

bre

ast cancer

mic

roarr

ay s

tud

ies.

Th

e d

ata

fro

m e

ach s

tud

y w

as c

onvert

ed to

raw

exp

ressio

n le

vels

of

PIK

3R

1 b

y takin

g th

e invers

e log2,

an

d th

en

norm

aliz

ed to

the

mean r

aw

PIK

3R

1 e

xpre

ssio

n le

vel of

the

no

rmal b

rea

st

tissue f

rom

that

sp

ecific

stu

dy.

Unpa

ire

d t

-test w

as

perf

orm

ed u

sin

g G

rap

hP

ad

Prism

6.0

to d

ete

rmin

e p

valu

es f

or

expre

ssio

n in

cancer

tissue a

s c

om

pare

d t

o n

orm

al bre

ast.

Page 73: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

56

RNAi-mediated PIK3R1 knockdown transforms human mammary epithelial cells

To study whether partial loss of p85g could contribute to transformation, we chose to use

human mammary epithelial cells (HMECs), a well-established system for the study of

PI3K-mediated transformation in vitro. In this system, HMECs are immortalized by

expression of the human telomerase reverse transcriptase (hTERT) catalytic subunit;

these cells have also been characterized to lack functional cyclin-dependent kinase 4

inhibitor A (p16INK4A) and have increased MYC expression (Romanov et al., 2001;

Utermark et al., 2007; Wang et al., 2000). HMECs expressing a dominant negative p53

mutant (DDp53) are unable to form colonies in agar (Zhao et al., 2003; Zhao et al.,

2005), an experimental measure of the transformation hallmark of anchorage-

independent growth (Hanahan and Weinberg, 2011). Activation of PI3K by expression of

myristoylated p110g or p110く, or mutant alleles of p110g found in cancers, transforms

these cells (Zhao et al., 2003; Zhao et al., 2005).

To address whether decreased PIK3R1 expression might play a role in the

transformation of mammary tissue, we generated polyclonal DDp53-HMEC lines with

stable RNAi-mediated knockdown of PIK3R1 (Figure 2.2 A). Two different shRNAs

targeting PIK3R1 reduced p85g protein levels by 76.6 ± 0.7% and 77.3 ± 2.9% in

comparison to the shControl (N = 3 for all) (Figure 2.2 B). While the positive control cells

expressing oncogenic p110g-H1047R (Zhao et al., 2003) showed dramatically increased

activation of AKT even after 4 hours of serum and growth factor starvation, with a fold

increase of 35.1 ± 9.0 in phosphorylation of AKT at S473 as compared to the shControl

line, PIK3R1 knockdown only slightly increased AKT S473 phosphorylation, by a fold

increase of 2.7 ± 0.4 and 4.0 ± 0.9 (N = 3 for all) (Figure 2.2 D). Although this increase

was statistically significant, it is clear that partial loss of p85g does not increase

constitutive PI3K pathway activation under un-stimulated conditions to a similar extent

Page 74: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

57

Figure 2.2: Generation of DDp53-HMECs with stable RNAi-mediated PIK3R1 knockdown. Human mammary epithelial cells (HMECs) expressing a dominant negative p53 mutant (DDp53) were infected with lentivirus containing a negative control scrambled shRNA or one of two distinct shRNAs targeting PIK3R1, or with retrovirus containing the cancer-associated p110g-H1047R mutant as a positive control. Following selection for stable polyclonal lines, cells were starved of serum and growth factors for 4 hours; protein lysates were collected and subjected to immunoblotting for PI3K pathway components. Membranes were stripped and re-probed for total proteins and vinculin as a loading control. (A) Representative immunoblot. (B-D) Bands from immunoblots of three independent sets of protein lysates were quantified by densitometry; protein levels were normalized first to the vinculin loading control (for p85g and PTEN) or the total unphosphorylated protein (for phospho-AKTS473), then to the corresponding mean value for the shControl. Means ± SEM are shown; N = 3 for each. Statistical significance was determined by unpaired t-test. Significance for comparison to the shControl is shown. *, P < 0.05; **, P < 0.01; ****, P < 0.0001; ns, not significant.

Page 75: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

58

as oncogenic mutant p110g.

Next we investigated whether PIK3R1 knockdown was sufficient to induce

transformation of DDp53-HMECs. It has previously been shown that DDp53-HMECs are

unable to form colonies in agar, but become transformed upon activation of PI3K (Zhao

et al., 2006; Zhao et al., 2003). We plated single-cell suspensions of our HMEC lines in

fully supplemented growth medium containing 0.3% agar and allowed them to grow for 4

weeks before scoring colonies. Consistent with previous reports, the shControl line was

unable to form colonies in agar, with only 16 ± 0.6 colonies per 50,000 cells plated, while

the positive control line expressing p110g-H1047R formed abundant colonies, with 909.7

± 35.0 colonies per 50,000 cells plated (N = 3 for both) (Figure 2.3 A). In comparison,

both shPIK3R1 lines readily formed colonies when grown in agar, with 633.7 ± 7.9 and

1,229 ± 39.0 colonies per 50,000 cells plated, respectively (N = 3 for both) (Figure 2.3

A). These results demonstrate that PIK3R1 knockdown is sufficient to transform DDp53-

HMECs in vitro.

To further assess the effects of PIK3R1 knockdown on PI3K signaling, we evaluated

activation of downstream effectors of this pathway in response to acute growth factor

stimulation. Cells were synchronized via a 4 hour serum and growth factor starvation,

and then stimulated with 20ng/ml EGF for timepoints up to 60 minutes; protein lysates

were collected and subjected to immunoblotting for phosphorylation at activation sites of

AKT and S6 ribosomal protein. Compared to the control line, PIK3R1 knockdown lines

exhibited both increased and sustained phosphorylation of AKT at activating residues

T308 and S473, and of S6 ribosomal protein at S235/236 (Figure 2.3 B). Comparable

increases in EGF-stimulated PI3K/AKT pathway activation were seen in DDp53-HMECs

expressing other shRNAs targeting PIK3R1, and a similar effect was seen upon

Page 76: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

59

Figure 2.3: PIK3R1 knockdown transforms DDp53-HMECs and increases growth factor-stimulated PI3K/AKT activation. (A) DDp53-HMECs with stable expression of shControl, shPIK3R1, or p110g-H1047R were tested for anchorage-independent growth. Single-cell suspensions were plated in 0.3% agar and grown for 4 weeks. 50,000 cells were plated per 6cm plate. The number of colonies was counted for three independent experiments of at least 3 plates each. Means ± SEM are shown; N = 3 for each group. Statistical significance was determined by unpaired t-test. Significance for comparison to the shControl is shown. ****, P < 0.0001. (B) The indicated cell lines were starved of serum and growth factors for 4 hours and then stimulated with 20ng/ml EGF for the indicated amounts of time. Protein lysates were collected and subjected to immunoblotting. Membranes were stripped and re-probed for total proteins and vinculin as a loading control. Bands were quantified by densitometry and normalized first to the corresponding total protein, then to the most intense band for the shControl. One representative experiment is shown.

Page 77: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

60

stimulation with insulin (data not shown). These results indicate that partial reduction of

p85g increases PIγK/AKT signaling in DDp5γ-HMECs, and that this increase is most

pronounced upon stimulation with growth factors.

To establish that the increased PI3K signaling seen in our PIK3R1 knockdown DDp53-

HMEC lines was specifically from decreased p85g expression and not an off-target

effect of stable shRNA introduction, we sought to rescue p85g protein levels in our

knockdown lines. For shRNAs targeting the γ’ UTR of PIK3R1, including shPIK3R1 #1, a

wildtype flag-tagged PIK3R1 construct was used; for shRNAs targeting the coding region

of PIK3R1, including shPIK3R1 #2, we introduced silent wobble mutations rendering

resistance to specific shRNAs into this wildtype construct. Stable expression of these

constructs in our PIK3R1 knockdown DDp53-HMECs restored p85g protein expression

to levels comparable to the shControl line (Figure 2.4 and data not shown). We then

evaluated PI3K/AKT pathway activation in shControl, shPIK3R1, and rescue cell lines in

response to acute growth factor stimulation. Cells were starved for 4 hours and then

stimulated with 20ng/ml EGF for timepoints up to one hour; protein lysates were

collected and subjected to immunoblotting for PI3K/AKT pathway components. As

expected, compared to the shControl line, shPIK3R1 cells exhibited increased and

sustained phosphorylation of AKT and S6 ribosomal protein; rescue of p85g protein

levels in our shPIK3R1 cells reduced AKT and S6 phosphorylation to levels comparable

to the shControl line (Figure 2.4). Similar results were seen with other shPIK3R1 rescue

DDp53-HMECs (data not shown). This data indicates that the augmented PI3K/AKT

signaling seen in our knockdown lines is likely due to decreased expression of p85g.

PIK3R1 knockdown augments HMEC transformation mediated by oncogenes

Because the increased PI3K/AKT signaling in PIK3R1 knockdown DDp53-HMECs was

Page 78: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

61

Figure 2.4: Augmented PI3K/AKT activation in PIK3R1 knockdown DDp53-HMECs is rescued by ectopic expression of PIK3R1. Silent wobble mutations were introduced into a wildtype, flag-tagged PIK3R1 construct to specifically render resistant to shPIK3R1 #2. This construct was stably introduced into DDp53-HMECs expressing shPIK3R1 #2 to rescue expression levels of p85g. The cell lines indicated were starved of serum and growth factors for 4 hours and then stimulated with 20ng/ml EGF for timepoints up to one hour. Protein lysates were collected and subjected to immunoblotting. Membranes were stripped and re-probed for total proteins and vinculin as a loading control. Bands were quantified by densitometry and normalized first to the corresponding total protein, then to the most intense band for the shControl. One representative experiment is shown.

Page 79: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

62

dependent on growth factors to initiate PI3K signaling, we were interested to know

whether reduced p85g could also cooperate with PIγK-activating oncogenes to augment

PI3K output and transformation. To address this, we generated DDp53-HMECs stably

expressing neuT, an activated rat form of HER2/neu, in addition to shRNA targeting

PIK3R1 (Figure 2.5 A). Similar to the DDp53-HMEC lines, shRNAs targeting PIK3R1

reduced p85g protein levels in neuT-expressing lines by 80.3 ± 0.4% and 70.6 ± 3.9%

compared to the shControl (N = 3 for all) (Figure 2.5 B). Expression of neuT increased

PI3K/AKT pathway activation even following overnight starvation, and as expected,

expression of p110g-H1047R further augmented AKT activation as assessed by S473

phosphorylation by 3.8 ± 0.5 fold, as compared to the shControl; both PIK3R1

knockdown lines also exhibited increased AKT S473 phosphorylation over the shControl

line, by 1.6 ± 0.1 and 1.6 ± 0.2 fold respectively (N = 3 for all) (Figure 2.5 D). These

results suggest that partial p85g loss may be able to cooperate with activated HER2/neu

to increase PI3K/AKT activation.

To study the potential synergy between activated HERβ/neu and reduced p85g, we

examined whether PIK3R1 knockdown could increase the in vitro transformation of

DDp53-HMECs by neuT. We plated single-cell suspensions of these HMEC lines in fully

supplemented growth medium containing 0.3% agar and allowed them to grow for 3

weeks before scoring colonies. While the neuT shControl line was able to form colonies

in agar, with 821.1 ± 69.2 colonies per 25,000 cells plated, additional expression of

p110g-H1047R increased colony formation, with 1,330 ± 190.3 colonies per 25,000 cells

plated (N = 4 for both) (Figure 2.6 A). Similarly, both neuT shPIK3R1 lines exhibited

significantly augmented colony formation over the neuT shControl line, with 1,227 ± 79.3

and 1,321 ± 125.6 colonies per 25,000 cells plated, respectively (N = 4 for both) (Figure

2.6 A). These results indicate that partial p85g loss can cooperate with oncogenic

Page 80: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

63

Figure 2.5: Generation of DDp53-HMECs with activated HER2/neu and RNAi-mediated PIK3R1 knockdown. DDp53-HMECs stably expressing neuT were infected with lentivirus containing a negative control scrambled shRNA or one of two distinct shRNAs targeting PIK3R1, or with retrovirus containing the cancer-associated p110g-H1047R mutant as a positive control. Following selection for stable polyclonal lines, cells were starved of serum and growth factors overnight; protein lysates were collected and subjected to immunoblotting for PI3K pathway components. Membranes were stripped and re-probed for total proteins and vinculin as a loading control. (A) Representative immunoblot. (B-D) Bands from immunoblots of three independent sets of protein lysates were quantified by densitometry; protein levels were normalized first to the vinculin loading control (for p85g and PTEN) or the total unphosphorylated protein (for phospho-AKTS473), then to the corresponding mean value for the shControl. Means ± SEM are shown; N = 3 for each. Statistical significance was determined by unpaired t-test. Significance for comparison to the shControl line is shown. *, P < 0.05; **, P < 0.01; ***, P < 0.001; ****, P < 0.0001; ns, not significant.

Page 81: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

64

Figure 2.6: PIK3R1 knockdown increases transformation and PI3K/AKT signaling driven by activated HER2/neu in DDp53-HMECs. (A) DDp53-HMECs with stable expression of neuT and shControl, shPIK3R1, or p110g-H1047R were tested for anchorage-independent growth. Single-cell suspensions were plated in 0.3% agar and grown for 3 weeks. 25,000 cells were plated per 6cm plate. The number of colonies was counted for four independent experiments of at least 3 plates each. Means ± SEM are shown; N = 4 for each group. Statistical significance was determined by unpaired t-test. Significance for comparison to the shControl is shown. *, P < 0.05; **, P < 0.01. (B) The indicated cell lines were starved of serum and growth factors overnight and then stimulated with 20ng/ml EGF for the indicated amounts of time. Protein lysates were collected and subjected to immunoblotting. Membranes were stripped and re-probed for total proteins and vinculin as a loading control. Bands were quantified by densitometry and normalized first to the corresponding total protein, then to the most intense band for the shControl line. One representative experiment is shown.

Page 82: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

65

HER2/neu to increase transformation of DDp53-HMECs.

We also assessed the ability of reduced p85g expression to increase PIγK/AKT

signaling in DDp53-HMECs with activated HER2/neu. DDp53-HMECs expressing neuT

and control or PIK3R1-targeting shRNAs were synchronized via overnight serum and

growth factor starvation, then stimulated with 20ng/ml EGF for timepoints up to 24 hours;

protein lysates were collected and subjected to immunoblotting for phosphorylation of

activation sites on AKT and S6 ribosomal protein. Compared to the shControl cells,

DDp53-HMECs expressing neuT and shPIK3R1 had increased phosphorylation of AKT

at both T308 and S473, and of S6 at S235/236, throughout the timecourse of EGF

stimulation (Figure 2.6 B). Similar results were seen with DDp53-HMECs expressing

neuT and other shRNAs targeting PIK3R1 (data not shown).Together this data indicates

that reduced p85g synergistically augments PIγK signaling driven by the pathway-

activating oncogenic HER2/neu in starved or growth factor-stimulated conditions.

To further explore the ability of partial p85g loss to synergize with oncogenes common to

breast cancer, we generated DDp53-HMECs stably expressing oncogenic p110g-

H1047R in addition to stable knockdown of PIK3R1. Compared to DDp53-HMECs

expressing p110g-H1047R and shControl, both PIK3R1 knockdown cell lines formed

approximately 4 fold more colonies when grown in 0.3% agar for 3 weeks (Figure 2.7

A). We also examined PI3K/AKT activation in these cells in response to EGF stimulation

following overnight starvation. In DDp53-HMECs expressing p110g-H1047R, PIK3R1

increased phosphorylation of AKT at both T308 and S473 in comparison to the

shControl (Figure 2.7 B). Together this data demonstrates that partial loss of p85g can

further augment transformation mediated by PI3K-activating oncogenes clinically

relevant to breast cancer, including HERβ/neu and p110g-H1047R.

Page 83: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

66

Figure 2.7: PIK3R1 knockdown increases transformation and PI3K/AKT signaling driven by p110g-H1047R in DDp53-HMECs. (A) DDp53-HMECs with stable expression of p110g-H1047R and shControl or shPIK3R1 were tested for anchorage-independent growth. Single-cell suspensions were plated in 0.3% agar and grown for 3 weeks. 25,000 cells were plated per 6cm plate. The number of colonies was counted for one independent experiment of at least 3 plates per cell line. Means ± SD are shown; N ≥ γ for each. (B) The indicated cell lines were starved of serum and growth factors overnight and then stimulated with 20ng/ml EGF for the indicated amounts of time. Protein lysates were collected and subjected to immunoblotting. Membranes were stripped and re-probed for vinculin as a loading control.

Page 84: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

67

Transformation driven by PIK3R1 knockdown is mediated by signaling through p110g

The PI3K catalytic isoforms play divergent roles in physiological and pathophysiological

signaling. Knowing which isoforms are functionally critical in different contexts can inform

future clinical use of isoform-selective inhibitors. Accordingly, we used pan-PI3K and

isoform-selective inhibitors to determine the contribution of different p110 isoforms to the

transformation of HMECs mediated by PIK3R1 knockdown. We plated DDp53-HMECs

expressing control or PIK3R1-targeting shRNAs in 0.3% agar, and grew cells for 4

weeks in the presence of either pan-PI3K or isoform-selective PI3K inhibitors. Either the

pan-PI3K inhibitor GDC0941 (Folkes et al., 2008) or the p110g-selective inhibitor

BYL719 (Furet et al., 2013) effectively inhibited colony formation of PIK3R1 knockdown

lines in a dose-dependent manner (Figure 2.8 A-B). Treatment with the p110く-selective

inhibitor TGX221 did not have a substantial effect on colony growth, even at doses well

above the IC50 of this compound (Jackson et al., 2005) (Figure 2.8 C). These results

suggest that transformation of DDp53-HMECs with partial p85g is mediated by PIγK

signaling through p110g and not p110く.

While it has been demonstrated that similar to our findings with PIK3R1 knockdown

DDp53-HMECs, HER2/neu-driven transformation is dependent on p110g (Utermark et

al., 2012), recent work has also demonstrated that the presence of co-existing

oncogenic events can shift PI3K isoform dependency (Schmit et al., 2014). Therefore,

we were interested to determine the isoform dependency of HER2/neu-driven

transformation in the context of partial p85g loss. We plated single-cell suspensions of

DDp53-HMECs with stable expression of neuT and either shControl, shPIK3R1, or

p110g-H1047R in 0.3% agar and allowed cells to grow for 3 weeks with different PI3K

inhibitors. Treatment with the pan-PIγK inhibitor GDC0941 or the p110g-selective

inhibitor BYL719 effectively inhibited colony formation of these cells (Figure 2.9 A-B),

Page 85: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

68

Figure 2.8: Transformation of PIK3R1 knockdown DDp53-HMECs is blocked by p110g-selective pharmacological inhibition. DDp53-HMECs with stable expression of shControl or shPIK3R1 were plated as single-cell suspensions in 0.3% agar and grown for 4 weeks. 9,000 cells were plated per well of a 12-well plate. Fresh growth medium containing either the pan-PI3K inhibitor GDC0941 (A), the p110g-selective inhibitor BYL719 (B), or the p110く-selective inhibitor TGX221 (C) was given every three days. Means ± SD are shown for triplicate wells from representative experiments.

Page 86: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

69

Figure 2.9: Transformation of PIK3R1 knockdown DDp53-HMECs with activated HER2/neu is blocked by p110g-selective inhibition. DDp53-HMECs with stable expression of neuT and shControl, shPIK3R1, or p110g-H1047R were plated as single-cell suspensions in 0.3% agar and grown for 3 weeks. 4,500 cells were plated per well of a 12-well plate. Fresh growth medium containing either the pan-PI3K inhibitor GDC0941 (A), the p110g-selective inhibitor BYL719 (B), or the p110く-selective inhibitor KIN193 (C) was given every three days. Means ± SD are shown for triplicate wells from representative experiments.

Page 87: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

70

while treatment with the p110く-selective inhibitor KIN193 did not substantially effect

colony growth (Figure 2.9 C), even at concentrations that completely block AKT

activation in PTEN-null cancer cell lines (Ni et al., 2012). Consistent with these findings,

either GDC0941 or BYL719 blocked PI3K/AKT pathway activation in DDp53-HMECs

expressing neuT and shPIK3R1 in a dose-dependent manner, while KIN193 only slightly

reduced AKT and S6 phosphorylation at the highest doses (Figure 2.10). We would note

that these inhibitor experiments are missing an important positive control for the p110く

inhibitors: to demonstrate that TGX221 and KIN193 are acting as expected, we plan to

carry out similar experiments using the PTEN null breast cancer cells HCC70, MDA-MB-

468, and BT-549; signaling in and transformation of these cells has been shown to be

largely reliant on p110く (Jia et al., 2008; Ni et al., 2012; Torbett et al., 2008; Wee et al.,

2008). Nonetheless, the results presented here indicate that the augmented PI3K/AKT

signaling in and transformation of PIK3R1 knockdown DDp53-HMECs with or without

expression of neuT is primarily mediated by p110g.

PIK3R1 knockdown does not affect PTEN levels or lipid phosphatase activity

Next we were interested in determining the molecular mechanism by which decreased

p85g levels could increase PIγK/AKT output and mediate transformation. Recent studies

have reported that p85g may directly bind PTEN (Chagpar et al., 2010; Rabinovsky et

al., 2009) and increase its phosphatase activity (Chagpar et al., 2010); this binding was

found to be dependent on EGF stimulation in one study (Chagpar et al., 2010). Other

publications have suggested that p85g may be important for PTEN stability at the mRNA

or protein level (Cheung et al., 2011; Taniguchi et al., 2010). We hypothesized that if

p85g had a stabilizing or activating function on this negative regulator of the PIγK/AKT

pathway, then partial p85g loss may lead to reduced levels or activity of PTEN, thereby

increasing PI3K/AKT signaling.

Page 88: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

71

Figure 2.10: PI3K/AKT signaling in PIK3R1 knockdown DDp53-HMECs with activated HER2/neu is blocked by p110g-selective inhibition. DDp53-HMECs stably expressing neuT and shPIK3R1 #2 were starved of serum and growth factors overnight, then stimulated with starvation medium containing 20ng/ml EGF and the indicated concentrations of PI3K inhibitors for 15 minutes. Protein lysates were collected and subjected to immunoblotting. Membranes were stripped and re-probed for total proteins and vinculin as a loading control. Bands were quantified by densitometry and normalized first to the corresponding total protein, then to the corresponding band for the DMSO control. One representative experiment is shown.

Page 89: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

72

Figure 2.11: Endogenous p85g and PTEN do not appear to interact in DDp53-HMECs. DDp53-HMECs with stable expression of scrambled control or PIK3R1-targeting shRNAs were starved of serum and growth factors for 4 hours and then given either fresh starvation medium or 20ng/ml EGF for the indicated amounts of time. Protein lysates were collected and subjected to immunoprecipitation reactions for either endogenous PTEN (A) or endogenous p85g (B). Samples were then analyzed by immunoblotting. Membranes were stripped and re-probed for vinculin as a loading control.

Page 90: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

73

We first sought to determine whether an interaction between endogenous p85g and

PTEN proteins occurred in our DDp53-HMECs. We synchronized shControl and

shPIK3R1 DDp53-HMECs by a 4 hour serum and growth factor starvation, stimulated

half of the cells with 20ng/ml EGF, collected protein lysates, and carried out

immunoprecipitation reactions for PTEN. Although we were able to achieve reliable

immunoprecipitation of endogenous PTEN from both cell lines, we could not detect any

evidence of p85g co-immunoprecipitation either under starved or stimulated conditions

(Figure 2.11 A). We also carried out similar experiments with immunoprecipitation of

p85g from synchronized or EGF-stimulated DDp53-HMECs. We consistently

immunoprecipitated endogenous p85g from shControl and shPIK3R1 cell lines, but while

we observed co-immunoprecipitation of endogenous p110g in these experiments, we did

not detect endogenous PTEN in immunoprecipitations from either starved or stimulated

cells (Figure 2.11 B). Despite altering a number of conditions for these

immunoprecipitation reactions, including the antibodies and buffers used, similar results

were seen across multiple experiments (data not shown). Thus we could not positively

demonstrate an interaction between endogenous p85g and PTEN in our HMECs.

We considered that a p85g-PTEN interaction might be easier to demonstrate in certain

cell types over others. Neither of the publications showing this interaction used HMECs;

however, both studies used HeLa cells among other cell types in their

immunoprecipitation experiments (Chagpar et al., 2010; Rabinovsky et al., 2009). We

therefore used HeLa cells in addition to wildtype HMECs without p53 inactivation. These

cell lines were starved of serum and growth factors for 4 hours, and then half of the cells

were stimulated with 20ng/ml EGF. Protein lysates were collected and subjected to

immunoprecipitation for endogenous p85g. Although p85g and p110g were reliably co-

immunoprecipitated from both cell lines, PTEN was not detected in these samples under

Page 91: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

74

Figure 2.12: PTEN and p85g do not appear to interact in a variety of cell types. (A) HeLa cells and HMECs without p53 inactivation were starved of serum and growth factors for 4 hours before being stimulated with 20ng/ml EGF for the indicated amounts of time. Protein lysates were collected and used in immunoprecipitation reactions for endogenous p85g or PTEN. (B) 293T cells were either mock transfected or transiently transfected with wildtype PIK3R1. Protein lysates were collected and subjected to immunoprecipitation for PTEN or p85g. Membranes were stripped and re-probed for vinculin as a loading control.

Page 92: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

75

either starved or stimulated conditions (Figure 2.12 A). Based on these results we were

unable to conclude that p85g and PTEN interact in HeLa cells.

Finally, we speculated that the amount of interacting p85g and PTEN might be a small

fraction of the total amount of these proteins, potentially making it difficult to detect

endogenous co-immunoprecipitation. We therefore transiently transfected 293T cells

with a wildtype PIK3R1 construct, collected protein lysates from these cells, and

subjected them to immunoprecipitation for either PTEN or p85g. Compared to the mock

transfected cells, transient transfection with PIK3R1 led to substantial overexpression of

p85g (Figure 2.12 B). Although immunoblots revealed a small amount of p85g present

in PTEN immunoprecipitates, a comparable amount of p85g was detected in the beads

only control, suggesting that when p85g was highly overexpressed it bound at a low,

non-specific level to the beads used for immunoprecipitation (Figure 2.12 B). Similar

results were seen in multiple independent experiments (data not shown). Consistent with

this conclusion, we did not observe any co-immunoprecipitated PTEN along with p85g in

PIK3R1-transfected 293T cells, despite achieving co-immunoprecipitation of both p110g

and p110く (Figure 2.12 B). Taken together, we were unable to satisfactorily

demonstrate an interaction between p85g and PTEN in a variety of cell types and

conditions.

While we were unable to positively shown an interaction between p85g and PTEN in our

experiments, it has been also been reported that mutation or loss of PIK3R1 might

destabilize PTEN, either at the level of transcription or translation (Cheung et al., 2011;

Taniguchi et al., 2010). To explore this possibility, we first examined PTEN protein levels

in our HMECs with PIK3R1 knockdown. We collected three independent sets of protein

lysates from cells synchronized by serum and growth factor starvation, subjected them

Page 93: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

76

to immunoblotting for PTEN, and used densitometry to quantify the intensity of each

band relative to the vinculin loading control. In DDp53-HMECs, stable PIK3R1

knockdown did not have a consistent effect on the steady-state PTEN levels; only one of

the shPIK3R1 lines had significantly reduced PTEN protein levels, of 29.8 ± 4.1%

compared to the shControl (N = 3 for both) (Figure 2.2 C). We also assessed PTEN

protein levels in our neuT-expressing DDp53-HMECs: compared to the shControl, only

one shPIK3R1 line showed a significant reduction in PTEN protein levels of 32.9 ± 4.0%;

the p110g-H1047R line also had a significant reduction of 17.1 ± 1.7% in PTEN levels (N

= 3 for all) (Figure 2.5 C). These results suggest that in HMECs, partial reduction of

p85g does not necessarily lead to destabilization of PTEN protein.

We also sought to confirm that PTEN transcripts were not affected by PIK3R1

knockdown in our HMEC lines. We used quantitative PCR (qPCR) to assess the levels

of PTEN mRNA in extracts from DDp53-HMECs stably expressing shControl or

shPIK3R1. As an additional control, we also examined mRNA levels of the INPP4B

phosphatase; INPP4B is a lipid phosphatase that removes phosphorylation of the 4’

hydroxyl group of PtdIns(3,4)P2, and has been identified as a tumor suppressor in breast

cancer in a screen using HMECs (Gewinner et al., 2009). Compared to the shControl

line, both shPIK3R1 lines had somewhat increased levels of INPP4B transcripts (Figure

2.13 A). In addition, while one shPIK3R1 line had somewhat decreased PTEN

transcripts compared to the shControl line, the other shPIK3R1 line showed slightly

elevated PTEN mRNA levels (Figure 2.13 A). Together these qPCR results suggest that

PIK3R1 expression is not a reliable predictor of the mRNA levels for either the INPP4B

or PTEN tumor suppressor genes.

Although we were unable to reproduce a p85g-PTEN interaction in our experiments, and

Page 94: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

77

Figure 2.13: PIK3R1 knockdown does not affect PTEN mRNA levels or lipid phosphatase activity in DDp53-HMECs. (A) Quantitative PCR (qPCR) was used to determine the levels of INPP4B and PTEN mRNA in extracts from DDp53-HMECs stably expressing scrambled control or PIK3R1-targeting shRNAs. Means ± SD are shown for one experiment performed in triplicate. Levels were normalized to the corresponding mean for shControl. (B-C) In vitro lipid phosphatase assays for PTEN immunoprecipitated from shControl or shPIK3R1 DDp53-HMECs. A malachite green reagent was used to detect phosphate in prepared standards (B) or released upon incubation of PtdIns(3,4,5)P3 substrate with PTEN immunoprecipitates. Three negative controls were used: PIP3 Only, assay performed using substrate but no source of PTEN; PTEN Only, assay performed using PTEN immunoprecipitated from shControl cells but with no substrate added; and Beads Only, assay performed using precipitates from shControl cells in which beads but no PTEN antibody was used. The amount of phosphate released in in vitro assays was interpolated from comparison to the phosphate standards. For (B), means ± SD are shown for standards in triplicate. For (C), means ± SEM are shown for assays performed using three independent immunoprecipitation reactions per group. Statistical analysis was performed using unpaired t-test as compared to the shControl. *, P < 0.05; ns, not significant.

Page 95: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

78

we did not find a consistent effect of PIK3R1 knockdown on the levels of PTEN mRNA or

protein, the possibility remained that partial p85g loss might affect PTEN lipid

phosphatase activity. To assess PTEN activity in our HMEC lines, we carried out PTEN

immunoprecipitations and performed in vitro lipid phosphatase assays using

PtdIns(3,4,5)P3 as a substrate. In these assays, a malachite green reagent was used to

detect phosphate liberated by lipid phosphatase activity towards the provided substrate.

We used three negative controls, consisting of assays performed either without PTEN

added (“PIPγ Only,” using substrate but no source of PTEN, and “Beads Only,” using

precipitate from shControl cells with beads but no antibody), or without substrate added

(“PTEN Only,” using PTEN immunoprecipitated from shControl cells). These controls

produced negligible phosphate release (Figure 2.13 C). Compared to these controls,

PTEN immunoprecipitated from shControl DDp53-HMECs was able to convert 25.0 ±

0.6% of the PtdIns(3,4,5)P3 substrate provided to PtdIns(4,5)P2 (N = 3) (Figure 2.13 C).

PTEN immunoprecipitated from one of the shPIK3R1 lines converted significantly more

substrate, with 30.5 ± 1.8% conversion, while PTEN immunoprecipitated from the other

shPIK3R1 line was similar to the control, with 25.2 ± 1.0% substrate conversion (N = 3

for both) (Figure 2.13 C). Admittedly, the conclusions from this experiment are limited by

the inability to control for the exact amount of immunoprecipitated PTEN in each in vitro

lipid phosphatase assay; however, previous experiments from these same cell lines

produced a comparable amount of PTEN protein in immunoprecipitation reactions

(Figure 2.11 A). These results suggest that in DDp53-HMECs, partial p85g loss does

not lead to reduced PTEN lipid phosphatase activity.

PIK3R1 knockdown does not increase RTK activation or alter RTK trafficking

Because we were unable to demonstrate an interaction between p85g and PTEN, and

did not find that PIK3R1 knockdown significantly impacted PTEN levels or lipid

Page 96: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

79

phosphatase activity, we looked for a mechanism of transformation mediated by partial

p85g loss that did not involve PTEN. To determine whether the augmented growth

factor-mediated PI3K signaling in PIK3R1 knockdown cells was due to increased

receptor activation, we assessed EGFR phosphorylation in response to EGF stimulation

using two different methods. First, we synchronized shControl and shPIK3R1 DDp53-

HMECs by a 4 hour serum and growth factor starvation, stimulated the cells with

20ng/ml EGF for timepoints up to 1 hour, and collected protein lysates. We then

performed immunoblotting using an antibody specific to EGFR phosphorylated on

Y1068, one of the three major activating autophosphorylation sites (Downward et al.,

1984). Although PIK3R1 knockdown cells exhibited significantly increased and sustained

phosphorylation of AKT and S6 ribosomal protein, phosphorylation of EGFR at Y1068

was largely unchanged (Figure 2.14 A). While Y1068 is one of the main residues

phosphorylated during EGFR activation, other phosphorylated tyrosines also contribute

to activation of this receptor (Downward et al., 1984). To completely assess EGFR

phosphorylation status, we next carried out the same EGF stimulation timecourse,

performed immunoprecipitations for tyrosine-phosphorylated proteins, and subjected

these samples to immunoblotting for total EGFR. Following EGF stimulation, PIK3R1

knockdown cells showed a similar or even slightly reduced amount of total tyrosine-

phosphorylated EGFR when compared to control cells (Figure 2.14 B). Together, these

results indicate that the increase in growth factor-stimulated PI3K signaling in PIK3R1

knockdown cells likely occurs at a step in pathway activation that is downstream of

receptor autophosphorylation and activation.

Following activation, RTKs are internalized in endosomes, and are then either recycled

back to the cell surface or targeted to lysosomes for degradation (Miaczynska, 2013).

Intracellular trafficking of receptors can control the magnitude and duration of signaling

Page 97: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

80

Figure 2.14: PIK3R1 knockdown does not increase growth factor-stimulated RTK phosphorylation in DDp53-HMECs. (A) The indicated cell lines were starved of serum and growth factors for 4 hours and then stimulated with 20ng/ml EGF for the indicated amounts of time. Protein lysates were collected and subjected to immunoblotting. Membranes were stripped and re-probed for vinculin as a loading control. One representative experiment is shown. (B) Growth factor stimulation was performed as in (A); protein lysates were collected and subjected to immunoprecipitation using beads conjugated to the 4G10 antibody recognizing tyrosine-phosphorylated residues. Immunoprecipitates were subjected to immunoblotting for phosphorylated tyrosines using the 4G10 antibody; membranes were stripped and re-probed for total EGFR and vinculin as a loading control. One representative experiment is shown.

Page 98: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

81

(Miaczynska, 2013), and is regulated by a number of Rab small GTPases (Jean and

Kiger, 2012; Stenmark, 2009). The BH domain of p85g reportedly has GAP activity

towards select small GTPases, including Rab4 and Rab5 (Chamberlain et al., 2004),

and disruption of this activity leads to increased PI3K/AKT activation and cellular

transformation (Chamberlain et al., 2004; Chamberlain et al., 2008), apparently due to

more rapid and sustained internalization and reduced degradation of RTKs

(Chamberlain et al., 2008; Chamberlain et al., 2010). Therefore, we were interested to

know whether the increased PI3K signaling and transformation in our PIK3R1

knockdown cells could be due to altered intracellular trafficking of receptors, as a

consequence of reduced p85g Rab-GAP function.

