the adaptor protein p66shc: roles in cell signaling ... · roles in cell signaling, metabolism and...

177
The Adaptor Protein p66SHC: Roles in Cell Signaling, Metabolism and Growth by Mohamed Ahmed Mohamed El-Badry Soliman A thesis submitted in conformity with the requirements for the Degree of Philosophy Graduate Department of Molecular Genetics University of Toronto © Copyright by Mohamed Soliman 2014

Upload: doanthu

Post on 16-Aug-2018

214 views

Category:

Documents


0 download

TRANSCRIPT

The Adaptor Protein p66SHC:

Roles in Cell Signaling, Metabolism and Growth

by

Mohamed Ahmed Mohamed El-Badry Soliman

A thesis submitted in conformity with the requirements

for the Degree of Philosophy

Graduate Department of Molecular Genetics

University of Toronto

© Copyright by Mohamed Soliman 2014

ii

The Adaptor Protein p66SHC: Roles in Cell Signaling,

Metabolism and Growth

Mohamed Soliman

Doctor of Philosophy

Department of Molecular Genetics

University of Toronto

2014

Abstract

Adaptor proteins link surface receptors to intracellular signaling pathways and control the way

cells respond to nutrient availability. Mice deficient in p66Shc, the most recently evolved

isoform of the Shc1 adaptor proteins and a mediator of receptor tyrosine kinase signaling,

display resistance to diabetes and obesity. Using quantitative mass spectrometry, I found that

p66Shc inhibited glucose metabolism. Depletion of p66Shc enhanced glycolysis and increased

the allocation of glucose-derived carbon into anabolic metabolism, characteristics of a metabolic

shift called the Warburg effect. This change in metabolism was mediated by the mammalian

target of rapamycin (mTOR) as inhibition of mTOR partly reversed the glycolytic phenotype

caused by p66Shc deficiency. Thus, unlike the other isoforms of Shc1, p66Shc appears to

antagonize insulin and mTOR signaling, which limits glucose uptake and metabolism. This study

identifies a critical inhibitory role for p66Shc in anabolic metabolism and insulin-mTOR

signaling.

iii

Happy is the one who has been able to learn the causes of things.

- Virgil, Georgics (II, 490)

iv

To the memory of my PhD advisor, Dr. Tony Pawson (1952-2013)

No longer mourn for me when I am dead

Then you shall hear the sullen surly bell

Give warning to the world that I am fled

- Shakespeare, Sonnet 71

v

Acknowledgments

Foremost is my gratitude to God for the shower of blessings throughout my PhD to complete my

research successfully. Especially, the blessings of health and determination were vital in

providing the physical and mental abilities to see this project through.

This work is dedicated to the memory of my PhD advisor, the late Dr. Tony Pawson, who passed

away months before the completion of this effort. I am grateful for the opportunity to train in one

of the best labs in the world in the field of cell signaling. I was inspired by his humbleness and

collegiality, and by his contagious enthusiasm about science. I am thankful for the exceptional

freedom and independence he gave me as a graduate student. I learned patience, perseverance,

and dedication; it was indeed a maturing journey.

I would like to express my deepest gratitude to my current PhD advisor, Dr. Jim Dennis, whose

expertise, understanding and knowledge added considerably to my graduate experience; to my

supervisory committee members, Drs. Charlie Boone, Anne-Claude Gingras, and Jane McGlade,

for the continuous advice they provided over the years at all levels of my research project; Drs.

Brendan Manning (Harvard University), Fritz Roth (University of Toronto) and Linda Penn

(University of Toronto) for taking the time out to serve as my examiners. My gratitude goes to

Dr. David Sabatini (MIT) for accommodating me in his lab as a visiting scientist to do the

mTOR experiments of my project.

vi

I have had the good fortune of keeping company with the members of the Pawson, Dennis and

Sabatini labs; they have provided me with novel perspectives on my project. Thank you all for

your advice, ideas, friendship and support during the certain and uncertain times in the lab. In

particular, I would like to thank Dr. Jerry Gish who has been an essential part of my education

and entertainment, Dr. Anas Abdel Rahman and Ms. Judy Pawling for all their effort with my

project, and Drs. Karen Colwill and Melissa Stacey for their insight and help over the years.

My gratitude goes to my research funding resouces: the University of Toronto, the Canadian

Institute of Heath Research, the Government of Ontario and the Vanier Canada Graduate

Scholarship program. Without their support, it would not have been possible to solely focus on

my research during my PhD tenure.

I am deeply indebted to my parents for always being supportive of my education. Although they

are so far away, they always provided the absolute encouragement that inspires me to follow my

career. Special thanks to my friends Drs. Hamza Jalal and Omer Yilmaz; your support and

understanding make you formidable friends.

Finally, my thanks go to each of those – unmentioned – who have made it possible for me to

reach this stage of my career, who shared my triumphs and frustration, who shaped my days and

continue to do so: I am grateful for having you in my life.

vii

List of Abbreviations ............................................................................................... x

List of Figures and Tables ................................................................................... xiii

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

1.1. Adaptor proteins in signal transduction ........................................................................... 2

1.1.1. Shc1: a prototype of adaptor proteins ....................................................................................4

1.1.2. p66Shc: an integrator of mitogenic and metabolic signaling ................................................9

1.1.3. Role of p66Shc in regulating oxidative stress .....................................................................10

1.1.4. p66Shc and energy metabolism ...........................................................................................12

1.2. Cell metabolism in normal and cancer cells................................................................... 13

1.2.1. The Warburg effect: glucose metabolism and anabolic demands of cell growth ................14

1.2.2. Glutamine: a metabolic fuel for proliferating cells ..............................................................17

1.3. Signaling Pathways and Regulation of Cellular Metabolism......................................... 18

1.3.1. Tyrosine kinase signaling and selective metabolic regulation in dividing cells ..................18

1.3.2. The PI3K-Akt-mTOR pathway ...........................................................................................20

1.3.3. Transcriptional regulators of anabolic metabolism .............................................................22

1.3.3.1. HIF .......................................................................................................................................22

1.3.3.2. Myc ......................................................................................................................................24

1.3.3.3. p53 .......................................................................................................................................24

1.3.4. Metabolic enzymes as oncogenes ........................................................................................25

1.4. mTOR: from signaling to metabolism ........................................................................... 28

1.4.1. Molecular components of mTOR ........................................................................................28

1.4.2. mTORC1 .............................................................................................................................31

1.4.2.1. Upstream regulators of mTORC1 ........................................................................................31

1.4.2.2. Downstream effectors of mTORC1 .....................................................................................35

1.4.3. mTORC2 .............................................................................................................................36

1.4.3.1. Upstream regulators of mTORC2 ........................................................................................37

1.4.3.2. Downstream effectors of mTORC2 .....................................................................................37

1.4.4. mTOR and metabolism ........................................................................................................38

1.4.4.1. Glucose metabolism ............................................................................................................38

1.4.4.2. Lipid synthesis .....................................................................................................................39

1.4.4.3. Protein synthesis ..................................................................................................................39

1.4.4.4. Nucleotide metabolism ........................................................................................................40

viii

1.5. Rationale and objectives of the study ............................................................................. 41

Chapter 2. Materials and Methods ....................................................................... 43

2.1. Cell culture and treatments ............................................................................................. 44

2.2. Freezing and thawing of cells......................................................................................... 44

2.3. Cell culture ..................................................................................................................... 45

2.4. Plasmid preparation and DNA constructs ...................................................................... 45

2.5. Cell transfection ............................................................................................................. 46

2.6. Retroviral production and cell infection......................................................................... 46

2.7. Metabolite extraction...................................................................................................... 47

2.8. Isotope labeling and kinetic profiling............................................................................. 48

2.9. [3H]-2-deoxy-D-glucose uptake assay ........................................................................... 48

2.10. Oxygen consumption rate measurement ........................................................................ 49

2.11. Cell size determination ................................................................................................... 49

2.12. Cell lysis and immunoblotting ....................................................................................... 49

2.13. Western blotting ............................................................................................................. 50

2.14. Mass spectrometry analysis of the p66Shc protein-interactions .................................... 51

2.15. RNA-seq ......................................................................................................................... 52

Chapter 3. p66Shc Inhibits Anabolic Metabolism .............................................. 53

3.1. Background .................................................................................................................... 54

3.2. Loss of p66Shc enhances glycolytic metabolism ........................................................... 55

3.3. Loss of p66Shc promotes glucose metabolism through the pentose phosphate and

hexosamine biosynthesis pathways ........................................................................................... 55

3.4. Restoring p66Shc expression inhibits glycolytic metabolism ....................................... 56

3.5. p66Shc expression inhibits amino acid biosynthesis and pyrimidine metabolism ........ 57

3.6. p66Shc regulates redox homeostasis .............................................................................. 58

3.7. p66Shc is necessary and sufficient to alter glucose uptake and metabolism ................. 59

3.8. Lack of p66Shc enhances glycolytic flux and anabolic metabolism ............................. 60

Chapter 4. p66Shc Inhibits Signaling to The Metabolic Sensor mTOR ........... 83

4.1. Background .................................................................................................................... 84

4.2. p66Shc inhibits growth factor signaling to the metabolic sensor mTOR ...................... 85

4.3. p66Shc expression causes a decrease in cell size ........................................................... 86

ix

4.4. Effects of p66Shc on glycolytic metabolism are mediated through mTOR .................. 86

Chapter 5. Discussion and Future Directions .................................................... 101

5.1. p66Shc signaling to mTOR: an open question ............................................................. 104

5.2. Competition between Shc1 isoforms in regulating PI3K-mTOR signaling ................. 106

5.3. Regulation of receptor and glucose transporter glycosylation by p66Shc ................... 107

5.4. p66Shc and fatty acid signaling to mTOR ................................................................... 108

5.5. Genes regulated through p66Shc expression ............................................................... 109

5.6. Negative regulation of insulin signaling by adaptor proteins ...................................... 110

5.7. Summary ...................................................................................................................... 111

Chapter 6. Appendix ............................................................................................ 113