To assess effects of PIK3R1 knockdown on RTK trafficking, we used biotin labeling to

track both internalization and degradation of surface proteins. First, to analyze receptor

internalization, shControl and shPIK3R1 cells were synchronized by overnight serum

and growth factor starvation, and then all surface proteins were labeled with biotin. Cells

were stimulated with 20ng/ml EGF to initiate receptor internalization, remaining surface

biotin was cleaved, and cells were lysed and subjected to immunoprecipitation with

streptavidin beads to capture internalized biotinylated proteins. Immunoblotting was then

used to assess the amount of total EGFR in the samples. Both shControl and shPIK3R1

cells had similar rates of EGFR internalization (Figure 2.15 A). Next, to analyze receptor

degradation, all surface proteins of synchronized shControl and shPIK3R1 cells were

labeled with biotin, cells were stimulated with EGF, and cells were lysed and subjected

to immunoprecipitation with streptavidin beads to capture all remaining biotinylated

proteins. Immunoblotting for total EGFR revealed that PIK3R1 knockdown cells had a

rate of EGFR degradation comparable to that of the control cells (Figure 2.15 B). These

experiments indicate that in DDp53-HMECs, partial loss of p85g does not have a

Page 99: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

82

Figure 2.15: PIK3R1 knockdown does not affect growth factor-stimulated RTK trafficking in DDp53-HMECs. (A) EGFR internalization in response to EGF. The indicated cell lines were starved of serum and growth factors overnight. Surface proteins were labeled with biotin, cells were stimulated with 20ng/ml EGF for the indicated amounts of time to initiate receptor internalization, and remaining surface biotin was cleaved. Protein lysates were generated and used in streptavidin immunoprecipitation to capture internalized biotinylated proteins. Immunoblotting was used to determine the amount of tyrosine-phosphorylated EGFR in the samples. Membranes were stripped and re-probed for total EGFR and vinculin as a loading control. T.S., total surface protein prior to stimulation. (B) EGFR degradation in response to EGF. The indicated cell lines were starved as in (A). Surface proteins were biotinylated, and then cells were stimulated with 20ng/ml EGF for the indicated amounts of time. Protein lysates were generated and then subjected to streptavidin immunoprecipitation to capture all remaining biotinylated proteins. Immunoblotting was used to determine the amount of total EGFR in the samples. Membranes were stripped and re-probed for vinculin as a loading control. T.S., total surface protein prior to stimulation. Representative experiments for (A) and (B) are shown.

Page 100: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

83

substantial effect on EGFR trafficking.

PIK3R1 knockdown increases the amount of p85-p110g bound to activated RTKs

Binding to the p85 regulatory subunit is required for p110 catalytic subunit stability (Yu et

al., 1998b). Furthermore, p85 serves as a necessary adaptor for recruiting p110 to

activated RTKs, where the p85-p110 heterodimer becomes activated (Rameh et al.,

1995; Yu et al., 1998a; Yu et al., 1998b). Research from the lab of Dr. Cantley has

suggested that in some tissues, p85 may be present in excess of p110, and that free

p85 might function as a negative regulator of PI3K signaling by competing with p85-p110

heterodimers for binding to phosphorylated RTKs (Luo and Cantley, 2005; Ueki et al.,

2003; Ueki et al., 2002b). Therefore, we were interested to know whether PIK3R1

knockdown in DDp53-HMECs might selectively reduce free p85g, allowing more p85-

p110 heterodimers to bind activated RTKs, thereby increasing PI3K/AKT signaling.

To test whether partial loss of p85 in our cells augments PI3K activation by selectively

depleting free p85, we generated polyclonal DDp53-HMEC lines with stable expression

of flag-tagged, activated ErbB3 (Flag-TEL-ErbB3) along with shRNA targeting PIK3R1.

Expression of Flag-TEL-ErbB3 lead to low-level colony formation of DDp53-HMECs, with

150.7 ± 18.8 colonies per 50,000 cells, while additional expression of p110g-H1047R

robustly increased transformation of these cells, to 891.5 ± 16.6 colonies per 50,000

cells (N = 3 for both) (Figure 2.16). Consistent with our previous data, knockdown of

PIK3R1 in these cells increased colony formation to a similar extent as p110g-H1047R,

with 685.8 ± 20.4 and 725.2 ± 53.9 colonies per 50,000 cells (N = 3 for both) (Figure

2.16). These findings suggest that in DDp53-HMECs, the expressed Flag-TEL-ErbB3 is

weakly activated and transforming, and can cooperate with partial p85g loss.

Page 101: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

84

Figure 2.16: PIK3R1 knockdown increases transformation of DDp53-HMECs expressing activated ErbB3. DDp53-HMECs stably expressing flag-tagged, activated ErbB3 (Flag-TEL-ErbB3) were infected with lentivirus containing a negative control scrambled shRNA or one of two distinct shRNAs targeting PIK3R1, or with retrovirus containing the cancer-associated p110g-H1047R mutant as a positive control. Following selection for stable polyclonal lines, these cells were tested for anchorage-independent growth. Single-cell suspensions were plated in 0.3% agar and grown for 4 weeks. 50,000 cells were plated per 6cm plate. The number of colonies was counted for three independent experiments of at least 3 plates each. Means ± SEM are shown; N = 3 for each group. Statistical significance was determined by unpaired t-test. Significance for comparison to the shControl is shown. ***, P < 0.001; ****, P < 0.0001.

Page 102: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

85

To assess the effect of PIK3R1 knockdown on binding of p85 and p110 to activated

RTKs in DDp53-HMECs, we starved the cells of serum and growth factors to limit RTK

activation to just the Flag-TEL-ErbB3, then used the Flag tag to immunoprecipitate this

activated receptor. Immunoblotting and densitometric quantification was then used to

determine the amounts of pan-p85 and p110g bound to Flag-TEL-ErbB3 (Figure 2.17

A). Compared to the shControl, both shPIK3R1 lines had significantly reduced pan-p85

levels, by 68.6 ± 3.7% and 69.3 ± 1.7% (N = 4 for all) (Figure 2.17 E). PIK3R1

knockdown reduced the amount of pan-p85 bound to Flag-TEL-ErbB3 by 43.6 ± 14.7%

and 53.4 ± 4.2% compared to the shControl (N = 4 for all) (Figure 2.17 B). Both

shPIKγR1 lines also exhibited a slight but significant reduction in p110g protein in whole

cell lysates as compared to the shControl line, of 20.9 ± 7.1% and 32.2 ± 6.4% (N = 4 for

all) (Figure 2.17 F); despite this, the amount of p110g associated with Flag-TEL-ErbB3

was actually increased in shPIK3R1 lines, with fold increases of 1.5 ± 0.3 and 1.4 ± 0.1

as compared to the shControl (N = 4 for all) (Figure 2.17 C). The net result was that

knockdown of PIK3R1 in DDp53-HMECs increased the ratio of p110g:pan-p85 bound to

activated ErbB3 by a fold of 3.7 ± 0.4 and 2.5 ± 0.1 as compared to the shControl (N = 4

for all) (Figure 2.17 D). These ratios correlated well with the fold increase in AKT

activation in these cells, as determined by phosphorylation at S473, of 2.4 ± 0.4 and 1.7

± 0.2 as compared to the shControl line (N = 4 for all) (Figure 2.17 G). These findings

were recapitulated with two additional PIK3R1 knockdown DDp53-HMEC lines, and also

confirmed using a different pan-p85 antibody (data not shown).

Together, the data presented in this chapter is consistent with the model of increased

PI3K/AKT signaling and transformation mediated by partial p85g loss depicted in Figure

2.18: in mammary epithelial cells with normal p85 levels, p85 is present in excess of

p110, and activated sites on receptors are able to be bound by both non-signaling free

Page 103: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

86

Figure 2.17: PIK3R1 knockdown increases the amount of p85-p110g bound to activated RTKs in DDp53-HMECs. DDp53-HMECs stably expressing Flag-TEL-ErbB3 and either shControl, shPIK3R1, or p110g-H1047R were starved of serum and growth factors for 4 hours; protein lysates were collected and subjected to immunoprecipitation for ErbB3 using beads conjugated to anti-flag antibody. (A) Representative immunoblot. Membranes were stripped and re-probed for total proteins and vinculin as a loading control. (B-D) Bands for proteins in ErbB3 immunoprecipitates from immunoblots were quantified by densitometry. For (B-C), protein levels were normalized first to the amount of ErbB3 in the immunoprecipitates, then to the corresponding mean for the shControl. For (D), protein levels of p110g were normalized to the levels of pan-p85 in immunoprecipitates, then to the corresponding mean for the shControl. (E-G) Bands for proteins in whole cell lysates from immunoblots were quantified by densitometry. Protein levels were normalized first to the vinculin loading control (for pan-85 and p110g) or the total unphosphorylated protein (for phospho-AKTS473), then to the corresponding mean value for the shControl. For (B-G), means ± SEM are shown for four independent experiments, N = 4 for each. Statistical significance was determined by unpaired t-test. Significance for comparison to the shControl is shown. *, P < 0.05; **, P < 0.01; ***, P < 0.001; ****, P < 0.0001; ns, not significant.

Page 104: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

87

Figure 2.18: Model: partial p85g loss leads to increased PI3K/AKT signaling and transformation. The data presented in Chapter 2 is consistent with a model in which p85 monomers compete with p85-p110 heterodimers to negatively regulate RTK-mediated PI3K/AKT signaling. Top: In mammary epithelial cells with normal p85g levels, p85 is present in excess of p110. Both monomeric p85 and heterodimeric p85-p110 can compete for binding to sites on activated RTKs, but only p85-p110 is capable of signaling. Bottom: In mammary epithelial cells with partial p85g reduction, the pool of monomeric p85g is selectively depleted, since p85 is required for p110 stability; more binding sites on activated RTKs are available for p85-p110 heterodimers, allowing for upregulated RTK-mediated PI3K signaling. While it is likely that reduced p85g levels similarly affect p110g and p110く, the p110く isoform has a minimal role in RTK-mediated signaling, possibly due to lower RTK-associated lipid kinase activity of p110く in comparison to p110g (Utermark et al., 2012).

Page 105: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

88

p85 and signaling-capable p85-p110 heterodimers; under conditions of partial p85g loss,

the pool of free p85 is selectively reduced, allowing for more p85-p110 heterodimers to

bind activated receptors, resulting in increased PI3K signal output. Although we expect

p85g reduction to similarly affect RTK signaling through p110g and p110く, since the

p110く isoform contributes little to PI3K signaling downstream of RTKs (Utermark et al.,

2012), it likely plays a minor role in this context. By this model, a balance between p85

and p110 subunits is expected to be critical for regulation of PI3K signaling in response

to inputs from activated RTKs.

Summary and discussion

In this chapter, we find that heterozygous deletion of PIK3R1 is frequent in human breast

cancers, and that furthermore PIK3R1 expression is significantly reduced in breast

tumors when compared to normal breast tissue. We show that RNAi-mediated

downregulation of p85g in DDp5γ-HMECs augments PI3K/AKT signaling in response to

growth factor stimulation, and increases colony formation in agar. PIK3R1 knockdown

also augments PI3K/AKT signaling and transformation driven by oncogenes common in

breast cancer, including p110g-H1047R and activated HER2/neu. Studies using pan-

PI3K and isoform-selective inhibitors suggest that HMEC transformation driven by partial

p85g loss is primarily mediated by signaling through p110g. Contrary to reports linking

p85g to PTEN stability or activity, we find that PIK3R1 downregulation in HMECs does

not substantially change steady-state PTEN mRNA or protein levels, or in vitro PTEN

lipid phosphatase activity. In addition, although p85g has been reported to have GAP

activity towards the Rab5 GTPase important for intracellular trafficking, we find that RTK

internalization and degradation is largely unchanged in PIK3R1 HMECs. Instead, we find

that partial reduction of p85g increases the amount of p85-p110g associated with

activated RTKs, suggesting a model where excess p85 monomers compete with p85-

Page 106: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

89

p110 heterodimers to negatively regulate RTK-mediated PI3K signaling.

The data presented in Chapter 2 strongly suggest that partial loss of p85g leads to an

increase in growth factor-stimulated PI3K/AKT pathway activation. However, a number

of additional experiments will help to confirm these findings. We have shown that

PIK3R1 knockdown increases phosphorylation of AKT at both activation sites in

response to growth factors, and while AKT phosphorylation is in general a good readout

for PI3K pathway activation, a more direct measure will be to assess production of

PtdIns(3,4,5)P3. We expect that in our HMECs, PtdIns(3,4,5)P3 levels will correlate well

with AKT phosphorylation, but in certain contexts it seems that this is not always the

case (Vora et al., 2014) and (J.A. Engelman, unpublished observations), so this will be

important to confirm. We also plan to extend our immunoblot experiments (Figure 2.2,

Figure 2.3, Figure 2.4, Figure 2.5, and Figure 2.6) to include other activated

downstream components of this pathway, in particular 4EBP1 phosphorylated at S65,

S6K phosphorylated at T389, PRAS40 phosphorylated at Tβ46, and GSKγく

phosphorylated at S9, to further confirm PI3K/AKT pathway upregulation in our PIK3R1

knockdown cells. Finally, we will expand on our pharmacological inhibition experiments

(Figure 2.8, Figure 2.9, and Figure 2.10) by using AKT-selective inhibitors such as the

ATP-competitive small molecule inhibitor GSK690693 (Rhodes et al., 2008) or the

allosteric inhibitor MK2206 (Hirai et al., 2010) to treat our PIK3R1 knockdown HMECs. If

transformation of these cells is indeed driven by PI3K/AKT pathway upregulation, we

expect that these agents will effectively block signaling in and transformation of these

cells in a manner similar to PI3K-selective inhibitors. Together, these proposed

experiments will further confirm that partial p85g loss transforms HMECs by increasing

PI3K/AKT pathway activation.

Page 107: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

90

Although the data presented in Chapter 2 are consistent with the competition model we

propose (Figure 2.18), additional experiments may substantially strengthen the

argument for this model. Using immunoblotting techniques, we have shown that PIK3R1

knockdown in HMECs increases the amount of p85-p110g bound to activated RTKs

(Figure 2.17 D). We plan to verify this result using mass spectrometry to determine the

number of p85 and p110 molecules bound to these activated RTKs. It will also be critical

to use in vitro lipid kinase assays to demonstrate that in these same cells, the PI3K

activity associated with activated RTKs is increased upon PIK3R1 knockdown. Our

model also relies on p85 being present in excess of p110 in HMECs. We performed

immunodepletion experiments similar to those published by other groups (Geering et al.,

2007; Mauvais-Jarvis et al., 2002; Ueki et al., 2002a) which indicated that this may in

fact be the case (data not shown), but technical difficulties precluded a definitive

conclusion from these results. An alternative approach will be to perform size-exclusion

chromatography on whole cell lysates from control and PIK3R1 knockdown HMECs,

followed by immunoblotting of fractions for p85 and p110 isoforms. We expect that if p85

is present in excess of p110, p85g will be detected in two distinct sets of fractions:

monomeric p85g without p110 will be found in earlier fractions of smaller molecular

weight, while heterodimeric p85g with p110 will be found in later fractions of higher

molecular weight. We additionally expect that compared to the control cells, PIK3R1

knockdown HMECs will have a reduced amount of monomeric p85g detected in earlier

fractions of lower molecular weight. Finally, to determine the absolute amounts of PI3K

subunits in our cells, we plan to perform mass spectrometry or immunoblotting of total

cell lysates in comparison to known amounts of recombinant p110 and p85 proteins.

Together, these experiments may strengthen the argument for the competition model we

have proposed here.

Page 108: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

91

One caveat to the immunoprecipitation experiments presented in this chapter to support

this competition model is the use of Flag-TEL-ErbB3 as an activated RTK probe for

bound PI3K isoforms. HER3/ErbB3 is generally considered to be a pseudokinase

because it lacks residues conserved among other HER/ErbB family RTKs thought to be

required for autophosphorylation and catalytic function (Guy et al., 1994; Jura et al.,

2009; Shi et al., 2010; Sierke et al., 1997); it instead propagates RTK signals by forming

heterodimers with other HER/ErbB family members, in particular EGFR/ErbB1 and

HER2/ErbB2 (Pinkas-Kramarski et al., 1996; Tzahar et al., 1996), and does not naturally

form homodimers (Berger et al., 2004). Here we use a fusion of ErbB3 to a TEL domain

that facilitates homodimerization and activation of RTKs (Carroll et al., 1996; Golub et

al., 1994; Jousset et al., 1997). Although it might be expected that ErbB3

homodimerization would therefore not activate RTK signaling, a recent report indicated

that ErbB3 might in fact retain weak catalytic activity (Shi et al., 2010). Our data

demonstrating that ectopic expression of Flag-TEL-ErbB3 in HMECs modestly increases

AKT phosphorylation under starvation conditions (Figure 2.17 A compared to Figure 2.2

A) and augments colony formation (Figure 2.16 compared to Figure 2.3 A) are

consistent with this report. Because ErbB3 possess direct p85-binding YXXM motifs

(Hellyer et al., 1998; Prigent and Gullick, 1994; Soltoff et al., 1994) while EGFR/ErbB1

and HER2/ErbB2 require adaptors to interact with p85 (Baselga and Swain, 2009), it

was an ideal probe for bound p85 and p110 in our system.

In addition to the competition model favored here, we explored other possible

explanations for the augmented PI3K/AKT signaling and transformation seen in HMECs

with reduced PIK3R1 expression. The BH domain of p85g has been reported to have

GAP activity towards Rab4 and Rab5, small GTPases critical for endosomal trafficking

(Chamberlain et al., 2004). Disruption of the Rab-GAP function of p85g leads to

Page 109: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

92

increased growth factor-stimulated PI3K/AKT activation and cellular transformation as a

result of more rapid and sustained Rab-mediated RTK internalization (Chamberlain et

al., 2004; Chamberlain et al., 2008; Chamberlain et al., 2010). Furthermore, a recent

report demonstrated that in MEFs, RNAi-mediated PIK3R1 knockdown increased the

amount of active GTP-bound Rab5, augmented PI3K/AKT pathway activation, and

induced autophagy (Dou et al., 2013). Accordingly, we examined whether p85g

downregulation affected growth factor-stimulated trafficking of RTKs in our HMEC lines.

We found that PIK3R1 knockdown did not have a substantial effect on EGFR

internalization or degradation in response to EGF stimulation (Figure 2.15). However,

we did not examine intracellular trafficking of EGFR by additional means, for example

immunofluorescence using fluorophore-conjugated EGF and known markers for different

intracellular compartments. We also did not explore whether PIK3R1 knockdown in

HMECs affected activation of Rab GTPases. Therefore, while our work suggests that the

increased growth factor-stimulated PI3K/AKT activation in these cells is likely not due to

altered RTK trafficking, we cannot completely rule out an effect on Rab GTPase

activation or function.

Based on published reports implicating p85g in PTEN binding (Chagpar et al., 2010;

Rabinovsky et al., 2009), stability (Cheung et al., 2011; Taniguchi et al., 2010), or lipid

phosphatase activity (Chagpar et al., 2010; Taniguchi et al., 2006), we also explored the

possibility that PIK3R1 knockdown had an effect on PTEN. Despite repeated attempts,

we were unable to confirm interaction of either endogenous or ectopically expressed

p85g with PTEN in HMECs or in a number of other cell types (Figure 2.11, Figure 2.12,

and data not shown). We also did not find that PIK3R1 knockdown substantially affected

PTEN mRNA or protein levels at the steady state (Figure 2.2 C, Figure 2.5 C, and

Figure 2.13 A). In addition, based on reports that in some contexts p85g is translocated

Page 110: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

93

to the nucleus (Chiu et al., 2014; Park et al., 2010; Winnay et al., 2010) and a multitude

of publications demonstrating a poorly-understood role for nuclear PTEN (reviewed in

(Planchon et al., 2008)), we examined whether PIK3R1 knockdown in HMECs affected

the subcellular localization of PTEN. Cellular fractionation experiments indicated that

p85g downregulation did not affect the proportion of PTEN associated with lipid

membrane, cytosolic, or nuclear fractions in either quiescent or EGF-stimulated HMECs

(data not shown). This is consistent with a recent study suggesting that p85g and PTEN

nuclear localization are not related (Chiu et al., 2014). Together our findings suggest that

in HMECs with PIK3R1 knockdown, upregulated PI3K/AKT signaling and transformation

may not be mediated by an effect of p85g directly on PTEN.

In summary, the data presented in this chapter demonstrates that the levels of p85g

modulate PI3K/AKT activation in mammary epithelial cells. RNAi-mediated PIK3R1

downregulation increases growth factor-stimulated PI3K signaling in and transformation

of HMECs. Partial reduction of p85g also synergizes with oncogenes common in breast

cancer, including p110g-H1047R and oncogenic HER2/neu. Both pan-PIγK and p110g-

selective pharmacological inhibitors are equally effective at blocking PI3K/AKT signaling

and colony formation mediated by PIK3R1 knockdown. Data from immunoprecipitation

of activated RTKs are consistent with a model where p85 is in excess of p110, and

monomeric p85 can compete with p85-p110 heterodimers for binding to RTKs to fine-

tune PIγK output. While others have reported a role for p85g in regulation of Rab-

mediated receptor trafficking, we find that PIK3R1 knockdown does not affect EGFR

internalization or degradation in HMECs. In contrast to published reports, we also do not

find that partial p85g reduction directly affects PTEN levels or lipid phosphatase activity.

The current literature on this topic, and the implications our work in this context, are

discussed in further detail in Chapter 4 of this dissertation.

Page 111: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

Chapter 3: PI3K regulatory subunit p85alpha plays a tumor suppressive role in

a genetically engineered mouse model of mammary tumorigenesis

Page 112: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

95

Acknowledgements

The HMS Rodent Histopathology Core prepared slides of fixed tissue for histological

analysis and performed hematoxylin and eosin (H&E) staining of these slides. Carolynn

Ohlson performed mammary tumor transplants, daily drug administrations, and

immunohistochemical staining of fixed tissue samples. Lauren Thorpe performed all

other experiments and data analysis.

I would additionally like to thank Lewis Cantley for the generous contribution of floxed

Pik3r1 mice, and William Muller for the generous contribution of MMTV-NIC mice. Thank

you to Thanh Von for sharing his extensive knowledge of in vivo techniques. Thank you

to Roderick Bronson of the HMS Rodent Histopathology Core for help in pathological

analysis. Thank you to Qi Wang for sharing protocols for mouse mammary epithelial cell

(MMEC) isolation, and Hailing Cheng for sharing protocols and reagents for the isolation

and culture of mouse tumor cells. Finally, I would like to thank Stephanie Santiago for

sharing tips and tricks for mouse mammary gland whole mount preparation.

Page 113: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

96

Preface

In this chapter, we use genetically engineered mouse models (GEMMs) to explore the

consequences of Cre/loxP-mediated conditional Pik3r1 ablation in the mouse mammary

epithelium. We find that p85g expression is not required for normal mouse mammary

gland development during puberty, pregnancy, or lactation. We then use a GEMM of

HER2/neu-driven breast cancer to demonstrate that Pik3r1 ablation significantly reduces

the latency of mammary tumor onset. When transplanted into recipient mice, the growth

of these tumors is blocked by treatment with pan-PIγK or p110g-selective inhibitors.

Together, these findings demonstrate the in vivo importance of p85g as a tumor

suppressor in the mammary epithelium, and suggest that isoform-selective PI3K

therapies may be effective in breast cancers characterized by a decrease in p85g.

Introduction

Genetically engineered mouse models are a convenient system in which to study

contributions of specific genes to both normal mammary gland development and

mammary tumorigenesis. Both mice and humans have three main stages of mammary

development: embryonic, pubertal, and adult (Watson and Khaled, 2008) (Figure 3.1).

During mouse embryonic development, five pairs of mammary buds form and undergo

limited initial branching, forming a rudimentary ductal tree. Following birth, terminal end

buds (TEBs) appear at the tips of the ducts and begin to invade the mammary fat pad;

during puberty, estrogen and other signals stimulate TEB proliferation and clefting,

resulting in ductal elongation and branching. By approximately 12 weeks after birth, the

mammary fat pad is filled, TEBs disappear, and ductal growth ceases. Adult mammary

glands undergo further change during pregnancy and lactation: progesterone and

prolactin induce side-branching and alveolar bud formation, differentiation into alveoli,

and milk secretion. After weaning, cell death, alveolar collapse, and remodeling occur

Page 114: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

97

Figure 3.1: Schematic of mammary gland development in the mouse. Mouse mammary gland development progresses through several distinct stages and is tightly regulated by a number of hormones and signaling pathways. At birth, a few rudimentary ducts surround the nipple. During puberty, these ducts undergo pronounced outgrowth and branching; at the end of puberty, the mammary fat pad is filled with a ductal network. During pregnancy, additional ductal branching occurs, along with extensive lobular-alveolar development; in preparation for lactation, the secretory epithelium undergoes functional differentiation, allowing milk to be produced and secreted. Following weaning, the alveolar compartment undergoes remodeling, termed involution, and the mammary gland returns to a pre-pregnancy-like state. Hormones and growth factors known to be important for transition from one state to another are indicated above blue arrows. Signaling pathways known to be important during certain stages are indicated above that stage. Adapted from (Hennighausen and Robinson, 2001) and (Hennighausen and Robinson, 1998).

Page 115: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

98

during a process called involution, returning the mammary gland to a pre-pregnancy-like

state (Hennighausen and Robinson, 1998). In addition to hormone signals, the different

stages of mammary gland development are tightly regulated by a number of other

signaling pathways (Hennighausen and Robinson, 2001; Hynes and Watson, 2010;

Watson and Khaled, 2008).

Several studies have established the important role of the PI3K pathway in mouse

mammary gland development and tumorigenesis. Conditional ablation of Pten in the

mammary epithelium led to accelerated ductal outgrowth and precocious lobulo-alveolar

development during puberty and pregnancy; these mice also frequently developed

mammary tumors with heterogeneous pathology as early as 2 months (Li et al., 2002).

Mice with mammary-specific expression of an inducible oncogenic p110g-H1047R

transgene developed mammary tumors with a mean latency of 7 months (Liu et al.,

2011), whereas Pik3ca ablation in the mammary epithelium significantly impaired

pubertal mammary gland branching and outgrowth and post-partum lactation, and also

blocked mammary tumor development driven by polyoma middle T antigen (pyMT) or

oncogenic HER2/neu (Utermark et al., 2012). Surprisingly, mammary-specific deletion of

Pik3cb resulted in modestly hypermorphic mammary gland development, with

precocious lobulo-alveolar growth and increased ductal branching, and moderately

accelerated pyMT- or HER2/neu-driven mammary tumor development (Utermark et al.,

2012). The distinct roles of p110g and p110く in the mammary gland were explained by a

proposed model in which p110g may have higher RTK-associated kinase activity than

p110く, allowing p110く to compete with p110g for binding sites on RTKs to regulate

PI3K output (Utermark et al., 2012) (Figure 1.6). This model was supported by extensive

biochemical experiments demonstrating increased RTK-bound p110g and RTK-

associated PIγK activity in mouse mammary epithelial cells with p110く knockout

Page 116: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

99

(Utermark et al., 2012). Together, these studies demonstrate that in mouse models,

genetic alterations that increase PI3K/AKT signaling lead to accelerated mammary gland

development and tumorigenesis, while alterations that reduce PI3K/AKT signaling

impede mammary gland development and block tumorigenesis in certain contexts.

A number of publications have suggested that in vivo, p85g expression levels modulate

PI3K/AKT activation. Mice with genetic ablation of the p85g isoform only (Terauchi et al.,

1999), of all three regulatory isoforms arising from Pik3r1 (p85g, p55g, and p50g)

(Fruman et al., 2000), or of p55g and p50g only (Chen et al., 2004) exhibited

hypoglycemia, increased insulin sensitivity, and increased PI3K/AKT pathway activation

upon insulin stimulation. Conversely, mice with increased p85g expression displayed

increased insulin resistance and reduced PtdIns(3,4,5)P3 production (Barbour et al.,

2005). Expression levels of p85g have also been shown to modulate pathophysiological

signals in mice. Heterozygous deletion of Pik3r1 increased the incidence of prostatic

intraepithelial neoplasia induced by heterozygous Pten knockout (Luo et al., 2005c) and

the number of lung tumors in a GEMM of lung cancer driven by oncogenic KRAS

(Engelman et al., 2008). Mice with liver-specific Pik3r1 ablation developed hepatitis and

dysplastic liver nodules by 6 months of age, and liver tumors resembling hepatocellular

carcinoma by 14 to 20 months (Taniguchi et al., 2010). In this study, liver lysates from

Pik3r1 knockout mice demonstrated upregulated PI3K/AKT activation and

PtdIns(3,4,5)P3 accumulation; liver tumors were found to have significantly reduced Pten

mRNA and protein levels (Taniguchi et al., 2010). This work indicates that reduced p85g

expression can increase PI3K/AKT signaling in vivo, and that additionally in some

tissues p85g downregulation augments tumorigenesis.

In this chapter, we examine the specific role of p85g in the mammary gland. We use

Page 117: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

100

conditional knockout techniques to determine the consequences of Pik3r1 ablation on

normal mouse mammary gland development. We then combine mammary-specific

Pik3r1 ablation with an established GEMM of HER2/neu-driven breast cancer to

examine the role of p85g in mammary tumor development. Finally, we explore the ability

of pan- and isoform-selective PI3K inhibitors to block in vivo mammary tumorigenesis in

the context of reduced p85g. As some of these agents are currently in early clinical trials

(Table 1.2 and Table 1.4), this work has important implications for therapeutic targeting

of breast cancers with reduced p85g expression.

Results

Pik3r1 expression is not required for mouse mammary gland development

The data presented in Chapter β indicates a tumor suppressive role for p85g in the in

vitro transformation of mammary epithelial cells. To determine whether these

observations hold true in vivo, we used GEMMs to evaluate the physiological and

pathophysiological consequences of p85g loss in this tissue. In mice, embryonic Pik3r1

knockout is lethal (Fruman et al., 2000). Therefore, we took advantage of the Cre/loxP

recombination system to conditionally ablate Pik3r1 in the mammary epithelium (Figure

3.2). Mice bearing a floxed Pik3r1 allele (Luo et al., 2005b) were backcrossed more than

ten generations to the FVB/N wildtype background to eliminate variations in mammary

development (MacLennan et al., 2011) and mammary tumor latency (Davie et al., 2007)

arising from different genetic backgrounds. To study the role of p85g in mouse

mammary gland development, these Pik3r1 floxed mice were then crossed with MMTV-

Cre transgenic mice, in which expression of the Cre recombinase is under control of the

MMTV LTR promoter and occurs in the secretory epithelium, including the mammary

gland, beginning at early stages of development (Wagner et al., 1997). The resulting

MMTV-Cre; Pik3r1+/loxP and MMTV-Cre; Pik3r1loxP/loxP mice have mosaic ablation of one

Page 118: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

101

Figure 3.2: Schematic of Pik3r1 conditional knockout allele and breeding scheme for mammary-specific Pik3r1 ablation. In mice, Pik3r1 generates three regulatory isoforms, p85g, p55g, and p50g, by alternative transcription initiation and splicing. The full-length isoform p85g has N-terminal SH3 and BH domains encoded by exons 1A through 6, while p55g and p50g have the unique first exons 1C and 1B, respectively. The first exon common to all three isoforms arising from Pik3r1 is exon 7. The conditional Pik3r1 knockout mouse used in these studies have knock-in of an engineered Pik3r1 allele in which exon 7 is flanked by loxP sites (Luo et al., 2005b); Cre recombinase mediates recombination of the loxP sites, resulting in deletion of exon 7. Transcription of the recombined gene produces a truncated p85g consisting of the SHγ and BH domains, but lacking the domains necessary for binding to p110g or activated RTKs, and essentially does not produce any p55g or p50g (Luo et al., 2005b). These floxed Pik3r1 mice were interbred with MMTV-Cre mice (Wagner et al., 1997), which express the Cre transgene under control of the MMTV LTR promoter, to achieve mammary-specific Pik3r1 ablation.

Page 119: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

102

or both Pik3r1 alleles mainly in luminal mammary epithelial cells, and will hereafter be

referred to as MMTV-Cre; Pik3r1+/- and MMTV-Cre; Pik3r1-/-.

To confirm the successful ablation of Pik3r1 in these mice, we isolated mouse mammary

epithelial cells (MMECs) from resected mammary glands of adult nulliparous female

MMTV-Cre; Pik3r1+/- and MMTV-Cre; Pik3r1-/- mice, generated protein lysates from

these cells, and used immunoblotting to determine expression of PI3K catalytic and

regulatory isoforms. Lysates of MMECs isolated from adult nulliparous female MMTV-

Cre mice were used as a wildtype control. MMECs derived from MMTV-Cre; Pik3r1+/-

and MMTV-Cre; Pik3r1-/- mice exhibited dramatically reduced p85g protein levels in

comparison to the MMTV-Cre control MMECs (Figure 3.3). In addition, as p85 is known

to be required for the stability of p110 (Fruman et al., 2000; Luo et al., 2005b; Yu et al.,

1998b), we were not surprised to find a reduction in p110g protein levels in MMTV-Cre;

Pik3r1+/- and MMTV-Cre; Pik3r1-/- MMECs. We also found that the level of p85g protein

varied slightly in each individual mouse, likely due to the mosaic nature of MMTV-Cre

expression (Wagner et al., 1997). Nonetheless, this protein analysis confirmed that this

GEMM could be used to determine the consequences of p85g loss in the mouse

mammary epithelium.

We then used whole mount techniques to analyze the effects of p85g loss in these mice

on the different stages of mammary gland development. The fourth inguinal mammary

glands from control and Pik3r1 floxed female mice were excised, fixed on slides, and

prepared with Carmine staining. Mammary gland development was assessed during

puberty in nulliparous female mice at 6 weeks of age, at the completion of puberty in 12-

week-old nulliparous females, during pregnancy at 14 days, and during lactation at 2

days postpartum (Figure 3.4). We found that there was no appreciable difference in the

Page 120: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

103

Figure 3.3: Transgenic MMTV-Cre ablates Pik3r1 expression in mouse mammary epithelial cells. MMECs were derived from the mammary glands of individual nulliparous MMTV-Cre, MMTV-Cre; Pik3r1+/-, and MMTV-Cre; Pik3r1-/- females. Protein lysates were prepared from these MMECs and subjected to immunoblotting for PI3K isoforms. The membrane was stripped and re-probed for vinculin as a loading control.

Page 121: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

104

Figure 3.4: Pik3r1 expression is not required for mouse mammary gland development. Whole mounts were prepared from the fourth inguinal mammary glands of MMTV-Cre; Pik3r1+/-, and MMTV-Cre; Pik3r1-/- female mice during puberty (A-C), at the end of puberty (D-F), during pregnancy (G-I), and during lactation (J-L). Mammary glands were stained with Carmine red to visualize ducts. Representative images from each genotype and stage are shown.

Page 122: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

105

mammary gland development of MMTV-Cre; Pik3r1+/- and MMTV-Cre; Pik3r1-/- mice as

compared to the MMTV-Cre control at all developmental stages examined. Halfway

through puberty, at 6 weeks of age, the duct outgrowth of all three genotypes was similar

relative to the lymph node (Figure 3.4 A-C). At the end of puberty at 12 weeks, the ducts

had completely filled the mammary fat pads in all three genotypes, and ductal branching

was comparable (Figure 3.4 D-F). Heterozygous or homozygous Pik3r1 ablation had no

substantial effect on lobulo-alveolar development during pregnancy (Figure 3.4 G-I) or

lactation (Figure 3.4 J-L). Together, these findings demonstrate that p85g is not

required for these stages of mammary gland development in the mouse.

Mammary-specific Pik3r1 ablation leads to spontaneous mammary tumor development

Although deletion of Pik3r1 did not have a substantial effect on mouse mammary gland

development, over a longer time frame nulliparous MMTV-Cre; Pik3r1+/- and MMTV-Cre;

Pik3r1-/- females eventually developed focal or multifocal spontaneous mammary

tumors. These tumors formed with an average latency of 14.1 months, an average

survival of 14.4 months, and a penetrance of 90% (9/10) (Table 3.1). In addition, we

found metastasis of the primary mammary tumor to the lungs in 11% (1/9) of the mice

with spontaneous mammary tumors. Many of the characteristics of mammary tumor

development in Pik3r1 knockout mice are comparable to those of other established

mouse models of breast cancer (Table 3.2).

To determine the expression of PI3K isoforms and activation of the PI3K/AKT pathway in

Pik3r1 knockout spontaneous mammary tumors, we generated protein lysates from

tumor tissue and subjected them to immunoblotting. Compared to whole mammary

gland lysates from MMTV-Cre nulliparous females, mammary tumors from MMTV-Cre;

Pik3r1+/- and MMTV-Cre; Pik3r1-/- mice had reduced p85g protein levels; tumors from

Page 123: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

106

Ta

ble

3.1

: N

ull

ipa

rou

s f

em

ale

mic

e w

ith

ma

mm

ary

-sp

ecif

ic P

ik3

r1 a

bla

tio

n

de

ve

lop

sp

on

tan

eo

us

ma

mm

ary

tu

mo

rs

Ge

no

typ

e

Mo

us

e ID

T

um

or

La

ten

cy

(mo

nth

s)

Su

rviv

al

(mo

nth

s)

Pri

ma

ry T

um

or

Pa

tho

log

y

Lu

ng

M

eta

sta

sis

?

Co

mm

en

ts

MM

TV

-Cre

; P

ik3

r1+

/-

A6

59

10

.7

11

.6

Un

diffe

ren

tia

ted

a

de

no

ca

rcin

om

a, sa

rco

ma

No

Fo

ca

l m

am

ma

ry tu

mo

r (1

)

MM

TV

-Cre

; P

ik3

r1+

/-

A6

09

12

.7

15

.9

Un

diffe

ren

tia

ted

a

de

no

ca

rcin

om

a, sa

rco

ma

No

Mu

ltifo

ca

l m

am

ma

ry tu

mors

(3

)

MM

TV

-Cre

; P

ik3

r1+

/-

A6

03

13

.5

14

.9

Un

diffe

ren

tia

ted

a

de

no

ca

rcin

om

a, sa

rco

ma

No

Mu

ltifo

ca

l m

am

ma

ry tu

mors

(2

)

MM

TV

-Cre

; P

ik3

r1+

/-

A6

05

17

.1

17

.8

ND

N

o

Mu

ltifo

ca

l m

am

ma

ry tu

mors

(2

)

MM

TV

-Cre

; P

ik3

r1+

/-

A6

07

17

.9

ND

N

D

No

MM

TV

-Cre

; P

ik3

r1-/

- A

66

0

ND

1

0.7

U

nd

iffe

ren

tia

ted

a

de

no

ca

rcin

om

a, sa

rco

ma

Ye

s

Fo

ca

l m

am

ma

ry tu

mo

r (1

)

MM

TV

-Cre

; P

ik3

r1-/

- A

60

6

12

.2

14

.1

Un

diffe

ren

tia

ted

a

de

no

ca

rcin

om

a, sa

rco

ma

No

Mu

ltifo

ca

l m

am

ma

ry tu

mors

(3

)

MM

TV

-Cre

; P

ik3

r1-/

- A

60

2

13

.9

15

.2

Ma

mm

ary

ade

noca

rcin

om

a,

ma

mm

ary

hyp

erp

lasia

N

o

Fo

ca

l m

am

ma

ry tu

mo

r (1

)

MM

TV

-Cre

; P

ik3

r1-/

- A

59

7

14

.7

14

.9

Un

diffe

ren

tia

ted

a

de

no

ca

rcin

om

a, sa

rco

ma

No

Mu

ltifo

ca

l m

am

ma

ry tu

mors

(2

)

MM

TV

-Cre

; P

ik3

r1-/

- A

61

8

NA

N

A

NA

N

A

Sa

cri

fice

d m

ou

se

at 2

2 m

on

ths

du

e t

o s

evere

de

rma

titis;

did

no

t h

ave

tu

mo

rs

ND

, not

de

term

ined;

NA

, no

t ap

plic

able

.

Page 124: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

107

Ta

ble

3.2

: C

om

pa

ris

on

of

ma

mm

ary

tu

mo

r d

ev

elo

pm

en

t in

Pik

3r1

k

no

ck

ou

t m

ice

to

oth

er

es

tab

lis

he

d G

EM

Ms

of

bre

as

t c

an

ce

r

Ge

no

typ

e

Tu

mo

r L

ate

nc

y

Tu

mo

r P

en

etr

an

ce

P

rim

ary

Tu

mo

r P

ath

olo

gy

Lu

ng

M

eta

sta

sis

?

Re

fere

nc

e

MM

TV

-Cre

; P

ik3

r1+

/- a

nd

M

MT

V-C

re;

Pik

3r1

-/-

14

.1 m

on

ths

9

0%

(9

/10

)

Mix

ed

(sa

rco

ma

, m

am

ma

ry

ad

en

oc

arc

ino

ma

)

11

% (

1/9

)

L.M

. T

ho

rpe

an

d J

.J. Z

hao

, u

np

ub

lis

he

d o

bs

erv

ati

on

s

MM

TV

-rtT

A; te

tO-P

ik3

ca

H1

04

7R

6.8

mo

nth

s

95

%

Mix

ed

ma

mm

ary

(a

den

oca

rcin

om

a,

ad

en

osq

ua

mo

us c

arc

ino

ma

)

NR

(L

iu e

t a

l., 2

01

1)

K1

4-C

re; p

53

-/- ;

Brc

a1

-/-

7.0

mo

nth

s

10

0%

M

am

mary

and

skin

(ID

C-n

os,

ca

rcin

osa

rcom

a,

ad

en

om

yo

ep

ith

elio

ma)

NR

(L

iu e

t a

l., 2

00

7)

MM

TV

-Cre

; E

rbB

2K

I 1

3.8

mo

nth

s

83

%

Fo

ca

l m

am

ma

ry c

om

ed

o-

ad

en

oca

rcin

om

as

6%

(A

nd

rech

ek e

t a

l., 2

00

0)

MM

TV

-NIC

6

.4 m

on

ths

10

0%

M

ultifo

ca

l, s

olid

no

du

lar

ma

mm

ary

ca

rcin

om

a

56

%

(Scha

de

et a

l., 2

00

9)

MM

TV

-Cre

; P

ten

+/-

15

.5 m

on

ths

75

%

So

lid n

od

ula

r m

am

mary

ca

rcin

om

a

18

% (

2/1

1)

(Do

urd

in e

t a

l., 2

00

8)

MM

TV

-Cre

; P

ten

-/-

11

mo

nth

s

75

%

Mix

ed

(fib

road

en

om

a,

ad

en

oca

rcin

om

a)

NR

(L

i e

t a

l., 2

00

2)

MM

TV

-Cre

; E

rbB

2K

I ; P

ten

+/-

6.5

mo

nth

s

10

0%

Mix

ed

ma

mm

ary

(so

lid n

odu

lar

ma

mm

ary

ca

rcin

om

a,

ad

en

om

yo

ep

ith

elio

ma

, a

de

no

sq

ua

mo

us c

arc

ino

ma

)

35

% (

6/1

7)

(Do

urd

in e

t a

l., 2

00

8)

N

R, not

rep

ort

ed.

Page 125: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

108

MMTV-Cre; Pik3r1-/- mice also generally had a greater reduction in total p85 levels when

compared to tumors from MMTV-Cre; Pik3r1+/- mice (Figure 3.5). Interestingly, though

PTEN loss has been reported in spontaneous liver tumors of mice with Pik3r1 knockout

(Taniguchi et al., 2010), we did not observe a pattern of PTEN protein reduction in

Pik3r1 knockout mammary tumors. In addition, we found that PI3K/AKT pathway

activation in these mammary tumors as assessed by phosphorylation of AKT and S6

ribosomal protein was highly variable (Figure 3.5). Thus while we find that mice with

mammary-specific Pik3r1 ablation develop spontaneous mammary tumors, our data is

inconclusive as to whether these tumors arise due to upregulated PI3K/AKT signaling.

We were further interested to know the pathology of spontaneous mammary tumors from

Pik3r1 knockout mice. Analysis of formalin-fixed tumor tissue by hematoxylin and eosin

(H&E) staining revealed that these tumors ranged in pathology, from sarcoma to

mammary adenocarcinoma (Figure 3.6 A-F). In addition, metastasis of the primary

mammary tumor to the lung was observed in 11% (1/9) of the mice with spontaneous

tumors (Figure 3.6 G-I). We also used whole mount techniques to examine the adjacent

mammary glands from Pik3r1 knockout mice with spontaneous mammary tumors.

Interestingly, non-tumor-bearing mammary glands from these mice displayed a slightly

hypermorphic phenotype with apparent lobulo-alveolar development and excessive

ductal branching (Figure 3.7) resembling mammary glands from pregnant females

(Figure 3.4 G-I). Together these findings suggest that Pik3r1 ablation alone is sufficient

for mammary tumor development in mice.

Pik3r1 ablation reduces the latency of HER2/neu-driven mouse mammary tumors

To further study the contribution of p85g loss to in vivo mammary tumorigenesis, we

combined a well-defined GEMM of HER2/neu-driven breast cancer with our conditional

Page 126: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

109

Figure 3.5: PI3K/AKT pathway activation in spontaneous mammary tumors from Pik3r1 knockout mice. Mammary tumor tissue was collected from random-fed MMTV-Cre; Pik3r1+/- and MMTV-Cre; Pik3r1-/- female mice. As a control, normal mammary gland tissue was collected from MMTV-Cre mice (“MG”). Protein lysates were generated and subjected to immunoblotting for PI3K/AKT pathway components and activation of PI3K/AKT pathway effectors. Membranes were stripped and re-probed for total proteins and vinculin as a loading control.

Page 127: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

110

Figure 3.6: Pathology of primary spontaneous mammary tumors and lung metastases from Pik3r1 knockout mice. (A-F) Representative images of formalin-fixed primary mammary tumor tissue from MMTV-Cre; Pik3r1+/- and MMTV-Cre; Pik3r1-/-

females stained with hematoxylin and eosin (H&E). These spontaneous tumors display a range of pathologies, from sarcoma and undifferentiated adenocarcinoma (A-C) to mammary hyperplasia and mammary adenocarcinoma (D-F). White arrowheads designate areas of the tumor where duct-like formations occur. Scale bars = 50たm. (G-I) Representative images of formalin-fixed lung tissue from one mouse in (A-F) stained with H&E. Black arrows designate metastases. Scale bars = 100たm.

Page 128: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

111

Figure 3.7: Adjacent mammary glands from Pik3r1 knockout mice with spontaneous mammary tumors have a hypermorphic phenotype. Mammary gland tissue from an approximately 1 year old nulliparous MMTV-Cre; Pik3r1-/- female with palpable mammary tumors was excised, fixed, and subjected to Carmine staining. The hypermorphic phenotype of these mammary glands, with what appears to be lobular-alveolar development and ductal branching similar to that occurring during pregnancy, can be seen clearly in (C) and (D), magnified images taken from the white boxes shown in (A) and (B) respectively.

Page 129: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

112

Pik3r1 knockout mice. We bred Pik3r1 floxed mice with MMTV-NIC mice, which express

a bicistronic transgene consisting of an activated HER2/neu allele and Cre recombinase

under control of the MMTV promoter (Schade et al., 2009; Ursini-Siegel et al., 2008)

(Figure 3.8). The resulting MMTV-NIC; Pik3r1+/loxP and MMTV-NIC; Pik3r1loxP/loxP mice

have ablation of one or both Pik3r1 alleles in the same luminal mammary epithelial cells

that express oncogenic HER2/neu. MMTV-NIC mice were used as a control. These mice

will hereafter be referred to as NIC, NIC; Pik3r1+/-, and NIC; Pik3r1-/-.

We monitored female cohorts of these mice for progression of mammary tumors. All

three genotypes developed multifocal mammary tumors with 100% penetrance. NIC

mice developed palpable tumors with a mean latency of 140 days (range 97-191, N =

25), consistent with other published information for this strain (Schade et al., 2009;

Ursini-Siegel et al., 2008; Utermark et al., 2012). Either heterozygous or homozygous

Pik3r1 ablation significantly reduced the time to tumor onset: NIC; Pik3r1+/- and NIC;

Pik3r1-/- mice developed tumors with mean latencies of 125 days (range 101-162, N =

38) and 126 days (range 86-155, N = 43), respectively (Figure 3.5 A). We additionally

determined the average number of mammary tumors, total tumor mass, and number of

lung metastases per mouse for all three genotypes five weeks after the onset of the first

palpable tumor. NIC mice developed on average 9.4 ± 1.5 tumors (N = 16), while NIC;

Pik3r1+/- mice had a significantly higher number of tumors, with on average 16.0 ± 1.2

tumors per mouse (N = 20); NIC; Pik3r1-/- mice developed a comparable number of

tumors to the NIC control, with on average 10.4 ± 4.3 tumors per mouse (N = 20)

(Figure 3.9 B). All three genotypes had comparable total tumor weight; NIC mice had an

average total tumor mass of 4.6 ± 0.8 grams (N = 16), while NIC; Pik3r1+/- and NIC;

Pik3r1-/- mice had an average total tumor mass of 5.4 ± 0.5 grams and 4.7 ± 0.4 grams,

respectively (N = 20 for both) (Figure 3.9 C). To determine the number of lung

Page 130: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

113

Figure 3.8: Schematic of the transgenic NIC allele and breeding scheme for mammary-specific HER2/neu expression and Pik3r1 ablation. MMTV-NIC mice are a well established genetically engineered model of HER2/neu-driven breast cancer (Schade et al., 2009; Ursini-Siegel et al., 2008). These mice express a bicistronic transgene consisting of a HER2/neu allele containing an activating in-frame deletion (Schade et al., 2009; Siegel et al., 1999; Ursini-Siegel et al., 2008) and the Cre recombinase gene, linked by an internal ribosomal entry site (IRES). This transgene is expressed under the control of the MMTV LTR promoter. MMTV-NIC mice develop multifocal mammary tumors with 100% penetrance and an average latency of 198 ± 43 days. Approximately 60% of MMTV-NIC mice also develop lung metastases. In this study, we interbred MMTV-NIC and conditional Pik3r1 knockout mice to achieve expression of activated HER2/neu and deletion of Pik3r1 in the same luminal mammary epithelial cells.

Page 131: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

114

Figure 3.9: Pik3r1 ablation reduces the latency of HER2/neu-driven mammary tumor development. (A) Cohorts of NIC, NIC; Pik3r1+/-, and NIC; Pik3r1-/- female mice were observed every three days for mammary tumor onset as determined by first palpation. Median tumor-free survival: NIC, 140 days (N = 25); NIC; Pik3r1+/-, 125 days (N = 38); NIC; Pik3r1-/-, 126 days (N = 43). Statistical significance was determined by Log-rank (Mantel-Cox) test. (B) The total number of mammary tumors per mouse was determined for NIC (N = 16), NIC; Pik3r1+/- (N = 20), and NIC; Pik3r1-/- (N = 20) females. (C) The total wet weight of mammary tumors plus associated mammary gland tissue per mouse was determined for NIC (N = 16), NIC; Pik3r1+/- (N = 20), and NIC; Pik3r1-/- (N = 20) females. (D) The number of lung metastases per mouse was determined for NIC, NIC; Pik3r1+/-, and NIC; Pik3r1-/- females (all groups N = 8). For (B-D), numbers were determined exactly 5 weeks after tumor onset. Means ± SEM are shown. Statistical significance was determined by unpaired t-test. Significance for NIC; Pik3r1+/- and NIC; Pik3r1-/- in comparison to the NIC control is shown. **, P < 0.01; ***, P < 0.001; ****, P < 0.0001; ns, not significant.

Page 132: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

115

metastases per mouse, we examined H&E-stained lung sections by microscopy.

Although the difference was not statistically significant, there was a trend towards higher

incidence of lung metastasis in mice with Pik3r1 ablation, with NIC mice having an

average of 2.7 ± 0.5 mets, and NIC; Pik3r1+/- and NIC; Pik3r1-/- having an average of 5.6

± 2.2 mets and 6.1 ± 4.9 mets, respectively (N = 8 for all groups) (Figure 3.9 D).

Together these results indicate that partial p85g loss reduces the latency of HER2/neu-

driven mammary tumorigenesis in mice, and may contribute to the severity of the

disease.

To analyze PI3K isoform expression and PI3K/AKT pathway activation in these tumors,

we isolated primary tumor tissue from random-fed mice, generated protein lysates, and

subjected them to immunoblotting (Figure 3.10). We also quantified the bands from

these blots by densitometry (Figure 3.11). Although there was some mouse-to-mouse

variation, likely owing to the mosaic expression of the MMTV LTR promoter (Wagner et

al., 1997), NIC; Pik3r1+/- mammary tumors had a 25.8 ± 15.1% reduction in p85g protein

levels, while NIC; Pik3r1-/- mammary tumors had a 83.9 ± 2.2% reduction in p85g protein

levels (N = 4 for both) (Figure 3.11 A). Expression of PI3K catalytic isoforms correlated

with the protein levels of p85g: NIC; Pik3r1+/- mammary tumors had a 46.5 ± 14.0%

reduction in p110g protein levels and a 9.8 ± 10.8% reduction in p110く levels, while

NIC; Pik3r1-/- mammary tumor cells had a 69.8 ± 4.8% reduction in p110g protein levels

and a 55.8 ± 5.4% reduction in p110く levels (N = 4 for all) (Figure 3.11 D-E). Notably,

although others have reported that p85g may important for the stability of PTEN

(Cheung et al., 2011; Taniguchi et al., 2010), we did not observe a significant change in

PTEN protein levels in NIC; Pik3r1+/- or NIC; Pik3r1-/- mammary tumors as compared to

the NIC control (Figure 3.11 F). Finally, we did not observe a significant difference in

activation of the PI3K/AKT pathway, as determined by phosphorylation of AKT and S6

Page 133: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

116

Figure 3.10: Effect of Pik3r1 ablation on PI3K/AKT pathway activation in HER2/neu-driven mammary tumors. Mammary tumor tissue was collected from random-fed NIC, NIC; Pik3r1+/-, and NIC; Pik3r1-/- female mice. Protein lysates were generated and subjected to immunoblotting for PI3K/AKT pathway components and activation of PI3K/AKT pathway effectors. Membranes were stripped and re-probed for total proteins and vinculin as a loading control.

Page 134: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

117

Figure 3.11: Quantification of the effect of Pik3r1 ablation on PI3K/AKT pathway activation in HER2/neu-driven mammary tumors. Bands in the immunoblots of NIC, NIC; Pik3r1+/-, and NIC; Pik3r1-/- mammary tumor lysates from Figure 3.10 were quantified by densitometry. Levels of p85g, pan-p85, p110g, p110く, and PTEN were normalized to the corresponding vinculin loading control on the re-probed membrane, then to the corresponding NIC control mean. Levels of phosphorylated HER2/neu, AKT, and S6 were normalized to the corresponding total proteins on the re-probed membrane, then to the corresponding NIC control mean. Means ± SEM are shown; N = 4 for each group. Statistical significance was determined by unpaired t-test. Significance is shown for NIC; Pik3r1+/- and NIC; Pik3r1-/- compared to the NIC control. *, P < 0.05; **, P < 0.01; ***, P < 0.001; ns, not significant.

Page 135: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

118

ribosomal protein (Figure 3.11 G-I). Thus although we have shown that partial p85g loss

reduces the latency of HER2/neu-driven mammary tumors, it is not clear from this data

whether PI3K signaling is upregulated in Pik3r1 knockout tumors.

We additionally examined the histology of formalin-fixed tumor tissue of each genotype.

Immunohistochemical (IHC) staining for p85g revealed that in NIC tumors, p85g was

mainly cytoplasmic, with strong and uniform signal throughout the tumor and perhaps

slightly elevated levels at the tumor edge (Figure 3.12 A); NIC; Pik3r1+/- tumors showed

a reduction in p85g, while NIC; Pik3r1-/- tumors had very little signal for p85g (Figure

3.12 B-C), correlating well with the immunoblot results (Figure 3.11 A). H&E staining of

tumor tissue showed that tumors from mice of all three genotypes had similar solid

nodular carcinoma histology (Figure 3.12 D-F); it is unsurprising that Pik3r1 ablation had

no effect on tumor pathology, since it has been shown previously that Pten ablation also

does not change the pathology of MMTV-NIC tumors (Schade et al., 2009). Finally, IHC

was used to stain for Ki67, a nuclear protein associated with cellular proliferation. Ki67

IHC revealed that compared to NIC tumors, which had 9.3 ± 0.9% proliferating cells,

NIC; Pik3r1+/- and NIC; Pik3r1-/- tumors had nearly double the proliferation indices, with

16.5 ± 2.1% and 17.5 ± 1.5% proliferating cells respectively (N = 12 for all groups)

(Figure 3.12 G-J). Together these results demonstrate that in mice, reduced p85g

expression significantly increases the proliferation of HER2/neu-driven mammary tumor

cells, correlating with a significant reduction in the latency of tumor onset.

Growth of Pik3r1 knockout tumors is blocked by p110g-selective inhibitors

Since mammary epithelial cell transformation driven by p85g loss is blocked in vitro by

pan-PI3K or p110g-selective inhibitors, we were interested to know whether these

agents could block growth of HER2/neu-driven mammary tumors with p85g loss in vivo.

Page 136: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

119

Figure 3.12: Effect of Pik3r1 ablation on tumor pathology and proliferation of HER2/neu-driven mammary tumors. Formalin-fixed tissue from NIC, NIC; Pik3r1+/-, and NIC; Pik3r1-/- mammary tumors was subjected to immunohistochemical staining for p85g (A-C), staining with hematoxylin and eosin (H&E) (D-F), or subjected to immunohistochemical staining for the proliferation marker Ki67 (G-I). The percentage of Ki67-positive nuclei (J) was calculated by dividing the number of positively-stained nuclei by the total number of nuclei in the field of view. Means ± SEM are shown; all groups N = 12. Statistical significance was determined by unpaired t-test. Significance is shown for NIC; Pik3r1+/- and NIC; Pik3r1-/- compared to the NIC control. **, P < 0.01; ***, P < 0.001. Scale bars = 50たm.

Page 137: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

120

Primary tumors were excised from a NIC; Pik3r1+/- donor female and orthotopically

transplanted into eight week old NcrNu female recipients. Recipient mice were randomly

assigned to four cohorts, and treated daily with methylcellulose vehicle, the pan-PI3K

inhibitor GDC0941 (125mg/kg), the p110g-selective inhibitor BYL719 (45mg/kg), or the

p110く-selective inhibitor KIN193 (20mg/kg); tumor size was measured every three days

using calipers. All agents were administered by oral gavage with the exception of

KIN193, which was administered by intraperitoneal injection. Compared to the vehicle

control, which had an average final tumor volume of 942.6 ± 222.9 mm3, KIN193 did not

have a substantial effect on tumor growth, with an average final tumor volume of 801.6 ±

111.1 mm3; treatment with either GDC0941 or BYL719 significantly blocked the growth

of transplanted tumors, with average final volumes of 299.5 ± 66.8 mm3 and 293.2 ±

85.0 mm3 respectively (N ≥ 10 for all cohorts) (Figure 3.13 A). Similar results were

obtained upon treatment of transplanted NIC; Pik3r1-/- tumors (data not shown).

Although all treatment groups maintained good body weight, signifying that none of the

drugs were excessively toxic over the course of treatment (Figure 3.13 B), it should be

noted that compared to the vehicle control, treatment with GDC0941 did lead to a slight

but significant percent decrease in body weight at later time points (Figure 3.13 C).

Together, these results indicate that pan-PIγK and p110g-selective inhibitors are

similarly able block in vivo tumor growth in the context of p85g loss.

To further study the effects of pan- and isoform-selective PI3K inhibitors on Pik3r1

ablated tumors in vivo, we analyzed activation of the PI3K pathway in tumors under each

treatment condition. Recipient mice were treated for four days, and tumors were excised

from recipient mice one hour after the last administration. A portion of each tumor was

used to generate protein lysates, which were then subjected to immunoblotting to assess

activation of components downstream of PI3K. Acute treatment with either

Page 138: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

121

Figure 3.13: Pan-PI3K or p110g-selective inhibitors block the growth of transplanted HER2/neu tumors with Pik3r1 ablation. NIC; Pik3r1+/- mammary tumors were orthotopically transplanted into female NcrNu mice. Recipients were randomly assigned to cohorts for once daily treatment with the pan-PI3K inhibitor GDC0941 (1β5mg/kg, oral gavage), the p110g-selective inhibitor BYL719 (45mg/kg, oral gavage), the p110く-selective inhibitor KIN193 (20mg/kg, intraperitoneal injection), or the vehicle control (methylcellulose, oral gavage). Tumor size was measured with calipers (A) and body weight was determined (B) every three days. Percent change in body weight (C) was calculated for each mouse relative to its weight at the start of treatment (day 0). Means ± SEM are shown; all treatment groups N ≥ 10. Statistical significance was determined by two-way ANOVA with Tukey’s multiple comparisons test. In (A), only significance for comparison of the final tumor volumes (day 18) of each treatment group to the vehicle control is shown. In (C), significance is shown for comparison of GDC0941 treatment to the vehicle control for days 12, 15, and 18; no other comparisons were statistically significant. *, P < 0.05; **, P < 0.01; ****, P < 0.0001; ns, not significant.

Page 139: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

122

GDC0941 or BYL719, but not KIN193, substantially reduced phosphorylation of AKT at

both activation sites and phosphorylation of S6 ribosomal protein at activation residues

S235/236 (Figure 3.14 A). The other portion of the excised treated tumor tissue was

fixed in formalin and used to assess histopathology. H&E staining revealed that the

transplanted tumors retained similar histology to the primary tumors (Figure 3.14 B-E).

IHC staining for AKT phosphorylated at S473 showed that tumors treated with vehicle or

KIN193 had moderate to strong nuclear and cytoplasmic AKT activation in most cells,

while tumors treated with GDC0941 or BYL719 had only slight AKT activation in a limited

number of cells (Figure 3.14 F-I). Similarly, IHC staining for S235/236-phosphorylated

S6 ribosomal protein showed strong cytoplasmic activated S6 signal in most cells of

tumors treated with vehicle or KIN193, while tumors treated with GDC0941 or BYL719

showed strong cytoplasmic S6 activation in only a few select cells (Figure 3.14 J-M).

These results demonstrate that pan-PI3K or p110g-selective inhibition, but not p110く-

selective inhibition, effectively decreases PI3K/AKT pathway signals in transplanted

HER2/neu-driven tumors with reduced p85g.

Finally, IHC staining was used to assess cellular proliferation and apoptosis in

transplanted tumors treated with pan-PI3K or isoform-selective inhibitors. Ki67 IHC was

used to visualize the nuclei of proliferating tumor cells. Compared to the vehicle-treated

sample with 34.2 ± 2.7% Ki67-positive nuclei, either GDC0941 or BYL719 treatment

significantly reduced the percentage of Ki67-positive nuclei to 9.7 ± 1.1% and 14.3 ±

1.4%, respectively, while KIN193 treatment had no effect on the proliferation index of

tumor cells, with 33.6 ± 2.9% Ki67-positive nuclei (N = 8 for all groups) (Figure 3.15 A-D

and Figure 3.15 I). IHC using the TUNEL method was performed to visualize the

fragmented DNA of cells undergoing apoptosis. While vehicle treatment did not induce

apoptosis of tumor cells, with only 1.2 ± 0.3% TUNEL-positive nuclei, either GDC0941

Page 140: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

123

Figure 3.14: Pan-PI3K or p110g-selective inhibitors suppress PI3K/AKT activation in transplanted HER2/neu tumors with Pik3r1 ablation. NIC; Pik3r1+/- mammary tumors orthotopically transplanted into NcrNu females and treated with the pan-PI3K inhibitor GDC0941, the p110g-selective inhibitor BYL719, the p110く-selective inhibitor KIN193, or the vehicle control as described in Figure 3.13. One hour after treatment on day 4, recipients were sacrificed and tumor tissue was collected. (A) Protein lysates from tumors in each treatment group were subjected to immunoblotting for PI3K/AKT pathway activation. Membranes were stripped and re-probed for total proteins and vinculin as a loading control. (B-M) Formalin-fixed tumor tissue was stained with hematoxylin and eosin (H&E) (B-E), or subjected to immunohistochemical staining for AKT phosphorylated at S473 (F-I) or S6 ribosomal protein phosphorylated at S235/236. Scale bars = 50たm.

Page 141: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

124

Figure 3.15: Pan-PI3K or p110g-selective inhibitors suppress proliferation and induce apoptosis in transplanted HER2/neu tumors with Pik3r1 ablation. NIC; Pik3r1+/- mammary tumors were orthotopically transplanted into NcrNu females. Recipients were treated with the pan-PI3K inhibitor GDC0941, the p110g-selective inhibitor BYL719, the p110く-selective inhibitor KIN193, or the vehicle control as described in Figure 3.13. One hour after treatment on day 4, recipients were sacrificed and tumor tissue was collected. Formalin-fixed tissue was subjected to immunohistochemical staining for the proliferation marker Ki67 (A-D) or for fragmented DNA using the TUNEL method (E-H). The percentage of Ki67-positive (I) or TUNEL-positive (J) nuclei was calculated by dividing the number of positively-stained nuclei by the total number of nuclei in the field of view. Means ± SEM are shown; all groups N = 8. Statistical significance was determined by unpaired t-test. Significance is shown for each treatment group compared to the vehicle control. **, P < 0.01, ****, P < 0.0001; ns, not significant. Scale bars = 50たm.

Page 142: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

125

or BYL719 treatment significantly increased the percentage of TUNEL-positive nuclei to

5.8 ± 1.0% and 4.8 ± 0.5%, respectively, while KIN193 treatment had no effect on tumor

cell apoptosis with 0.9 ± 0.3% TUNEL-positive nuclei (N = 8 for all groups) (Figure 3.15

E-H and Figure 3.15 J). Together this data demonstrates that pan-PI3K or p110g-

selective inhibitors, but not p110く-selective inhibitors, effectively block PI3K pathway

activation in mammary tumors in vivo in the context of p85g loss, and furthermore

reduce proliferation and induce apoptosis of mammary tumor cells.

Summary and discussion

In this chapter, we use Cre/loxP-mediated conditional deletion to study the effects of

Pik3r1 ablation on the mouse mammary gland. We find that Pik3r1 expression is not

required for mammary development during puberty, pregnancy, or lactation. However,

mice with mammary-specific Pik3r1 knockout develop spontaneous mammary tumors

with a mean latency of 14.1 months. We furthermore show that when combined with an

established GEMM of HER2/neu-driven breast cancer, mammary-specific Pik3r1

ablation significantly reduces the latency and increases the cellular proliferation of

mammary tumors. Growth of these tumors is equally and effectively blocked by pan-

PIγK or p110g-selective pharmacological inhibitors. Pan-PIγK or p110g-selective

therapeutics also block PI3K/AKT signaling, reduce proliferation, and induce apoptosis in

mammary tumors with Pik3r1 knockout. These findings help elucidate the previously

unstudied role of p85g in the mammary epithelium, and also have important implications

for therapeutic targeting of breast cancers with decreased p85g.

Because PI3K signaling is known to be important for mouse mammary gland

development, we were surprised to find that Pik3r1 ablation did not have an appreciable

effect on the morphology of this tissue. Previous studies have shown that increased

Page 143: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

126

PI3K activation via Pten ablation leads to hypermorphic mammary gland development

during puberty and pregnancy (Li et al., 2002), while reduced PI3K signaling due to

Pik3ca ablation significantly impairs pubertal mammary development and post-partum

lactation (Utermark et al., 2012). Based on our data in Chapter 2 demonstrating that

partial p85g loss leads to upregulated PIγK signaling, we expected to find mildly

accelerated mammary gland development in mice with conditional Pik3r1 ablation.

Although we do not have a definitive explanation for our observations, we propose that

because partial p85g loss has a subtle effect on PIγK activation in comparison to strong

oncogenes such as p110g-H1047R (Figure 2.2 D), it may not upregulate PI3K signaling

strongly enough in this tissue during development to substantially affect its morphology.

This idea is supported by the published result that Pten ablation leads to tumor

development in mice as early as 2 months (Li et al., 2002), while we find that the earliest

onset of spontaneous tumors in mice with mammary-specific Pik3r1 ablation is about

10.7 months (Table 3.1). Analysis of adjacent mammary glands from these mice using

whole mount techniques suggested that by the time spontaneous mammary tumors

developed, the remaining mammary glands without tumors showed signs of alveolar

differentiation (Figure 3.7) similar to normal mammary gland development during

pregnancy (Figure 3.4 G-I). It is likely that in mice with mammary-specific Pik3r1

ablation, PI3K/AKT signaling upregulation is modest and does not substantially affect

mammary gland development during puberty or pregnancy, but over a longer time period

leads to a slightly hypermorphic phenotype and spontaneous tumor development.

An additional confounding factor to our examination of the effect of conditional Pik3r1

ablation on mammary gland development is the levels of other PI3K regulatory and

catalytic isoforms. Although we were unable to identify an antibody capable of selectively

recognizing murine p85く by immunoblot (data not shown), we expect that MMTV-Cre;

Page 144: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

127

Pik3r1+/- and MMTV-Cre; Pik3r1-/- mice should have comparable levels of p85く to the

MMTV-Cre control. Retained Pik3r2 expression is expected to be critical for the normal

mammary gland morphology in our conditional Pik3r1 knockout mice, as p85く is

necessary to sustain PIγK signaling in conditions of p85g depletion (Ueki et al., 2002a);

ablation of both Pik3r1 and Pik3r2 would likely lead to substantially impaired mammary

gland development. This hypothesis is supported by the finding that conditional Pik3r1

ablation increases insulin-stimulated PI3K/AKT activity in the heart, while double Pik3r1

and Pik3r2 knockout largely blocks PI3K signaling (Luo et al., 2005a). Although Pik3r1

ablation should not affect p85く levels, we do find that it leads to significant reductions in

protein levels of both the p110g and p110く catalytic isoforms (Figure 3.3, Figure 3.5,

Figure 3.10, and Figure 3.11 D-E). Since it has previously been demonstrated that

Pik3ca ablation impairs mammary gland development during puberty and pregnancy,

while Pik3cb ablation leads to precocious lobulo-alveolar development (Utermark et al.,

2012), p110g and p110く downregulation likely contribute to the mammary phenotype of

mice with Pik3r1 ablation. The overall effect of mammary-specific Pik3r1 ablation on

PI3K/AKT activation and mammary gland development in these mice is probably a

combination of the changes to all class IA PI3K isoforms in this tissue.

Although Pik3r1 ablation does not have a substantial impact on mouse mammary gland

development, we find that 90% of MMTV-Cre; Pik3r1+/- and MMTV-Cre; Pik3r1-/- females

develop spontaneous mammary tumors by an average of 14.1 months (Table 3.1). This

timeline to spontaneous tumor development is comparable to the 14 to 20 month latency

for development of spontaneous liver tumors resembling hepatocellular carcinoma in

mice with liver-specific Pik3r1 ablation (Taniguchi et al., 2010), and to other GEMMs of

breast cancer (Table 3.2). While tumor lysates prepared from random-fed mice were

confirmed to have reduced p85g protein levels by immunoblotting, they exhibited

Page 145: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

128

variable PI3K/AKT activation (Figure 3.5). While we believe that these spontaneous

tumors likely arose due to persistent low-level upregulation of PI3K, which may be

difficult to demonstrate in tumor tissue from random-fed mice, it will also be interesting to

determine whether these tumors are prone to the accumulation of other genetic

alterations, as has been observed with other GEMM models of breast cancer (Liu et al.,

2011). Further study of these spontaneous Pik3r1 knockout mammary tumors using

RNA sequencing techniques will provide insight into this question.

Another puzzling result from our in vivo studies is the finding that in mice, heterozygous

or homozygous Pik3r1 ablation has a nearly identical effect on the latency of HER2/neu-

driven mammary tumors (Figure 3.9). Because heterozygous PIK3R1 loss is much more

frequent than homozygous deletion in human breast cancers (Figure 2.1 A), we

expected that in our mouse model heterozygous Pik3r1 ablation would be more

tumorigenic. While a number of studies have shown that either heterozygous (Mauvais-

Jarvis et al., 2002) or homozygous (Fruman et al., 2000; Terauchi et al., 1999) Pik3r1

ablation leads to hypoglycemia and enhanced insulin sensitivity as compared to wildtype

controls, few have made direct comparisons between the consequences of

heterozygous or homozygous loss of this gene on PI3K/AKT signaling. One publication

compared wildtype, heterozygous Pik3r1 knockout, and homozygous Pik3r1 knockout

MEFs, and found that while heterozygous knockout increased IGF1-stimulated PI3K

activity associated with p110g or tyrosine-phosphorylated proteins and robustly

enhanced IGF1-stimulated PtdIns(3,4,5)P3 production in comparison to wildtype MEFs,

homozygous Pik3r1 knockout had a less pronounced effect (Ueki et al., 2002a). Thus it

was unclear from the literature why heterozygous or homozygous Pik3r1 ablation would

similarly increase tumorigenesis in the mouse mammary gland.