6.1. LC-MS/MS transitions for the metabolites measured in this study ............................. 114

6.2. LC-MS/MS transitions for 1,2-13

C2 Glucose intermediates. ........................................ 121

6.3. List of gene differentially regulated by p66Shc expression ......................................... 122

6.4. List of identified p66Shc-interacting proteins .............................................................. 138

x

List of Abbreviations

1,3BPG 1,3-bisphosphoglycerate

2-DG 2-deoxy-D-glucose

2-HG 2-hydroxyglutarate

3PG 3-phosphoglycerate

4E-BP Eukaryotic translation initiation factor 4E-binding protein

ACoA Acetyl-CoA

AMPK AMP-activated protein kinase

Ang II Angiotensin II

CAD Carbamoyl-phosphate synthetase 2, aspartate transcarbamylase, and

dihydroorotase

CH1 Collagen homology 1

CH2 Collagen homology 2

CTP Cytidine triphosphate

Deptor DEP domain-containing mTOR-interacting protein

DHAP Dihydroxyacetone phosphate

E4P Erythrose-4-phosphate

EGFR Epidermal growth factor receptor

ErbB2 V-erb-b2 avian erythroblastic leukemia viral oncogene homolog 2

Erk Extracellular signal-regulated kinase

F1,6BP Fructose-1,6-bisphosphate

F6P Fructose-6-phosphate

Fes Feline sarcoma oncogene

FoxO Forkhead box O

G3P Glycerol-3-phosphate

G6P Glucose-6-phosphate

Gab Grb2-associated binding protein

GADP Glyceraldehyde-3-phosphate

GAP GTPase-activating protein

GAPDH Glyceraldehyde-3-phosphate dehydrogenase

xi

GlcNAcP N-acetylglucosamine-6-phosphate

GN6P Glucosamine-6-phosphate

Glut Glucose transporter

Grb2 Growth factor receptor-bound protein 2

GSH Glutathione (reduced)

GSK3β Glycogen synthase kinase 3β

GSSG Glutathione (oxidized)

HIF Hypoxia-inducible factor

IDH Isocitrate dehydrogenase

IGF-1 Insulin-like growth factor-1

IMP Inositol polyphosphate multikinase

IRS Insulin receptor substrate

KO Knockout

LC-MS/MS Liquid chromatography–tandem mass spectrometry

MAPK Mitogen-activated protein kinases

MCoA Malonyl-CoA

MEFs Murine embryonic fibroblasts

Mgat5 Mannosyl (α-1,6-)-Glycoprotein

β-1,6-N-Acetyl-Glucosaminyltransferase

mTORC1 Mammalian target of rapamycin complex 1

mTORC2 Mammalian target of rapamycin complex 2

NDRG N-myc downstream regulated gene

OAA Oxaloacetate

PEP Phosphoenolpyruvate

PI3K Phosphoinositide-3-kinase

PIP3 Phosphatidylinositol (3,4,5)-triphosphate

PK Pyruvate kinase

PKC Protein kinase C

PPP Pentose phosphate pathway

PRAS40 40 kDa pro-rich akt substrate

Protor Protein observed with rictor

xii

PTB Phospho-tyrosine binding

PTEN Phosphatase and tensin homolog

PTPN12 Protein tyrosine phosphatase, non-receptor type 12

R5P Ribose-5-phosphate

Rac1 Ras-related C3 botulinum toxin substrate 1

Raptor Regulatory-associated protein of mTOR

Ras Rat sarcoma oncogene

Rheb Ras homolog enriched in brain

Rictor Rapamycin-insensitive companion of mTOR

ROS Reactive oxygen species

RTK Receptor tyrosine kinase

SGK Serum- and glucocorticoid-induced protein kinase

SH2 Src homology 2

SH3 Src homology 3

SHC Src-homology collagen-containing protein

shRNA short hairpin RNA

SIN1 Stress-activated map-kinase-interacting protein 1

SOD Superoxide dismutase

Sos Son of sevenless homolog

SREBP Sterol regulatory element-binding protein

TCA cycle Tricarboxylic acid cycle

TSC1/2 Tuberous sclerosis 1 and 2

UDP-GlcNAc Uridine-diphosphate N-acetylglucosamine

UTP Uridine triphosphate

VHL Von Hippel–Lindau

X5P Xylulose-5-phosphate

xiii

List of Figures and Tables

Fig. 1.1 Schematic diagram of the Shc1 proteins. 6

Fig. 1.2 Role of Shc1 in signaling downstream of RTK. 7

Fig. 1.3 Cancer metabolism: the Warburg effect. 16

Fig. 1.4 mTOR signaling pathway. 30

Fig. 3.1 Effect of p66Shc on glycolytic metabolism. 61

Fig. 3.2 Lack of p66Shc enhances glycolytic metabolism. 62

Fig. 3.3 p66Shc deficiency increases the levels of the pentose phosphate

and the hexosamine pathway intermediates. 63

Fig. 3.4 Levels of Shc1 isoforms in p66Shc KO and p66+ MEFs. 64

Fig. 3.5 Unsupervised principal component analysis for targeted metabolomics

screen in p66Shc KO and p66+ MEFs. 65

Fig. 3.6 p66Shc expression decreases the levels of glycolytic intermediates. 66

Fig. 3.7 p66Shc inhibits fatty acid biosynthesis. 67

Fig. 3.8 p66Shc expression decreases the levels of the pentose phosphate

and the hexosamine pathway intermediates. 68

Fig. 3.9 Deficiency of p66Shc inhibits oxygen consumption rate and lowers

AMP/ATP ratio. 69

Fig. 3.10 p66Shc expression inhibits the synthesis of nonessential amino acids. 70

Fig. 3.11 Tracing of 15

N-labeled-amino acids in p66Shc KO and p66+ cells. 71

Fig. 3.12 p66Shc inhibits de novo pyrimidine synthesis intermediates. 72

Fig. 3.13 p66Shc regulates redox homeostasis. 73

xiv

Fig. 3.14 Lack of p66Shc enhances 2-DG uptake. 74

Fig. 3.15 p66Shc inhibits cellular lactate secretion. 75

Fig. 3.16 Abundance of Glut1 in p66Shc-deficient and p66Shc-competent cells. 76

Fig. 3.17 Isotope-tracing of 13

C-labeled glucose in p66Shc-deficient and

p66Shc-competent HeLa cells. 77

Fig. 3.18 Isotope-tracing of 13

C-labeled hexosamine pathway intermediates in

p66Shc-deficient and p66Shc-competent HeLa cells. 78

Fig. 3.19 Isotope-tracing of 13

C-labeled glycolytic intermediates in p66Shc KO

and p66+ MEFs. 79

Fig. 3.20 Isotope-tracing of 13C-labeled nonessential amino acids in p66Shc KO

and p66+ MEFs 80

Fig. 3.21 Flux analysis of 13

C-labeled glucose in p66Shc KO and p66+ MEFs. 81

Table 3.1 Fold change of the most significantly p66Shc-inhibited metabolites 82

Fig. 4.1 p66Shc inhibits mTORC1 and mTORC2 activation following serum

stimulation. 88

Fig. 4.2 p66Shc inhibits insulin signaling to mTOR. 89

Fig. 4.3 p66Shc inhibits mTORC1 activation in response to 90

amino acid stimulation.

Fig. 4.4 p66Shc expression inhibits mTOR activation in response to insulin 91

and IGF1, but not to EGF, stimulation.

Fig. 4.5 p66Shc expression inhibits mTOR activation in 92

response to serum stimulation.

Fig. 4.6 p66Shc expression inhibits mTOR activation in 93

response to amino acid stimulation.

Fig. 4.7 Stable expression of p66Shc, but not p52Shc, in p66Shc KO cells

inhibits the mTOR pathway. 94

Fig. 4.8 Stable expression of p66Shc, but not p52Shc, in p66Shc KO 95

cells inhibits glycolytic metabolism.

xv

Fig. 4.9 p66Shc expression decreases cell size. 96

Fig. 4.10 p66Shc mediates cell growth. 97

Fig. 4.11 Effect of rapamycin on mTOR signaling in p66Shc KO and

p66+ MEFs. 98

Fig. 4.12 mTOR mediates the effects of p66Shc on glycolytic

and pyrimidine metabolism. 99

Fig. 4.13 Effect of Akt inhibition on the abundance of glycolytic

metabolites in p66Shc-competent and p66Shc-deficient MEFs. 100

1

Chapter 1. Introduction

2

1.1. Adaptor proteins in signal transduction

Membrane receptors sample the extracellular environment. When activated by threshold levels of

cognate ligands, receptors stimulate a signaling cascade that leads to precise biological

responses. Signaling proteins contain catalytic and adaptor functions that typically reside in

discrete, independently folded domains. Most adaptor domains display binding specificity for

peptide motifs in other signaling proteins that transmit and codify the signals (Pawson, 2007).

For example, Src kinase has a canonical kinase fold as well as a Src homology 2 (SH2) domain

that bind to specific phosphotyrosine residues on activated receptors, and a Src homology 3

(SH3) domain that binds to polyproline (Pro-X-X-Pro) motifs. SH2 and SH3 domains are found

in many other proteins and function to recruit signaling proteins into complexes where catalytic

efficiency is greatly enhanced.

Analysis of the molecular evolution of genomes suggests that phenotypic diversity can

frequently be attributed to new combinations of existing protein domains, rather than from the

creation of completely new proteins. Gene duplication and shuffling of modular domains, also

results in the emergence of novel connectivity between existing proteins, expanding the

information flow through regulatory pathways and the complexity of responses. This could

explain the less-than-expected number of protein-coding genes and protein domains upon

sequencing of animal genomes more than a decade ago (Bhattacharyya et al., 2006). Thus

evolutionary innovation is achieved by domain duplication, recombination and adaptation of

domain affinities for novel binding partners (Jin and Pawson, 2012). As independent folds within

the same protein sequence, the catalytic and recruitment domains can evolve at different rates.

Disconnecting catalytic and recruitment domains into separate genes allows independent

3

regulation by transcription, splicing and translation, and thereby more diversity and evolutionary

possibilities for increased complexity in signaling networks (Bhattacharyya et al., 2006).

The terms adaptors and scaffolds are used interchangeably in the literature for non-catalytic

proteins that cross-link and promote the assembly of specific signaling complexes (Pawson and

Scott, 1997). Herein, we refer to these proteins as “adaptor proteins.” In mammals they include

the growth factor receptor-bound protein 2 (Grb2)/ Grb2-related adapter protein (Grap)/ Grb2-

related adaptor downstream of Shc (Gads), Grb7/10/14, SH2B/ adapter protein with pleckstrin

homology (APS)/ lymphocyte adaptor protein (Lnk), SH2D1-4, Shc1-4, SHB/SHD/SHE/SHF,

and the non catalytic region of tyrosine kinase (Nck)1/2 gene families, among others.

Differential expression or posttranslational modifications of an adaptor can determine whether a

pathway will function in a particular cell type. For example, the SH2-containing collagen-related

(Shc), alternative splice variants can have different interaction and pathway output, depending on

their levels of expression in various tissues. In addition, relocalization of adaptors to a specific

cellular compartment in a timely fashion is often a requirement of signal transduction fidelity

(Scott and Pawson, 2009). For example, Shc1 very rapidly recruits proteins associated with acute

stimulation of epidermal growth factor receptor (EGFR), such as Grb2-associated binding

protein 1/2 (Gab1/2) and Grb2-son of sevenless homolog 1 (Sos1) complex which promotes

exchange of Rat sarcoma oncogene (Ras)-bound GDP by GTP, while slowly recruits signaling

proteins that mediate negative regulation of EGFR signaling, such as protein tyrosine

phosphatase, non-receptor type 12 (PTPN12) (Zheng et al., 2013).

4

1.1.1. Shc1: a prototype of adaptor proteins

The Shc adaptor proteins were first identified by screening a human cDNA library for sequences

complementary to the SH2 domain of the feline sarcoma oncogene (c-fes) tyrosine kinase

(Pelicci et al., 1992). Following this initial screen, three sequence-related Shc-like transcripts and

proteins were identified (Luzi et al., 2000). The mammalian Shc gene family comprises four

members: ShcA, B, C and D. In addition, alternative splicing of ShcA and ShcC transcripts

results in multiple protein isoforms. While ShcA is expressed in almost all tissues, ShcB

(Sck/Sli/Shc2) and ShcC (Rai/N-Shc/Shc3) are found predominantly in the brain and ShcD is

expressed mainly in brain and muscle tissues (Wills and Jones, 2012). All of the Shc members

are structurally characterized by the unique modular arrangement of a phospho-tyrosine binding

(PTB) domain, a collagen homology 1 (CH1) region followed by a SH2 domain (Luzi et al.,

2000). The PTB and SH2 domains independently bind motifs containing phosphorylated tyrosine

residues. The domains are separated by the CH1 region which contains three consensus tyrosine

residues that are phosphorylated by tyrosine kinases (Luzi et al., 2000). The phosphotyrosine

residues subsequently serve as recognition motifs for the SH2 domain of proximal signaling

molecules including Grb2. The C. elegans Shc homologue lacks the phosphotyrosine sites in the

CH1 region, whereas the D. melanogaster Shc homologue contains two of the three consensus

sites, consistent with an expanded role for phosphotyrosine signaling with metazoan evolution

(Lim and Pawson, 2010).

The ShcA gene (hereafter named Shc1) locus in mammals encodes three isoforms; p46Shc,

p52Shc and p66Shc (Fig. 1.1) (Ravichandran, 2001). p52Shc is the most extensively studied

isoform and the one traditionally referred to as Shc1 in the literature. p46Shc and p52Shc

5

originate from different translation initiation sites within the same mRNA (Pelicci et al., 1992).

p52/p46Shc bind to autophosphorylated tyrosine residues in activated receptor and cytoplasmic

tyrosine kinases, which in turn phosphorylate p52/p46Shc on three tyrosine residues (239, 240,

and 317) located in the CH1 region, enabling Shc proteins to recruit the Grb2-SOS complex that

activates the GTPase Ras and the Grb2-Gab2-phosphoinositide-3-kinase (PI3K) complex (Fig.

1.2) (Ravichandran, 2001; Wills and Jones, 2012). Rapid tyrosine phosphorylation of these three

residues is necessary for transmitting the RTK-mediated mitogenic and cell survival signals to

downstream targets. For example, recent quantitative proteomic analysis of Shc1 signaling

showed that phosphorylation of EGFR Tyr1148 and Tyr1173, the main binding sites of Shc1,

occurs within seconds of EGF ligand binding, allowing rapid recruitment and tyrosine

phosphorylation of Shc1 (Dengjel et al., 2007). The SH2 domain of Grb2 binds to the

phosphorylated tyrosine residues in the CH1 region of Shc1, promoting signaling to the Ras-

mitogen-activated protein kinases (MAPK) and the PI3K-protein kinase B (PKB/Akt) cascades.

Shc1, however, is not necessary to activate the Ras-MAPK pathway as Grb2 can bind directly to

EGFR (Batzer et al., 1994). Shc1 mainly functions to sensitize cells to EGF stimulation and

MAPK activation as deletion of all Shc1 isoforms diminishes, but does not abolish, MAPK

phorphorylation (Lai and Pawson, 2000). Since EGFR has been extensively studied as a model

for RTK-Shc1 signaling to Ras-MAPK activation, it will be discussed in further detail.

6

Fig. 1.1. Schematic diagram of the Shc1 proteins.

All three Shc1 isoforms share the same domain architecture. The PTB and SH2 domains of Shc1

bind to phosphotyrosine-containing sequences. The PTB preferentially binds to NPXpY motifs

and specificity is determined by residues N-terminal to the pY. The SH2 domain binds to pY-

hydrophobic-X-hydrophobic motif and the specificity is determined by residues C-terminal to the

pY. The CH1 region contains three phospho-tyrosine sites: The Y239/Y240 twin tyrosines and

Y317. The SH2 domain of Grb2 binds to the CH1 pY residues, coupling Shc1 to MAPK

activation.

p52Shc is the isoform most extensively studied in the context of growth factor signaling. p46Shc

lacks the first 46 amino acids within the PTB domain and its function is not clearly understood.

The p66Shc isoform possesses an additional N-terminus CH2 region of 110 amino acids

containing a S36 phosphorylation site that has been implicated in mediating oxidative stress

signaling.

7

Fig. 1.2: Role of Shc1 in signaling downstream of RTK.

Ligand binding causes RTKs to dimerize and autophosphorylate specific tyrosine residues in their

cytoplasmic tail, which serve as docking sites for the Shc1 PTB domain. The EGFR kinase

phosphorylates tyrosine in the CH1 region, to which the Grb2 adaptor bind and initiate the Ras/MAPK

cascade. Shc1 can also be phosphoryalted by cytoplamic tyrosine kinases such as Src. Grb2 and its

constitutive binding partner, the guanine nucleotide exchange factor (GEF) SOS, activates the

membrane-associated Ras GTPase, which in turn activates the Raf Ser/Thr kinase which activates the

Erk MAPK. The Ras-MAPK pathway promotes proliferation, differentiation and survival. Like

p52Shc, p66Shc also interacts with Grb2, however it exerts an inhibitory or no effect on the MAPK

pathway. Shc1 can also activate the PI3k/Akt pathway through binding to the adaptor Gab2. Following

RTK activation, Gab2 interacts with the p85 regulatory subunit of PI3K. The catalytic p110 subunit of

PI3K can then phosphorylate membrane phospholipids, generating PIP3, which recruits the Ser/Thr

kinase Akt. Akt activates several downstream targets to enhance cell survival and growth.

8

The Shc1 proteins mediate signal transduction to downstream effectors through complex

phosphorylation/dephosphorylation dynamics. For example, following EGF stimulation, p52Shc

is phosphorylated on both Tyr and Ser/Thr residues but at different time points. Initially, as

mentioned above, p52Shc is phosphorylated on Tyr239, Tyr240 and Tyr317 (1–2 minutes after

EGF stimulation). Afterwards, Ser29 (3 minutes), Thr214 (5 minutes) and Ser335 (20 minutes)

undergo phosphorylation (Zheng et al., 2013). Each phosphorylation event leads to the

association of p52Shc with specific proteins (Zheng et al., 2013). The first phosphorylation event

leads to the recruitment of the Grb2 adaptor to p52Shc1 pTyr sites, activating pro-mitogenic and

survival pathways. Ser29/Thr214 phosphorylation recruits the tyrosine phosphatase PTPN12

which dephosphorylates p52Shc1 leading to loss of Grb2 binding. Ser335 phosphorylation of

p52Shc is required for cytoskeletal reorganization through recruiting phosphatases and GTPase-

activating proteins which downregulate Ras-MAPK signaling (Zheng et al., 2013).

Deletion of all Shc1 isoforms (p66Shc, p52Shc and p46Shc) is embryonic lethal by E11.5 due to

cardiovascular defects (Lai and Pawson, 2000). Knock-in embryos expressing Shc1 with a non-

functional PTB domain (R175Q substitution) die at approximately E11.5 and exhibit cardiac

defects reminiscent of full Shc1 KO, indicating that Shc1 regulates heart development by a PTB-

dependent mechanism (Hardy et al., 2007). However, Knock-in embryos with a non-

phosphotyrosine binding SH2 domain mutation (R397K) or all three phosphotyrosine residues in

the CH1 region substituted by phenylalanine (3Y to 3F) can survive to birth, but demonstrate

impaired motor coordination due to abnormal development of muscle spindles (Hardy et al.,

2007). Thus it appears that heart morphogenesis requires Shc1 with a functional PTB domain,

but occurs independently of phosphorylation of the CH1 tyrosine residues, the primary Grb2

9

binding sites. Biochemical data from experiments in D. melanogaster suggest that this process is

conserved as the Drosophila Shc1 ortholog, dShc, requires the PTB domain for signaling

downstream of the RTKs DER and Torso (Lai et al., 1995; Li et al., 1996).

1.1.2. p66Shc: an integrator of mitogenic and metabolic signaling

The p66Shc isoform has emerged with vertebrates (Luzi et al., 2000; Migliaccio et al., 1997). It

has an additional unique N-terminal 110-amino-acid collagen-homology region 2 (CH2). While

p46Shc and p52Shc are ubiquitously expressed in various tissues and developmental stages,

p66Shc is expressed in a tissue-specific fashion, mainly in liver, lungs, skin and heart. It is

notably absent from some tissues, including haematopoietic cells and brain (Lebiedzinska et al.,

2009). p66Shc fails to enhance MAPK phosphorylation upon EGF stimulation, despite being

phosphoryaled by EGFR and binding Grb2 (Migliaccio et al., 1997). Unlike p52Shc,

overexpressing p66Shc does not transform NIH3T3 mouse fibroblasts (Migliaccio et al., 1997).