Page 146: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

129

To explain our observations, we considered that based on our model from Chapter 2

(Figure 2.18) heterozygous Pik3r1 ablation in the mouse mammary epithelium might

partially reduce free p85g, while homozygous Pik3r1 ablation may completely reduce

free p85g and also partially reduce p85g-p110 heterodimers. Conceivably these two

scenarios could result in similar levels of PI3K/AKT signaling upregulation (Figure 3.16).

We sought to address this by immunoprecipitating endogenous HER3/ErbB3 from NIC,

NIC; Pik3r1+/-, and NIC; Pik3r1-/- mammary tumor lysates, and using immunoblotting to

assess the amount of p85 and p110 bound to these RTKs. Unfortunately, we found that

although we could achieve robust HER3/ErbB3 immunoprecipitation from tumor lysates,

the tissue homogenization process apparently disrupted all protein-protein interactions,

as no PI3K isoforms were detectable in the immunoprecipitates by immunoblot (data not

shown). We also cultured cells from these mammary tumors, generated protein lysates,

and again performed HER3/ErbB3 immunoprecipitations; this protocol resulted in

successful p85 and p110 pulldown, but the background on these immunoblots was too

high to reliably quantify the protein bands (data not shown). Verification of this proposed

explanation for our findings will likely require the use of other techniques such as mass

spectrometry.

In summary, the data presented in this chapter demonstrates the important role of p85g

in the mammary epithelium. While p85g expression is not required for normal mouse

mammary gland development during puberty, pregnancy, or lactation, mammary-specific

Pik3r1 ablation leads to spontaneous mammary tumor development accompanied by a

hypermorphic mammary gland phenotype within about one year. Pik3r1 ablation also

significantly reduces the latency of mammary tumor development driven by HER2/neu in

an established GEMM. Either pan-PIγK or p110g-selective inhibitors effectively block

growth of transplanted HER2/neu-driven mammary tumors with Pik3r1 ablation and

Page 147: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

130

Figure 3.16: Model: heterozygous or homozygous Pik3r1 ablation has a similar effect on HER2/neu-driven tumorigenesis. Pik3r1 ablation reduces the latency of HER2/neu-driven mammary tumor development in the NIC mouse model. Surprisingly, NIC; Pik3r1+/- and NIC; Pik3r1-/- female mice have nearly identical average time to tumor development. Here we show one possible explanation for these observations. Top: In NIC mice, p85 is present in excess of p110; p85 monomers and p85-p110 heterodimers compete for binding to activated RTKs, but only p85-p110 heterodimers can signal. Middle: Heterozygous Pik3r1 ablation partially depletes the pool of free p85, but some monomers remain. Bottom: Homozygous Pik3r1 ablation completely depletes the pool of free p85, but also reduces the number of p85-p110 heterodimers. By this model, heterozygous or homozygous Pik3r1 ablation could result in similar low-level upregulation of RTK-mediated PI3K signaling. While we expect that partial p85g loss similarly affects signaling by p110g and p110く, the p110く isoform is believed to have minimal contribution to RTK-mediated PI3K signaling due to substantially lower RTK-associated lipid kinase activity (Utermark et al., 2012).

Page 148: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

131

PI3K/AKT signaling in these tumors. The implications of these findings, along with the

data presented in Chapter 2, for the clinical application of PI3K-targeted therapies in

breast cancers with reduced p85g are discussed in Chapter 4 of this dissertation.

Page 149: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

Chapter 4: Summary, discussion, and future directions

Page 150: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

133

Summary

Research in the past decade has established that PI3K isoforms play critical but

divergent roles in both normal cellular signaling and in cancer. Chapter 1 discusses the

current state of this field: the distinct roles of PI3K isoforms in different signaling

contexts, the ways in which different PI3K isoforms are altered in cancer, and the

promise of emerging isoform-selective therapeutics in the clinic. The roles of class I PI3K

catalytic isoforms in cancer are relatively well studied; oncogenic mutation of p110g is

frequent in human cancers, while p110く, p110h, and p110け are rarely mutated but can

be overexpressed (Table 1.1 and Appendix A). Although PI3K regulatory isoforms had

not previously been considered to contribute to oncogenesis, recent studies have

converged to implicate p85g mutation or loss of expression in certain cancers. In this

dissertation we find that heterozygous deletion of PIK3R1 is a frequent event in human

breast cancers, and furthermore that PIK3R1 expression is significantly reduced in

breast tumors when compared to normal breast tissue. Based on these findings, we use

both in vitro and in vivo approaches to explore the role of p85g as a tumor suppressor in

the transformation of mammary epithelial cells.

In Chapter 2, we use RNAi-mediated knockdown of PIK3R1 to assess the effects of

p85g loss on human mammary epithelial cells (HMECs) in vitro. We find that partial

p85g reduction leads to increased growth factor-stimulated PI3K signaling in and

transformation of these cells. We further show that PIK3R1 knockdown augments HMEC

transformation by oncogenes, including activated HER2/neu. Using pharmacological

inhibitors, we demonstrate that the increased PI3K signaling and transformation driven

by partial p85g loss is largely mediated by p110g, and can be equally blocked by either

the pan-PI3K inhibitor GDC0941 or the p110g-selective inhibitor BYL719. Although

others have reported a role for p85g in the stabilization or activation of the PTEN

Page 151: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

134

phosphatase opposing PI3K, we were unable to demonstrate an effect of PIK3R1

knockdown on PTEN mRNA or protein levels, or on in vitro PTEN lipid phosphatase

activity. We also did not detect a change in RTK endocytosis or degradation, counter to

reports that p85g plays an important role in activation of Rabs critical for intracellular

trafficking. Instead, we find that partial loss of p85g increases the amount of p85-p110g

bound to activated RTKs. This result is consistent with a model in which p85 is in excess

of p110 in wildtype HMECs, allowing monomeric p85 to compete with p85-p110 for

binding to activated RTKs and negatively regulate PI3K signaling. PIK3R1 knockdown

might selectively reduce the pool of monomeric p85g, allowing more p85-p110

heterodimers to bind RTKs, increasing PI3K signaling and transformation (Figure 2.18).

In Chapter 3, we use the Cre/loxP system to study the effects of mammary-specific

Pik3r1 deletion on both normal mouse mammary gland development and mammary

oncogenesis. Surprisingly, we find that Pik3r1 expression is not required for mouse

mammary gland development during puberty, pregnancy, or lactation. However, mice

with mammary-specific Pik3r1 ablation develop spontaneous mammary tumors of

heterogeneous pathology with a mean latency of 14.1 months. Furthermore, Pik3r1

ablation significantly reduces the latency and increases the cellular proliferation of

mammary tumors in a mouse model of HER2/neu-driven breast cancer. Either the pan-

PI3K inhibitor GDC0941 or the p110g-selective inhibitor BYL719 effectively blocked

growth of transplanted Pik3r1 knockout tumors. Treatment of these tumors with

GDC0941 or BYL719 also considerably reduced PI3K/AKT activation, and significantly

reduced cellular proliferation and increased apoptosis. These findings indicate an

important role for p85g as a tumor suppressor in mammary tumor formation in vivo, and

furthermore suggest that pan-PI3K or p110g-selective inhibitors might be effective

therapeutics in breast cancers characterized by p85g loss.

Page 152: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

135

Discussion and future directions

In this dissertation, we present in vitro data using HMECs and in vivo data using GEMMs

demonstrating a tumor suppressive role for PIγK regulatory subunit p85g in mammary

epithelial cells. Our data is consistent with a model where p85 is present in excess of

p110 in normal mammary epithelial cells, allowing p85 monomers to compete with p85-

p110 heterodimers for binding sites on activated RTKs to modulate PI3K signaling. We

additionally show that in the context of transformation mediated by partial p85g loss, the

efficacy of p110g-selective inhibitors is comparable to that of pan-PI3K inhibitors. Our

findings have several important implications for the current literature regarding p85g-

mediated transformation, the regulation of RTK-mediated PIγK signals by p85g, and the

therapeutic targeting of breast cancers characterized by partial p85g loss. They also

raise interesting questions about the role of differential p85g expression in various

tissues and metabolic contexts. These concepts will be important to address in

continuing work on this project.

Transformation mediated by partial p85g loss does not appear to be through PTEN

A number of recent reports link p85g to PTEN stability or lipid phosphatase activity, and

some studies have demonstrated an interaction between these two proteins. This

proposed connection between p85g and PTEN has been invoked to explain increased

PI3K/AKT signaling and transformation in the context of p85g mutation or

downregulation (Cheung et al., 2011; Taniguchi et al., 2006; Taniguchi et al., 2010). The

results presented in this dissertation suggest that further study is needed to conclude

whether PTEN contributes to transformation mediated by partial reduction of p85g. At a

minimum, our data suggest that these reported effects of p85g on PTEN may not be

operative in all tissues.

Page 153: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

136

An interaction between p85g and PTEN has been reported by two separate groups. In

one study p85g, p85く, and p110く were found to participate in a complex with

unphosphorylated PTEN (Rabinovsky et al., 2009); a second study demonstrated direct

binding between the C-terminal SHγ and BH domains of p85g and PTEN that was

dependent on EGF stimulation only in certain cell types (Chagpar et al., 2010). Thus the

exact nature of the proposed p85g-PTEN interaction and the contexts in which these

proteins bind are still not clear. Although we were unable to positively show co-

immunoprecipitation of p85g and PTEN in a number of different cell types and conditions

(Figure 2.11, Figure 2.12, and data not shown), we cannot absolutely rule out this

interaction. It is possible that the antibodies we used for our co-immunoprecipitation

assays obscured protein domains necessary for binding; although we tried many of the

antibodies used in these two publications, at least one was not a commercial reagent

(D.H. Anderson, personal communication). However, we would note that at least one

other study reported being unable to demonstrate binding of p85g and PTEN (Taniguchi

et al., 2006).

Other publications have indicated that p85g is important for PTEN mRNA or protein

stability. Ectopic expression of the cancer-associated truncation mutant p85g-E160*

reportedly led to PTEN protein destabilization via ubiquitin-mediated proteasomal

degradation, while overexpression of wildtype p85g was shown to stabilize PTEN protein

levels (Cheung et al., 2011). In another study, mice with liver-specific Pik3r1 ablation

developed liver tumors resembling hepatocellular carcinoma within 14 to 20 months;

compared to liver lysates from these mice at 6 months of age, livers from mice aged 16

to 18 months exhibited upregulated AKT phosphorylation and PtdIns(3,4,5)P3

production, and a corresponding reduction in both PTEN protein and mRNA levels

(Taniguchi et al., 2010). In HMECs, we found that RNAi-mediated PIK3R1 knockdown

Page 154: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

137

did not have a substantial effect on steady-state PTEN mRNA levels (Figure 2.13 A) or

protein levels (Figure 2.2 C and Figure 2.5 C). We also did not find that ablation of one

or both Pik3r1 alleles had a consistent effect on PTEN protein levels either in

spontaneous mammary tumors as compared to wildtype mammary glands (Figure 3.5)

or in mammary tumors driven by HER2/neu (Figure 3.10 and Figure 3.11 F). These

findings corroborate a report that liver-specific Pik3r1 ablation had no effect on PTEN

protein levels in this tissue (Taniguchi et al., 2006). Together our results indicate that

p85g may not be important in the regulation of PTEN mRNA or protein levels in

mammary epithelial cells. Our data furthermore suggests that PTEN mRNA and protein

reduction found in spontaneous Pik3r1 knockout liver tumors may be a secondary event

that is not necessarily linked to p85g downregulation.

Finally, p85g has been reported to regulate PTEN lipid phosphatase activity. In liver

extracts from mice with liver-specific Pik3r1 ablation, PtdIns(3,4,5)P3 production and

AKT activation were found to be upregulated; although Pik3r1 knockout had no effect on

PTEN protein levels, it significantly reduced both basal and insulin-stimulated PTEN lipid

phosphatase activity (Taniguchi et al., 2006). Another study reported that in in vitro

assays, addition of purified p85g increased PTEN lipid phosphatase activity in a

concentration-dependent manner (Chagpar et al., 2010). However, we did not find a

difference in the in vitro lipid phosphatase activity associated with PTEN

immunoprecipitates from shControl and shPIK3R1 HMEC lines (Figure 2.13 C). This

data suggests that at least in this cell type, p85g may not modulate PTEN activity.

Together our results indicate that in both mouse and human mammary epithelial cells,

partial p85g loss does not substantially alter PTEN levels. They also suggest that in

HMECs, p85g may not have a significant effect on PTEN activity. We conclude it is

Page 155: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

138

unlikely that the observed transformation mediated by p85g downregulation in our model

systems is due to reduced PTEN function. It would be especially interesting to carry out

experiments examining the ability of partial p85g loss to increase transformation of

PTEN-null mammary epithelial cells. One approach would be to use RNAi techniques to

silence PIK3R1 in PTEN-null breast cancer cell lines or in HMECs with CRISPR-

mediated PTEN deletion. A parallel approach would be to use Cre/loxP-mediated

conditional ablation to knock out both Pik3r1 and Pten in the mouse mammary

epithelium. The ability of reduced p85g to augment transformation and tumorigenesis in

the complete absence of PTEN would further support the idea that this mechanism is

independent of proposed effects of p85g on PTEN stability or activity.

Implications of the competition model for transformation mediated by p85g loss

We present data demonstrating that in HMECs, PIK3R1 knockdown increases PI3K/AKT

pathway signaling in response to growth factors, cellular transformation, and the amount

of p85-p110g associated with activated RTKs. Our results are consistent with a model

where partial p85g loss selectively reduces a pool of monomeric p85 which competes

with p85-p110 heterodimers for binding activated RTKs to negatively regulate PI3K/AKT

signaling downstream of RTKs (Figure 2.18). While there are a number of remaining

experiments that would more conclusively support this model, discussed in Chapter 2 of

this dissertation, it has several important implications for p85-mediated regulation of

PI3K signaling, and the general role of p85 regulatory isoforms in transformation.

Although our work here has focused on PIK3R1, there are three genes encoding class

IA regulatory isoforms: PIK3R1, encoding p85g and its splicing variants p55g and p50g,

PIK3R2, encoding p85く, and PIK3R3, encoding p55け. These five isoforms are

collectively called p85 type regulatory subunits. All p85 isoforms possess the iSH2

Page 156: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

139

domain responsible for binding class IA catalytic isoforms, and the SH2 domains

responsible for binding to phosphorylated YXXM motifs on activated RTKs or their

adaptors (Figure 1.1). The experiments presented in this dissertation reduced or ablated

expression of all three p85 isoforms arising from PIK3R1, leaving expression of PIK3R2

and PIK3R3 intact. We expect that the remaining p85 isoforms are necessary to sustain

PI3K signaling in the context of reduced PIK3R1 expression.

Our model predicts that modulation of the ratio of p85 monomers to p85-p110

heterodimers might produce a range of RTK-mediated PI3K activity (Figure 4.1). When

p85 is present in excess of p110, p85 monomers might negatively regulate RTK-

mediated PI3K output. According to this model, PI3K signaling downstream of RTKs

would be maximal when p85 isoform expression is reduced exactly to the level where

the number of p85 subunits is equal to the number of p110 subunits; further reduction in

p85 expression would decrease the number of signaling-capable p85-p110

heterodimers, reducing RTK-mediated PI3K signaling. Consistent with this idea, others

have shown that ablation of both Pik3r1 and Pik3r2 significantly impairs PI3K lipid kinase

activity and AKT phosphorylation in mouse cardiac tissue (Luo et al., 2005b). In a

separate study using a GEMM of KRAS-driven lung cancer, Pik3r1+/-; Pik3r2-/- mice

developed a greater number of lung tumors in comparison to Pik3r2-/- mice, while tumor

development was nearly completely blocked in Pik3r1-/-; Pik3r2-/- mice (Engelman et al.,

2008). These results support the idea that a balance between p85 and p110 isoforms is

critical for regulation of PI3K signaling.

If our model is correct, it might also be expected that partial loss of any p85 isoforms

could increase RTK-mediated PI3K signaling. We have performed preliminary

experiments indicating that in HMECs, RNAi-mediated PIK3R2 knockdown increases

Page 157: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

140

Fig

ure

4.1

: M

od

el:

mod

ulat

ion

of p

85g

leve

ls m

ight

pro

du

ce

a r

an

ge o

f R

TK

-me

dia

ted

PI3

K a

cti

vit

y.

Le

ft:

Wh

en

the a

mount

of

p85

is v

ery

lo

w, p1

10 le

vels

are

concom

ita

ntly r

educed,

sin

ce

p8

5 is r

equir

ed

for

p110

sta

bili

ty; fe

w p

85

-p1

10

he

tero

dim

ers

are

availa

ble

fo

r bin

din

g t

o a

ctivate

d R

TK

s.

In t

his

conte

xt, R

TK

-media

ted

PI3

K s

ignalin

g m

ight

be

min

imal. M

iddle

: W

he

n p

85

an

d p

110

are

pre

se

nt in

eq

ual am

oun

ts, th

ere

is n

o m

onom

eri

c p

85 t

o c

om

pete

with

he

tero

dim

eric p

85

-p1

10 f

or

bin

din

g to a

ctivate

d R

TK

s; all

bin

din

g s

ite

s c

an

be

occu

pie

d b

y s

ign

alin

g p

85

-p110

he

tero

dim

ers

. In

th

is c

on

text,

RT

K-m

edia

ted

PI3

K s

ignalin

g m

ight

be m

axim

al.

Rig

ht:

Wh

en

p8

5 le

vels

are

very

hig

h, p

85 r

eg

ula

tory

isofo

rms m

ay b

e in

exce

ss o

f p110

cata

lytic isofo

rms; both

mono

meri

c p

85

an

d

hete

rodim

eric p

85

-p11

0 c

om

pete

for

bin

din

g t

o a

ctivate

d R

TK

s,

bu

t only

p85

-p1

10

he

tero

dim

ers

ca

n s

igna

l. In

this

co

nte

xt, R

TK

-m

edia

ted

PI3

K s

ignalin

g m

igh

t be

lo

w.

In a

ll ca

ses, w

hile

p85

leve

ls a

re e

xpec

ted

to s

imila

rly m

odul

ate

RTK

bin

ding

of p

110g

- a

nd

p1

10く-

co

nta

inin

g P

I3K

he

tero

dim

ers

, p11

0g is

the

mai

n ca

taly

tic is

ofo

rm m

edia

ting P

I3K

activation d

ow

nstr

ea

m o

f R

TK

in

puts

(U

term

ark

et

al., 2

012

).

Page 158: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

141

growth factor-stimulated PI3K/AKT activation and anchorage-independent growth (data

not shown). Others have shown that Pik3r2-/- mice display increased insulin sensitivity,

and cells derived from these mice showed heightened insulin-stimulated AKT

phosphorylation (Ueki et al., 2003; Ueki et al., 2002b). However, two recent publications

demonstrated that PIK3R2 overexpression in chicken embryo fibroblasts (CEFs) or

immortalized murine fibroblasts (NIH 3T3) increases PI3K/AKT activation and cellular

transformation (Cortes et al., 2012; Ito et al., 2014). Furthermore, these studies

suggested that PIK3R2 is overexpressed in human breast, colon, and ovarian cancers

(Cortes et al., 2012; Ito et al., 2014). It has been proposed that p85く-p110g has greater

kinase activity towards PtdIns(3,4,5)P3 than p85g-p110g (Cortes et al., 2012) or that

p85く is a less effective inhibitor of p110 catalytic activity (Ito et al., 2014). We queried

https://www.oncomine.org and http://www.cbioportal.org, the same online databases we

used to detect significant PIK3R1 underexpression in breast cancers (Figure 2.1 and

Table 2.1), to determine whether PIK3R2 alterations were a common event in human

cancers. However, we did not identify a significant trend towards either overexpression

or underexpression of PIK3R2 (data not shown). We conclude that further study is

needed to clarify the role of PIK3R2 expression in transformation.

Our model also relies on an imbalance in the number of class IA p85 regulatory and

p110 catalytic subunits. The presence of free p85 is currently a source of controversy in

the field, with publications arguing both for and against the existence of p85 monomers.

One study used quantitative mass spectrometry to determine the absolute number of

p85 and p110 molecules in murine B lymphocytes (WEHI-231), murine fibroblasts (NIH

3T3), and various homogenized mouse tissues (Geering et al., 2007). This work

demonstrated that in WEHI-231 and NIH 3T3 cells, and in mouse muscle, liver, fat, and

spleen tissue, p85 and p110 are present in approximately equal amounts; only in the

Page 159: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

142

mouse brain was p85 found to be present in excess (Geering et al., 2007). In this study,

it was concluded that class IA PI3Ks are obligate p85-p110 heterodimers.

In contrast, two studies from the Cantley and Kahn groups used immunodepletion

techniques on wildtype and Pik3r1 ablated mouse cells and tissues to demonstrate the

presence of p85 monomers (Mauvais-Jarvis et al., 2002; Ueki et al., 2002a). From these

results it was estimated that the ratio of p85-p110 dimers to p85 monomers was 2:1 in

wildtype MEFs, 3:1 in heterozygous Pik3r1 knockout MEFs, and 7:1 in homozygous

Pik3r1 knockout MEFs (Ueki et al., 2002a); in mouse livers, the ratio of p85-p110 dimers

to p85 monomers was estimated to be 2:1 in wildtype mice and 4:1 in heterozygous

Pik3r1 knockout mice (Mauvais-Jarvis et al., 2002). However, none of these studies

examined mammary cells or tissues. We used similar immunodepletion techniques in

our HMECs, but while these results suggested that p85 might be in excess in these

cells, our findings were inconclusive (data not shown). It is possible that the relative

levels of p85 and p110 may vary depending on the cell type or tissue. In fact, in

preliminary studies we found that RNAi-mediated downregulation of p85g in MCF10A

human mammary epithelial cells or in HC11 murine mammary epithelial cells increased

PI3K/AKT signaling in response to EGF or insulin, while Pik3r1 knockdown in wildtype

MEFs had no effect on growth factor-stimulated PI3K/AKT activation (data not shown). It

will be critical to assess the relative numbers of p85 and p110 isoforms in different cell

and tissue types using sensitive techniques such as quantitative mass spectrometry.

If correct, our model also has several implications for RTKs relative to PI3K subunits.

First, the ratio of available RTKs to PI3K molecules will be important. It is likely that in

mammary epithelial cells, the number of p85 monomers plus p85-p110 heterodimers is

in excess in comparison to the number of RTKs, even when Flag-TEL-ErbB3 is

Page 160: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

143

ectopically expressed. If instead RTKs were in excess, we would expect that partial

reduction of p85g would have no effect on PIγK/AKT signaling, since all p85-p110

heterodimers would already be able to complex with activated RTKs in this instance.

Second, our model makes the assumption that p85 monomers and p85-p110

heterodimers bind to RTKs with approximately the same affinity, allowing monomers and

heterodimers to compete for the same active sites. We could begin to address this by

performing immunoprecipitations of Flag-TEL-ErbB3 similar to those shown in Chapter

2, and comparing the ratio of p85 to p110 in immunoprecipitates in comparison to the

non-immunoprecipitated supernatant; in this experiment, we would expect the ratio of

p85 to p110 to be similar in both the immunoprecipitate and supernatant fractions.

However, a more definitive demonstration will require careful biochemical experiments

with purified proteins.

Finally, our data indicates that increased PIγK/AKT signaling in contexts of partial p85g

reduction may be specific to RTK inputs. Our model predicts that p85g downregulation

might only affect PI3K signaling downstream of receptors for which p85 serves as a

necessary binding adaptor for the PI3K heterodimer. Growth factor-stimulated PI3K

signaling relies on binding of the SH2 domains of p85 subunits to phosphorylated motifs

on activated RTKs, recruiting class IA p110 catalytic subunits and activating their lipid

kinase activity (Figure 1.2). A number of studies have demonstrated that p110g is the

primary catalytic subunit involved in growth factor-stimulated PI3K activation (Foukas et

al., 2006; Graupera et al., 2008; Knight et al., 2006; Utermark et al., 2012; Zhao et al.,

2006). Consistent with this, we find that increased signaling and transformation mediated

by partial p85g loss is blocked by p110g-selective pharmacological inhibition. It is likely

that p85g depletion similarly affects both p110g and p110く, but since p110く plays a

minor role in RTK-mediated PIγK signaling, effects on p110く are difficult to detect.

Page 161: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

144

Interestingly, of the class IA catalytic isoforms only p110く has been shown to also

mediate GPCR-activated PI3K signaling (Ciraolo et al., 2008; Guillermet-Guibert et al.,

2008; Jia et al., 2008). A recent study demonstrated that a unique region of p110く

directly binds GPCR-associated Gくけ subunits, facilitating p110く-mediated PI3K signaling

downstream of GPCRs (Figure 1.5) (Dbouk et al., 2012). This suggests that p85

subunits may not function as adaptors for PI3K signaling downstream of GPCRs,

consistent with our data demonstrating that p110く-selective inhibition does not suppress

PI3K/AKT signaling and transformation brought about by a partial reduction in p85g. It

may be that p85g downregulation uniquely synergizes with RTK-activating signals. It will

be interesting to test this idea using GPCR agonists such as LPA to stimulate our control

and PIK3R1 knockdown HMECs. If correct, we would expect to see no difference in

LPA-stimulated PI3K/AKT activation between these cell lines.

Implications for therapeutic targeting of PI3K in cancers with p85g loss

The discovery of frequent PI3K activation in cancers has made this pathway an

attractive target for small-molecule therapeutics. The first PI3K inhibitors to enter clinical

trials were pan-PI3K inhibitors (Table 1.2) and dual pan-PI3K/mTOR inhibitors (Table

1.3). Unfortunately, many of these drugs have had only modest single-agent success in

the clinic (Rodon et al., 2013). This is in part due to compensatory feedback signaling

networks which facilitate the development of resistance to targeted therapies; one major

focus in the field is to identify these resistance pathways and develop rational

combination therapy strategies to circumvent them (Figure 1.8). Another limiting aspect

of many pan-PI3K inhibitors is their broad spectrum of off-target effects on PI3K-related

kinases and other cellular components (Fruman and Rommel, 2014). This is likely

because inhibitors targeting the active site of all class I PI3Ks will necessarily be

promiscuous (Knight and Shokat, 2005). Isoform-selective PI3K inhibitors are now

Page 162: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

145

emerging in clinical trials (Table 1.4), and should theoretically have both fewer off-target

effects on molecules outside the PI3K family, and fewer on-target toxicities by sparing

PI3K isoforms that are not contributing to tumorigenesis. Some p110g-selective

inhibitors have shown promising success in early clinical trials (Juric et al., 2013b), and

the p110h-selective inhibitor idelalisib has recently become the first FDA-approved PI3K

therapy due to its significant success in patients with B cell malignancies. Thus another

major focus of the field is to identify the contexts in which isoform-selective PI3K

inhibitors will be successful (Figure 1.7).

Using isoform-selective PI3K inhibitors, we demonstrate that the transformation of

mammary epithelial cells with partial p85g loss is primarily mediated by catalytic isoform

p110g. Either the pan-PI3K inhibitor GDC0941 or the p110g-selective inhibitor BYL719

effectively inhibited colony formation of PIK3R1 knockdown DDp53-HMECs, while

p110く-selective inhibition had no effect (Figure 2.8). GDC0941 and BYL719 also

substantially reduced PI3K/AKT activation in and colony formation of DDp53-HMECs

expressing oncogenic HER2/neu in conjunction with knockdown of PIK3R1 (Figure 2.9

and Figure 2.10). Pan-PIγK or p110g-selective pharmacological inhibition of p110g

blocked the in vivo growth of transplanted tumors with Pik3r1 ablation, while the p110く-

selective inhibitor KIN193 did not affect tumor growth (Figure 3.13). IHC and

immunoblot analysis of tumor tissue revealed that GDC0941 or BYL719 treatment

suppressed PI3K/AKT pathway activation in Pik3r1 knockout tumors, while the effect of

KIN193 treatment on PI3K/AKT signaling was indistinguishable from that of the vehicle

control (Figure 3.14). Pan-PIγK and p110g-selective inhibition also significantly reduced

cellular proliferation and increased the percentage of cells undergoing apoptosis in

transplanted tumors (Figure 3.15). These results demonstrate that in mammary

epithelial cells with partial p85g loss, the efficacy of p110g-selective pharmacological

Page 163: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

146

inhibition at blocking transformation and tumorigenesis is comparable to that of a pan-

PI3K agent.

It is perhaps unsurprising that transformation in this context is governed by PI3K

signaling through p110g. If our model for increased PIγK signaling mediated by partial

p85g loss (Figure 2.18) is correct, downregulation of p85g should only augment PIγK

signaling through RTKs, and not through other inputs such as GPCRs. A number of

studies using pharmacological inhibition and genetic inactivation or ablation have

demonstrated that PI3K signaling in response to growth factors is principally through

p110g (Foukas et al., 2006; Graupera et al., 2008; Knight et al., 2006; Sopasakis et al.,

2010; Utermark et al., 2012; Zhao et al., 2006) and not p110く (Ciraolo et al., 2008;

Guillermet-Guibert et al., 2008; Jia et al., 2008). Furthermore, Pik3ca ablation or p110g-

selective inhibition was sufficient to block mouse mammary tumorigenesis driven by

HER2/neu in a recent study from our group (Utermark et al., 2012). These findings were

explained by a proposed model where p110g may have higher RTK-associated lipid

kinase activity than p110く, making p110g the primary catalytic isoform mediating PI3K

signaling downstream of RTKs (Utermark et al., 2012) (Figure 1.6). Based on this

model, oncogenic lesions activating RTK signaling will be effectively targeted by p110g-

selective inhibitors (Figure 1.7).

Early clinical data has reported promising activity of p110g-selective agents BYL719 or

GDC00γβ in advanced breast tumors with p110g activation (Juric et al., 2013b). The

data presented in this dissertation suggests that such inhibitors may also be successful

in treatment of breast tumors with reduced expression of PIK3R1. Furthermore, another

preclinical study has demonstrated that p110g- but not p110く-selective inhibitors block

both PI3K/AKT signaling in and cellular transformation of cells with ectopic expression of

Page 164: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

147

cancer-associated p85g mutants (Sun et al., 2010). Together this work highlights the

potential for use of p110g-selective agents in treatment of cancers with p85g loss or

mutation. Since loss of PIK3R1 expression is common in breast cancers (Figure 2.1 and

Table 2.1) (Cizkova et al., 2013) and certain other cancer types (Taniguchi et al., 2010),

and PIK3R1 mutation is a frequent event particularly in endometrial and pancreatic

cancers (Table 1.1 and Appendix A), our findings along with other recent publications

emphasize the need to consider alterations in PIK3R1 as a diagnostic marker in

cancers, and underline the importance of evaluating p110g-selective agents in the

treatment of cancers with PIK3R1 alterations in a clinical setting.

Potential role of differential p85g expression in various metabolic contexts

The work presented in this dissertation supports the concept that modulation of p85g

levels could provide a mechanism to fine-tune activation of the PI3K/AKT pathway

(Figure 4.1). We and others have shown that reduced p85g expression leads to

enhanced growth factor-stimulated PI3K/AKT activation (Fruman et al., 2000; Mauvais-

Jarvis et al., 2002; Taniguchi et al., 2010; Terauchi et al., 1999; Ueki et al., 2002a), while

several publications have demonstrated that increased p85g expression reduces

PI3K/AKT output (Barbour et al., 2005; Luo et al., 2005a; Ueki et al., 2000). Together

these findings suggest that differential p85g expression could serve as a physiological

mechanism to control the extent of PI3K/AKT signaling.

One potential application for PIγK/AKT pathway regulation by p85g levels might be in

different metabolic contexts. Under conditions of glucose elevation, for example as a

result of ingested nutrients, reduced p85g expression could increase sensitivity to insulin

and reduce blood glucose levels. Conversely, in conditions of reduced glucose, such as

during nutrient deprivation, elevated p85g levels could reduce insulin sensitivity and

Page 165: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

148

favor fatty acid metabolism. Consistent with these ideas, one publication found that in

GEMMs of genetic insulin resistance via heterozygous deletion of IR or IRS1,

heterozygous Pik3r1 ablation improved insulin sensitivity and glucose homeostasis, and

protected mice from the development of diabetes (Mauvais-Jarvis et al., 2002). A second

study demonstrated that Pik3r1 expression was highly induced in adipose tissue from

mice with obesity induced by a high fat diet; heterozygous Pik3r1 ablation preserved

insulin sensitivity in these mice (McCurdy et al., 2012). Thus it will be especially

interesting to explore whether p85g expression levels are dynamic, particularly in those

tissues critical for metabolism, depending on nutrient conditions and disease states.

It will also be important to identify the mechanisms governing p85g expression. Several

recent publications have identified microRNAs (miRNAs) which functionally regulate

Pik3r1 expression (Huang et al., 2014; Huang et al., 2012; Nicoli et al., 2012; Peng et

al., 2013; Tian et al., 2013; Zheng et al., 2012). A separate study reported PIK3R1

transcriptional upregulation in adipocytes by direct binding of peroxisome proliferator-

activated receptor gamma (PPARけ) to two peroxisome proliferator response elements

(PPREs) in the PIK3R1 promoter (Kim et al., 2014). However, the regulation of PIK3R1

expression remains an underexplored area of study. If dynamic changes in p85g levels

in fact contribute to physiological processes such as regulation of metabolism, it will be

critical to identify the different mechanisms which control expression of PIK3R1.

Conclusions and perspective

A tremendous amount of work using RNAi-mediated gene silencing, genetically

engineered mouse models, and emerging isoform-selective pharmacological inhibitors

has begun to elucidate the distinct roles of PI3K catalytic isoforms in different signaling

contexts. Large-scale sequencing efforts in the past decade have also identified

Page 166: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

149

frequently occurring oncogenic mutations in catalytic isoform p110g in a range of human

cancers; catalytic isoforms p110く, p110h, and p110け are rarely mutated but can be

overexpressed. While the roles of PI3K catalytic isoforms in signaling and cancer are

beginning to be understood, the roles of the regulatory isoforms are less well studied.

The work presented in this dissertation adds to a growing number of recent publications

indicating that PIγK regulatory isoform p85g may also contribute to tumorigenesis.

Frequent mutations in PIK3R1 have been identified in certain cancer types; studies have

shown that some of these cancer-associated p85g mutants can still bind but not inhibit

p110, leading to enhanced PI3K/AKT activation. One preclinical study has also indicated

a potential role for p85g as a tumor suppressor in the liver. Here we show that partial

reduction of p85g increases PIγK/AKT signaling in and transformation of human

mammary epithelial cells in vitro and contributes to mammary tumor formation in vivo.

Our findings are consistent with a model in which excess monomeric p85 competes with

p85-p110 heterodimers to negatively regulate PI3K signaling downstream of RTKs. This

work begins to address the role of p85g in different physiological and pathophysiological

PI3K signaling contexts, and highlights the potential for PIK3R1 mutation or

underexpression as a diagnostic and therapeutic marker for breast cancer.

Page 167: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

Materials and Methods

Page 168: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

151

Publically available clinical data

The Oncomine database version 4.4.4.3 (https://www.oncomine.org/) was queried for

expression of PIK3R1 in breast cancer microarray studies. The data from each study

was converted to raw expression levels of PIK3R1 by taking the inverse log2, and then

normalized to the mean raw PIK3R1 expression level of the normal breast tissue from

that specific study.

Vectors

Lentiviral pLKO.1-puromycin shRNA vectors were obtained from the RNAi Consortium at

the Broad Institute, Cambridge, MA. The sequences for these vectors were as follows:

shControl scrambled vector 5’-TCC TAA GGT TAA GTC GCC CTC G-γ’, shPIKγR1 #1

(TRCN0000039903, targeting the γ’ UTR of PIK3R1) 5’-GCG CTA TGC AAT TCT TAA

TTT-γ’, and shPIK3R1 #2 (TRCN0000033284, targeting the coding sequence of

PIK3R1) 5’-CCT TCA GTT CTG TGG TTG AAT-γ’. Retroviral vectors were as follows:

pWZL-blasticidin-p53DD, pBabe-neomycin-p53DD, pBabe-blasticidin-neuT, pWZL-

neomycin-HA-PIK3CAH1047R, pBabe-puromycin-HA-PIK3CAH1047R, and pWZL-neomycin-

Flag-tel-ErbB3. To generate constructs for rescue of PIK3R1 knockdown, a pCMV6

entry vector containing the cDNA ORF of wildtype human PIK3R1 with C-terminal Myc

and Flag tags was obtained from Origene (RC210544) and cloned into the pWZL-

blasticidin retroviral vector. This vector was used for rescue of shPIK3R1 #1. For rescue

of shPIK3R1 #2, the QuikChange II XL site-directed mutagenesis kit (Agilent) was used

to make wobble point mutations T1197C and G1203A with primers 5'-CTC TGA CCC

ATT AAC CTT CAG CTC TGT AGT TGA ATT AAT AAA CCA CTA CC-3' and 5’-GGT

AGT GGT TTA TTA ATT CAA CTA CAG AGC TGA AGG TTA ATG GGT CAG AG-γ’.

Page 169: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

152

Cell culture and transduction

hTERT-immortalized HMECs were cultured in HMEC Growth Medium (DMEM/F12

GlutaMAX [Gibco] supplemented with 0.01たg/ml EGF [Sigma], 10たg/ml insulin [Gibco],

0.0β5たg/ml hydrocortisone [Sigma], 1ng/ml cholera toxin [Sigma], 0.6% FBS [Gibco],

penicillin/streptomycin [Gibco], and antimycotic [Gibco]). 239T and phoenix cells used for

virus production were cultured in DMEM 10% FBS (DMEM supplemented with 5% FBS,

penicillin/streptomycin, and antimycotic, all from Gibco). All cells were maintained at

37°C and 5% CO2.

VSV lentivirus was produced by transfecting 293T cells with the pMD2-VSV-G and

pCMV-〉R8.91 packaging vectors along with a pLKO.1 vector encoding the shRNA of

interest. Retrovirus was produced by transfecting phoenix cells with the retroviral vector

encoding the gene of interest. All transfections were carried out for 20 minutes at room

temperature using the FuGene6 transfection reagent (Promega) in Opti-MEM reduced

serum media (Gibco). Virus was harvested by passing culture supernatants through a

0.45たm filter β-3 days post-transfection.

Stable HMEC lines were generated by infecting cells with lentiviral or retroviral

supernatants in the presence of 4たg/ml polybrene (Milipore) overnight. After infection,

successfully transduced stable polyclonal lines selected for several days in HMEC

Growth Medium containing the appropriate antibiotic (β50たg/ml neomycin [Gibco],

β.5たg/ml blasticidin [Invitrogen], or 1.5たg/ml puromycin [Calbiochem]) until a control

plate of non-transduced cells were completely killed.

Growth factor stimulation timecourse assays

Cells were rinsed twice with PBS (Gibco) and starved in HMEC Starvation Medium

Page 170: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

153

(DMEM/F12 GlutaMAX with penicillin/streptomycin and antimycotic [all from Gibco]) for

either 4 hours or overnight as indicated. For EGF timecourses, human recombinant EGF

(Sigma) was prepared in HMEC Starvation Medium to a final concentration of 20mg/ml,

and used to stimulate cells at 37°C and 5% CO2 for the indicated amounts of time.

Protein lysate preparation and immunoblotting

All protein lysates were prepared by scraping plates of cells on ice using NP40 lysis

buffer (137mM NaCl, 20mM Tris-HCl [pH 8.0], 0.2mM EDTA, 10% glycerol, 1% NP40)

with protease inhibitors (Roche) and phosphatase inhibitors (Thermo Scientific). Whole

cell lysates were prepared by taking the supernatant following centrifugation at

14000rpm and 4°C. Protein concentrations of total cell lysates were determined by

Bradford assay (Bio-Rad) and then 4X SDS/DTT sample buffer (40% glycerol, 250mM

Tris-HCl [pH 6.8], 8% SDS, 0.04% bromophenol blue) was added to a final concentration

of 1X. Samples were boiled at 100°C for 10 minutes and stored at -80°C.

For immunoblotting, proteins in the samples were separated by SDS-PAGE on 8%,

10%, or 1β% polyacrylamide gels and electrotransferred onto 0.45たm NitroBind

nitrocellulose membranes (Maine Manufacturing). Membranes were blocked for 1 hour

at room temperature in a solution of 5% nonfat dry milk (Bio-Rad) in TBS (50mM Tris-

HCl [pH 7.5], 150mM NaCl). Primary antibodies were diluted in a solution of 5% BSA

(Research Products International) in TBST (50mM Tris-HCl [pH 7.5], 150mM NaCl,

0.15% Tween-20) and incubated with membranes at 4°C overnight. The following

primary antibodies were used for immunoblotting at the dilutions specified: vinculin

(Sigma V9131 1:10000), pan-p85 (Millipore Absβγ4 1:600), p85g (Millipore 05β1β

1:1000), p85g (Millipore 0440γ 1:600), p85く (Santa Cruz sc569γ4 1:100), p110g (Cell

Signaling 4β49 1:1000), p110く (Cell Signaling γ011 1:1000), p110く (Santa Cruz sc60β

Page 171: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

154

1:100), Flag (Origene TA50011 1:1000), Flag (Sigma F1804 1:1000), HA (Cell Signaling

2367 1:1000), ErbB2 (Calbiochem OP15 1:100), ErbB3 (Cell Signaling 4754 1:1000),

PTEN (Cell Signaling 9552 1:1000), phospho-AKTT308 (Cell Signaling 4056 1:1000),

phospho-AKTS473 (Cell Signaling 9271 1:1000), total AKT (Cell Signaling 9271 1:1000),

phospho-ERKT202/Y204 (Cell Signaling 9101 1:1000), total ERK (Cell Signaling 9102

1:1000), phospho-S6S235/236 (Cell Signaling 2211 1:1000), total S6 (Cell Signaling 2217

1:1000), phospho-EGFRY1068 (Cell Signaling 3777 1:1000), and EGFR (Cell Signaling

4267 1:1000). Anti-mouse IgG IRDye800-conjugated (Rockland) and anti-rabbit IgG

AlexaFluor680-conjugated (Invitrogen) secondary antibodies were used at a dilution of

1:5000 in a solution of 1.36% nonfat dry milk (Bio-Rad) and 0.01% SDS (Invitrogen) in

TBST, and were incubated with membranes for 1 hour at room temperature. Fluorescent

protein signals were detected and quantified using a LI-COR Odyssey CLx imaging

system and the accompanying ImageStudio software, version 3.1.4.

Anchorage-independent growth assays

Single-cell suspensions were plated in a solution of 0.3% agar in HMEC Growth Medium

on top of a base layer of 0.6% agar in DMEM (Gibco). To prevent agar from drying out,

HMEC Growth Medium with or without PI3K inhibitors was added fresh every 3 days.

For HMECs stably expressing neuT, 2.5x104 viable cells were plated per 60mm dish or

4.5x103 viable cells per well of a 12-well plate, and grown for 3 weeks at 37°C and 5%

CO2; for all other HMEC lines, 5.0x104 viable cells were plated per 60mm dish or 9.0x103

viable cells per well of a 12-well plate, and grown for 4 weeks at 37°C and 5% CO2.

Pictures of unstained colonies were taken at 2X magnification using a Nikon SMZ-U

dissecting microscope with a SPOT Flex 15.2 64Mp Shifting Pixel camera (Diagnostic

Instruments) and the accompanying SPOT advanced software, version 4.5.9.1. Plates

were then stained overnight at 37°C and 5% CO2 with a solution of 0.5mg/ml

Page 172: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

155

iodonitrotetrazolium chloride (Sigma) in HMEC Growth Medium; pictures of stained

plates were taken using visualized using an AlphaImager EP transilluminator (Alpha

Innotech) and the accompanying AlphaView software, version 1.3.0.7.

Proliferation assays

Single-cell suspensions in HMEC Growth Medium were generated, and 1.25x104 viable

cells were plated per well of 24-well plates and allowed to attach overnight. For cells

grown under minimal growth conditions, wells were washed once with PBS (Gibco) and

then given Minimal Growth Medium (HMEC Growth Medium with 0.5% of the normal

growth factor and serum supplements). Fresh HMEC Growth Medium or Minimal Growth

Medium was given every other day. At each time point, wells were washed once with

PBS, cells were fixed for 15 minutes at room temperature with 10% phosphate-buffered

formalin (Fisher), washed twice with ddH2O, stained for 30 minutes with 0.1% crystal

violet (Sigma), washed three times with ddH2O, and allowed to dry completely. Cell-

associated dye was extracted with 1ml of 10% acetic acid per well, and optical density at

590nm was read using a Benchmark Plus microplate spectrophotometer (Bio-Rad) and

the accompanying Microplate Manager III software, version 1.133 for Mac OSX.

Immunoprecipitation assays

For immunoprecipitation of Flag-ErbB3 from HMECs, cells were rinsed twice with PBS

and starved for 4 hours in HMEC Starvation Medium before being lysed in NP40 lysis

buffer. For each reaction, 40たl of anti-Flag beads (Sigma) were washed 3 times with

NP40 lysis buffer, and then mixed with 1mg total protein from the appropriate whole cell

lysate in a total volume of 500たl NP40 lysis buffer. Reactions were carried out on a

rotator at 4°C for 1 hour. Beads were washed γ times with β50たl NP40 lysis buffer and

then resuspended in 2X SDS/DTT sample buffer and boiled at 100°C for 10 minutes.

Page 173: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

156

For immunoprecipitation of endogenous ErbB3 from cultured mouse mammary tumor

cells, cells were rinsed twice with PBS and starved overnight in DMEM containing

penicillin/streptomycin and antimycotic before being lysed in NP40 lysis buffer. For each

reaction, 500たg total protein from the appropriate whole cell lysate was mixed with 4たg

anti-ErbB3 antibody (Millipore 05-390) and incubated on a rotator at 4°C for 1 hour.

Then, 40たl of Protein A/G beads (Santa Cruz) were added and reactions were carried

out on a rotator at 4°C for an additional 1 hour. The beads were washed 3 times with

β50たl NP40 lysis buffer and then resuspended in βX SDS/DTT sample buffer and boiled

at 100°C for 10 minutes.

For immunoprecipitation of phospho-tyrosine-containing proteins, HMECs were rinsed

twice with PBS and given HMEC Starvation Medium for 4 hours before being lysed in

NP40 lysis buffer. Each reaction consisted of 10たl 4G10 anti-phospho-tyrosine

sepharose bead slurry (Millipore) and 100たg total protein from the appropriate whole cell

lysate in a final volume of β50たl. Reactions were carried out on a rotator overnight at

4°C, and then the beads were washed γ times with β50たl NP40 lysis buffer and then

resuspended in 2X SDS/DTT sample buffer and boiled at 100°C for 10 minutes.

PTEN lipid phosphatase activity assay

HMECs were grown to confluency in 15cm tissue culture dishes. Asynchronous cells

were lysed in NP40 lysis buffer. Endogenous PTEN was immunoprecipitated from cell

lysates in triplicate reactions. For “Beads Only” reaction, 2mg total protein from

shControl whole cell lysate was diluted to a final volume of 400たl in NP40 lysis buffer.

For “PTEN Only” reaction, 2mg total protein from shControl whole cell lysate was mixed

with 8たl anti-PTEN antibody (Cell Signaling 9559) and diluted to a final volume of 400たl

in NP40 lysis buffer. For all other reactions, 2mg total protein from the appropriate whole

Page 174: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

157

cell lysate was mixed with 8たl anti-PTEN antibody (Cell Signaling 9559) and diluted to a

final volume of 400たl in NP40 lysis buffer. All tubes were incubated on a rotator

overnight at 4°C. Then, 40たl of Protein A/G beads (Santa Cruz) were added to each

tube, and reactions were incubated on a rotator for 3 hours at 4°C. The beads were

washed 3 times with 400たl NP40 lysis buffer and then washed once with PTEN Reaction

Buffer (TBS [25 mM Tris-HCl (pH 7.5), 140mM NaCl, 2.7mM KCl] with 10mM DTT added

fresh). The beads were resuspended in γ0たl PTEN Reaction Buffer and used in in vitro

PTEN lipid phosphatase assays (Echelon Biosciences, method adapted from provided

documents and (Song et al., 2011)). Briefly, phosphate standards from 0-2,000pmol

were prepared using provided reagents. Then βたl of 1mM diC8PtdIns(3,4,5)P3 was

added to each of the standards, a “PIPγ Substrate Only” control (consisting of PTEN

Reaction Buffer only), and the immunoprecipitation reactions and accompanying

controls, and brought to a final volume of 50たl with PTEN Reaction Buffer. Reactions

were incubated at γ7°C for γ0 minutes, and then 100たl Malachite Green Reagent was

added to each reaction and incubated for 10 minutes at room temperature. The optical

density at 620nm, corresponding to phosphate released by PTEN enzymatic activity,

was read using a Benchmark Plus microplate spectrophotometer (Bio-Rad) and the

accompanying Microplate Manager III software, version 1.133 for Mac OSX.

Receptor internalization and degradation assays

To track internalization of EGFR, HMECs were rinsed with PBS and starved overnight in

HMEC Starvation Medium. Plates were washed with PBS-CM (PBS with 1mM MgCl2

and 0.1mM CaCl2 added), then incubated on a rocker for 40 minutes at 4°C with 5ml

PBS-CM containing 0.5mg/ml Sulfo-NHS-SS-Biotin (Pierce) to label surface proteins.

Plates were washed with PBS-CM, then incubated on a rocker for 10 minutes at 4°C

with 10ml PBS-CM containing 50mM NH4Cl (Sigma) to inactivate excess unbound sulfo-

Page 175: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

158

NHS-SS-biotin. Plates were washed with PBS-CM. One plate per cell line was then

lysed with TritonX100 Lysis Buffer (200mM NaCl, 75mM Tris-HCl [pH 7.5], 15mM NaF,

2.5mM EDTA, 2.5mM EGTA, 1.5% TritonX100, 0.75% NP40, 0.1% SDS) for

determination of total surface EGFR. The remaining plates were stimulated with

20mg/ml human recombinant EGF (Sigma) in HMEC Starvation Medium at 37°C and 5%

CO2 for the indicated amounts of time to induce receptor internalization. At the

appropriate time, plates were rinsed with PBS-CM, washed twice with 10ml Glutathione

Buffer (90mM NaCl, 1mM MgCl2, 0.1mM CaCl2, 50mM reduced glutathione, 60mM

NaOH) to cleave any sulfo-NHS-SS-biotin remaining on the cell surface, and rinsed with

10ml PBS-IAA (50mM iodoacetamide in PBS-CM) on a rocker at 4°C for 15 minutes, to

quench excess glutathione. After rinsing with PBS-CM, plates were lysed with

TritonX100 lysis buffer and protein samples were prepared as described above. A

portion of these samples was reserved for whole cell lysates, and the rest was used in

Streptavidin immunoprecipitation. First, 1mg total protein per sample was pre-cleared by

incubating with pansorbin (Calbiochem) on a rotator at 4°C for 1 hour. The pansorbin

was removed by centrifugation, and 40ul Streptavidin beads (Pierce) were added for

overnight immunoprecipitation reactions. Beads were rinsed 4 times with TritonX100

lysis buffer before being boiled in 2X SDS/DTT sample buffer for 10 minutes. Samples

were then subjected to electrophoresis and immunoblotting as described above.

To track degradation of EGFR, HMECs were rinsed with PBS and starved overnight in

HMEC Starvation Medium. Plates were washed with PBS-CM, then incubated on a

rocker for 40 minutes at 4°C with 5ml PBS-CM containing 0.5mg/ml Sulfo-NHS-LC-LC-

Biotin (Pierce) to label surface proteins. Plates were washed with PBS-CM, then

incubated on a rocker for 10 minutes at 4°C with 10ml PBS-CM containing 50mM NH4Cl

(Sigma) to inactivate excess unbound sulfo-NHS-LC-LC-biotin. Plates were washed with

Page 176: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

159

PBS-CM. One plate per cell line was then lysed with TritonX100 Lysis Buffer for

determination of total surface EGFR. The remaining plates were stimulated with

20mg/ml human recombinant EGF (Sigma) in HMEC Starvation Medium at 37°C and 5%

CO2 for the indicated amounts of time to induce receptor internalization. At the

appropriate time, plates were rinsed with PBS-CM and lysed with TritonX100 lysis buffer,

and protein samples were prepared as described above. A portion of these samples was

reserved for whole cell lysates, and the rest was used in Streptavidin

immunoprecipitation. First, 1mg total protein per sample was pre-cleared by incubating

with pansorbin (Calbiochem) on a rotator at 4°C for 1 hour. The pansorbin was removed

by centrifugation, and 40ul Streptavidin beads (Pierce) were added for overnight

immunoprecipitation reactions. Beads were rinsed 4 times with TritonX100 lysis buffer

before being boiled in 2X SDS/DTT sample buffer for 10 minutes. Samples were then

subjected to electrophoresis and immunoblotting as described above.

Animal husbandry and breeding strategy

MMTV-Cre (Wagner et al., 1997), NIC (Schade et al., 2009; Ursini-Siegel et al., 2008),

and Pik3r1 floxed (Luo et al., 2005b) mice were backcrossed to the FVB/N wildtype

background at least 10 generations. To generate the female mice used in these studies,

male MMTV-Cre or NIC mice heterozygous for the floxed Pik3r1 allele were crossed with

female mice heterozygous for the floxed Pik3r1 allele. For orthotopic tumor

transplantations, eight-week-old NCrNu female mice (Harlan) were used. Transplant

recipient mice were treated daily with vehicle control (0.5% [w/V] methylcellulose

[Sigma], administered by oral gavage at 1ml/kg body weight), BYL719 (in 0.5%

methylcellulose, administered by oral gavage at 45mg/kg), GDC0941 (in 0.5%

methylcellulose, administered by oral gavage at 125mg/kg), or KIN193 (in a solution of

7.5% NMP [1-methyl-2-pyrrolidinone, Sigma] and 40% PEG400 [Sigma], administered

Page 177: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

160

by intraperitoneal injection at 20mg/kg). All animals were housed and treated in

accordance with protocols approved by the Institutional Animal Care and Use

Committees of Dana-Farber Cancer Institute and Harvard Medical School.

Genotyping

Genomic DNA was extracted from 2-3mm of tail tissue by boiling in 150ul of an alkaline

buffer containing 2mM EDTA and 25mM NaOH at 100C for 60 minutes, followed by

neutralization with 150ul of a buffer containing 40mM Tris Base. PCR was then

performed on the DNA extracts using GoTaq DNA polymerase (Promega) as follows for

each gene: Pik3r1, primers 5'-CAC CGA GCA CTG GAG CAC TG-3' and 5'-CCA GTT

ACT TTC AAA TCA GCA CAG-3', wildtype Pik3r1 allele generates a fragment of 252bp,

while floxed Pik3r1 allele generates a 301bp fragment; NIC transgene, primers 5'-TTC

CGG AAC CCA CAT CAG GCC-3' and 5'-GTT TCC TGC AGC AGC CTA CGC-3',

transgene generates a 630bp fragment; MMTV-Cre transgene, primers 5’-CTG ATC

TGA GCT CTG AGT G-γ’, 5’-CAT CAC TCG TTG CAT CGA CC-γ’, transgene

generates a 250bp fragment. PCR samples were then resolved by electrophoresis

through a 2% agarose gel with SYBR Safe (Invitrogen) and visualized under UV light

using an AlphaImager EP transilluminator (Alpha Innotech) and the accompanying

AlphaView software, version 1.3.0.7.

Mammary whole mount preparation

The fourth mammary gland tissue was excised, spread on glass slides, and fixed

overnight in Carnoy’s Fixative (a 1:γ [V/V] solution of glacial acetic acid and 100%

ethanol). Slides were then washed for 30 minutes in 70% ethanol and 30 minutes in

ddH2O, and then stained overnight in Carmine Alum (1g carmine [Sigma] dissolved in

500ml ddH2O). Slides were then washed for 30 minutes in 70% ethanol, 30 minutes in

Page 178: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

161

95% ethanol, and 30 minutes 100% ethanol before being immersed in toluene. All steps

were carried out at room temperature. Images of fixed mammary tissue were taken with

a Nikon SMZ-U dissecting microscope with a SPOT Flex 15.2 64Mp Shifting Pixel

camera (Diagnostic Instruments) and the accompanying SPOT advanced software,

version 4.5.9.1.

Mouse mammary epithelial cell isolation

For each mouse, the third and fourth mammary glands were excised, washed in PBS

(Gibco), finely chopped, combined, and digested overnight at 37°C and 5% CO2 in 5ml

Digestion Medium (DMEM/F12 GlutaMAX [Gibco] with penicillin/streptomycin [Gibco]

and collagenase/hyaluronidase [Stem Cell Technologies] added to a final concentration

of 1X). Following digestion, the tissue was pelleted by centrifugation, the supernatant

removed, and the tissue resuspended in Red Blood Cell Lysis Buffer (a 1:4 [V/V] solution

of cold HF Solution [Hank’s Balanced Salt Solution (Stem Cell Technologies) with β%

FBS (Gibco) and penicillin/streptomycin (Gibco)] and ammonium chloride [Stem Cell

Technologies]). The tissue was then pelleted by centrifugation, the supernatant

removed, and the pellet resuspended in 2ml pre-warmed 0.25% Trypsin-EDTA (Gibco)

for 1-3 minutes, followed by addition of 3ml FBS and 10ml cold HF Solution. The tissue

was then pelleted by centrifugation, the supernatant removed, the pellet resuspended in

1ml Dispase and 100たl DNAseI (both Stem Cell Technologies) for 1 minute followed by

addition of 5-10ml cold HF Solution, and then passed through a 40たm cell strainer.

Analysis of mammary tumors and lung metastases

Cohorts of female mice were examined every 3 days for the onset of tumors, defined by

the first palpation. Five weeks after tumor onset, the mice were sacrificed by CO2 in

accordance with protocols approved by the Institutional Animal Care and Use

Page 179: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

162

Committees of Dana-Farber Cancer Institute and Harvard Medical School. The total

tumor burden per mouse was determined by taking the wet weight of excised mammary

tissue and associated tumors. The total number of tumors per mouse was determined by

counting the number of distinct excised tumors with a diameter of at least 3mm. To

determine the total number of lung metastases per mouse, lungs were excised and fixed

overnight in 10% phosphate-buffered formalin (Fisher). All lung lobes were embedded in

paraffin, and three sections 50たm apart were mounted on slides and stained with

hematoxylin and eosin. Metastases were visualized and quantified using a Nikon Eclipse

E600 microscope. For each mouse, the total number of lung metastases was taken to be

that of the section with the highest count.

Histology and immunohistochemistry

Tissue was fixed overnight in 10% phosphate-buffered formalin (Fisher) and then

transferred to 70% ethanol. Formalin-fixed tissue was embedded in paraffin blocks and

mounted on slides by the HMS Rodent Histopathology Core. Hematoxylin and eosin

staining was performed by the HMS Rodent Histopathology Core. Immunohistochemical

staining of tissue sections was performed using a sodium citrate antigen retrieval

method. Slides were deparaffinized in xylene and hydrated in ethanol washes of

decreasing concentration. Antigen retrieval was then performed by boiling slides in

10mM sodium citrate (pH 6.0) (Boston BioProducts) for 20 minutes, followed by 30

minutes of cooling. Endogenous peroxidases were blocked by soaking the slides in 3%

hydrogen peroxide (Sigma) for 10 minutes at room temperature, and then slides were

further blocked by a 1 hour incubation with 5% serum in IHC-TBST (50mM Tris-HCl [pH

7.5], 150mM NaCl, 0.1% Tween-20) at room temperature. Slides were then incubated

with primary antibodies overnight. Antibodies from Cell Signaling were diluted in the

buffer provided by Cell Signaling, and all other antibodies were diluted in 5% serum in

Page 180: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

163

IHC-TBST. The following primary antibodies were used: p85g (EMD Millipore 04403

1:500), phospho-AKTS473 (Cell Signaling 3787 1:50), phospho-ERKT202/Y204 (Cell

Signaling 4376 1:400), phospho-S6S235/236 (Cell Signaling 4858 1:400), and Ki67 (Vector

Laboratories VPK451 1:1000). Slides were then incubated for 30 minutes at room

temperature with biotinylated goat anti-rabbit IgG secondary antibody (Vector

Laboratories BA1000 1:250-1:2000) followed by a 30-minute incubation at room

temperature with ABC reagent (Vector Laboratories). Antibody signal was then

developed by brief incubation with DAB horseradish peroxidase substrate (Vector

Laboratories) and slides were counterstained with hematoxylin (Vector Laboratories).

For TUNEL, the ApopTag Plus Peroxidase In Situ Apoptosis Detection Kit was used

(EMD Millipore) followed by a counterstain with methyl green (Vector Laboratories).

Tissue sections were then dehydrated by washes in increasing concentrations of ethanol

followed by washes in xylene. Coverslips were mounted using Cytoseal XYL (Thermo

Scientific) and slides were visualized using a Nikon Eclipse E600 microscope, and

images were obtained with a SPOT Flex 15.2 64Mp Shifting Pixel camera (Diagnostic

Instruments) and the accompanying SPOT advanced software, version 4.5.9.1.

Preparation of mouse tissue protein lysates

Tissue pieces were snap frozen in liquid nitrogen. To generate protein lysates, tissue

was homogenized in NP40 lysis buffer using 0.5mm zirconium oxide beads (Next

Advance Inc.) in an air-cooling bullet blender (Next Advance). Once homogenized,

lysates were cleared and analyzed as described above.

Isolation and culture of mouse mammary tumor cells

Mouse mammary tumor cells were cultured in DMEM 5% FBS (DMEM supplemented

with 5% fetal bovine serum, penicillin/streptomycin, and antimycotic, all from Gibco) at

Page 181: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

164

37°C and 5% CO2. To isolate cells for culture, tumor pieces of approximately 0.5cm3

were finely chopped, resuspended in 6-8ml DMEM 5% FBS containing 1mg/ml

collagenase (Roche), and incubated on a rotator for 20 minutes at 37°C. Cells were then

pelleted by a 5 minute centrifugation at 1200rpm, washed with PBS, pelleted again,

resuspended in 10ml DMEM 5% FBS, and plated in 10cm tissue culture dishes.

Graphing software and statistical analysis

All graphs and statistical analysis were done using GraphPad Prism 6.0 for Mac OS X.

Page 182: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

References

Page 183: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

166

Ali, K., Bilancio, A., Thomas, M., Pearce, W., Gilfillan, A.M., Tkaczyk, C., Kuehn, N., Gray, A., Giddings, J., Peskett, E., et al. (2004). Essential role for the p110delta phosphoinositide 3-kinase in the allergic response. Nature 431, 1007-1011.

Ali, K., Soond, D.R., Pineiro, R., Hagemann, T., Pearce, W., Lim, E.L., Bouabe, H., Scudamore, C.L., Hancox, T., Maecker, H., et al. (2014). Inactivation of PI(3)K p110delta breaks regulatory T-cell-mediated immune tolerance to cancer. Nature.

Alimonti, A., Carracedo, A., Clohessy, J.G., Trotman, L.C., Nardella, C., Egia, A., Salmena, L., Sampieri, K., Haveman, W.J., Brogi, E., et al. (2010). Subtle variations in Pten dose determine cancer susceptibility. Nature genetics 42, 454-458.

Andrechek, E.R., Hardy, W.R., Siegel, P.M., Rudnicki, M.A., Cardiff, R.D., and Muller, W.J. (2000). Amplification of the neu/erbB-2 oncogene in a mouse model of mammary tumorigenesis. Proceedings of the National Academy of Sciences of the United States of America 97, 3444-3449.

Angulo, I., Vadas, O., Garcon, F., Banham-Hall, E., Plagnol, V., Leahy, T.R., Baxendale, H., Coulter, T., Curtis, J., Wu, C., et al. (2013). Phosphoinositide 3-kinase delta gene mutation predisposes to respiratory infection and airway damage. Science 342, 866-871.

Backer, J.M. (2008). The regulation and function of Class III PI3Ks: novel roles for Vps34. The Biochemical journal 410, 1-17.

Ballou, L.M., Chattopadhyay, M., Li, Y., Scarlata, S., and Lin, R.Z. (2006). Galphaq binds to p110alpha/p85alpha phosphoinositide 3-kinase and displaces Ras. The Biochemical journal 394, 557-562.

Ballou, L.M., Lin, H.Y., Fan, G., Jiang, Y.P., and Lin, R.Z. (2003). Activated G alpha q inhibits p110 alpha phosphatidylinositol 3-kinase and Akt. The Journal of biological chemistry 278, 23472-23479.

Barbour, L.A., Mizanoor Rahman, S., Gurevich, I., Leitner, J.W., Fischer, S.J., Roper, M.D., Knotts, T.A., Vo, Y., McCurdy, C.E., Yakar, S., et al. (2005). Increased P85alpha is a potent negative regulator of skeletal muscle insulin signaling and induces in vivo insulin resistance associated with growth hormone excess. The Journal of biological chemistry 280, 37489-37494.

Baselga, J., and Swain, S.M. (2009). Novel anticancer targets: revisiting ERBB2 and discovering ERBB3. Nature reviews. Cancer 9, 463-475.

Beeton, C.A., Chance, E.M., Foukas, L.C., and Shepherd, P.R. (2000). Comparison of the kinetic properties of the lipid- and protein-kinase activities of the p110alpha and p110beta catalytic subunits of class-Ia phosphoinositide 3-kinases. The Biochemical journal 350 Pt 2, 353-359.

Berenjeno, I.M., Guillermet-Guibert, J., Pearce, W., Gray, A., Fleming, S., and Vanhaesebroeck, B. (β01β). Both p110g and p110く isoforms of PIγK can modulate the impact of loss-of-function of the PTEN tumour suppressor. The Biochemical journal 442, 151-159.

Page 184: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

167

Berger, M.B., Mendrola, J.M., and Lemmon, M.A. (2004). ErbB3/HER3 does not homodimerize upon neuregulin binding at the cell surface. FEBS letters 569, 332-336.

Bi, L., Okabe, I., Bernard, D.J., and Nussbaum, R.L. (2002). Early embryonic lethality in mice deficient in the p110beta catalytic subunit of PI 3-kinase. Mammalian genome : official journal of the International Mammalian Genome Society 13, 169-172.

Bi, L., Okabe, I., Bernard, D.J., Wynshaw-Boris, A., and Nussbaum, R.L. (1999). Proliferative defect and embryonic lethality in mice homozygous for a deletion in the p110alpha subunit of phosphoinositide 3-kinase. The Journal of biological chemistry 274, 10963-10968.

Biswas, K., Yoshioka, K., Asanuma, K., Okamoto, Y., Takuwa, N., Sasaki, T., and Takuwa, Y. (2013). Essential role of class II phosphatidylinositol-3-kinase-C2alpha in sphingosine 1-phosphate receptor-1-mediated signaling and migration in endothelial cells. The Journal of biological chemistry 288, 2325-2339.

Blondeau, F., Laporte, J., Bodin, S., Superti-Furga, G., Payrastre, B., and Mandel, J.L. (2000). Myotubularin, a phosphatase deficient in myotubular myopathy, acts on phosphatidylinositol 3-kinase and phosphatidylinositol 3-phosphate pathway. Human molecular genetics 9, 2223-2229.

Bohnacker, T., Marone, R., Collmann, E., Calvez, R., Hirsch, E., and Wymann, M.P. (2009). PI3Kgamma adaptor subunits define coupling to degranulation and cell motility by distinct PtdIns(3,4,5)P3 pools in mast cells. Science signaling 2, ra27.

Bollag, G., Hirth, P., Tsai, J., Zhang, J., Ibrahim, P.N., Cho, H., Spevak, W., Zhang, C., Zhang, Y., Habets, G., et al. (2010). Clinical efficacy of a RAF inhibitor needs broad target blockade in BRAF-mutant melanoma. Nature 467, 596-599.

Britschgi, A., Andraos, R., Brinkhaus, H., Klebba, I., Romanet, V., Muller, U., Murakami, M., Radimerski, T., and Bentires-Alj, M. (2012). JAK2/STAT5 inhibition circumvents resistance to PI3K/mTOR blockade: a rationale for cotargeting these pathways in metastatic breast cancer. Cancer cell 22, 796-811.

Brock, C., Schaefer, M., Reusch, H.P., Czupalla, C., Michalke, M., Spicher, K., Schultz, G., and Nurnberg, B. (2003). Roles of G beta gamma in membrane recruitment and activation of p110 gamma/p101 phosphoinositide 3-kinase gamma. The Journal of cell biology 160, 89-99.

Burke, J.E., Perisic, O., Masson, G.R., Vadas, O., and Williams, R.L. (2012). Oncogenic mutations mimic and enhance dynamic events in the natural activation of phosphoinositide 3-kinase p110alpha (PIK3CA). Proceedings of the National Academy of Sciences of the United States of America 109, 15259-15264.

Busaidy, N.L., Farooki, A., Dowlati, A., Perentesis, J.P., Dancey, J.E., Doyle, L.A., Brell, J.M., and Siu, L.L. (2012). Management of metabolic effects associated with anticancer agents targeting the PI3K-Akt-mTOR pathway. Journal of clinical oncology : official journal of the American Society of Clinical Oncology 30, 2919-2928.

Page 185: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

168

Cancer Genome Atlas, N. (2012). Comprehensive molecular portraits of human breast tumours. Nature 490, 61-70.

Cancer Genome Atlas Research, N. (2008). Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 455, 1061-1068.

Cao, C., Backer, J.M., Laporte, J., Bedrick, E.J., and Wandinger-Ness, A. (2008). Sequential actions of myotubularin lipid phosphatases regulate endosomal PI(3)P and growth factor receptor trafficking. Molecular biology of the cell 19, 3334-3346.

Carroll, M., Tomasson, M.H., Barker, G.F., Golub, T.R., and Gilliland, D.G. (1996). The TEL/platelet-derived growth factor beta receptor (PDGF beta R) fusion in chronic myelomonocytic leukemia is a transforming protein that self-associates and activates PDGF beta R kinase-dependent signaling pathways. Proceedings of the National Academy of Sciences of the United States of America 93, 14845-14850.

Castellano, E., Sheridan, C., Thin, M.Z., Nye, E., Spencer-Dene, B., Diefenbacher, M.E., Moore, C., Kumar, M.S., Murillo, M.M., Gronroos, E., et al. (2013). Requirement for interaction of PI3-kinase p110alpha with RAS in lung tumor maintenance. Cancer cell 24, 617-630.

Cerami, E., Gao, J., Dogrusoz, U., Gross, B.E., Sumer, S.O., Aksoy, B.A., Jacobsen, A., Byrne, C.J., Heuer, M.L., Larsson, E., et al. (2012). The cBio cancer genomics portal: an open platform for exploring multidimensional cancer genomics data. Cancer discovery 2, 401-404.

Chagpar, R.B., Links, P.H., Pastor, M.C., Furber, L.A., Hawrysh, A.D., Chamberlain, M.D., and Anderson, D.H. (2010). Direct positive regulation of PTEN by the p85 subunit of phosphatidylinositol 3-kinase. Proceedings of the National Academy of Sciences of the United States of America 107, 5471-5476.

Chamberlain, M.D., Berry, T.R., Pastor, M.C., and Anderson, D.H. (2004). The p85alpha subunit of phosphatidylinositol 3'-kinase binds to and stimulates the GTPase activity of Rab proteins. The Journal of biological chemistry 279, 48607-48614.

Chamberlain, M.D., Chan, T., Oberg, J.C., Hawrysh, A.D., James, K.M., Saxena, A., Xiang, J., and Anderson, D.H. (2008). Disrupted RabGAP function of the p85 subunit of phosphatidylinositol 3-kinase results in cell transformation. The Journal of biological chemistry 283, 15861-15868.

Chamberlain, M.D., Oberg, J.C., Furber, L.A., Poland, S.F., Hawrysh, A.D., Knafelc, S.M., McBride, H.M., and Anderson, D.H. (2010). Deregulation of Rab5 and Rab4 proteins in p85R274A-expressing cells alters PDGFR trafficking. Cellular signalling 22, 1562-1575.

Chandarlapaty, S., Sawai, A., Scaltriti, M., Rodrik-Outmezguine, V., Grbovic-Huezo, O., Serra, V., Majumder, P.K., Baselga, J., and Rosen, N. (2011). AKT inhibition relieves feedback suppression of receptor tyrosine kinase expression and activity. Cancer cell 19, 58-71.

Page 186: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

169

Chang, H.W., Aoki, M., Fruman, D., Auger, K.R., Bellacosa, A., Tsichlis, P.N., Cantley, L.C., Roberts, T.M., and Vogt, P.K. (1997). Transformation of chicken cells by the gene encoding the catalytic subunit of PI 3-kinase. Science 276, 1848-1850.