In contrast, stable knockdown of p66Shc in L6 skeletal muscle myoblasts caused increased basal

activation of the Erk MAPK (Natalicchio et al., 2004). This observation is intriguing given that

p66Shc shares identical PTB sequences, phosphotyrosine sites on the CH1 region and SH2

sequence with the shorter isoforms. Negative regulation of EGFR signaling by p66Shc has been

attributed to MAPK-mediated phosphorylation of Ser36 in the CH2 region unique to p66Shc

(Okada et al., 1997). This serine phosphorylation event is thought to destabilize EGFR-p66Shc

interaction, thereby decoupling p66Shc from Ras-MAPK activation (Okada et al., 1997). Since

p66Shc and p52Shc compete for a limited pool of Grb2, p66Shc may sequester the Grb2-SOS

complex away from RTKs at the membrane, inhibiting Ras-MAPK activation.

10

Unlike p52Shc, the abundance of p66Shc is substantially decreased in breast cancer cell lines

expressing the oncogene v-erb-b2 avian erythroblastic leukemia viral oncogene homolog 2

(ErbB2), suggesting that p66Shc may function as a tumor suppressor (Stevenson and Frackelton,

1998). p66Shc expression correlates with a favorable outcome for breast cancer, particularly in

the context of the ratio of total phospho-Tyr317 Shc1 to the levels of p66Shc isoform (pTyr

Shc/p66Shc); higher levels of phosphorylated Shc1 relative to the abundance of the p66Shc

isoform are associated with aggressive neoplasms and increased risk of relapse (Davol et al.,

2003). This antagonistic effect of p66Shc on RTK signaling occurs, most likely, due to the

competition between p66Shc and p52Shc for common effectors of RTK signaling (Okada et al.,

1997).

1.1.3. Role of p66Shc in regulating oxidative stress

p66Shc has been portrayed as both a sensor and a proponent of reactive oxygen species (ROS)

production, promoting oxidative stress and pro-apoptotic signaling (Pinton and Rizzuto, 2008).

The role of p66Shc in the generation of ROS involves at least three mechanisms: a) p66Shc

activation leads to inhibition of the Forkhead box O (FoxO) transcription factors via Akt-

mediated phosphorylation. This results in a decrease in FoxO-dependent expression of ROS

scavenging enzymes, such as superoxide dismutase (SOD) and catalase (Nemoto and Finkel,

2002), b) At the plasma membrane, p66Shc promotes the activation of the Ras-related C3

botulinum toxin substrate 1 (Rac1) GTPase, triggering ROS production by the membrane-bound

NADPH oxidase (Khanday et al., 2006), and c) p66Shc acts in the mitochondrial intermembrane

space (IMS) where it interacts with cytochrome c (Giorgio et al., 2005). While all three isoforms

of Shc1 are predominantly cytoplasmic and a fraction translocate to the plasma membrane

following growth factor stimulation, approximately 10-20% of p66Shc also relocates to the

11

mitochondrial intermembrane space in response to oxidative stress (Orsini et al., 2004). Under

stress conditions (e.g. hydrogen peroxide treatment), Ser36 in the CH2 region is phosphorylated

by protein kinase C β (PKCβ) resulting in a phospho-Ser36-Pro37 motif recognized by the prolyl

isomerase Pin1 which induces isomerization around the Ser-Pro bond, targeting p66Shc to the

mitochondria (Pinton et al., 2007). Inside the IMS, p66Shc interacts with cytochrome c,

promoting transfer of electrons to oxygen and generating hydrogen peroxide (Giorgio et al.,

2005). The increase in ROS production leads to swelling and rupture of mitochondria and release

of pro-apoptotic factors into the cytoplasm. In this context, p66Shc uses the reducing equivalents

of the mitochondrial electron transfer chain through the oxidation of cytochrome c, leading to the

activation of programmed cell death (Giorgio et al., 2005). Factors regulating p66Shc expression

under stress conditions are still unknown. One potential mechanism involves the tumor

suppressor protein p53, where activated p53 upregulates p66Shc expression to induce apoptosis

in murine embryonic fibroblasts (MEFs) and endothelial cells under oxidative stress conditions

(Kim et al., 2008a; Trinei et al., 2002).

Consistent with its role in mediating oxidative stress and programmed cell death, p66Shc

knockout MEFs show increased resistance to apoptosis in response to various conditions,

including treatment with UV, taxol or amyloid β-peptide (Migliaccio et al., 1999; Smith et al.,

2005b; Yang and Horwitz, 2000). Additionally, MEFs lacking p66Shc show lower levels of

systemic and tissue oxidative stress markers, such as 8-oxoguanine, compared to p66Shc-

competent cells (Napoli et al., 2003; Nemoto and Finkel, 2002; Trinei et al., 2002). p66Shc

mediates cell death in cardiac cells following chronic exposure to angiotensin II (Ang II), a

secreted factor that shows high abundance in patients with hypertension, atherosclerosis, and

12

diabetes. Selective disruption of the p66Shc isoform in mice caused resistance to Ang II-

mediated hypertrophy and apoptosis in cardiomyocytes and endothelial cells (Graiani et al.,

2005). Deletion of p66Shc also reduces vascular cell apoptosis and early atherogenesis in mice

fed a high-fat diet (Napoli et al., 2003). These phenotypes have been partly attributed to

increased resistance of the p66Shc KO cells to oxidative stress-mediated cell damage (Migliaccio

et al., 1999).

1.1.4. p66Shc and energy metabolism

p66Shc fulfills a function in energy metabolism, as indicated by its role in oxidative stress and

insulin signaling. p66Shc has been shown to inhibit insulin-like growth factor-1 (IGF-1)-Akt

signaling in vascular smooth muscle cells (Xi et al., 2010a; Xi et al., 2010b). In addition, L6

myoblasts overexpressing p66Shc displayed reduced rates of basal glucose uptake and a

reduction in the protein abundance of glucose transporters (Natalicchio et al., 2009). In animal

models, inactivation of p66Shc in mice improved glucose tolerance and insulin sensitivity

(Ranieri et al., 2010; Tomilov et al., 2011). In obese mice lacking the hormone leptin (ob/ob

mice), fasting glycemia values were significantly lower in p66Shc KO than in p66Shc WT mice,

despite no difference in plasma insulin levels (Ranieri et al., 2010). These studies suggest that

p66Shc may suppress glucose metabolism by dampening insulin signaling (Giorgio et al., 2012)

13

1.2. Cell metabolism in normal and cancer cells

The metabolic program of resting non-dividing cells provides the energy required for

maintaining ATP production. On the contrary, proliferating cells must generate sufficient energy

to support cell division, cellular biosynthesis and maintain cellular redox homeostasis (Cairns et

al., 2011). Reprogramming of cellular metabolism towards anabolism, in response to growth

factor signaling, is crucial for supplying nucleotides, proteins and lipids needed for a cell to

double its size before dividing into two daughter cells (Ward and Thompson, 2012).

The first observation regarding the atypical metabolic demands of rapidly dividing cells can be

traced back to the pioneering work of Otto Warburg (Warburg, 1956). In the presence of oxygen,

most normal tissues break down glucose to generate pyruvate, which is largely oxidized to

carbon dioxide in the mitochondria through the tricarboxylic acid (TCA) cycle followed by

oxidative phosphorylation. This process produces a net of 36 ATPs. However, under anaerobic

conditions, pyruvate is shunted away from mitochondria and gets reduced to lactate, generating

only 2 ATPs. Warburg’s studies on rapidly dividing ascites cancer cells showed that tumors

displayed unusually high rates of glycolysis despite the low ATP yield of converting glucose to

lactate. He proposed that tumor cells have impaired mitochondria and, hence, depend on

fermenting glucose to meet their energy demand. This phenomenon, named “the Warburg

effect,” has been confirmed in other cancer types and has been extended beyond glycolysis to

include increased deployment of carbon into anabolic metabolism that supports cell growth

(Vander Heiden et al., 2009). The reliance of cancer on higher glucose uptake has been clinically

used for tumor detection and monitoring through the implementation of 18

F-deoxyglucose-

positron emission tomography (PDG-PET) imaging.

14

1.2.1. The Warburg effect: glucose metabolism and anabolic demands of cell

growth

The selective advantage that glycolytic metabolism provides for proliferating tumor cells has

been debatable for years (Ward and Thompson, 2012). The initial proposal by Warburg that

mitochondrial dysfunction in tumor cells required aerobic glycolysis to cope with low

competence of mitochondrial ATP generation was not fully correct. Mitochondrial respiration

occurs in cancer cells and remains the primary source of ATP production in most tumors (Ward

and Thompson, 2012). For example, an oncogenic mutant of KRas induced aerobic glycolysis as

evidenced by ~ 2-fold increase in glucose uptake and lactate secretion. Yet, most cellular ATP (~

60%) is still produced by mitochondrial oxidation in these cells (Fan et al., 2013). In addition,

mitochondrial function in KRas-mediated tumorigenesis may be crucial for transformation and

tumor progression (Weinberg et al., 2010). Another possible explanation for the Warburg effect

is that heightened glycolytic metabolism takes place as an adaptive response to hypoxic

conditions in the early phases of tumor development prior to vascularization, driven in part by

acidic microenvironment resulting from excess lactate production (Gatenby and Gillies, 2004).

An alternative and more plausible explanation is that the preferential metabolism of glucose

through glycolysis serves to provide precursor molecules, such as amino acids, for biomass

synthesis and to maintain redox balance in proliferating cells (Fig. 1.3) (Cairns et al., 2011;

Vander Heiden et al., 2009). There are several examples supporting this concept. The glycolytic

intermediates fructose-6-phosphate (F6P) and glyceraldehyde-3-phosphate (GADP) can be

shunted into the non-oxidative branch of the pentose phosphate pathway, generating ribose-5-

phosphate (R5P), a critical component of nucleotide biosynthesis. Similarly, glucose-6-

15

phosphate (G6P), the first intermediate in the glycolytic pathway, could feed the oxidative arm of

the pentose phosphate pathway to generate NADPH, a coenzyme used in anabolic reactions,

such as lipid and nucleic acid synthesis, and in maintaining the reductive environment of the cell.