Chapman, P.B., Hauschild, A., Robert, C., Haanen, J.B., Ascierto, P., Larkin, J., Dummer, R., Garbe, C., Testori, A., Maio, M., et al. (2011). Improved survival with vemurafenib in melanoma with BRAF V600E mutation. The New England journal of medicine 364, 2507-2516.

Chaussade, C., Rewcastle, G.W., Kendall, J.D., Denny, W.A., Cho, K., Gronning, L.M., Chong, M.L., Anagnostou, S.H., Jackson, S.P., Daniele, N., et al. (2007). Evidence for functional redundancy of class IA PI3K isoforms in insulin signalling. The Biochemical journal 404, 449-458.

Chen, D., Mauvais-Jarvis, F., Bluher, M., Fisher, S.J., Jozsi, A., Goodyear, L.J., Ueki, K., and Kahn, C.R. (2004). p50alpha/p55alpha phosphoinositide 3-kinase knockout mice exhibit enhanced insulin sensitivity. Molecular and cellular biology 24, 320-329.

Cheung, L.W., Hennessy, B.T., Li, J., Yu, S., Myers, A.P., Djordjevic, B., Lu, Y., Stemke-Hale, K., Dyer, M.D., Zhang, F., et al. (2011). High frequency of PIK3R1 and PIK3R2 mutations in endometrial cancer elucidates a novel mechanism for regulation of PTEN protein stability. Cancer discovery 1, 170-185.

Chiu, Y.H., Lee, J.Y., and Cantley, L.C. (2014). BRD7, a tumor suppressor, interacts with p85alpha and regulates PI3K activity. Molecular cell 54, 193-202.

Ciraolo, E., Iezzi, M., Marone, R., Marengo, S., Curcio, C., Costa, C., Azzolino, O., Gonella, C., Rubinetto, C., Wu, H., et al. (2008). Phosphoinositide 3-kinase p110beta activity: key role in metabolism and mammary gland cancer but not development. Science signaling 1, ra3.

Ciraolo, E., Morello, F., Hobbs, R.M., Wolf, F., Marone, R., Iezzi, M., Lu, X., Mengozzi, G., Altruda, F., Sorba, G., et al. (2010). Essential role of the p110beta subunit of phosphoinositide 3-OH kinase in male fertility. Molecular biology of the cell 21, 704-711.

Cizkova, M., Vacher, S., Meseure, D., Trassard, M., Susini, A., Mlcuchova, D., Callens, C., Rouleau, E., Spyratos, F., Lidereau, R., et al. (2013). PIK3R1 underexpression is an independent prognostic marker in breast cancer. BMC cancer 13, 545.

Clayton, E., Bardi, G., Bell, S.E., Chantry, D., Downes, C.P., Gray, A., Humphries, L.A., Rawlings, D., Reynolds, H., Vigorito, E., et al. (2002). A crucial role for the p110delta subunit of phosphatidylinositol 3-kinase in B cell development and activation. The Journal of experimental medicine 196, 753-763.

Cortes, I., Sanchez-Ruiz, J., Zuluaga, S., Calvanese, V., Marques, M., Hernandez, C., Rivera, T., Kremer, L., Gonzalez-Garcia, A., and Carrera, A.C. (2012). p85beta phosphoinositide 3-kinase subunit regulates tumor progression. Proceedings of the National Academy of Sciences of the United States of America 109, 11318-11323.

Page 187: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

170

Cully, M., You, H., Levine, A.J., and Mak, T.W. (2006). Beyond PTEN mutations: the PI3K pathway as an integrator of multiple inputs during tumorigenesis. Nature reviews. Cancer 6, 184-192.

Davie, S.A., Maglione, J.E., Manner, C.K., Young, D., Cardiff, R.D., MacLeod, C.L., and Ellies, L.G. (2007). Effects of FVB/NJ and C57Bl/6J strain backgrounds on mammary tumor phenotype in inducible nitric oxide synthase deficient mice. Transgenic research 16, 193-201.

Dbouk, H.A., Khalil, B.D., Wu, H., Shymanets, A., Nurnberg, B., and Backer, J.M. (2013). Characterization of a tumor-associated activating mutation of the p110beta PI 3-kinase. PloS one 8, e63833.

Dbouk, H.A., Pang, H., Fiser, A., and Backer, J.M. (2010). A biochemical mechanism for the oncogenic potential of the p110beta catalytic subunit of phosphoinositide 3-kinase. Proceedings of the National Academy of Sciences of the United States of America 107, 19897-19902.

Dbouk, H.A., Vadas, O., Shymanets, A., Burke, J.E., Salamon, R.S., Khalil, B.D., Barrett, M.O., Waldo, G.L., Surve, C., Hsueh, C., et al. (2012). G protein-coupled receptor-mediated activation of p110beta by Gbetagamma is required for cellular transformation and invasiveness. Science signaling 5, ra89.

Delgado, P., Cubelos, B., Calleja, E., Martinez-Martin, N., Cipres, A., Merida, I., Bellas, C., Bustelo, X.R., and Alarcon, B. (2009). Essential function for the GTPase TC21 in homeostatic antigen receptor signaling. Nature immunology 10, 880-888.

Delmore, J.E., Issa, G.C., Lemieux, M.E., Rahl, P.B., Shi, J., Jacobs, H.M., Kastritis, E., Gilpatrick, T., Paranal, R.M., Qi, J., et al. (2011). BET bromodomain inhibition as a therapeutic strategy to target c-Myc. Cell 146, 904-917.

Denley, A., Gymnopoulos, M., Kang, S., Mitchell, C., and Vogt, P.K. (2009). Requirement of phosphatidylinositol(3,4,5)trisphosphate in phosphatidylinositol 3-kinase-induced oncogenic transformation. Molecular cancer research : MCR 7, 1132-1138.

Denley, A., Kang, S., Karst, U., and Vogt, P.K. (2008). Oncogenic signaling of class I PI3K isoforms. Oncogene 27, 2561-2574.

Di Cristofano, A., Pesce, B., Cordon-Cardo, C., and Pandolfi, P.P. (1998). Pten is essential for embryonic development and tumour suppression. Nature genetics 19, 348-355.

Diouf, B., Cheng, Q., Krynetskaia, N.F., Yang, W., Cheok, M., Pei, D., Fan, Y., Cheng, C., Krynetskiy, E.Y., Geng, H., et al. (2011). Somatic deletions of genes regulating MSH2 protein stability cause DNA mismatch repair deficiency and drug resistance in human leukemia cells. Nature medicine 17, 1298-1303.

Dou, Z., Pan, J.A., Dbouk, H.A., Ballou, L.M., DeLeon, J.L., Fan, Y., Chen, J.S., Liang, Z., Li, G., Backer, J.M., et al. (2013). Class IA PI3K p110beta subunit promotes autophagy through Rab5 small GTPase in response to growth factor limitation. Molecular cell 50, 29-42.

Page 188: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

171

Dourdin, N., Schade, B., Lesurf, R., Hallett, M., Munn, R.J., Cardiff, R.D., and Muller, W.J. (2008). Phosphatase and tensin homologue deleted on chromosome 10 deficiency accelerates tumor induction in a mouse model of ErbB-2 mammary tumorigenesis. Cancer research 68, 2122-2131.

Downward, J., Parker, P., and Waterfield, M.D. (1984). Autophosphorylation sites on the epidermal growth factor receptor. Nature 311, 483-485.

Durand, C.A., Hartvigsen, K., Fogelstrand, L., Kim, S., Iritani, S., Vanhaesebroeck, B., Witztum, J.L., Puri, K.D., and Gold, M.R. (2009). Phosphoinositide 3-kinase p110 delta regulates natural antibody production, marginal zone and B-1 B cell function, and autoantibody responses. Journal of immunology 183, 5673-5684.

Elis, W., Triantafellow, E., Wolters, N.M., Sian, K.R., Caponigro, G., Borawski, J., Gaither, L.A., Murphy, L.O., Finan, P.M., and Mackeigan, J.P. (2008). Down-regulation of class II phosphoinositide 3-kinase alpha expression below a critical threshold induces apoptotic cell death. Molecular cancer research : MCR 6, 614-623.

Elkabets, M., Vora, S., Juric, D., Morse, N., Mino-Kenudson, M., Muranen, T., Tao, J., Campos, A.B., Rodon, J., Ibrahim, Y.H., et al. (2013). mTORC1 inhibition is required for sensitivity to PI3K p110alpha inhibitors in PIK3CA-mutant breast cancer. Science translational medicine 5, 196ra199.

Engelman, J.A. (2009). Targeting PI3K signalling in cancer: opportunities, challenges and limitations. Nature reviews. Cancer 9, 550-562.

Engelman, J.A., Chen, L., Tan, X., Crosby, K., Guimaraes, A.R., Upadhyay, R., Maira, M., McNamara, K., Perera, S.A., Song, Y., et al. (2008). Effective use of PI3K and MEK inhibitors to treat mutant Kras G12D and PIK3CA H1047R murine lung cancers. Nature medicine 14, 1351-1356.

Engelman, J.A., Luo, J., and Cantley, L.C. (2006). The evolution of phosphatidylinositol 3-kinases as regulators of growth and metabolism. Nature reviews. Genetics 7, 606-619.

Falasca, M., Hughes, W.E., Dominguez, V., Sala, G., Fostira, F., Fang, M.Q., Cazzolli, R., Shepherd, P.R., James, D.E., and Maffucci, T. (2007). The role of phosphoinositide 3-kinase C2alpha in insulin signaling. The Journal of biological chemistry 282, 28226-28236.

Falasca, M., and Maffucci, T. (2012). Regulation and cellular functions of class II phosphoinositide 3-kinases. The Biochemical journal 443, 587-601.

Fedele, C.G., Ooms, L.M., Ho, M., Vieusseux, J., O'Toole, S.A., Millar, E.K., Lopez-Knowles, E., Sriratana, A., Gurung, R., Baglietto, L., et al. (2010). Inositol polyphosphate 4-phosphatase II regulates PI3K/Akt signaling and is lost in human basal-like breast cancers. Proceedings of the National Academy of Sciences of the United States of America 107, 22231-22236.

Folkes, A.J., Ahmadi, K., Alderton, W.K., Alix, S., Baker, S.J., Box, G., Chuckowree, I.S., Clarke, P.A., Depledge, P., Eccles, S.A., et al. (2008). The identification of 2-(1H-indazol-4-yl)-6-(4-methanesulfonyl-piperazin-1-ylmethyl)-4-morpholin-4-yl-t hieno[3,2-

Page 189: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

172

d]pyrimidine (GDC-0941) as a potent, selective, orally bioavailable inhibitor of class I PI3 kinase for the treatment of cancer. Journal of medicinal chemistry 51, 5522-5532.

Foukas, L.C., Claret, M., Pearce, W., Okkenhaug, K., Meek, S., Peskett, E., Sancho, S., Smith, A.J., Withers, D.J., and Vanhaesebroeck, B. (2006). Critical role for the p110alpha phosphoinositide-3-OH kinase in growth and metabolic regulation. Nature 441, 366-370.

Franco, I., Gulluni, F., Campa, C.C., Costa, C., Margaria, J.P., Ciraolo, E., Martini, M., Monteyne, D., De Luca, E., Germena, G., et al. (2014). PI3K class II alpha controls spatially restricted endosomal PtdIns3P and Rab11 activation to promote primary cilium function. Developmental cell 28, 647-658.

Fritsch, R., de Krijger, I., Fritsch, K., George, R., Reason, B., Kumar, M.S., Diefenbacher, M., Stamp, G., and Downward, J. (2013). RAS and RHO families of GTPases directly regulate distinct phosphoinositide 3-kinase isoforms. Cell 153, 1050-1063.

Fruman, D.A., and Cantley, L.C. (2014). Idelalisib--a PI3Kdelta inhibitor for B-cell cancers. The New England journal of medicine 370, 1061-1062.

Fruman, D.A., Mauvais-Jarvis, F., Pollard, D.A., Yballe, C.M., Brazil, D., Bronson, R.T., Kahn, C.R., and Cantley, L.C. (2000). Hypoglycaemia, liver necrosis and perinatal death in mice lacking all isoforms of phosphoinositide 3-kinase p85 alpha. Nature genetics 26, 379-382.

Fruman, D.A., and Rommel, C. (2014). PI3K and cancer: lessons, challenges and opportunities. Nature reviews. Drug discovery 13, 140-156.

Furet, P., Guagnano, V., Fairhurst, R.A., Imbach-Weese, P., Bruce, I., Knapp, M., Fritsch, C., Blasco, F., Blanz, J., Aichholz, R., et al. (2013). Discovery of NVP-BYL719 a potent and selective phosphatidylinositol-3 kinase alpha inhibitor selected for clinical evaluation. Bioorganic & medicinal chemistry letters 23, 3741-3748.

Furman, R.R., Sharman, J.P., Coutre, S.E., Cheson, B.D., Pagel, J.M., Hillmen, P., Barrientos, J.C., Zelenetz, A.D., Kipps, T.J., Flinn, I., et al. (2014). Idelalisib and rituximab in relapsed chronic lymphocytic leukemia. The New England journal of medicine 370, 997-1007.

Gao, J., Aksoy, B.A., Dogrusoz, U., Dresdner, G., Gross, B., Sumer, S.O., Sun, Y., Jacobsen, A., Sinha, R., Larsson, E., et al. (2013). Integrative analysis of complex cancer genomics and clinical profiles using the cBioPortal. Science signaling 6, pl1.

Garrett, J.T., Olivares, M.G., Rinehart, C., Granja-Ingram, N.D., Sanchez, V., Chakrabarty, A., Dave, B., Cook, R.S., Pao, W., McKinely, E., et al. (2011). Transcriptional and posttranslational up-regulation of HER3 (ErbB3) compensates for inhibition of the HER2 tyrosine kinase. Proceedings of the National Academy of Sciences of the United States of America 108, 5021-5026.

Garrett, J.T., Sutton, C.R., Kurupi, R., Bialucha, C.U., Ettenberg, S.A., Collins, S.D., Sheng, Q., Wallweber, J., Defazio-Eli, L., and Arteaga, C.L. (2013). Combination of

Page 190: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

173

antibody that inhibits ligand-independent HER3 dimerization and a p110alpha inhibitor potently blocks PI3K signaling and growth of HER2+ breast cancers. Cancer research 73, 6013-6023.

Geering, B., Cutillas, P.R., Nock, G., Gharbi, S.I., and Vanhaesebroeck, B. (2007). Class IA phosphoinositide 3-kinases are obligate p85-p110 heterodimers. Proceedings of the National Academy of Sciences of the United States of America 104, 7809-7814.

Gewinner, C., Wang, Z.C., Richardson, A., Teruya-Feldstein, J., Etemadmoghadam, D., Bowtell, D., Barretina, J., Lin, W.M., Rameh, L., Salmena, L., et al. (2009). Evidence that inositol polyphosphate 4-phosphatase type II is a tumor suppressor that inhibits PI3K signaling. Cancer cell 16, 115-125.

Golub, T.R., Barker, G.F., Lovett, M., and Gilliland, D.G. (1994). Fusion of PDGF receptor beta to a novel ets-like gene, tel, in chronic myelomonocytic leukemia with t(5;12) chromosomal translocation. Cell 77, 307-316.

Gonzalez-Billalabeitia, E., Seitzer, N., Song, S.J., Song, M.S., Patnaik, A., Liu, X.S., Epping, M.T., Papa, A., Hobbs, R.M., Chen, M., et al. (2014). Vulnerabilities of PTEN-p53-deficient prostate cancers to compound PARP/PI3K inhibition. Cancer discovery.

Gopal, A.K., Kahl, B.S., de Vos, S., Wagner-Johnston, N.D., Schuster, S.J., Jurczak, W.J., Flinn, I.W., Flowers, C.R., Martin, P., Viardot, A., et al. (2014). PI3Kdelta inhibition by idelalisib in patients with relapsed indolent lymphoma. The New England journal of medicine 370, 1008-1018.

Graupera, M., Guillermet-Guibert, J., Foukas, L.C., Phng, L.K., Cain, R.J., Salpekar, A., Pearce, W., Meek, S., Millan, J., Cutillas, P.R., et al. (2008). Angiogenesis selectively requires the p110alpha isoform of PI3K to control endothelial cell migration. Nature 453, 662-666.

Gritsman, K., Yuzugullu, H., Von, T., Yan, H., Clayton, L., Fritsch, C., Maira, S.M., Hollingworth, G., Choi, C., Khandan, T., et al. (2014). Hematopoiesis and RAS-driven myeloid leukemia differentially require PI3K isoform p110alpha. The Journal of clinical investigation 124, 1794-1809.

Gruber Filbin, M., Dabral, S.K., Pazyra-Murphy, M.F., Ramkissoon, S., Kung, A.L., Pak, E., Chung, J., Theisen, M.A., Sun, Y., Franchetti, Y., et al. (2013). Coordinate activation of Shh and PI3K signaling in PTEN-deficient glioblastoma: new therapeutic opportunities. Nature medicine 19, 1518-1523.

Guillermet-Guibert, J., Bjorklof, K., Salpekar, A., Gonella, C., Ramadani, F., Bilancio, A., Meek, S., Smith, A.J., Okkenhaug, K., and Vanhaesebroeck, B. (2008). The p110beta isoform of phosphoinositide 3-kinase signals downstream of G protein-coupled receptors and is functionally redundant with p110gamma. Proceedings of the National Academy of Sciences of the United States of America 105, 8292-8297.

Gupta, S., Ramjaun, A.R., Haiko, P., Wang, Y., Warne, P.H., Nicke, B., Nye, E., Stamp, G., Alitalo, K., and Downward, J. (2007). Binding of ras to phosphoinositide 3-kinase p110alpha is required for ras-driven tumorigenesis in mice. Cell 129, 957-968.

Page 191: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

174

Guy, P.M., Platko, J.V., Cantley, L.C., Cerione, R.A., and Carraway, K.L., 3rd (1994). Insect cell-expressed p180erbB3 possesses an impaired tyrosine kinase activity. Proceedings of the National Academy of Sciences of the United States of America 91, 8132-8136.

Gymnopoulos, M., Elsliger, M.A., and Vogt, P.K. (2007). Rare cancer-specific mutations in PIK3CA show gain of function. Proceedings of the National Academy of Sciences of the United States of America 104, 5569-5574.

Hanahan, D., and Weinberg, R.A. (2011). Hallmarks of cancer: the next generation. Cell 144, 646-674.

Hao, Y., Wang, C., Cao, B., Hirsch, B.M., Song, J., Markowitz, S.D., Ewing, R.M., Sedwick, D., Liu, L., Zheng, W., et al. (2013). Gain of interaction with IRS1 by p110alpha-helical domain mutants is crucial for their oncogenic functions. Cancer cell 23, 583-593.

Harada, K., Truong, A.B., Cai, T., and Khavari, P.A. (2005). The class II phosphoinositide 3-kinase C2beta is not essential for epidermal differentiation. Molecular and cellular biology 25, 11122-11130.

Harris, D.P., Vogel, P., Wims, M., Moberg, K., Humphries, J., Jhaver, K.G., DaCosta, C.M., Shadoan, M.K., Xu, N., Hansen, G.M., et al. (2011). Requirement for class II phosphoinositide 3-kinase C2alpha in maintenance of glomerular structure and function. Molecular and cellular biology 31, 63-80.

Hellyer, N.J., Cheng, K., and Koland, J.G. (1998). ErbB3 (HER3) interaction with the p85 regulatory subunit of phosphoinositide 3-kinase. The Biochemical journal 333 ( Pt 3), 757-763.

Hennighausen, L., and Robinson, G.W. (1998). Think globally, act locally: the making of a mouse mammary gland. Genes & development 12, 449-455.

Hennighausen, L., and Robinson, G.W. (2001). Signaling pathways in mammary gland development. Developmental cell 1, 467-475.

Hirai, H., Sootome, H., Nakatsuru, Y., Miyama, K., Taguchi, S., Tsujioka, K., Ueno, Y., Hatch, H., Majumder, P.K., Pan, B.S., et al. (2010). MK-2206, an allosteric Akt inhibitor, enhances antitumor efficacy by standard chemotherapeutic agents or molecular targeted drugs in vitro and in vivo. Molecular cancer therapeutics 9, 1956-1967.

Hirsch, D.S., Shen, Y., Dokmanovic, M., and Wu, W.J. (2010). pp60c-Src phosphorylates and activates vacuolar protein sorting 34 to mediate cellular transformation. Cancer research 70, 5974-5983.

Hodgson, M.C., Shao, L.J., Frolov, A., Li, R., Peterson, L.E., Ayala, G., Ittmann, M.M., Weigel, N.L., and Agoulnik, I.U. (2011). Decreased expression and androgen regulation of the tumor suppressor gene INPP4B in prostate cancer. Cancer research 71, 572-582.

Hon, W.C., Berndt, A., and Williams, R.L. (2012). Regulation of lipid binding underlies the activation mechanism of class IA PI3-kinases. Oncogene 31, 3655-3666.

Page 192: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

175

Huang, C.H., Mandelker, D., Schmidt-Kittler, O., Samuels, Y., Velculescu, V.E., Kinzler, K.W., Vogelstein, B., Gabelli, S.B., and Amzel, L.M. (2007). The structure of a human p110alpha/p85alpha complex elucidates the effects of oncogenic PI3Kalpha mutations. Science 318, 1744-1748.

Huang, S.M., Mishina, Y.M., Liu, S., Cheung, A., Stegmeier, F., Michaud, G.A., Charlat, O., Wiellette, E., Zhang, Y., Wiessner, S., et al. (2009). Tankyrase inhibition stabilizes axin and antagonizes Wnt signalling. Nature 461, 614-620.

Huang, X., Hou, J., Shen, X., Huang, C., Zhang, X., Xie, Y., and Luo, X. (2014). MicroRNA-486-5p, which is downregulated in hepatocellular carcinoma, suppresses tumor growth by targeting PIK3R1. The FEBS journal.

Huang, X., Shen, Y., Liu, M., Bi, C., Jiang, C., Iqbal, J., McKeithan, T.W., Chan, W.C., Ding, S.J., and Fu, K. (2012). Quantitative proteomics reveals that miR-155 regulates the PI3K-AKT pathway in diffuse large B-cell lymphoma. The American journal of pathology 181, 26-33.

Hynes, N.E., and Watson, C.J. (2010). Mammary gland growth factors: roles in normal development and in cancer. Cold Spring Harbor perspectives in biology 2, a003186.

Ibrahim, Y.H., Garcia-Garcia, C., Serra, V., He, L., Torres-Lockhart, K., Prat, A., Anton, P., Cozar, P., Guzman, M., Grueso, J., et al. (2012). PI3K inhibition impairs BRCA1/2 expression and sensitizes BRCA-proficient triple-negative breast cancer to PARP inhibition. Cancer discovery 2, 1036-1047.

Ilic, N., Utermark, T., Widlund, H.R., and Roberts, T.M. (2011). PI3K-targeted therapy can be evaded by gene amplification along the MYC-eukaryotic translation initiation factor 4E (eIF4E) axis. Proceedings of the National Academy of Sciences of the United States of America 108, E699-708.

Isakoff, S.J., Engelman, J.A., Irie, H.Y., Luo, J., Brachmann, S.M., Pearline, R.V., Cantley, L.C., and Brugge, J.S. (2005). Breast cancer-associated PIK3CA mutations are oncogenic in mammary epithelial cells. Cancer research 65, 10992-11000.

Ito, Y., Hart, J.R., Ueno, L., and Vogt, P.K. (2014). Oncogenic activity of the regulatory subunit p85beta of phosphatidylinositol 3-kinase (PI3K). Proceedings of the National Academy of Sciences of the United States of America.

Jackson, S.P., Schoenwaelder, S.M., Goncalves, I., Nesbitt, W.S., Yap, C.L., Wright, C.E., Kenche, V., Anderson, K.E., Dopheide, S.M., Yuan, Y., et al. (2005). PI 3-kinase p110beta: a new target for antithrombotic therapy. Nature medicine 11, 507-514.

Jaiswal, B.S., Janakiraman, V., Kljavin, N.M., Chaudhuri, S., Stern, H.M., Wang, W., Kan, Z., Dbouk, H.A., Peters, B.A., Waring, P., et al. (2009). Somatic mutations in p85alpha promote tumorigenesis through class IA PI3K activation. Cancer cell 16, 463-474.

Jean, S., and Kiger, A.A. (2012). Coordination between RAB GTPase and phosphoinositide regulation and functions. Nature reviews. Molecular cell biology 13, 463-470.

Page 193: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

176

Jia, S., Gao, X., Lee, S.H., Maira, S.M., Wu, X., Stack, E.C., Signoretti, S., Loda, M., Zhao, J.J., and Roberts, T.M. (2013). Opposing effects of androgen deprivation and targeted therapy on prostate cancer prevention. Cancer discovery 3, 44-51.

Jia, S., Liu, Z., Zhang, S., Liu, P., Zhang, L., Lee, S.H., Zhang, J., Signoretti, S., Loda, M., Roberts, T.M., et al. (2008). Essential roles of PI(3)K-p110beta in cell growth, metabolism and tumorigenesis. Nature 454, 776-779.

Jou, S.T., Carpino, N., Takahashi, Y., Piekorz, R., Chao, J.R., Carpino, N., Wang, D., and Ihle, J.N. (2002). Essential, nonredundant role for the phosphoinositide 3-kinase p110delta in signaling by the B-cell receptor complex. Molecular and cellular biology 22, 8580-8591.

Jousset, C., Carron, C., Boureux, A., Quang, C.T., Oury, C., Dusanter-Fourt, I., Charon, M., Levin, J., Bernard, O., and Ghysdael, J. (1997). A domain of TEL conserved in a subset of ETS proteins defines a specific oligomerization interface essential to the mitogenic properties of the TEL-PDGFR beta oncoprotein. The EMBO journal 16, 69-82.

Jura, N., Shan, Y., Cao, X., Shaw, D.E., and Kuriyan, J. (2009). Structural analysis of the catalytically inactive kinase domain of the human EGF receptor 3. Proceedings of the National Academy of Sciences of the United States of America 106, 21608-21613.

Juric, D., Argiles, G., Burris, H., Gonzalez-Angulo, A., Saura, C., Quadt, C., Douglas, M., Demanse, D., De Buck, S., and Baselga, J. (2012). Phase I study of BYL719, an alpha-specific PI3K inhibitor, in patients with PIK3CA mutant advanced solid tumors: preliminary efficacy and safety in patients with PIK3CA mutant ER-positive (ER+) metastatic breast cancer (MBC). Cancer research 72, P6-10-07.

Juric, D., Gonzalez-Angulo, A., Burris, H., Schuler, M., Schellens, J., Berlin, J., Gupta, A., Seggewiss-Bernhardt, R., Adamo, B., Gil-Martin, M., et al. (2013a). Abstract P2-16-14: Preliminary safety, pharmacokinetics and anti-tumor activity of BYL719, an alpha-specific PI3K inhibitor in combination with fulvestrant: Results from a phase I study. Cancer research 73, P2-16-14.

Juric, D., Saura, C., Cervantes, A., Kurkjian, C., Patel, M., Sachdev, J., Mayer, I., Krop, I., Oliveira, M., Sanabria, S., et al. (2013b). Abstract PD1-3: Ph1b study of the PI3K inhibitor GDC-0032 in combination with fulvestrant in patients with hormone receptor-positive advanced breast cancer. Cancer research 73, PD1-3.

Juvekar, A., Burga, L.N., Hu, H., Lunsford, E.P., Ibrahim, Y.H., Balmana, J., Rajendran, A., Papa, A., Spencer, K., Lyssiotis, C.A., et al. (2012). Combining a PI3K inhibitor with a PARP inhibitor provides an effective therapy for BRCA1-related breast cancer. Cancer discovery 2, 1048-1063.

Kan, Z., Jaiswal, B.S., Stinson, J., Janakiraman, V., Bhatt, D., Stern, H.M., Yue, P., Haverty, P.M., Bourgon, R., Zheng, J., et al. (2010). Diverse somatic mutation patterns and pathway alterations in human cancers. Nature 466, 869-873.

Kang, S., Bader, A.G., and Vogt, P.K. (2005). Phosphatidylinositol 3-kinase mutations identified in human cancer are oncogenic. Proceedings of the National Academy of Sciences of the United States of America 102, 802-807.

Page 194: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

177

Kang, S., Denley, A., Vanhaesebroeck, B., and Vogt, P.K. (2006). Oncogenic transformation induced by the p110beta, -gamma, and -delta isoforms of class I phosphoinositide 3-kinase. Proceedings of the National Academy of Sciences of the United States of America 103, 1289-1294.

Katso, R.M., Pardo, O.E., Palamidessi, A., Franz, C.M., Marinov, M., De Laurentiis, A., Downward, J., Scita, G., Ridley, A.J., Waterfield, M.D., et al. (2006). Phosphoinositide 3-Kinase C2beta regulates cytoskeletal organization and cell migration via Rac-dependent mechanisms. Molecular biology of the cell 17, 3729-3744.

Kim, Y.J., Kim, H.J., Chung, K.Y., Choi, I., and Kim, S.H. (2014). Transcriptional activation of PIK3R1 by PPARgamma in adipocytes. Molecular biology reports 41, 5267-5272.

Kinross, K.M., Montgomery, K.G., Kleinschmidt, M., Waring, P., Ivetac, I., Tikoo, A., Saad, M., Hare, L., Roh, V., Mantamadiotis, T., et al. (2012). An activating Pik3ca mutation coupled with Pten loss is sufficient to initiate ovarian tumorigenesis in mice. The Journal of clinical investigation 122, 553-557.

Klarlund, J.K., Guilherme, A., Holik, J.J., Virbasius, J.V., Chawla, A., and Czech, M.P. (1997). Signaling by phosphoinositide-3,4,5-trisphosphate through proteins containing pleckstrin and Sec7 homology domains. Science 275, 1927-1930.

Klippel, A., Reinhard, C., Kavanaugh, W.M., Apell, G., Escobedo, M.A., and Williams, L.T. (1996). Membrane localization of phosphatidylinositol 3-kinase is sufficient to activate multiple signal-transducing kinase pathways. Molecular and cellular biology 16, 4117-4127.

Knight, Z.A., Gonzalez, B., Feldman, M.E., Zunder, E.R., Goldenberg, D.D., Williams, O., Loewith, R., Stokoe, D., Balla, A., Toth, B., et al. (2006). A pharmacological map of the PI3-K family defines a role for p110alpha in insulin signaling. Cell 125, 733-747.

Knight, Z.A., and Shokat, K.M. (2005). Features of selective kinase inhibitors. Chemistry & biology 12, 621-637.

Knobbe, C.B., and Reifenberger, G. (2003). Genetic alterations and aberrant expression of genes related to the phosphatidyl-inositol-3'-kinase/protein kinase B (Akt) signal transduction pathway in glioblastomas. Brain pathology 13, 507-518.

Krugmann, S., Anderson, K.E., Ridley, S.H., Risso, N., McGregor, A., Coadwell, J., Davidson, K., Eguinoa, A., Ellson, C.D., Lipp, P., et al. (2002). Identification of ARAP3, a novel PI3K effector regulating both Arf and Rho GTPases, by selective capture on phosphoinositide affinity matrices. Molecular cell 9, 95-108.

Kubo, H., Hazeki, K., Takasuga, S., and Hazeki, O. (2005). Specific role for p85/p110beta in GTP-binding-protein-mediated activation of Akt. The Biochemical journal 392, 607-614.

Kurek, K.C., Luks, V.L., Ayturk, U.M., Alomari, A.I., Fishman, S.J., Spencer, S.A., Mulliken, J.B., Bowen, M.E., Yamamoto, G.L., Kozakewich, H.P., et al. (2012). Somatic

Page 195: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

178

mosaic activating mutations in PIK3CA cause CLOVES syndrome. American journal of human genetics 90, 1108-1115.

Kurig, B., Shymanets, A., Bohnacker, T., Prajwal, Brock, C., Ahmadian, M.R., Schaefer, M., Gohla, A., Harteneck, C., Wymann, M.P., et al. (2009). Ras is an indispensable coregulator of the class IB phosphoinositide 3-kinase p87/p110gamma. Proceedings of the National Academy of Sciences of the United States of America 106, 20312-20317.

Kurosu, H., Maehama, T., Okada, T., Yamamoto, T., Hoshino, S., Fukui, Y., Ui, M., Hazeki, O., and Katada, T. (1997). Heterodimeric phosphoinositide 3-kinase consisting of p85 and p110beta is synergistically activated by the betagamma subunits of G proteins and phosphotyrosyl peptide. The Journal of biological chemistry 272, 24252-24256.

Kwabi-Addo, B., Giri, D., Schmidt, K., Podsypanina, K., Parsons, R., Greenberg, N., and Ittmann, M. (2001). Haploinsufficiency of the Pten tumor suppressor gene promotes prostate cancer progression. Proceedings of the National Academy of Sciences of the United States of America 98, 11563-11568.

Li, G., Robinson, G.W., Lesche, R., Martinez-Diaz, H., Jiang, Z., Rozengurt, N., Wagner, K.U., Wu, D.C., Lane, T.F., Liu, X., et al. (2002). Conditional loss of PTEN leads to precocious development and neoplasia in the mammary gland. Development 129, 4159-4170.

Liaw, D., Marsh, D.J., Li, J., Dahia, P.L., Wang, S.I., Zheng, Z., Bose, S., Call, K.M., Tsou, H.C., Peacocke, M., et al. (1997). Germline mutations of the PTEN gene in Cowden disease, an inherited breast and thyroid cancer syndrome. Nature genetics 16, 64-67.

Liu, P., Cheng, H., Roberts, T.M., and Zhao, J.J. (2009). Targeting the phosphoinositide 3-kinase pathway in cancer. Nature reviews. Drug discovery 8, 627-644.

Liu, P., Cheng, H., Santiago, S., Raeder, M., Zhang, F., Isabella, A., Yang, J., Semaan, D.J., Chen, C., Fox, E.A., et al. (2011). Oncogenic PIK3CA-driven mammary tumors frequently recur via PI3K pathway-dependent and PI3K pathway-independent mechanisms. Nature medicine 17, 1116-1120.

Liu, P., Morrison, C., Wang, L., Xiong, D., Vedell, P., Cui, P., Hua, X., Ding, F., Lu, Y., James, M., et al. (2012). Identification of somatic mutations in non-small cell lung carcinomas using whole-exome sequencing. Carcinogenesis 33, 1270-1276.

Liu, X., Holstege, H., van der Gulden, H., Treur-Mulder, M., Zevenhoven, J., Velds, A., Kerkhoven, R.M., van Vliet, M.H., Wessels, L.F., Peterse, J.L., et al. (2007). Somatic loss of BRCA1 and p53 in mice induces mammary tumors with features of human BRCA1-mutated basal-like breast cancer. Proceedings of the National Academy of Sciences of the United States of America 104, 12111-12116.

Lu, N., Shen, Q., Mahoney, T.R., Neukomm, L.J., Wang, Y., and Zhou, Z. (2012). Two PI 3-kinases and one PI 3-phosphatase together establish the cyclic waves of phagosomal PtdIns(3)P critical for the degradation of apoptotic cells. PLoS biology 10, e1001245.

Page 196: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

179

Lucas, C.L., Kuehn, H.S., Zhao, F., Niemela, J.E., Deenick, E.K., Palendira, U., Avery, D.T., Moens, L., Cannons, J.L., Biancalana, M., et al. (2014). Dominant-activating germline mutations in the gene encoding the PI(3)K catalytic subunit p110delta result in T cell senescence and human immunodeficiency. Nature immunology 15, 88-97.

Luo, J., and Cantley, L.C. (2005). The negative regulation of phosphoinositide 3-kinase signaling by p85 and it's implication in cancer. Cell cycle 4, 1309-1312.

Luo, J., Field, S.J., Lee, J.Y., Engelman, J.A., and Cantley, L.C. (2005a). The p85 regulatory subunit of phosphoinositide 3-kinase down-regulates IRS-1 signaling via the formation of a sequestration complex. The Journal of cell biology 170, 455-464.

Luo, J., McMullen, J.R., Sobkiw, C.L., Zhang, L., Dorfman, A.L., Sherwood, M.C., Logsdon, M.N., Horner, J.W., DePinho, R.A., Izumo, S., et al. (2005b). Class IA phosphoinositide 3-kinase regulates heart size and physiological cardiac hypertrophy. Molecular and cellular biology 25, 9491-9502.