In addition to their role in glycolysis, G6P and F6P are precursors of the hexosamine

biosynthesis pathway which provides uridine-diphosphate N-acetylglucosamine (UDP-GlcNAc)

as a substrate for O-GlcNAcylation of cytosolic proteins and N-linked and O-linked

glycosylation of proteins produced in the secretory pathway. N-glycans branching provides

positive feedback to enhance nutrient transport and receptors residency at the cell surface (Lau et

al., 2007; Ohtsubo et al., 2005; Partridge et al., 2004). The glycolytic intermediates also serve as

precursors for protein synthesis: 3-phosphoglycerate (3PG) can be converted to serine and

pyruvate can be transaminated to alanine. In addition, glycolytic metabolism provides the

building blocks for lipid synthesis. Reduction of the glucose-derived dihydroxyacetone (DHAP)

to glycerol-3-phosphate (G3P) endows proliferating cells with an essential precursor for

phospholipids and triacylglycerol biosynthesis. In addition, pyruvate can contribute to the

production of mitochondrial citrate, which can then be exported to the cytoplasm to be utilized in

de novo synthesis of fatty acids (Hatzivassiliou et al., 2005).

16

Fig. 1.3. Cancer metabolism: the Warburg effect.

Rapidly dividing cells needs sufficient energy to support cell division, increase cell biomass and maintain

redox homeostasis. Reprogramming of cellular metabolism towards glycolysis provides the building

blocks required for the synthesis of nucleotides, proteins and lipids. Preferential diversion of pyruvate into

lactate production allows proliferating cells to shunt glycolytic intermediates into branching anabolic

pathways, including the pentose phosphate and hexosamine pathways, and lipid and protein biosynthesis.

17

1.2.2. Glutamine: a metabolic fuel for proliferating cells

Glucose metabolism supplies cancer cells with essential anabolic building blocks (Hanahan and

Weinberg, 2011), yet cannot explain all the metabolic changes needed to support cell growth.

The significance of other nutrients, such as the amino acid glutamine, as essential fuel for cancer

growth and survival has become better understod in recent years (DeBerardinis and Cheng,

2010). Cancer cells are in high demand of nitrogenous compounds, including nucleotides,

nonessential amino acids, and hexosamines. Glutamine, the most abundant free amino acid in

human blood, acts as an obligate nitrogen donor in purine and pyrimidine synthesis, and as the

primary nitrogen source in anabolism of nonessential amino acids from α-keto acids

(DeBerardinis and Cheng, 2010). Glutamine is not an essential amino acid, but high growth rates

in embryonic and cancer cells depend on glutamine as a nitrogen source and also in conversion

to α-ketoglutarate which supports the TCA cycle and ATP production. This allows more of the

glycolytic intermediates to be utilized in anabolic pathways. Real-time 13

C NMR demonstrates

that glutamine carbon can be converted into lactic acid and secreted from cancer cells in a

process termed “glutaminolysis” (DeBerardinis et al., 2007).

Glutamine also functions as an essential carbon source for replenishing depleted TCA cycle

intermediates, a process known as “anaplerosis.” It gets deaminated to the amino acid glutamate

which is converted to the mitochondrial intermediate α-ketoglutarate and then to citrate.

Glutamine contributes to the carbons of mitochondrial citrate, which can be exported to the

cytoplasm and converted to acetyl-CoA, a precursor for fatty acid synthesis, and oxaloacetate

(OAA). OAA can be shuttled back to the mitochondria and metabolized through multiple steps

18

to yield NADPH, providing the reducing substrates required for lipid synthesis and regeneration

of reduced glutathione needed in rapidly dividing cells (DeBerardinis et al., 2007). In addition,

OAA can be transaminated to aspartate, which can be utilized as a carbon source in nucleotide

biosynthesis. Further studies are needed to dissect the exact mechanisms regulating glutamine

metabolism and its impact on proliferating cells.

1.3. Signaling pathways and regulation of cellular metabolism

Signaling downstream of growth factor receptors not only mediates cell proliferation, but also

alters cellular metabolism. As stated earlier, this metabolic rewiring is critical for the cell to meet

the energy and anabolic requirements associated with cell growth and division. For example,

receptor tyrosine kinase signaling mediated by the PI3K/Akt pathway plays a central role in

regulating glucose uptake, expression and activation of glycolytic enzymes, as well as regulating

glucose carbon flux into lipid synthesis (Elstrom et al., 2004). In addition, tyrosine kinase-

dependent regulation of glycolytic enzymes, such as pyruvate kinase, reroutes glucose

metabolism into anabolic pathways (Christofk et al., 2008a; Christofk et al., 2008b). These

signaling pathways, among others, allow cells to coordinate growth and division with their

metabolic activity.

1.3.1. Tyrosine kinase signaling and selective metabolic regulation in

dividing cells

Several metabolic enzymes become tyrosine phosphorylated following growth factor stimulation,

where signaling is initiated by phosphorylation of activated receptors and downstream effectors

19

on specific tyrosine residues. An example is pyruvate kinase (PK), the enzyme that catalyzes the

final irreversible step of glycolysis, converting phosphoenolpyruvate (PEP) to pyruvate with the

concomitant generation of ATP. There are four members of the PK family of proteins: PKL,

PKR, PKM1 and PKM2. PKL and PKR expression is restricted to liver and red blood cells,

respectively. PKM1 and PKM2 are splice variants encoded by the PKM gene. These two

isoforms differ only by one exon: inclusion of exon 9 and exclusion of exon 10 for PKM1 and

vice versa for PKM2, rsulting in a difference of 23 amino acids at their carboxy terminal. Most

cells in adult tissues predominantly express PKM1. PKM2, on the other hand, is found in cells

with self-renewal capacity, including stem cells and embryonic cells. In addition, this isoform is

highly expressed in many tumors (Mazurek et al., 2005).

Despite the seemingly paradoxical observation that PKM2 possesses lower specific activity than

PKM1, cells expressing PKM2 have a selective growth advantage over cells expressing PKM1

(Christofk et al., 2008a; Christofk et al., 2008b). Unlike the constitutively active PKM1 and

other PK variants, PKM2 is sensitive to growth factor signaling. This isoform binds to tyrosine

phosphorylated proteins and is tyrosine phosphorylated, leading to inhibition of PKM2 activity

(Hitosugi et al., 2009), allowing PKM2 to act as a gatekeeper for the glucose carbon metabolic

fate. By slowing glycolysis, PKM2 enables cells to shunt glycolytic intermediates into branching

anabolic pathways, such as serine and glycine synthesis, pentose phosphate and hexosamine

pathways, rather than sending pyruvate to the TCA cycle.

The oncoprotein Myc promotes the expression of PKM2 at the expense of PKM1 by altering

exon splicing through upregulating the expression of heterogeneous nuclear ribonucleoproteins

20

(hnRNPs) (David et al., 2010). This observation is well in line with the role of Myc in

stimulating glycolysis and inducing the expression of hypoxia-inducible factor 1 (HIF-1) to

support cell growth of rapidly dividing cells (Gordan et al., 2007). PKM2 may also have roles

beyond regulating PK activity. For example, PKM2 interacts with HIF-1α, stimulating HIF-1-

mediated transactivation of glycolytic genes (Luo et al., 2011). In addition, PKM2, upon

translocation to the nucleus, activates β-catenin (Yang et al., 2011) where PKM2 binds to

tyrosine-phosphorylated β-catenin in the nucleus and contributes to β-catenin-mediated

transactivation of cyclin D and Myc, promoting tumor progression (Yang et al., 2011). Further

investigation is needed to determine how the glycolytic and non-glycolytic functions of PKM2

are regulated in both resting and proliferating cells.

1.3.2. The PI3K-Akt-mTOR pathway

One of the most highly conserved signal transduction pathways downstream of growth factor

receptors is PI3K-Akt-mTOR (Hemmings and Restuccia, 2012). PI3K is activated when growth

factors bind to receptor tyrosine kinases, G-protein coupled receptors and cytokine receptors

(Vanhaesebroeck et al., 2012). PI3K not only provides growth and survival signals, but redirects

cellular metabolism to meet the cellular demands for growth. Not surprisingly, the PI3k-Akt

pathway is one of the most commonly mutated signaling nodes in human cancers (Rodon et al.,

2013). Activation can occur due to mutations in PI3K, its negative regulator phosphatase and

tensin homolog (PTEN), or abnormal signaling from upstream receptors (Engelman et al., 2006).

21

When activated, PI3K phosphorylates membrane lipids to generate phosphatidylinositol (3,4,5)-

triphosphate (PIP3) which subsequently leads to recruitment and activation of Ser/Thr kinases

with PH domains that bind PIP3 (Engelman et al., 2006). The best characterized effector

downstream of PI3K is Akt (also known as PKB). Akt drives glycolytic metabolism by

increasing the expression of the glucose transporter 1 (Glut1) and membrane translocation of

adipose- and striated muscle-specific glucose transporter 4 (Glut4) (Robey and Hay, 2009). In

addition, Akt phosphorylates and activates key glycolytic enzymes, such as hexokinase (Gottlob

et al., 2001). Akt also directly phosphorylates the enzyme phosphofructokinase 2, increasing the

levels of fructose 2,6-bisphosphate which enhances the activity of the rate-limiting glycolytic

enzyme phosphofructokinase 1 (Deprez et al., 1997).

The prolonged Akt signaling associated with transformation inhibits the FoxO transcription

factors, resulting in transcriptional changes that lead to enhanced glycolytic metabolism (Khatri

et al., 2010). When FoxO1 is phosphorylated by Akt on Thr24, Ser256, and Ser319, it is

excluded from the nucleus, ubiquitilated and degraded. Phosphorylation of FoxO1 by Akt

decreases the hepatic glucose production through decreasing the transcription of glucose 6-

phosphatase (Nakae et al., 2001). Akt also stimulates anabolic pathways downstream of

glycolysis. For example, Akt phosphorylates and activates ATP citrate lyase, an enzyme that

catalyzes the conversion of citrate to acetyl-CoA required for fatty acid synthesis (Berwick et al.,

2002). Furthermore, Akt activates the mammalian target of rapamycin complex 1 (mTORC1) by

phosphorylating and inhibiting its negative regulator tuberous sclerosis 2 (TSC2) (Manning et

al., 2002; Tee et al., 2002). mTOR functions as a key metabolic regulator, stimulating glycolytic

metabolism, ribogenesis, protein and lipid synthesis, and cell growth in response to growth

22

signals and nutrient availability (Laplante and Sabatini, 2012). In addition, mTOR activates

transcription factors, such as HIF1, which increases the capacity of cells to carry out glycolysis

(Hudson et al., 2002). Upstream and downstream modulators of the mTOR signaling pathway

will be discussed in a separate section.

1.3.3. Transcriptional regulators of anabolic metabolism

Several transcriptional regulators have been implicated in regulating cell metabolism. Due to

their fundamental roles in controlling cell proliferation, the metabolic functions of HIF, Myc and

p53 will be covered in this section.