Luo, J., Sobkiw, C.L., Logsdon, N.M., Watt, J.M., Signoretti, S., O'Connell, F., Shin, E., Shim, Y., Pao, L., Neel, B.G., et al. (2005c). Modulation of epithelial neoplasia and lymphoid hyperplasia in PTEN+/- mice by the p85 regulatory subunits of phosphoinositide 3-kinase. Proceedings of the National Academy of Sciences of the United States of America 102, 10238-10243.

Ma, X.J., Dahiya, S., Richardson, E., Erlander, M., and Sgroi, D.C. (2009). Gene expression profiling of the tumor microenvironment during breast cancer progression. Breast cancer research : BCR 11, R7.

MacLennan, M.B., Anderson, B.M., and Ma, D.W. (2011). Differential mammary gland development in FVB and C57Bl/6 mice: implications for breast cancer research. Nutrients 3, 929-936.

Maffucci, T., Cooke, F.T., Foster, F.M., Traer, C.J., Fry, M.J., and Falasca, M. (2005). Class II phosphoinositide 3-kinase defines a novel signaling pathway in cell migration. The Journal of cell biology 169, 789-799.

Maier, U., Babich, A., and Nurnberg, B. (1999). Roles of non-catalytic subunits in gbetagamma-induced activation of class I phosphoinositide 3-kinase isoforms beta and gamma. The Journal of biological chemistry 274, 29311-29317.

Maira, S.M., Stauffer, F., Brueggen, J., Furet, P., Schnell, C., Fritsch, C., Brachmann, S., Chene, P., De Pover, A., Schoemaker, K., et al. (2008). Identification and characterization of NVP-BEZ235, a new orally available dual phosphatidylinositol 3-kinase/mammalian target of rapamycin inhibitor with potent in vivo antitumor activity. Molecular cancer therapeutics 7, 1851-1863.

Mandelker, D., Gabelli, S.B., Schmidt-Kittler, O., Zhu, J., Cheong, I., Huang, C.H., Kinzler, K.W., Vogelstein, B., and Amzel, L.M. (2009). A frequent kinase domain mutation that changes the interaction between PI3Kalpha and the membrane. Proceedings of the National Academy of Sciences of the United States of America 106, 16996-17001.

Page 197: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

180

Martin, A.L., Schwartz, M.D., Jameson, S.C., and Shimizu, Y. (2008). Selective regulation of CD8 effector T cell migration by the p110 gamma isoform of phosphatidylinositol 3-kinase. Journal of immunology 180, 2081-2088.

Mauvais-Jarvis, F., Ueki, K., Fruman, D.A., Hirshman, M.F., Sakamoto, K., Goodyear, L.J., Iannacone, M., Accili, D., Cantley, L.C., and Kahn, C.R. (2002). Reduced expression of the murine p85alpha subunit of phosphoinositide 3-kinase improves insulin signaling and ameliorates diabetes. The Journal of clinical investigation 109, 141-149.

McCurdy, C.E., Schenk, S., Holliday, M.J., Philp, A., Houck, J.A., Patsouris, D., MacLean, P.S., Majka, S.M., Klemm, D.J., and Friedman, J.E. (2012). Attenuated Pik3r1 expression prevents insulin resistance and adipose tissue macrophage accumulation in diet-induced obese mice. Diabetes 61, 2495-2505.

Meier, T.I., Cook, J.A., Thomas, J.E., Radding, J.A., Horn, C., Lingaraj, T., and Smith, M.C. (2004). Cloning, expression, purification, and characterization of the human Class Ia phosphoinositide 3-kinase isoforms. Protein expression and purification 35, 218-224.

Mellor, P., Furber, L.A., Nyarko, J.N., and Anderson, D.H. (2012). Multiple roles for the p85alpha isoform in the regulation and function of PI3K signalling and receptor trafficking. The Biochemical journal 441, 23-37.

Miaczynska, M. (2013). Effects of membrane trafficking on signaling by receptor tyrosine kinases. Cold Spring Harbor perspectives in biology 5, a009035.

Miled, N., Yan, Y., Hon, W.C., Perisic, O., Zvelebil, M., Inbar, Y., Schneidman-Duhovny, D., Wolfson, H.J., Backer, J.M., and Williams, R.L. (2007). Mechanism of two classes of cancer mutations in the phosphoinositide 3-kinase catalytic subunit. Science 317, 239-242.

Muellner, M.K., Uras, I.Z., Gapp, B.V., Kerzendorfer, C., Smida, M., Lechtermann, H., Craig-Mueller, N., Colinge, J., Duernberger, G., and Nijman, S.M. (2011). A chemical-genetic screen reveals a mechanism of resistance to PI3K inhibitors in cancer. Nature chemical biology 7, 787-793.

Muranen, T., Selfors, L.M., Worster, D.T., Iwanicki, M.P., Song, L., Morales, F.C., Gao, S., Mills, G.B., and Brugge, J.S. (2012). Inhibition of PI3K/mTOR leads to adaptive resistance in matrix-attached cancer cells. Cancer cell 21, 227-239.

Murga, C., Fukuhara, S., and Gutkind, J.S. (2000). A novel role for phosphatidylinositol 3-kinase beta in signaling from G protein-coupled receptors to Akt. The Journal of biological chemistry 275, 12069-12073.

Murphy, G.A., Graham, S.M., Morita, S., Reks, S.E., Rogers-Graham, K., Vojtek, A., Kelley, G.G., and Der, C.J. (2002). Involvement of phosphatidylinositol 3-kinase, but not RalGDS, in TC21/R-Ras2-mediated transformation. The Journal of biological chemistry 277, 9966-9975.

Ni, J., Liu, Q., Xie, S., Carlson, C., Von, T., Vogel, K., Riddle, S., Benes, C., Eck, M., Roberts, T., et al. (2012). Functional characterization of an isoform-selective inhibitor of PI3K-p110beta as a potential anticancer agent. Cancer discovery 2, 425-433.

Page 198: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

181

Nicoli, S., Knyphausen, C.P., Zhu, L.J., Lakshmanan, A., and Lawson, N.D. (2012). miR-221 is required for endothelial tip cell behaviors during vascular development. Developmental cell 22, 418-429.

Nobusawa, S., Lachuer, J., Wierinckx, A., Kim, Y.H., Huang, J., Legras, C., Kleihues, P., and Ohgaki, H. (2010). Intratumoral patterns of genomic imbalance in glioblastomas. Brain pathology 20, 936-944.

Norris, F.A., Atkins, R.C., and Majerus, P.W. (1997). The cDNA cloning and characterization of inositol polyphosphate 4-phosphatase type II. Evidence for conserved alternative splicing in the 4-phosphatase family. The Journal of biological chemistry 272, 23859-23864.

O'Reilly, K.E., Rojo, F., She, Q.B., Solit, D., Mills, G.B., Smith, D., Lane, H., Hofmann, F., Hicklin, D.J., Ludwig, D.L., et al. (2006). mTOR inhibition induces upstream receptor tyrosine kinase signaling and activates Akt. Cancer research 66, 1500-1508.

Okkenhaug, K., Bilancio, A., Farjot, G., Priddle, H., Sancho, S., Peskett, E., Pearce, W., Meek, S.E., Salpekar, A., Waterfield, M.D., et al. (2002). Impaired B and T cell antigen receptor signaling in p110delta PI 3-kinase mutant mice. Science 297, 1031-1034.

Okkenhaug, K., and Vanhaesebroeck, B. (2003). PI3K in lymphocyte development, differentiation and activation. Nature reviews. Immunology 3, 317-330.

Orloff, M.S., He, X., Peterson, C., Chen, F., Chen, J.L., Mester, J.L., and Eng, C. (2013). Germline PIK3CA and AKT1 mutations in Cowden and Cowden-like syndromes. American journal of human genetics 92, 76-80.

Pacold, M.E., Suire, S., Perisic, O., Lara-Gonzalez, S., Davis, C.T., Walker, E.H., Hawkins, P.T., Stephens, L., Eccleston, J.F., and Williams, R.L. (2000). Crystal structure and functional analysis of Ras binding to its effector phosphoinositide 3-kinase gamma. Cell 103, 931-943.

Papakonstanti, E.A., Zwaenepoel, O., Bilancio, A., Burns, E., Nock, G.E., Houseman, B., Shokat, K., Ridley, A.J., and Vanhaesebroeck, B. (2008). Distinct roles of class IA PI3K isoforms in primary and immortalised macrophages. Journal of cell science 121, 4124-4133.

Park, S.W., Zhou, Y., Lee, J., Lu, A., Sun, C., Chung, J., Ueki, K., and Ozcan, U. (2010). The regulatory subunits of PI3K, p85alpha and p85beta, interact with XBP-1 and increase its nuclear translocation. Nature medicine 16, 429-437.

Parsons, D.W., Jones, S., Zhang, X., Lin, J.C., Leary, R.J., Angenendt, P., Mankoo, P., Carter, H., Siu, I.M., Gallia, G.L., et al. (2008). An integrated genomic analysis of human glioblastoma multiforme. Science 321, 1807-1812.

Parsons, R. (2004). Human cancer, PTEN and the PI-3 kinase pathway. Seminars in cell & developmental biology 15, 171-176.

Peng, Y., Dai, Y., Hitchcock, C., Yang, X., Kassis, E.S., Liu, L., Luo, Z., Sun, H.L., Cui, R., Wei, H., et al. (2013). Insulin growth factor signaling is regulated by microRNA-486,

Page 199: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

182

an underexpressed microRNA in lung cancer. Proceedings of the National Academy of Sciences of the United States of America 110, 15043-15048.

Philp, A.J., Campbell, I.G., Leet, C., Vincan, E., Rockman, S.P., Whitehead, R.H., Thomas, R.J., and Phillips, W.A. (2001). The phosphatidylinositol 3'-kinase p85alpha gene is an oncogene in human ovarian and colon tumors. Cancer research 61, 7426-7429.

Pinkas-Kramarski, R., Soussan, L., Waterman, H., Levkowitz, G., Alroy, I., Klapper, L., Lavi, S., Seger, R., Ratzkin, B.J., Sela, M., et al. (1996). Diversification of Neu differentiation factor and epidermal growth factor signaling by combinatorial receptor interactions. The EMBO journal 15, 2452-2467.

Planchon, S.M., Waite, K.A., and Eng, C. (2008). The nuclear affairs of PTEN. Journal of cell science 121, 249-253.

Prigent, S.A., and Gullick, W.J. (1994). Identification of c-erbB-3 binding sites for phosphatidylinositol 3'-kinase and SHC using an EGF receptor/c-erbB-3 chimera. The EMBO journal 13, 2831-2841.

Rabinovsky, R., Pochanard, P., McNear, C., Brachmann, S.M., Duke-Cohan, J.S., Garraway, L.A., and Sellers, W.R. (2009). p85 Associates with unphosphorylated PTEN and the PTEN-associated complex. Molecular and cellular biology 29, 5377-5388.

Rahmani, M., Aust, M.M., Attkisson, E., Williams, D.C., Jr., Ferreira-Gonzalez, A., and Grant, S. (2013). Dual inhibition of Bcl-2 and Bcl-xL strikingly enhances PI3K inhibition-induced apoptosis in human myeloid leukemia cells through a GSK3- and Bim-dependent mechanism. Cancer research 73, 1340-1351.

Rameh, L.E., Chen, C.S., and Cantley, L.C. (1995). Phosphatidylinositol (3,4,5)P3 interacts with SH2 domains and modulates PI 3-kinase association with tyrosine-phosphorylated proteins. Cell 83, 821-830.

Rao, S.K., Edwards, J., Joshi, A.D., Siu, I.M., and Riggins, G.J. (2010). A survey of glioblastoma genomic amplifications and deletions. Journal of neuro-oncology 96, 169-179.

Raynaud, F.I., Eccles, S.A., Patel, S., Alix, S., Box, G., Chuckowree, I., Folkes, A., Gowan, S., De Haven Brandon, A., Di Stefano, F., et al. (2009). Biological properties of potent inhibitors of class I phosphatidylinositide 3-kinases: from PI-103 through PI-540, PI-620 to the oral agent GDC-0941. Molecular cancer therapeutics 8, 1725-1738.

Reif, K., Okkenhaug, K., Sasaki, T., Penninger, J.M., Vanhaesebroeck, B., and Cyster, J.G. (2004). Cutting edge: differential roles for phosphoinositide 3-kinases, p110gamma and p110delta, in lymphocyte chemotaxis and homing. Journal of immunology 173, 2236-2240.

Rhodes, D.R., Kalyana-Sundaram, S., Mahavisno, V., Varambally, R., Yu, J., Briggs, B.B., Barrette, T.R., Anstet, M.J., Kincead-Beal, C., Kulkarni, P., et al. (2007). Oncomine 3.0: genes, pathways, and networks in a collection of 18,000 cancer gene expression profiles. Neoplasia 9, 166-180.

Page 200: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

183

Rhodes, D.R., Yu, J., Shanker, K., Deshpande, N., Varambally, R., Ghosh, D., Barrette, T., Pandey, A., and Chinnaiyan, A.M. (2004). ONCOMINE: a cancer microarray database and integrated data-mining platform. Neoplasia 6, 1-6.

Rhodes, N., Heerding, D.A., Duckett, D.R., Eberwein, D.J., Knick, V.B., Lansing, T.J., McConnell, R.T., Gilmer, T.M., Zhang, S.Y., Robell, K., et al. (2008). Characterization of an Akt kinase inhibitor with potent pharmacodynamic and antitumor activity. Cancer research 68, 2366-2374.

Richardson, A.L., Wang, Z.C., De Nicolo, A., Lu, X., Brown, M., Miron, A., Liao, X., Iglehart, J.D., Livingston, D.M., and Ganesan, S. (2006). X chromosomal abnormalities in basal-like human breast cancer. Cancer cell 9, 121-132.

Rios, J.J., Paria, N., Burns, D.K., Israel, B.A., Cornelia, R., Wise, C.A., and Ezaki, M. (2013). Somatic gain-of-function mutations in PIK3CA in patients with macrodactyly. Human molecular genetics 22, 444-451.

Riviere, J.B., Mirzaa, G.M., O'Roak, B.J., Beddaoui, M., Alcantara, D., Conway, R.L., St-Onge, J., Schwartzentruber, J.A., Gripp, K.W., Nikkel, S.M., et al. (2012). De novo germline and postzygotic mutations in AKT3, PIK3R2 and PIK3CA cause a spectrum of related megalencephaly syndromes. Nature genetics 44, 934-940.

Rodon, J., Dienstmann, R., Serra, V., and Tabernero, J. (2013). Development of PI3K inhibitors: lessons learned from early clinical trials. Nature reviews. Clinical oncology 10, 143-153.

Rodriguez-Viciana, P., Sabatier, C., and McCormick, F. (2004). Signaling specificity by Ras family GTPases is determined by the full spectrum of effectors they regulate. Molecular and cellular biology 24, 4943-4954.

Rodriguez-Viciana, P., Warne, P.H., Dhand, R., Vanhaesebroeck, B., Gout, I., Fry, M.J., Waterfield, M.D., and Downward, J. (1994). Phosphatidylinositol-3-OH kinase as a direct target of Ras. Nature 370, 527-532.

Rodriguez-Viciana, P., Warne, P.H., Vanhaesebroeck, B., Waterfield, M.D., and Downward, J. (1996). Activation of phosphoinositide 3-kinase by interaction with Ras and by point mutation. The EMBO journal 15, 2442-2451.

Romanov, S.R., Kozakiewicz, B.K., Holst, C.R., Stampfer, M.R., Haupt, L.M., and Tlsty, T.D. (2001). Normal human mammary epithelial cells spontaneously escape senescence and acquire genomic changes. Nature 409, 633-637.

Rubio, I., Rodriguez-Viciana, P., Downward, J., and Wetzker, R. (1997). Interaction of Ras with phosphoinositide 3-kinase gamma. The Biochemical journal 326 ( Pt 3), 891-895.

Sampath, D., Nannini, M., Jane, G., Hong, R., Parsons, K., Belvin, M., Friedman, L., and Wallin, J. (2013). Abstract P4-15-02: The PI3K inhibitor GDC-0032 enhances the efficacy of standard of care therapeutics in PI3K alpha mutant breast cancer models. Cancer research 73, P4-15-02.

Page 201: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

184

Samuels, Y., Diaz, L.A., Jr., Schmidt-Kittler, O., Cummins, J.M., Delong, L., Cheong, I., Rago, C., Huso, D.L., Lengauer, C., Kinzler, K.W., et al. (2005). Mutant PIK3CA promotes cell growth and invasion of human cancer cells. Cancer cell 7, 561-573.

Samuels, Y., Wang, Z., Bardelli, A., Silliman, N., Ptak, J., Szabo, S., Yan, H., Gazdar, A., Powell, S.M., Riggins, G.J., et al. (2004). High frequency of mutations of the PIK3CA gene in human cancers. Science 304, 554.

Sasaki, T., Irie-Sasaki, J., Jones, R.G., Oliveira-dos-Santos, A.J., Stanford, W.L., Bolon, B., Wakeham, A., Itie, A., Bouchard, D., Kozieradzki, I., et al. (2000). Function of PI3Kgamma in thymocyte development, T cell activation, and neutrophil migration. Science 287, 1040-1046.

Saudemont, A., Garcon, F., Yadi, H., Roche-Molina, M., Kim, N., Segonds-Pichon, A., Martin-Fontecha, A., Okkenhaug, K., and Colucci, F. (2009). p110gamma and p110delta isoforms of phosphoinositide 3-kinase differentially regulate natural killer cell migration in health and disease. Proceedings of the National Academy of Sciences of the United States of America 106, 5795-5800.

Sawyer, C., Sturge, J., Bennett, D.C., O'Hare, M.J., Allen, W.E., Bain, J., Jones, G.E., and Vanhaesebroeck, B. (2003). Regulation of breast cancer cell chemotaxis by the phosphoinositide 3-kinase p110delta. Cancer research 63, 1667-1675.

Schade, B., Rao, T., Dourdin, N., Lesurf, R., Hallett, M., Cardiff, R.D., and Muller, W.J. (2009). PTEN deficiency in a luminal ErbB-2 mouse model results in dramatic acceleration of mammary tumorigenesis and metastasis. The Journal of biological chemistry 284, 19018-19026.

Schmid, M.C., Avraamides, C.J., Dippold, H.C., Franco, I., Foubert, P., Ellies, L.G., Acevedo, L.M., Manglicmot, J.R., Song, X., Wrasidlo, W., et al. (2011). Receptor tyrosine kinases and TLR/IL1Rs unexpectedly activate myeloid cell PI3kgamma, a single convergent point promoting tumor inflammation and progression. Cancer cell 19, 715-727.

Schmid, M.C., Franco, I., Kang, S.W., Hirsch, E., Quilliam, L.A., and Varner, J.A. (2013). PI3-kinase gamma promotes Rap1a-mediated activation of myeloid cell integrin alpha4beta1, leading to tumor inflammation and growth. PloS one 8, e60226.

Schmit, F., Utermark, T., Zhang, S., Wang, Q., Von, T., Roberts, T.M., and Zhao, J.J. (2014). PI3K isoform dependence of PTEN-deficient tumors can be altered by the genetic context. Proceedings of the National Academy of Sciences of the United States of America 111, 6395-6400.

Schu, P.V., Takegawa, K., Fry, M.J., Stack, J.H., Waterfield, M.D., and Emr, S.D. (1993). Phosphatidylinositol 3-kinase encoded by yeast VPS34 gene essential for protein sorting. Science 260, 88-91.

Sergina, N.V., Rausch, M., Wang, D., Blair, J., Hann, B., Shokat, K.M., and Moasser, M.M. (2007). Escape from HER-family tyrosine kinase inhibitor therapy by the kinase-inactive HER3. Nature 445, 437-441.

Page 202: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

185

Serra, V., Scaltriti, M., Prudkin, L., Eichhorn, P.J., Ibrahim, Y.H., Chandarlapaty, S., Markman, B., Rodriguez, O., Guzman, M., Rodriguez, S., et al. (2011). PI3K inhibition results in enhanced HER signaling and acquired ERK dependency in HER2-overexpressing breast cancer. Oncogene 30, 2547-2557.

Shi, F., Telesco, S.E., Liu, Y., Radhakrishnan, R., and Lemmon, M.A. (2010). ErbB3/HER3 intracellular domain is competent to bind ATP and catalyze autophosphorylation. Proceedings of the National Academy of Sciences of the United States of America 107, 7692-7697.

Shymanets, A., Prajwal, Bucher, K., Beer-Hammer, S., Harteneck, C., and Nurnberg, B. (2013). p87 and p101 subunits are distinct regulators determining class IB phosphoinositide 3-kinase (PI3K) specificity. The Journal of biological chemistry 288, 31059-31068.

Siegel, P.M., Ryan, E.D., Cardiff, R.D., and Muller, W.J. (1999). Elevated expression of activated forms of Neu/ErbB-2 and ErbB-3 are involved in the induction of mammary tumors in transgenic mice: implications for human breast cancer. The EMBO journal 18, 2149-2164.

Sierke, S.L., Cheng, K., Kim, H.H., and Koland, J.G. (1997). Biochemical characterization of the protein tyrosine kinase homology domain of the ErbB3 (HER3) receptor protein. The Biochemical journal 322 ( Pt 3), 757-763.

Soltoff, S.P., Carraway, K.L., 3rd, Prigent, S.A., Gullick, W.G., and Cantley, L.C. (1994). ErbB3 is involved in activation of phosphatidylinositol 3-kinase by epidermal growth factor. Molecular and cellular biology 14, 3550-3558.

Song, M.S., Carracedo, A., Salmena, L., Song, S.J., Egia, A., Malumbres, M., and Pandolfi, P.P. (2011). Nuclear PTEN regulates the APC-CDH1 tumor-suppressive complex in a phosphatase-independent manner. Cell 144, 187-199.

Song, M.S., Salmena, L., and Pandolfi, P.P. (2012). The functions and regulation of the PTEN tumour suppressor. Nature reviews. Molecular cell biology 13, 283-296.

Sopasakis, V.R., Liu, P., Suzuki, R., Kondo, T., Winnay, J., Tran, T.T., Asano, T., Smyth, G., Sajan, M.P., Farese, R.V., et al. (2010). Specific roles of the p110alpha isoform of phosphatidylinsositol 3-kinase in hepatic insulin signaling and metabolic regulation. Cell metabolism 11, 220-230.

Sorlie, T., Perou, C.M., Tibshirani, R., Aas, T., Geisler, S., Johnsen, H., Hastie, T., Eisen, M.B., van de Rijn, M., Jeffrey, S.S., et al. (2001). Gene expression patterns of breast carcinomas distinguish tumor subclasses with clinical implications. Proceedings of the National Academy of Sciences of the United States of America 98, 10869-10874.

Sorlie, T., Tibshirani, R., Parker, J., Hastie, T., Marron, J.S., Nobel, A., Deng, S., Johnsen, H., Pesich, R., Geisler, S., et al. (2003). Repeated observation of breast tumor subtypes in independent gene expression data sets. Proceedings of the National Academy of Sciences of the United States of America 100, 8418-8423.

Page 203: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

186

Stambolic, V., Suzuki, A., de la Pompa, J.L., Brothers, G.M., Mirtsos, C., Sasaki, T., Ruland, J., Penninger, J.M., Siderovski, D.P., and Mak, T.W. (1998). Negative regulation of PKB/Akt-dependent cell survival by the tumor suppressor PTEN. Cell 95, 29-39.

Stenmark, H. (2009). Rab GTPases as coordinators of vesicle traffic. Nature reviews. Molecular cell biology 10, 513-525.

Stephens, L.R., Eguinoa, A., Erdjument-Bromage, H., Lui, M., Cooke, F., Coadwell, J., Smrcka, A.S., Thelen, M., Cadwallader, K., Tempst, P., et al. (1997). The G beta gamma sensitivity of a PI3K is dependent upon a tightly associated adaptor, p101. Cell 89, 105-114.

Stjernstrom, A., Karlsson, C., Fernandez, O.J., Soderkvist, P., Karlsson, M.G., and Thunell, L.K. (2014). Alterations of INPP4B, PIK3CA and pAkt of the PI3K pathway are associated with squamous cell carcinoma of the lung. Cancer medicine.

Stoyanov, B., Volinia, S., Hanck, T., Rubio, I., Loubtchenkov, M., Malek, D., Stoyanova, S., Vanhaesebroeck, B., Dhand, R., Nurnberg, B., et al. (1995). Cloning and characterization of a G protein-activated human phosphoinositide-3 kinase. Science 269, 690-693.

Subramaniam, P.S., Whye, D.W., Efimenko, E., Chen, J., Tosello, V., De Keersmaecker, K., Kashishian, A., Thompson, M.A., Castillo, M., Cordon-Cardo, C., et al. (2012). Targeting nonclassical oncogenes for therapy in T-ALL. Cancer cell 21, 459-472.

Suire, S., Coadwell, J., Ferguson, G.J., Davidson, K., Hawkins, P., and Stephens, L. (2005). p84, a new Gbetagamma-activated regulatory subunit of the type IB phosphoinositide 3-kinase p110gamma. Current biology : CB 15, 566-570.

Suire, S., Condliffe, A.M., Ferguson, G.J., Ellson, C.D., Guillou, H., Davidson, K., Welch, H., Coadwell, J., Turner, M., Chilvers, E.R., et al. (2006). Gbetagammas and the Ras binding domain of p110gamma are both important regulators of PI(3)Kgamma signalling in neutrophils. Nature cell biology 8, 1303-1309.

Suire, S., Hawkins, P., and Stephens, L. (2002). Activation of phosphoinositide 3-kinase gamma by Ras. Current biology : CB 12, 1068-1075.

Sun, M., Hillmann, P., Hofmann, B.T., Hart, J.R., and Vogt, P.K. (2010). Cancer-derived mutations in the regulatory subunit p85alpha of phosphoinositide 3-kinase function through the catalytic subunit p110alpha. Proceedings of the National Academy of Sciences of the United States of America 107, 15547-15552.

Taniguchi, C.M., Tran, T.T., Kondo, T., Luo, J., Ueki, K., Cantley, L.C., and Kahn, C.R. (2006). Phosphoinositide 3-kinase regulatory subunit p85alpha suppresses insulin action via positive regulation of PTEN. Proceedings of the National Academy of Sciences of the United States of America 103, 12093-12097.

Taniguchi, C.M., Winnay, J., Kondo, T., Bronson, R.T., Guimaraes, A.R., Aleman, J.O., Luo, J., Stephanopoulos, G., Weissleder, R., Cantley, L.C., et al. (2010). The phosphoinositide 3-kinase regulatory subunit p85alpha can exert tumor suppressor

Page 204: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

187

properties through negative regulation of growth factor signaling. Cancer research 70, 5305-5315.

Tao, J.J., Castel, P., Radosevic-Robin, N., Elkabets, M., Auricchio, N., Aceto, N., Weitsman, G., Barber, P., Vojnovic, B., Ellis, H., et al. (2014). Antagonism of EGFR and HER3 enhances the response to inhibitors of the PI3K-Akt pathway in triple-negative breast cancer. Science signaling 7, ra29.

Tenbaum, S.P., Ordonez-Moran, P., Puig, I., Chicote, I., Arques, O., Landolfi, S., Fernandez, Y., Herance, J.R., Gispert, J.D., Mendizabal, L., et al. (2012). beta-catenin confers resistance to PI3K and AKT inhibitors and subverts FOXO3a to promote metastasis in colon cancer. Nature medicine 18, 892-901.

Terauchi, Y., Tsuji, Y., Satoh, S., Minoura, H., Murakami, K., Okuno, A., Inukai, K., Asano, T., Kaburagi, Y., Ueki, K., et al. (1999). Increased insulin sensitivity and hypoglycaemia in mice lacking the p85 alpha subunit of phosphoinositide 3-kinase. Nature genetics 21, 230-235.

Thomas, R.K., Baker, A.C., Debiasi, R.M., Winckler, W., Laframboise, T., Lin, W.M., Wang, M., Feng, W., Zander, T., MacConaill, L., et al. (2007). High-throughput oncogene mutation profiling in human cancer. Nature genetics 39, 347-351.

Tian, L., Fang, Y.X., Xue, J.L., and Chen, J.Z. (2013). Four microRNAs promote prostate cell proliferation with regulation of PTEN and its downstream signals in vitro. PloS one 8, e75885.

Torbett, N.E., Luna-Moran, A., Knight, Z.A., Houk, A., Moasser, M., Weiss, W., Shokat, K.M., and Stokoe, D. (2008). A chemical screen in diverse breast cancer cell lines reveals genetic enhancers and suppressors of sensitivity to PI3K isoform-selective inhibition. The Biochemical journal 415, 97-110.

Tzahar, E., Waterman, H., Chen, X., Levkowitz, G., Karunagaran, D., Lavi, S., Ratzkin, B.J., and Yarden, Y. (1996). A hierarchical network of interreceptor interactions determines signal transduction by Neu differentiation factor/neuregulin and epidermal growth factor. Molecular and cellular biology 16, 5276-5287.

Ueki, K., Algenstaedt, P., Mauvais-Jarvis, F., and Kahn, C.R. (2000). Positive and negative regulation of phosphoinositide 3-kinase-dependent signaling pathways by three different gene products of the p85alpha regulatory subunit. Molecular and cellular biology 20, 8035-8046.

Ueki, K., Fruman, D.A., Brachmann, S.M., Tseng, Y.H., Cantley, L.C., and Kahn, C.R. (2002a). Molecular balance between the regulatory and catalytic subunits of phosphoinositide 3-kinase regulates cell signaling and survival. Molecular and cellular biology 22, 965-977.

Ueki, K., Fruman, D.A., Yballe, C.M., Fasshauer, M., Klein, J., Asano, T., Cantley, L.C., and Kahn, C.R. (2003). Positive and negative roles of p85 alpha and p85 beta regulatory subunits of phosphoinositide 3-kinase in insulin signaling. The Journal of biological chemistry 278, 48453-48466.

Page 205: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

188

Ueki, K., Yballe, C.M., Brachmann, S.M., Vicent, D., Watt, J.M., Kahn, C.R., and Cantley, L.C. (2002b). Increased insulin sensitivity in mice lacking p85beta subunit of phosphoinositide 3-kinase. Proceedings of the National Academy of Sciences of the United States of America 99, 419-424.

Urick, M.E., Rudd, M.L., Godwin, A.K., Sgroi, D., Merino, M., and Bell, D.W. (2011). PIK3R1 (p85alpha) is somatically mutated at high frequency in primary endometrial cancer. Cancer research 71, 4061-4067.

Ursini-Siegel, J., Hardy, W.R., Zuo, D., Lam, S.H., Sanguin-Gendreau, V., Cardiff, R.D., Pawson, T., and Muller, W.J. (2008). ShcA signalling is essential for tumour progression in mouse models of human breast cancer. The EMBO journal 27, 910-920.

Utermark, T., Rao, T., Cheng, H., Wang, Q., Lee, S.H., Wang, Z.C., Iglehart, J.D., Roberts, T.M., Muller, W.J., and Zhao, J.J. (2012). The p110alpha and p110beta isoforms of PI3K play divergent roles in mammary gland development and tumorigenesis. Genes & development 26, 1573-1586.

Utermark, T., Schaffhausen, B.S., Roberts, T.M., and Zhao, J.J. (2007). The p110alpha isoform of phosphatidylinositol 3-kinase is essential for polyomavirus middle T antigen-mediated transformation. Journal of virology 81, 7069-7076.

Vadas, O., Dbouk, H.A., Shymanets, A., Perisic, O., Burke, J.E., Abi Saab, W.F., Khalil, B.D., Harteneck, C., Bresnick, A.R., Nurnberg, B., et al. (2013). Molecular determinants of PI3Kgamma-mediated activation downstream of G-protein-coupled receptors (GPCRs). Proceedings of the National Academy of Sciences of the United States of America 110, 18862-18867.

Vanhaesebroeck, B., Guillermet-Guibert, J., Graupera, M., and Bilanges, B. (2010). The emerging mechanisms of isoform-specific PI3K signalling. Nature reviews. Molecular cell biology 11, 329-341.

Vanhaesebroeck, B., Jones, G.E., Allen, W.E., Zicha, D., Hooshmand-Rad, R., Sawyer, C., Wells, C., Waterfield, M.D., and Ridley, A.J. (1999). Distinct PI(3)Ks mediate mitogenic signalling and cell migration in macrophages. Nature cell biology 1, 69-71.

Vanhaesebroeck, B., and Khwaja, A. (2014). PI3Kdelta inhibition hits a sensitive spot in B cell malignancies. Cancer cell 25, 269-271.

Vanhaesebroeck, B., Welham, M.J., Kotani, K., Stein, R., Warne, P.H., Zvelebil, M.J., Higashi, K., Volinia, S., Downward, J., and Waterfield, M.D. (1997). P110delta, a novel phosphoinositide 3-kinase in leukocytes. Proceedings of the National Academy of Sciences of the United States of America 94, 4330-4335.

Velichkova, M., Juan, J., Kadandale, P., Jean, S., Ribeiro, I., Raman, V., Stefan, C., and Kiger, A.A. (2010). Drosophila Mtm and class II PI3K coregulate a PI(3)P pool with cortical and endolysosomal functions. The Journal of cell biology 190, 407-425.

Vogt, P.K. (2011). PI3K p110beta: more tightly controlled or constitutively active? Molecular cell 41, 499-501.

Page 206: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

189

Voigt, P., Dorner, M.B., and Schaefer, M. (2006). Characterization of p87PIKAP, a novel regulatory subunit of phosphoinositide 3-kinase gamma that is highly expressed in heart and interacts with PDE3B. The Journal of biological chemistry 281, 9977-9986.

Volinia, S., Dhand, R., Vanhaesebroeck, B., MacDougall, L.K., Stein, R., Zvelebil, M.J., Domin, J., Panaretou, C., and Waterfield, M.D. (1995). A human phosphatidylinositol 3-kinase complex related to the yeast Vps34p-Vps15p protein sorting system. The EMBO journal 14, 3339-3348.

Vora, S.R., Juric, D., Kim, N., Mino-Kenudson, M., Huynh, T., Costa, C., Lockerman, E.L., Pollack, S.F., Liu, M., Li, X., et al. (2014). CDK 4/6 Inhibitors Sensitize PIK3CA Mutant Breast Cancer to PI3K Inhibitors. Cancer cell 26, 136-149.

Wagner, K.U., Wall, R.J., St-Onge, L., Gruss, P., Wynshaw-Boris, A., Garrett, L., Li, M., Furth, P.A., and Hennighausen, L. (1997). Cre-mediated gene deletion in the mammary gland. Nucleic acids research 25, 4323-4330.

Wang, J., Hannon, G.J., and Beach, D.H. (2000). Risky immortalization by telomerase. Nature 405, 755-756.

Wang, Q., Von, T., Bronson, R., Ruan, M., Mu, W., Huang, A., Maira, S.M., and Zhao, J.J. (2013a). Spatially distinct roles of class Ia PI3K isoforms in the development and maintenance of PTEN hamartoma tumor syndrome. Genes & development 27, 1568-1580.

Wang, Q., Weisberg, E., and Zhao, J.J. (2013b). The gene dosage of class Ia PI3K dictates the development of PTEN hamartoma tumor syndrome. Cell cycle 12, 3589-3593.

Wang, S., Gao, J., Lei, Q., Rozengurt, N., Pritchard, C., Jiao, J., Thomas, G.V., Li, G., Roy-Burman, P., Nelson, P.S., et al. (2003). Prostate-specific deletion of the murine Pten tumor suppressor gene leads to metastatic prostate cancer. Cancer cell 4, 209-221.

Watson, C.J., and Khaled, W.T. (2008). Mammary development in the embryo and adult: a journey of morphogenesis and commitment. Development 135, 995-1003.

Wee, S., Wiederschain, D., Maira, S.M., Loo, A., Miller, C., deBeaumont, R., Stegmeier, F., Yao, Y.M., and Lengauer, C. (2008). PTEN-deficient cancers depend on PIK3CB. Proceedings of the National Academy of Sciences of the United States of America 105, 13057-13062.

Wei, Y., Zou, Z., Becker, N., Anderson, M., Sumpter, R., Xiao, G., Kinch, L., Koduru, P., Christudass, C.S., Veltri, R.W., et al. (2013). EGFR-mediated Beclin 1 phosphorylation in autophagy suppression, tumor progression, and tumor chemoresistance. Cell 154, 1269-1284.

Weigelt, B., Warne, P.H., Lambros, M.B., Reis-Filho, J.S., and Downward, J. (2013). PI3K pathway dependencies in endometrioid endometrial cancer cell lines. Clinical cancer research : an official journal of the American Association for Cancer Research 19, 3533-3544.

Page 207: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

190

Welch, H.C., Coadwell, W.J., Ellson, C.D., Ferguson, G.J., Andrews, S.R., Erdjument-Bromage, H., Tempst, P., Hawkins, P.T., and Stephens, L.R. (2002). P-Rex1, a PtdIns(3,4,5)P3- and Gbetagamma-regulated guanine-nucleotide exchange factor for Rac. Cell 108, 809-821.

Will, M., Qin, A.C., Toy, W., Yao, Z., Rodrik-Outmezguine, V., Schneider, C., Huang, X., Monian, P., Jiang, X., de Stanchina, E., et al. (2014). Rapid induction of apoptosis by PI3K inhibitors is dependent upon their transient inhibition of RAS-ERK signaling. Cancer discovery 4, 334-347.

Winnay, J.N., Boucher, J., Mori, M.A., Ueki, K., and Kahn, C.R. (2010). A regulatory subunit of phosphoinositide 3-kinase increases the nuclear accumulation of X-box-binding protein-1 to modulate the unfolded protein response. Nature medicine 16, 438-445.

Wong, K.K., Engelman, J.A., and Cantley, L.C. (2010). Targeting the PI3K signaling pathway in cancer. Current opinion in genetics & development 20, 87-90.

Wu, H., Shekar, S.C., Flinn, R.J., El-Sibai, M., Jaiswal, B.S., Sen, K.I., Janakiraman, V., Seshagiri, S., Gerfen, G.J., Girvin, M.E., et al. (2009). Regulation of Class IA PI 3-kinases: C2 domain-iSH2 domain contacts inhibit p85/p110alpha and are disrupted in oncogenic p85 mutants. Proceedings of the National Academy of Sciences of the United States of America 106, 20258-20263.

Wu, R., Baker, S.J., Hu, T.C., Norman, K.M., Fearon, E.R., and Cho, K.R. (2013). Type I to type II ovarian carcinoma progression: mutant Trp53 or Pik3ca confers a more aggressive tumor phenotype in a mouse model of ovarian cancer. The American journal of pathology 182, 1391-1399.

Yeung, W.W., and Wong, Y.H. (2010). Galpha16 interacts with Class IA phosphatidylinositol 3-kinases and inhibits Akt signaling. Cellular signalling 22, 1379-1387.

Yoshioka, K., Yoshida, K., Cui, H., Wakayama, T., Takuwa, N., Okamoto, Y., Du, W., Qi, X., Asanuma, K., Sugihara, K., et al. (2012). Endothelial PI3K-C2alpha, a class II PI3K, has an essential role in angiogenesis and vascular barrier function. Nature medicine 18, 1560-1569.

Yu, J., Wjasow, C., and Backer, J.M. (1998a). Regulation of the p85/p110alpha phosphatidylinositol 3'-kinase. Distinct roles for the n-terminal and c-terminal SH2 domains. The Journal of biological chemistry 273, 30199-30203.

Yu, J., Zhang, Y., McIlroy, J., Rordorf-Nikolic, T., Orr, G.A., and Backer, J.M. (1998b). Regulation of the p85/p110 phosphatidylinositol 3'-kinase: stabilization and inhibition of the p110alpha catalytic subunit by the p85 regulatory subunit. Molecular and cellular biology 18, 1379-1387.

Yuan, W., Stawiski, E., Janakiraman, V., Chan, E., Durinck, S., Edgar, K.A., Kljavin, N.M., Rivers, C.S., Gnad, F., Roose-Girma, M., et al. (2013). Conditional activation of Pik3ca(H1047R) in a knock-in mouse model promotes mammary tumorigenesis and emergence of mutations. Oncogene 32, 318-326.

Page 208: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

191

Yum, H.K., Arcaroli, J., Kupfner, J., Shenkar, R., Penninger, J.M., Sasaki, T., Yang, K.Y., Park, J.S., and Abraham, E. (2001). Involvement of phosphoinositide 3-kinases in neutrophil activation and the development of acute lung injury. Journal of immunology 167, 6601-6608.

Zhang, X., Vadas, O., Perisic, O., Anderson, K.E., Clark, J., Hawkins, P.T., Stephens, L.R., and Williams, R.L. (2011). Structure of lipid kinase p110beta/p85beta elucidates an unusual SH2-domain-mediated inhibitory mechanism. Molecular cell 41, 567-578.

Zhao, J.J., Cheng, H., Jia, S., Wang, L., Gjoerup, O.V., Mikami, A., and Roberts, T.M. (2006). The p110alpha isoform of PI3K is essential for proper growth factor signaling and oncogenic transformation. Proceedings of the National Academy of Sciences of the United States of America 103, 16296-16300.

Zhao, J.J., Gjoerup, O.V., Subramanian, R.R., Cheng, Y., Chen, W., Roberts, T.M., and Hahn, W.C. (2003). Human mammary epithelial cell transformation through the activation of phosphatidylinositol 3-kinase. Cancer cell 3, 483-495.

Zhao, J.J., Liu, Z., Wang, L., Shin, E., Loda, M.F., and Roberts, T.M. (2005). The oncogenic properties of mutant p110alpha and p110beta phosphatidylinositol 3-kinases in human mammary epithelial cells. Proceedings of the National Academy of Sciences of the United States of America 102, 18443-18448.

Zhao, L., and Vogt, P.K. (2008). Helical domain and kinase domain mutations in p110alpha of phosphatidylinositol 3-kinase induce gain of function by different mechanisms. Proceedings of the National Academy of Sciences of the United States of America 105, 2652-2657.

Zhao, L., and Vogt, P.K. (2010). Hot-spot mutations in p110alpha of phosphatidylinositol 3-kinase (pI3K): differential interactions with the regulatory subunit p85 and with RAS. Cell cycle 9, 596-600.

Zheng, Y., Yin, L., Chen, H., Yang, S., Pan, C., Lu, S., Miao, M., and Jiao, B. (2012). miR-376a suppresses proliferation and induces apoptosis in hepatocellular carcinoma. FEBS letters 586, 2396-2403.

Page 209: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

192

Appendix A: Supplemental table of class I PI3K alterations in cancer, with complete references

Alteration Type Cancer Type Frequency of

Alteration Sample Size Range

References

Class IA

PIK3CA (p110g) Mutation Endometrial 10.3-53.0% 29-232 1, 2 Breast 7.1-35.5% 65-507 3-9 Ovarian (CC) 33.0% 97 10 Colorectal 16.9

†-30.6%

72-195 11, 12

Bladder 5.0-20.0% 20-130 13-16 Lung (SCC) 20.0% 5 17 Lung (SQCC) 2.9-16.8% 35-178 17, 18 Lung (LCC) 11.9% 9 17 Lung (ADC) 0.6-4.3% 57-183 3, 17, 19, 20 Cervical 13.6% 22 2 Glioblastoma 4.3-11.0% 91-291 21-24 Head and neck 8.1-9.4%

32-74 25, 26

Esophageal 5.5%

145 27 Melanoma 5.0%

121

28

Prostate 1.3-3.6%

55-156 3, 29-31 Sarcoma 2.9%

207 32

Renal (CC) 1.0-2.9%

98-417 33, 34 Liver (HCC) 1.6% 125 35 Megalencephaly

‡ 48.0% 50 36

Copy number Head and neck 9.1-100% 11-117 37-39 gain/amplification Cervical 9.1-76.4% 22-55 2, 40 Lung (SQCC) 42.9-69.6% 28-52 17, 41, 42 Lung (SCC) 33.3-66.7% 3-12 17, 41 Lung (LCC) 16.7-37.5% 6-16 17, 41 Lung (ADC) 9.5-19.1% 47-74 17, 41 Lung (NSCLC) 12.0% 92 43 Lymphoma (MCL) 68.2% 22 44 Lymphoma (DLBCL) 16.7% 60 45 Ovarian 39.8% 93 46 Ovarian (Serous) 13.3-24.3% 60-74 47, 48 Gastric 36.4% 55 49 Thyroid 30.0% 110 50 Prostate 28.1% 32 16 Breast 8.7-13.4% 92-209 8, 9 Glioblastoma 1.9-12.2% 139-206 21, 22 Endometrial 10.3% 29 2 Thyroid 9.4% 128 51 Esophageal 5.7% 87 52 Leukemia (CLL) 5.6% 161 53

Increased expression Prostate 40.0% 25 16

Page 210: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

193

Appendix A: Supplemental table of class I PI3K alterations in cancer, with complete references (continued)

Alteration Type Cancer Type Frequency of

Alteration Sample Size Range

References

Class IA

PIK3CB (p110く) Mutation Breast 0.5% 183 3, 54

Copy number Lung (SQCC) 56.5% 46 42 gain/amplification Thyroid 42.3% 97 50 Ovarian 5-26.9% NA-93 46, 55

Lymphoma (DLBCL) 20.0% 60 45

Glioblastoma 5.8% 103 56 Breast 4.9-5% NA-81

55, 57

Increased expression Prostate 46.7% 30 58 Glioblastoma 3.9% 103 56

PIK3CD (p110h) Copy number gain Glioblastoma 40.0% 10 59

Increased expression Neuroblastoma 52.6% 19 60 Glioblastoma 5.8% 103 56

PIK3R1 (p85g, p55g, p50g) Mutation Endometrial 19.8-32.8% 108-243 1, 61, 62 Pancreatic 16.7% 6 63 Glioblastoma 7.6-11.3% 91-291 22-24 Colorectal 4.6

†-8.3% 108-195 11, 63

Melanoma 4.4% 68 64 Ovarian 3.8% 80 65 Esophageal 3.4%

145 27

Breast 1.1-2.8% 62-507 3, 4, 63, 66 Colon 1.7% 60 65

Decreased expression Breast 61.8% 458 66 Prostate 17-75%* NA 67 Lung 19-46%* NA 67 Ovarian 22%* NA 67 Breast 18%* NA 67 Bladder 18%* NA 67

Copy number loss Ovarian 21.5% 93 46

PIK3R2 (p85く) Mutation Endometrial 4.9% 243 61 Colorectal 0.9% 108 63 Megalencephaly

‡ 22.0% 50 36

Amplification Lymphoma (DLBCL) 23.3% 60 45

Increased expression Colon 55.0% 20 68 Breast 45.7% 35 68

PIK3R3 (p55け) Copy number gain Ovarian 15.0% 93 46

Page 211: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

194

Appendix A: Supplemental table of class I PI3K alterations in cancer, with complete references (continued)

Alteration Type Cancer Type Frequency of

Alteration Sample Size Range

References

Class IB

PIK3CG (p110け) Copy number gain Ovarian 19.3% 93 46

Increased expression Breast 77.5% 40 69

Prostate 72.4% 29 70

Medulloblastoma 52.9% 17 71

PIK3R5 (p101)

Mutation Melanoma 38.2% 68 64 Gastric 2.7% 37 63

CC, clear cell; SCC, small cell carcinoma; SQCC, squamous cell carcinoma; ADC, adenocarcinoma; LCC, large cell carcinoma; NSCLC, non-small cell lung carcinoma; MCL, mantle cell lymphoma; CLL, chronic lymphocytic leukemia; DLBCL, diffuse large B cell lymphoma; HCC, hepatocellular carcinoma. ‡ Megalencephaly syndromes are a collection of sporadic overgrowth disorders characterized by enlarged

brain size and other distinct features. † Combined number of hypermutated and non-hypermutated colon and colorectal patient samples with

mutations in the indicated gene. * Represents the percent reduction in gene expression. NA Sample size not available for this study.

Page 212: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

195

Appendix A references:

1. Cancer Genome Atlas Research, N. et al. Integrated genomic characterization of endometrial carcinoma. Nature 497, 67-73 (2013).

2. Miyake, T. et al. PIK3CA gene mutations and amplifications in uterine cancers, identified by methods that avoid confounding by PIK3CA pseudogene sequences. Cancer Lett 261, 120-6 (2008).

3. Kan, Z. et al. Diverse somatic mutation patterns and pathway alterations in human cancers. Nature 466, 869-73 (2010).

4. Cancer Genome Atlas, N. Comprehensive molecular portraits of human breast tumours. Nature 490, 61-70 (2012).

5. Shah, S.P. et al. The clonal and mutational evolution spectrum of primary triple-negative breast cancers. Nature 486, 395-9 (2012).

6. Banerji, S. et al. Sequence analysis of mutations and translocations across breast cancer subtypes. Nature 486, 405-9 (2012).

7. Stephens, P.J. et al. The landscape of cancer genes and mutational processes in breast cancer. Nature 486, 400-4 (2012).

8. Lopez-Knowles, E. et al. PI3K pathway activation in breast cancer is associated with the basal-like phenotype and cancer-specific mortality. Int J Cancer 126, 1121-31 (2010).

9. Wu, G. et al. Somatic mutation and gain of copy number of PIK3CA in human breast cancer. Breast Cancer Res 7, R609-16 (2005).

10. Kuo, K.T. et al. Frequent activating mutations of PIK3CA in ovarian clear cell carcinoma. Am J Pathol 174, 1597-601 (2009).

11. Cancer Genome Atlas, N. Comprehensive molecular characterization of human colon and rectal cancer. Nature 487, 330-7 (2012).

12. Seshagiri, S. et al. Recurrent R-spondin fusions in colon cancer. Nature 488, 660-4 (2012).

13. The Cancer Genome Atlas Research, N. Comprehensive molecular characterization of urothelial bladder carcinoma. Nature (2014).

14. Iyer, G. et al. Prevalence and co-occurrence of actionable genomic alterations in high-grade bladder cancer. J Clin Oncol 31, 3133-40 (2013).

15. Guo, G. et al. Whole-genome and whole-exome sequencing of bladder cancer identifies frequent alterations in genes involved in sister chromatid cohesion and segregation. Nat Genet 45, 1459-63 (2013).

16. Agell, L. et al. PI3K signaling pathway is activated by PIK3CA mRNA overexpression and copy gain in prostate tumors, but PIK3CA, BRAF, KRAS and AKT1 mutations are infrequent events. Mod Pathol 24, 443-52 (2011).

Page 213: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

196

17. Okudela, K. et al. PIK3CA mutation and amplification in human lung cancer. Pathol Int 57, 664-71 (2007).

18. Cancer Genome Atlas Research, N. Comprehensive genomic characterization of squamous cell lung cancers. Nature 489, 519-25 (2012).

19. Ding, L. et al. Somatic mutations affect key pathways in lung adenocarcinoma. Nature 455, 1069-75 (2008).

20. Imielinski, M. et al. Mapping the hallmarks of lung adenocarcinoma with massively parallel sequencing. Cell 150, 1107-20 (2012).

21. Kita, D., Yonekawa, Y., Weller, M. & Ohgaki, H. PIK3CA alterations in primary (de novo) and secondary glioblastomas. Acta Neuropathol 113, 295-302 (2007).

22. Cancer Genome Atlas Research, N. Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 455, 1061-8 (2008).

23. Parsons, D.W. et al. An integrated genomic analysis of human glioblastoma multiforme. Science 321, 1807-12 (2008).

24. Brennan, C.W. et al. The somatic genomic landscape of glioblastoma. Cell 155, 462-77 (2013).

25. Stransky, N. et al. The mutational landscape of head and neck squamous cell carcinoma. Science 333, 1157-60 (2011).

26. Agrawal, N. et al. Exome sequencing of head and neck squamous cell carcinoma reveals inactivating mutations in NOTCH1. Science 333, 1154-7 (2011).

27. Dulak, A.M. et al. Exome and whole-genome sequencing of esophageal adenocarcinoma identifies recurrent driver events and mutational complexity. Nat Genet 45, 478-86 (2013).

28. Hodis, E. et al. A landscape of driver mutations in melanoma. Cell 150, 251-63 (2012).

29. Taylor, B.S. et al. Integrative genomic profiling of human prostate cancer. Cancer Cell 18, 11-22 (2010).

30. Baca, S.C. et al. Punctuated evolution of prostate cancer genomes. Cell 153, 666-77 (2013).

31. Barbieri, C.E. et al. Exome sequencing identifies recurrent SPOP, FOXA1 and MED12 mutations in prostate cancer. Nat Genet 44, 685-9 (2012).

32. Barretina, J. et al. Subtype-specific genomic alterations define new targets for soft-tissue sarcoma therapy. Nat Genet 42, 715-21 (2010).

33. Cancer Genome Atlas Research, N. Comprehensive molecular characterization of clear cell renal cell carcinoma. Nature 499, 43-9 (2013).

34. Guo, G. et al. Frequent mutations of genes encoding ubiquitin-mediated proteolysis pathway components in clear cell renal cell carcinoma. Nat Genet 44, 17-9 (2012).

Page 214: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

197

35. Guichard, C. et al. Integrated analysis of somatic mutations and focal copy-number changes identifies key genes and pathways in hepatocellular carcinoma. Nat Genet 44, 694-8 (2012).

36. Riviere, J.B. et al. De novo germline and postzygotic mutations in AKT3, PIK3R2 and PIK3CA cause a spectrum of related megalencephaly syndromes. Nat Genet 44, 934-40 (2012).

37. Fenic, I., Steger, K., Gruber, C., Arens, C. & Woenckhaus, J. Analysis of PIK3CA and Akt/protein kinase B in head and neck squamous cell carcinoma. Oncol Rep 18, 253-9 (2007).

38. Pedrero, J.M. et al. Frequent genetic and biochemical alterations of the PI 3-K/AKT/PTEN pathway in head and neck squamous cell carcinoma. Int J Cancer 114, 242-8 (2005).

39. Woenckhaus, J. et al. Genomic gain of PIK3CA and increased expression of p110alpha are associated with progression of dysplasia into invasive squamous cell carcinoma. J Pathol 198, 335-42 (2002).

40. Ma, Y.Y. et al. PIK3CA as an oncogene in cervical cancer. Oncogene 19, 2739-44 (2000).

41. Massion, P.P. et al. Early involvement of the phosphatidylinositol 3-kinase/Akt pathway in lung cancer progression. Am J Respir Crit Care Med 170, 1088-94 (2004).

42. Massion, P.P. et al. Genomic copy number analysis of non-small cell lung cancer using array comparative genomic hybridization: implications of the phosphatidylinositol 3-kinase pathway. Cancer Res 62, 3636-40 (2002).

43. Kawano, O. et al. PIK3CA gene amplification in Japanese non-small cell lung cancer. Lung Cancer 58, 159-60 (2007).

44. Psyrri, A. et al. Phosphatidylinositol 3'-kinase catalytic subunit alpha gene amplification contributes to the pathogenesis of mantle cell lymphoma. Clin Cancer Res 15, 5724-32 (2009).

45. Cui, W. et al. Frequent copy number variations of PI3K/AKT pathway and aberrant protein expressions of PI3K subunits are associated with inferior survival in diffuse large B cell lymphoma. J Transl Med 12, 10 (2014).

46. Huang, J. et al. Frequent genetic abnormalities of the PI3K/AKT pathway in primary ovarian cancer predict patient outcome. Genes Chromosomes Cancer 50, 606-18 (2011).

47. Nakayama, K. et al. Sequence mutations and amplification of PIK3CA and AKT2 genes in purified ovarian serous neoplasms. Cancer Biol Ther 5, 779-85 (2006).

48. Nakayama, K. et al. Amplicon profiles in ovarian serous carcinomas. Int J Cancer 120, 2613-7 (2007).

49. Byun, D.S. et al. Frequent monoallelic deletion of PTEN and its reciprocal associatioin with PIK3CA amplification in gastric carcinoma. Int J Cancer 104, 318-27 (2003).

Page 215: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

198

50. Liu, Z. et al. Highly prevalent genetic alterations in receptor tyrosine kinases and phosphatidylinositol 3-kinase/akt and mitogen-activated protein kinase pathways in anaplastic and follicular thyroid cancers. J Clin Endocrinol Metab 93, 3106-16 (2008).

51. Wu, G. et al. Uncommon mutation, but common amplifications, of the PIK3CA gene in thyroid tumors. J Clin Endocrinol Metab 90, 4688-93 (2005).

52. Miller, C.T. et al. Gene amplification in esophageal adenocarcinomas and Barrett's with high-grade dysplasia. Clin Cancer Res 9, 4819-25 (2003).

53. Brown, J.R. et al. Integrative genomic analysis implicates gain of PIK3CA at 3q26 and MYC at 8q24 in chronic lymphocytic leukemia. Clin Cancer Res 18, 3791-802 (2012).

54. Dbouk, H.A. et al. Characterization of a tumor-associated activating mutation of the p110beta PI 3-kinase. PLoS One 8, e63833 (2013).

55. Brugge, J., Hung, M.C. & Mills, G.B. A new mutational AKTivation in the PI3K pathway. Cancer Cell 12, 104-7 (2007).

56. Knobbe, C.B. & Reifenberger, G. Genetic alterations and aberrant expression of genes related to the phosphatidyl-inositol-3'-kinase/protein kinase B (Akt) signal transduction pathway in glioblastomas. Brain Pathol 13, 507-18 (2003).

57. Crowder, R.J. et al. PIK3CA and PIK3CB inhibition produce synthetic lethality when combined with estrogen deprivation in estrogen receptor-positive breast cancer. Cancer Res 69, 3955-62 (2009).

58. Zhu, Q. et al. Phosphoinositide 3-OH kinase p85alpha and p110beta are essential for androgen receptor transactivation and tumor progression in prostate cancers. Oncogene 27, 4569-79 (2008).

59. Mizoguchi, M., Nutt, C.L., Mohapatra, G. & Louis, D.N. Genetic alterations of phosphoinositide 3-kinase subunit genes in human glioblastomas. Brain Pathol 14, 372-7 (2004).

60. Boller, D. et al. Targeting the phosphoinositide 3-kinase isoform p110delta impairs growth and survival in neuroblastoma cells. Clin Cancer Res 14, 1172-81 (2008).

61. Cheung, L.W. et al. High frequency of PIK3R1 and PIK3R2 mutations in endometrial cancer elucidates a novel mechanism for regulation of PTEN protein stability. Cancer Discov 1, 170-85 (2011).

62. Urick, M.E. et al. PIK3R1 (p85alpha) is somatically mutated at high frequency in primary endometrial cancer. Cancer Res 71, 4061-7 (2011).

63. Jaiswal, B.S. et al. Somatic mutations in p85alpha promote tumorigenesis through class IA PI3K activation. Cancer Cell 16, 463-74 (2009).

64. Shull, A.Y. et al. Novel somatic mutations to PI3K pathway genes in metastatic melanoma. PLoS One 7, e43369 (2012).

65. Philp, A.J. et al. The phosphatidylinositol 3'-kinase p85alpha gene is an oncogene in human ovarian and colon tumors. Cancer Res 61, 7426-9 (2001).

Page 216: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

199

66. Cizkova, M. et al. PIK3R1 underexpression is an independent prognostic marker in breast cancer. BMC Cancer 13, 545 (2013).

67. Taniguchi, C.M. et al. The phosphoinositide 3-kinase regulatory subunit p85alpha can exert tumor suppressor properties through negative regulation of growth factor signaling. Cancer Res 70, 5305-15 (2010).

68. Cortes, I. et al. p85beta phosphoinositide 3-kinase subunit regulates tumor progression. Proc Natl Acad Sci U S A 109, 11318-23 (2012).

69. Xie, Y. et al. Identification of upregulated phosphoinositide 3-kinase gamma as a target to suppress breast cancer cell migration and invasion. Biochem Pharmacol 85, 1454-62 (2013).

70. Edling, C.E. et al. Key role of phosphoinositide 3-kinase class IB in pancreatic cancer. Clin Cancer Res 16, 4928-37 (2010).

71. Guerreiro, A.S. et al. A sensitized RNA interference screen identifies a novel role for the PI3K p110gamma isoform in medulloblastoma cell proliferation and chemoresistance. Mol Cancer Res 9, 925-35 (2011).

Page 217: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

200

Appendix B: Supplemental table of genetically engineered mouse models of PI3K isoforms in cancer, with complete references.

Genotype Phenotype Ref

PIK3CA (p110g)

KRasLA2

; Pik3caRBD/RBD Protected from KRas-induced lung

tumors 1

Rosa26-Cre; KRasLA2

; Pik3caRBD/flox

Partial regression of KRas-induced lung tumors

2

MMTV-Neu-IRES-Cre; Pik3caflox/flox

Protected from Her2/neu-driven mammary tumors

3

Pb-Cre; Ptenflox/flox

; Pik3caflox/flox No effect on high-grade PIN driven by

Pten loss 4

Mx1-Cre; KRasG12D

; Pik3caflox/flox Protection from MPN induced by

oncogenic KRas 5

Mx1-Cre, LSL-Shp2GOF/+

; Pik3caflox/flox No effect on MPN induced by Shp2

GOF 6

Pten−/+

; Pik3caKD/+

Increased endometrial hyperplasia; reduced pheochromocytoma and thyroid tumors

7

CCSP-rtTA; Tet-op-PIK3CAH1047R

Develop lung tumors within 3 months 8

MMTV-rtTA; tetO-PIK3CAH1047R

Develop mammary tumors within 7 months

9

MMTV-Cre; LSL-PIK3CAH1047R Surviving mice develop mammary

tumors within 7 months 10

MMTV-Cre; Pik3cae20H1047R/+ Develop mammary tumors within 16

months 11

WAP-Cre; LSL-PIK3CAH1047R Develop mammary tumors within 36

days post-partum 10

WAP-Cre; LSL-PIK3CAE545K

Develop mammary tumors within 80 days post-partum

12

MMTV-Cre; p53flox/+

; Rosa26-Pik3caH1047R

Develop mammary tumors within 5 months

13

MMTV-rtTA; tetO-Cre; ErbB3flox/flox

; tetO-PIK3CAH1047R

Delayed mammary hyperplasia but no effect on mammary tumor formation driven by PIK3CA

H1047R

14

MMTV-rtTA; MMTV-Her2; tetO-PIK3CAH1047R

Accelerated mammary tumor formation and increased lung metastasis compared to Her2 or PIK3CA

H1047R alone

15

Ptenflox/flox

; Pik3caLat-H1047R/+ Develop ovarian tumors within 16

weeks 16

Gpa33-CrePR2; APCLOF/LOF

; Pik3caLat-H1047R/+

Accelerated development of intestinal tumors compared to Pik3ca

H1047R or

APCLOF

alone 17

Fabp1-Cre; ApcMin/+

; Rosa26-Pik3ca*

Increased number and size of intestinal tumors compared to Pik3ca* or Apc

Min/+ alone

18

Fabp1-Cre; Apcflox/+

; Rosa26-Pik3ca*

Increased number and size of intestinal tumors compared to Pik3ca* or Apc

flox/+ alone

18

Page 218: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

201

Appendix B: Supplemental table of genetically engineered mouse models of PI3K isoforms in cancer, with complete references (continued).

Genotype Phenotype Ref

PIK3CB (p110く)

MMTV-Her2/neuT; Pik3cbKD/KD

Reduced number of mammary tumors driven by Her2/neuT

19

Pb-Cre; Ptenflox/flox

; Pik3cbflox/flox Protection from high-grade PIN driven

by Pten loss 4

Pten−/+

; Pik3cbKD/+ Reduced PIN and prostate cancer

driven by Pten loss 7

(ARR)2PB-Pik3cbCA Develop VP PIN by 10 weeks and

DLP PIN by 60 weeks 20

MMTV-Neu-IRES-Cre; Pik3cbflox/flox

Accelerated mammary tumor formation and increased tumor burden driven by Her2/neu

3

PIK3CA (p110g) and PIK3CB (p110く)

K14-Cre; Ptenflox/flox

; Pik3caflox/flox

; Pik3cbflox/flox

Loss of ¾ alleles of Pik3ca and Pik3cb blocks skin lesions and mammary hyperplasia driven by Pten loss

21, 22

PIK3CD (p110h)

Pik3cdKD

Reduced trafficking of NK cells; reduced NK cell extravasation to tumor cells

23

Mx1-Cre, LSL-Shp2GOF/+

;Pik3cdKD/KD

Reduced MPN induced by Shp2 GOF 6

Lck-Cre; Ptenflox/flox

; Pik3cd−/−

No effect on development of T-ALL driven by Pten loss

24

PIK3CG (p110け)

Lck-Cre; Ptenflox/flox

; Pik3cg−/−

No effect on development of T-ALL driven by Pten loss

24

PIK3CD (p110h) and PIK3CG (p110け)

Lck-Cre; Ptenflox/flox

; Pik3cd−/−

; Pik3cg−/−

Delayed development of T-ALL driven by Pten loss

24

PIK3R1 (p85g, p55g, p50g)

CD19-Cre; Pik3r1flox/flox

Reduced B-cell leukemia development driven by ex vivo infection with BCR-ABL

25

Albumin-Cre; Pik3r1flox/flox

Develop liver tumors within 20 months 26

Pten−/+

; Pik3r1−/+ Increased intestinal polyps but no

change in PIN driven by Pten loss 27

PIK3R2 (p85く)

Pik3r2−/− Decreased number of colon tumors

induced by AOM/DSS 28

Pten−/+

; Pik3r2−/− No change in intestinal polyps or PIN

driven by Pten loss 27

CD19-Cre; Pik3r2−/−

No effect on B-cell leukemia development driven by ex vivo infection with BCR-ABL

25

Page 219: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

202

Appendix B: Supplemental table of genetically engineered mouse models of PI3K isoforms in cancer, with complete references (continued).

Genotype Phenotype Ref

PIK3R1 (p85g, p55g, p50g) and PIK3R2 (p85く) CCSP-rtTA; tetO-KRas

G12D; Pik3r1

flox/flox; Pik3r2

−/−

LSL-KRasG12D

; Pik3r1flox/flox

; Pik3r2−/−

Decreased incidence of lung tumors driven by KRas

8

CD19-Cre; Pik3r1flox/flox

; Pik3r2−/−

Blocked B-cell leukemia development driven by ex vivo infection with BCR-ABL

25

CCSP-rtTA; tetO-KRasG12D

; Pik3r1flox/+

; Pik3r2−/−

LSL-KRasG12D

; Pik3r1flox/+

; Pik3r2−/−

Increased incidence of lung tumors driven by KRas

8

PIK3C2A (PI3K-Cβg)

Cdh5(PAC)-CreERT2

; Pik3c2aflox/flox Decreased microvessel density and

tumor burden of implanted tumors 29

RBD, Ras binding domain mutant; KD, kinase dead mutant; CA, constitutively active; Tg, transgene; PIN, prostate intraepithelial neoplasia; AOM/DSS, azoxymethane/dextran sodium sulfate; LOF, loss of function; GOF, gain of function; VP, ventral prostate; DLP, dorsal/lateral prostate; MPN, myoproliferative neoplasia.

Page 220: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

203

Appendix B references:

1. Gupta, S. et al. Binding of ras to phosphoinositide 3-kinase p110alpha is required for ras-driven tumorigenesis in mice. Cell 129, 957-68 (2007).

2. Castellano, E. et al. Requirement for interaction of PI3-kinase p110alpha with RAS in lung tumor maintenance. Cancer Cell 24, 617-30 (2013).

3. Utermark, T. et al. The p110alpha and p110beta isoforms of PI3K play divergent roles in mammary gland development and tumorigenesis. Genes Dev 26, 1573-86 (2012).

4. Jia, S. et al. Essential roles of PI(3)K-p110beta in cell growth, metabolism and tumorigenesis. Nature 454, 776-9 (2008).

5. Gritsman, K. et al. Hematopoiesis and RAS-driven myeloid leukemia differentially require PI3K isoform p110alpha. J Clin Invest 124, 1794-809 (2014).

6. Goodwin, C.B. et al. PI3K p110delta uniquely promotes gain-of-function Shp2-induced GM-CSF hypersensitivity in a model of JMML. Blood 123, 2838-42 (2014).

7. Berenjeno, I.M. et al. Both p110alpha and p110beta isoforms of PI3K can modulate the impact of loss-of-function of the PTEN tumour suppressor. Biochem J 442, 151-9 (2012).

8. Engelman, J.A. et al. Effective use of PI3K and MEK inhibitors to treat mutant Kras G12D and PIK3CA H1047R murine lung cancers. Nat Med 14, 1351-6 (2008).

9. Liu, P. et al. Oncogenic PIK3CA-driven mammary tumors frequently recur via PI3K pathway-dependent and PI3K pathway-independent mechanisms. Nat Med 17, 1116-20 (2011).

10. Meyer, D.S. et al. Luminal expression of PIK3CA mutant H1047R in the mammary gland induces heterogeneous tumors. Cancer Res 71, 4344-51 (2011).

11. Yuan, W. et al. Conditional activation of Pik3ca(H1047R) in a knock-in mouse model promotes mammary tumorigenesis and emergence of mutations. Oncogene 32, 318-26 (2013).

12. Meyer, D.S. et al. Expression of PIK3CA mutant E545K in the mammary gland induces heterogeneous tumors but is less potent than mutant H1047R. Oncogenesis 2, e74 (2013).

13. Adams, J.R. et al. Cooperation between Pik3ca and p53 mutations in mouse mammary tumor formation. Cancer Res 71, 2706-17 (2011).

14. Young, C.D. et al. Conditional loss of ErbB3 delays mammary gland hyperplasia induced by mutant PIK3CA without affecting mammary tumor latency, gene expression, or signaling. Cancer Res 73, 4075-85 (2013).

15. Hanker, A.B. et al. Mutant PIK3CA accelerates HER2-driven transgenic mammary tumors and induces resistance to combinations of anti-HER2 therapies. Proc Natl Acad Sci U S A 110, 14372-7 (2013).

16. Kinross, K.M. et al. An activating Pik3ca mutation coupled with Pten loss is sufficient to initiate ovarian tumorigenesis in mice. J Clin Invest 122, 553-7 (2012).

Page 221: PI3K regulatory subunit p85alpha plays a tumor suppressive role in the transformation of mammary

204

17. Hare, L.M. et al. Physiological expression of the PI3K-activating mutation Pik3ca(H1047R) combines with Apc loss to promote development of invasive intestinal adenocarcinomas in mice. Biochem J 458, 251-8 (2014).

18. Deming, D.A. et al. PIK3CA and APC mutations are synergistic in the development of intestinal cancers. Oncogene 33, 2245-54 (2014).

19. Ciraolo, E. et al. Phosphoinositide 3-kinase p110beta activity: key role in metabolism and mammary gland cancer but not development. Sci Signal 1, ra3 (2008).

20. Lee, S.H. et al. A constitutively activated form of the p110beta isoform of PI3-kinase induces prostatic intraepithelial neoplasia in mice. Proc Natl Acad Sci U S A 107, 11002-7 (2010).

21. Wang, Q. et al. Spatially distinct roles of class Ia PI3K isoforms in the development and maintenance of PTEN hamartoma tumor syndrome. Genes Dev 27, 1568-80 (2013).

22. Wang, Q., Weisberg, E. & Zhao, J.J. The gene dosage of class Ia PI3K dictates the development of PTEN hamartoma tumor syndrome. Cell Cycle 12, 3589-93 (2013).

23. Saudemont, A. et al. p110gamma and p110delta isoforms of phosphoinositide 3-kinase differentially regulate natural killer cell migration in health and disease. Proc Natl Acad Sci U S A 106, 5795-800 (2009).

24. Subramaniam, P.S. et al. Targeting nonclassical oncogenes for therapy in T-ALL. Cancer Cell 21, 459-72 (2012).

25. Kharas, M.G. et al. Ablation of PI3K blocks BCR-ABL leukemogenesis in mice, and a dual PI3K/mTOR inhibitor prevents expansion of human BCR-ABL+ leukemia cells. J Clin Invest 118, 3038-50 (2008).

26. Taniguchi, C.M. et al. The phosphoinositide 3-kinase regulatory subunit p85alpha can exert tumor suppressor properties through negative regulation of growth factor signaling. Cancer Res 70, 5305-15 (2010).

27. Luo, J. et al. Modulation of epithelial neoplasia and lymphoid hyperplasia in PTEN+/- mice by the p85 regulatory subunits of phosphoinositide 3-kinase. Proc Natl Acad Sci U S A 102, 10238-43 (2005).

28. Cortes, I. et al. p85beta phosphoinositide 3-kinase subunit regulates tumor progression. Proc Natl Acad Sci U S A 109, 11318-23 (2012).

29. Yoshioka, K. et al. Endothelial PI3K-C2alpha, a class II PI3K, has an essential role in angiogenesis and vascular barrier function. Nat Med 18, 1560-9 (2012).