1.3.3.1. HIF

Under hypoxic conditions, mammalian cells undergo a metabolic shift towards increasing

glucose consumption and redirecting glycolytic pyruvate to lactate (Greer et al., 2012). This

adjustment is mediated by the HIF1 and HIF2 complexes which are the major transcription

factors expressed in response to low oxygen conditions. The two exist as heterodimers of the

constitutively expressed HIF1β subunit and either HIF1α or HIF2α subunits that are stabilized

under hypoxic conditions (Bertout et al., 2008). Under normoxic conditions, the alpha subunits

of HIF are hydroxylated at conserved proline residues by HIF prolyl hydroxylases, resulting in

their recognition and subsequent ubiquitilation by the Von Hippel–Lindau (VHL) tumor

suppressor E3 ubiquitin ligase which tags them for rapid degradation by the proteasome

(Maxwell et al., 1999). HIF prolyl hydroxylases are inhibited in hypoxic environments, since

they utilize oxygen as a cosubstrate (Semenza, 2004). While HIF1α is ubiquitously expressed,

HIF2α is limited to specific tissues (Bertout et al., 2008). Although both transcription factors

23

activate overlapping sets of genes, most studies have focused on the role of HIF1 in regulating

metabolism.

When oxygen is limited, continuous mitochondrial oxidative phosphorylation might cause

mitochondrial redox stress. Under such conditions, HIF1 induces expression of genes supporting

anaerobic glucose metabolism, including glucose transporters and glycolytic enzymes to increase

the capacity of cells to carry out glycolysis at the expense of the TCA cycle. The HIF1-mediated

increased expression of lactate dehydrogenase and pyruvate dehydrogenase kinase 1 divert

pyruvate from entering the mitochondria to lactate production. This reduction in pyruvate flux

into the TCA cycle decreases oxidative phosphorylation and oxygen consumption, enabling ATP

production in rapidly dividing cells in an oxygen-independent mechanism (Greer et al., 2012).

In addition, HIF1 can be stabilized in proliferating cells under normoxic conditions. For

example, activated Akt, through mTOR signaling, increases HIF1α mRNA translation (Hudson

et al., 2002). Furthermore, in the presence of normal oxygen levels, the increased glucose uptake

in proliferating cells can also inhibit the hydroxylation and subsequent degradation of HIF1α.

This occurs through enhanced glucose-mediated production of ROS which act as strong

inhibitors of the HIF-targeting prolyl hydroxylases (Chandel et al., 2000). Mutations in the TCA

cycle enzymes succinate dehydrogenase and fumarate hydratase lead to the accumulation of their

substrates, succinate and fumarate, respectively, which can also inhibit prolyl hydroxylases

(Isaacs et al., 2005; Selak et al., 2005). Inhibition of prolyl hydroxylases by ROS, succinate or

fumarate represents a feedback mechanism decreasing the flow of glucose carbon into

mitochondria and reinforcing a glycolytic phenotype in tumors.

24

1.3.3.2. Myc

Like HIF1, Myc increases the expression of genes involved in glucose uptake and glycolytic

metabolism (Osthus et al., 2000). For example, Myc induction of lactate dehydrogenase diverts

pyruvate away from mitochondria, enhancing the glycolytic pathway (Osthus et al., 2000). It also

stimulates the expression of genes supporting anabolic utilization of glutamine, where Myc

directly induces the expression of glutamine transporters SLC5A1 and SLC7A1 (Gao et al.,

2009). In addition, Myc increases the levels of glutaminase 1, the first enzyme of glutaminolysis,

by inhibiting the expression of its negative regulators mir-23A and mir-23B (Gao et al., 2009).

Myc-transformed cells cannot survive in the absence of exogenous glutamine which functions as

a critical carbon source for anapleuretic reactions in these cells (Wise et al., 2008; Yuneva et al.,

2007). Myc also induces the expression of amino acid (serine hydroxymethyltransferase) and

fatty acid (fatty acid synthase) biosynthesis genes (O'Connell et al., 2003).

1.3.3.3. p53

The transcriptional factor and tumor suppressor p53 is a vital gatekeeper against cellular stresses.

It activates a myriad of cell defense pathways, including cell cycle arrest, DNA damage repair

and apoptosis (Bieging and Attardi, 2012). More recently, several lines of evidence suggest a

critical role for p53 in regulating metabolism (Vousden and Ryan, 2009). p53 inhibits glycolytic

metabolism by repressing the expression of the glucose transporters Glut1 and Glut4

(Schwartzenberg-Bar-Yoseph et al., 2004) and the glycolytic enzyme phosphoglycerate mutase

(Kondoh et al., 2005). In addition, p53 increases the levels of the TP-53-induced glycolysis and

apoptosis regulator gene (TIGER), a negative regulator of glycolytic metabolism (Bensaad et al.,

2006). Importantly, p53 mutated at three critical Lys sites (p533KR/3KR

) is deficient in p53-

25

dependent regulation of cell-cycle arrest, apoptosis and senescence, yet retains its tumor

suppressor activity. The p533KR/3KR

mutation continues to regulate a subset of metabolic genes,

notably suppressing the expression of Glut3 (Li et al., 2012).

p53 promotes oxidative phosphorylation by activating the expression of cytochrome oxidase c 2

(SCO2) which is required for the assembly of the cytochrome c complex in the electron transport

chain (Matoba et al., 2006). p53 can also alter metabolism in a transcription-independent

manner. It has been reported that p53 directly binds to and inactivates the enzyme glucose-6-

phosphate dehydrogenase, inhibiting glucose metabolism through the pentose phosphate

pathway. Indeed, p53-deficient cells have higher levels of R5P and NADPH compared to p53-

competent cells (Jiang et al., 2011). These observations suggest that p53 mutation or deletion

could act as a driving force supporting a glycolytic phenotype in cancer.

1.3.4. Metabolic enzymes as oncogenes

Not all cancer-associated mutations influencing cellular metabolism are connected to augmenting

anabolic demands of dividing cells (Gottlieb and Tomlinson, 2005). Oncogenic mutations could

provide a metabolic enzyme with a neomorphic activity to drive tumorigenesis.

An example is the driving mutation discovered in the cytosolic isocitrate dehydrogenase 1

(IDH1). Whole genome sequencing identified somatic mutations in IDH1 in a subset of gliomas

and acute myeloid leukemia (AML). These mutations are remarkably restricted to specific

arginine residues required for IDH binding to its substrate isocitrate (Mardis et al., 2009; Parsons

26

et al., 2008). All affected patients were heterozygous for the mutations, retaining a wild-type

IDH1 allele. The heterozygous nature of the mutation has been shown to dominantly inhibit

wild-type IDH1 in cells (Zhao et al., 2009). The IDH1 R132 mutation endows IDH1 with a new

reductive activity to convert α-ketoglutarate to 2-hydroxyglutarate (2-HG), a rare metabolite

found in trace amounts in cells under normal conditions (Dang et al., 2009; Ward et al., 2010).

Studies of mutations in IDH2, the mitochondrial homologue of IDH1, demonstrated that 2-HG is

a pathogenic “oncometabolite” rather than a byproduct of a loss-of-function mutation. Mutations

in R172 of IDH2, the analogous residue of R132 in IDH1, also resulted in elevated levels of 2-

HG in AML and glioma patients (Yan et al., 2009). However, not all AML samples with

elevated 2-HG have neomorphic mutations in either IDH1 R132 or IDH2 R172 (Ward et al.,

2010). A more comprehensive sequencing of these samples revealed an additional critical

mutation of IDH2, R140 (Ward et al., 2010), emphasizing that the main feature selected for by

this mutation is 2-HG production. While much work is needed to utilize 2-HG measurement in

diagnostic tests, recent data showed that 1H NMR can be applied for 2-HG detection of glioma in

vivo (Andronesi et al., 2012; Choi et al., 2012).

The observations that IDH mutations are selected for early during tumor progression (Watanabe

et al., 2009), and that 2-HG is not a mutagen (Mardis et al., 2009) suggested a specific role of 2-

HG in altering cancer metabolism. Several reports showed that 2-HG can inhibit the ten-eleven

translocation (TET) family of enzymes which oxidize 5-methylcytosine to 5-

hydroxymethylcytosine, a key intermediate in DNA demethylation (Figueroa et al., 2010; Turcan

et al., 2012; Xu et al., 2011). TET2 loss-of-function mutations and IDH1 or IDH2 neomorphic

27

mutations are mutually exclusive in AML (Figueroa et al., 2010). Furthermore, knocking down

TET2 recapitulated the effect of overexpressing IDH mutants in preventing hematopoietic cell

differentiation (Figueroa et al., 2010).

Inhibition of histone lysine demethylases is a potential mechanism by which 2-HG mediates its

oncogenic effect (Chowdhury et al., 2011). Expression of mutant IDH or treatment with cell-

permeable 2-HG repressed the expression of lineage-specific differentiation genes in 3T3-L cells

and blocked cellular differentiation. This correlated with a significant increase in repressive

histone methylation marks (Lu et al., 2012). The 2-HG-mediated alterations in histone and DNA

methylation are likely synergistic. The precise mechanisms connecting the effect of 2-HG on

DNA and histone methylation are still unknown. Further investigation is needed to clarify

whether epigenetic regulation is the only mechanism mediating the oncogenic effects of 2-HG.

28

1.4. mTOR: from signaling to metabolism

The mTOR pathway integrates signals for cell growth and energy metabolism. It enables

unicellular organisms such as yeast to sense nutrient availability and to support cell growth under

favorable environmental conditions. With emerging complexity in multicellullar organisms,

mTOR acquired additional roles such as regulation of immunity and neurogenesis. Hence,

dysregulation of mTOR signaling has been implicated in health conditions, including cancer,

Alzheimer’s and metabolic syndrome (Zoncu et al., 2011).

1.4.1. Molecular components of mTOR

mTOR was discovered in a yeast genetic screen for molecular targets of the immunosuppressant

rapamycin, yielding two genes, TOR1 and TOR2 that mediated the toxic effects of the drug

(Cafferkey et al., 1993; Kunz et al., 1993). Shortly afterwards, the mammalian homolog of TOR

(mTOR) was isolated as the direct target of rapamycin (Brown et al., 1994; Sabatini et al., 1994;

Sabers et al., 1995). mTOR is a Ser/Thr protein kinase that belongs to the PI3K-related kinases

(PIKK) family. It interacts with several distinct proteins to form complexes named mTOR

complex 1 (mTORC1) and 2 (mTORC2). The two complexes respond differently to rapamycin

and have different upstream regulators and downstream effectors (Zoncu et al., 2011). The

regulatory-associated protein of mTOR (Raptor) and the rapamycin-insensitive companion of

mTOR (Rictor) characterize mTORC1 and mTORC2, respectively. These two proteins function

as scaffolds for substrate binding and complex assembly (Zoncu et al., 2011).

29

Both mTOR complexes share 2 proteins: GβL (mLST8) (Jacinto et al., 2004; Kim et al., 2003)

and DEP domain-containing mTOR-interacting protein (Deptor) (Peterson et al., 2009) (Fig.

1.4). Unique components of mTORC1 include a negative regulator, 40 kDa proline-rich Akt

substrate (PRAS40) (Sancak et al., 2007), whereas mTORC2 contains stress-activated map-

kinase-interacting protein 1 (mSin1) (Frias et al., 2006; Jacinto et al., 2006; Yang et al., 2006)

and protein observed with rictor 1 and 2 (Protor 1 and 2) (Pearce et al., 2007). Structural and

biochemical studies suggest that mTORC1 functions as an obligate dimer where the dimeric

interfaces are formed by interactions between the mTOR and raptor subunits. Biochemical

analysis suggest that mTORC2 functions as oligomer, most likely as a TORC2-TORC2 dimer

(Wullschleger et al., 2005; Yip et al., 2010).

Rapamycin binds the prolyl isomerase FK506-binding protein (FKBP12) forming the rapamycin

-FKBP12 complex which binds to and inhibits mTOR (Brown et al., 1994; Chen et al., 1995;

Sabatini et al., 1994). It has been suggested that rapamycin might disrupt mTOR-Raptor

association, hindering mTOR from binding to its substrates (Kim et al., 2002; Yip et al., 2010).

However, the exact mechanism of mTORC1 inhibition by the FKBP12-rapamycin complex is

still unclear. Prolonged treatment with rapamycin can lead to a complete loss of intact mTORC2

in a subset of cell lines (Sarbassov et al., 2006). The effect of rapamycin on mTORC2 could be

attributed to rapamycin-FKBP12 mediated sequestration of the mTOR pool in the cell, thus

decreasing the availability of free mTOR for assembly into mTORC2.

30

Fig 1.4. mTOR signaling pathway.

Growth factors, such as insulin, stimulate PI3K to generate PIP3 to which PH domain-containing

Ser/Thr kinases, such as Akt and PDK1 bind. Activated by mTORC2 and PDK1, Akt

phosphorylates TSC1/2 on multiple sites and inhibits its GAP activity towards the GTPase Rheb.

GTP-loaded Rheb then activates mTORC1 which in turn phosphorylates several downstream targets

including 4E-BP1 to enhance protein synthesis, ULK1 to inhibit autophagy and S6K1 to enhance

ribogenesis and cell growth. Kinases other than Akt, such as Erk1/2 and RSK1, can also

phosphorylate TSC1/2 and inhibit its activity. Low energy (high AMP/ATP ratio) activates AMPK

which phosphorylates Raptor and TSC1/2, inhibiting mTORC1 activation. In addition to

phosphorylating Akt, mTORC2 also phosphorylates PKCα and SGK1. The number of phosphate

groups in the figure does not represent the actual number of phosphorylation sites on the proteins

indicated.

31

1.4.2. mTORC1

mTORC1 is the more extensively characterized of the two complexes. It can sense inputs mainly

from growth factors, amino acids, stress and energy status of the cell (Dibble and Manning,

2013). This allows mTORC1 to control several cellular functions including autophagy, protein

synthesis and cell survival (Dibble and Manning, 2013).

1.4.2.1. Upstream regulators of mTORC1

1) Growth factor signaling to mTORC1: One key component of growth factor signaling to

mTORC1 is the heterodimer of TSC1 and TSC2. The TSC1/2 dimer stably associates with TBC1

domain family member 7 (TBC1D7), the third core subunit of the complex, which is needed for

stabilization and full GTPase-activating protein (GAP) activity of TSC1/2 (Dibble et al., 2012).

The complex functions as a GAP for the Ras homolog enriched in brain (Rheb) GTPase (Garami

et al., 2003; Inoki et al., 2003; Tee et al., 2003; Zhang et al., 2003). The active GTP-loaded form

of Rheb directly interacts with and activates mTORC1 (Long et al., 2005). Indeed, Rheb is

essential for mTORC1 activation as its loss prevents mTOR activation by growth factors and

nutrients. Conversely, overexpression of Rheb constitutively activates mTORC1 even in the

absence of growth factors and nutrients (Saucedo et al., 2003; Stocker et al., 2003). By

converting Rheb into an inactive GDP-bound state, the TSC1/2 complex inhibits mTORC1

activation.

32

The TSC1/2 complex integrates signals from several mTORC1 upstream, including those arising

from growth factors, such as insulin. Akt, activated downstream of the insulin receptor, directly

phosphorylates and inactivates TSC1/2 leading to mTORC1 activation (Fig. 1.4) (Inoki et al.,

2002; Manning et al., 2002; Potter et al., 2002). Akt can also activate mTORC1 in a TSC1/2-

independent manner by phosphorylating the mTOR negative regulator PRAS40, causing its

dissociation from mTORC1 (Kovacina et al., 2003). Kinases other than Akt, such as Erk1/2 (Ma

et al., 2005) and ribosomal S6 kinase (RSK1) (Roux et al., 2004), can also phosphorylate TSC1/2

to inhibit it. Moreover, TSC1/2 can be phosphorylated and inactivated by glycogen synthase

kinase 3β (GSK3β) downstream of Wnt signaling causing mTORC1 activation (Castilho et al.,

2009; Inoki et al., 2006; Yang et al., 2006). The convergence of multiple growth signaling inputs

on mTORC1 allows it to act as a central signaling hub in several developmental stages. This

might explain the absolute requirement of mTORC1 signaling in early embryonic development

(Guertin et al., 2006).

2) Stress signaling to mTORC1: In addition to receiving inputs from growth factors, TSC1/2

also integrates signals from cell energy status associated with nutrient deprivation, low oxygen

and DNA damage (Dibble and Manning, 2013). mTORC1 indirectly senses low ATP levels

through the AMP-activated protein kinase (AMPK). When AMP/ATP ratio is high, under

conditions of nutrient deprivation, AMPK phosphorylates TSC2 (Corradetti et al., 2004; Dibble

and Manning, 2013; Inoki et al., 2006; Inoki et al., 2003). AMPK-mediated phosphorylation

activates TSC2 towards Rheb, inhibiting mTORC1. Under low energy stress conditions, AMPK

also phosphorylates the mTORC1 scaffold protein Raptor, inducing its binding to the regulatory

14-3-3 proteins which leads to inhibition of mTORC1 signaling (Gwinn et al., 2008).

33

In contrast to energy depletion, hypoxia can also promote TSC1/2 activation and, hence,

mTORC1 suppression in an AMPK-independent manner. Hypoxia-induced HIF1 stabilization

increases the expression of the gene regulated in development and DNA damage response 1

(REDD1), which promotes TSC1/2 activation through a mechanism that is yet to be elucidated

(Brugarolas et al., 2004; DeYoung et al., 2008; Reiling and Hafen, 2004). The HIF1-mediated

induction of REDD1 is controlled by signaling from the DNA damage repair kinase ataxia

telangiectasia mutated (ATM) (Cam et al., 2010). Like hypoxia, DNA damage inhibits mTORC1

activation. This occurs through p53-dependent induction of sestrin 1 and sestrin 2 which activate

AMPK (Budanov and Karin, 2008; Jones et al., 2005). DNA damage also induces the expression

of TSC2, PTEN and AMPK, suppressing the PI3K-mTORC1 pathway (Feng et al., 2007).

3) Nutrients signaling to mTORC1: Amino acids, particularly leucine and arginine, are

indispensable for mTORC1 signaling regardless of the upstream stimulating signal (Hara et al.,

1998; Wang et al., 1998). The most studied connection between amino acid stimulation and

mTORC1 is the Rag family of GTPases (Kim et al., 2008b; Sancak et al., 2008). Rags form

obligate heterodimers of either RagA/RagB with either RagC/D. The two units of the

heterodimer have opposite nucleotide loading states: when RagA/B is GDP-bound, RagC/D is

bound to GTP and vice versa. The yeast Gtr1 (RagA/B homolog) and Gtr2 (RagC/D homolog)

have been shown to functionally and genetically interact with mTORC1 (Urano et al., 2000).

Through currently uncharacterized mechanisms, amino acids cause the Rag heterodimer to

switch to the active form, where RagA/B becomes loaded with GTP and the RagC/D is bound to

GDP. This allows the Rags to interact with Raptor, promoting mTORC1 clustering at the surface

of lysosomes and late endosomes where the mTOR activators, Rag and Rheb GTPases, reside

34

(Sancak et al., 2008). Expression of constitutively GTP-bound RagA/B mutants renders

mTORC1 resistant to amino acid starvation (Kim et al., 2008b; Sancak et al., 2010). Inhibition of

glutaminolysis prevented GTP loading of RagB and subsequent lysosomal translocation and

mTORC1 activation (Duran et al., 2012). The mTORC1-Rag complex then anchors to a protein

complex called Ragulator, essential for Rag tethering to the lysosomal surface and amino acids-

mediated activation of mTORC1 (Sancak et al., 2010). In addition to its role as a scaffold

protein, Ragulator serves as a GEF, activating RagA/B following amino acid stimulation (Bar-

Peled et al., 2012).

After the Rag-Ragulator complex brings mTORC1 to the lysosomal surface, Rheb can bind to

and activate mTORC1 (Kim et al., 2008b; Sancak et al., 2008). RagA/B gets inactivated by a

tumor suppressor protein complex named GATOR that has a GAP activity (Bar-Peled et al.,

2013; Panchaud et al., 2013) while RagC/D are deactivated by the tumor suppressor folliculin1/2

which possesses GAP activity (Tsun et al., 2013). While amino acids can activate mTORC1

independently of TSC1/2 complex (Smith et al., 2005a), new reports show that TSC2 plays an

essential role in deactivation of mTORC1 following amino acid withdrawal (Demetriades et al.,

2014; Menon et al., 2014).

In addition to the Rag GTPases, several pathways have been linked to amino acid signaling

upstream of mTORC1, including mitogen-activated protein kinase kinase kinase kinase 3

(MAP4K3) (Findlay et al., 2007; Yan et al., 2010), the PI3K catalytic subunit type 3 (VPS34)

(Gulati et al., 2008; Nobukuni et al., 2005), the autophagy protein p62 (Duran et al., 2011),

leucyl-tRNA synthetase (Han et al., 2012) and inositol polyphosphate multikinase (IMP) (Kim et

35

al., 2011). The mechanisms by which these proteins signal to and interact with the Rag-

Ragulator complex are not fully characterized.

Amino acids are not the only nutrients that can activate mTORC1. Genetic studies suggest a role

for Rags in mTORC1 sensing of glucose. Knock-in embryos expressing constitutively GTP-

bound RagA (RagAGTP

) are resistant to amino acid or glucose withdrawal effects on mTORC1

signaling (Efeyan et al., 2013), and glucose, like amino acids, controls mTORC1 recruitment to

the lysosomal surface (Efeyan et al., 2013). In addition, phosphatidic acid has been reported to

activate mTORC1 (Fang et al., 2001). Indeed, over expressing the phosphatidic acid-producing

enzymes phospholipase D1 and D2 activates mTORC1 (Fang et al., 2003).

1.4.2.2. Downstream effectors of mTORC1

mTORC1 phosphorylates a myriad of downstream targets (Hsu et al., 2011; Yu et al., 2011). For

the sake of conciseness, the focus will be on the most characterized effectors of mTORC1 and

how they mediate their cellular functions.

mTORC1 enhances protein synthesis through phosphorylating S6K1 and the eukaryotic

translation initiation factor (eIF4E)-binding protein (4E-BP1). Both proteins associate with the

control of mRNA translation initiation and regulation of the protein synthesis rate (Ma and

Blenis, 2009). Following its phosphorylation by mTORC1, 4E-BP1 dissociates from eIF4E,

enabling it to form the eIF4E initiation complex necessary for initiation of cap-dependent

translation (Gingras et al., 1999; Haghighat et al., 1995; Hara et al., 1997). On the other hand,

36

S6K phosphorylates ribosomal protein S6 (component of the 40S ribosomal subunit) at five

residues (Ferrari et al., 1991). Knock-in mice with the five Ser targets in S6 mutated to Ala are

defective for ribosomal biogenesis and mRNA expression (Ruvinsky et al., 2005). In addition,

phosphorylated S6K1 enhances mRNA translation initiation and progression by associating with

several downstream targets including eukaryotic elongation factor 2 kinase (eEF2K) (Wang et

al., 2001), S6K1 Aly/REF-like target (SKAR) (Ma et al., 2008), eIF4B (Wilson et al., 2000),

among others. In addition, S6K1 upregulates the transcriptional activity of RNA polymerase I

and, hence, increases the expression of ribosomal RNA (rRNA) (Mayer et al., 2004).

In addition to protein synthesis, mTORC1 regulates autophagy, a catabolic cellular recycling

process that involves cell degradation of dysfunctional cellular components through the actions

of lysosomes (Rabinowitz and White, 2010). In mammalian cells, mTORC1 phosphorylates and

inhibits ULK1/Atg13/FIP200 kinase complex required for autophagy initiation (Ganley et al.,

2009; Hosokawa et al., 2009; Jung et al., 2009). In addition, mTORC1 directly phosphorylates

and inhibits the transcription factor EB (TFEB), a key regulator of lysosomal biogenesis (Pena-

Llopis et al., 2011; Settembre et al., 2012). By inhibiting autophagy, mTORC1 ensures the

utilization of available cellular energy resources towards cell growth.

1.4.3. mTORC2

mTORC2 functions as an important regulator of the actin cytoskeleton through activating the

Rac1 and Rho GTPases, and PKCα (Jacinto et al., 2004; Laplante and Sabatini, 2012; Sarbassov

37

et al., 2004). In addition, mTORC2 phosphorylates Akt at Ser473 (Sarbassov et al., 2005) as well

as Thr450 (Oh et al., 2010), leading to full activation of Akt.

1.4.3.1. Upstream regulators of mTORC2

Compared to mTORC1, little is known about upstream activators of mTORC2. The complex

responds to stimulation by growth factors such as insulin, but not to amino acid treatment

(Nicklin et al., 2009; Nobukuni et al., 2005). Insulin-dependent activation involves a physical

interaction of mTORC2 with ribosomes in a PI3K-dependent fashion (Zinzalla et al., 2011).

TSC1/2 might also be involved in proper activation of mTORC2 and its downstream target Akt

(Huang et al., 2008).

1.4.3.2. Downstream effectors of mTORC2

The mTORC2 complex phosphorylates and activates several members of the AGC family

kinases including Akt (Sarbassov et al., 2005), PKC (Ikenoue et al., 2008), and serum- and

glucocorticoid-induced protein kinase 1 (SGK1) (Garcia-Martinez and Alessi, 2008) in a motif

required for maximum kinase activation. Knocking down mTORC2 inhibits the phosphorylation

of some Akt targets, such as FoxO1/3a, but not other targets including TSC1/2 (Guertin et al.,

2006; Jacinto et al., 2006). Mechanisms regulating mTORC2-Akt substrate specificity are

currently unknown. Unlike Akt, SGK1 activity is completely abolished by mTORC2 loss, and

phosphorylation of NDRG1/2 (N-myc downstream regulated gene 1 and 2), physiological

substrates of SGK1, is dimished in Rictor- or Sin1-deficient fibroblasts (Garcia-Martinez and

38

Alessi, 2008). Finally, mTORC2-mediated phosphorylation of PKC plays a role in regulating

actin cytoskeleton and cell morphology (Jacinto et al., 2004; Sarbassov et al., 2004).

1.4.4. mTOR and metabolism

As mentioned earlier, mTOR integrates nutrient availability and signals from growth factors,

playing a central role in regulating metabolism in cells (Shimobayashi and Hall, 2014). In

animals, the transition between fasting and fed states changes the circulating amounts of

nutrients and growth factors. In turn, these changes affect how tissues balance catabolic and

anabolic processes. In fed states, high levels of nutrients and growth factors increase lipogenesis

in adipose tissues, glycogenesis in liver and muscle, and inhibit gluconeogenesis in liver.

1.4.4.1. Glucose metabolism

Transcriptomic analysis of Tsc1 and Tsc2 KO MEFs revealed that mTORC1 increases

transcription of genes involved in glycolysis, the pentose phosphate pathway and lipogenesis

(Duvel et al., 2010). Metabolomic analysis of the same cells mirrored changes observed at the

transcriptional levels. Chronic activation of mTORC1 in Tsc2 KO MEFs enhanced glucose

uptake, lactate secretion, glycolytic metabolism, and biosynthesis of the pentose phosphate

pathway and lipid intermediates. Rapamycin treatment of Tsc2 KO MEFs identified nine

metabolites positively regulated by mTORC1 (Ben-Sahra et al., 2013). Consistent with previous

observations (Duvel et al., 2010), most of these metabolites were part of the pentose phosphate

pathway. mTORC2 has also been shown to regulate glucose metabolism. Liver-specific Rictor

knockout mice displayed reduced glucokinase activity in the liver, leading to constitutive

39

gluconeogenesis and impaired glycolysis, and expression of constitutively active Akt2 in

mTORC2-deficient hepatocytes restored both glucose flux and lipogenesis (Hagiwara et al.,

2012). This suggests that mTORC2 regulates hepatic glucose via insulin-induced Akt signaling.

In addition, mTORC2-Akt signaling is required for Wnt3A-mediated reprogramming of glucose

metabolism towards glycolysis (Esen et al., 2013). In summary, these observations confirm that

both mTORC1 and mTORC2 are involved in regulating glucose metabolism.

1.4.4.2. Lipid synthesis

mTORC1 promotes lipogenesis through activating the sterol regulatory element-binding protein

1/2 (SREBP1/2) transcription factors which control expression of several fatty acid and

cholesterol synthesis genes. In response to insulin or sterol deficiency, SREBPs undergo

proteolytic cleavage and the active form translocates to the nucleus to activate transcription.

mTORC1 induces SREBP1/2, activating the expression of lipogenesis genes in both S6K-

dependent and independent mechanisms (Duvel et al., 2010; Porstmann et al., 2008). mTORC2

also plays a role in lipogenesis. SREBP1c activity is reduced in the liver-specific Rictor

knockout mice and is restored following the expression of constitutively active Akt2 (Hagiwara

et al., 2012). Therefore, mTORC1, mTORC2 and Akt are required for activation of SREBP and

lipogenesis.

1.4.4.3. Protein synthesis

As discussed in a previous section, mTORC1 plays an integral role in ribosomal biogenesis and

mRNA translation (Ma and Blenis, 2009). In contract to mTORC1, the role of mTORC2 in

40

protein synthesis is less-defined. Recently, it has been shown that mTORC2 associates with

actively translating ribosomes to phosphorylate its substrates. For example, mTORC2 co-

translationally phosphorylates Akt at Thr450 which prevents ubiquitilation and increases the

stability of Akt (Oh et al., 2010). mTORC2 also co-translationaly phosphorylates IGF2 mRNA-

binding protein 1 (IMP1), strongly enhancing IMP1 binding to the IGF2 mRNA translational

initiation region. This increases IGF2 translation, stimulating cell growth (Dai et al., 2013). The

molecular mechanism of mTORC2-ribosome association is still unclear.

1.4.4.4. Nucleotide metabolism

mTORC1 stimulates the expression of the pentose phosphate pathway which provides the ribose

moiety for nucleotide synthesis (Duvel et al., 2010). Metabolomic and phosphoproteomic

analyses revealed that mTORC1 stimulates de novo pyrimidine biosynthesis in S6K1-dependent

mechanism (Ben-Sahra et al., 2013; Robitaille et al., 2013). This occurs through mTORC1-

mediated phosphorylation and activation of the carbamoyl-phosphate synthetase 2, aspartate

transcarbamylase, and dihydroorotas (CAD) protein, which catalyzes the first steps in the

synthesis of the pyrimidine ring. The phosphorylation of CAD at Ser1859 by mTOR stimulates

its oligomerization and activation (Robitaille et al., 2013). Whether mTORC1 also controls

purine synthesis and whether mTORC2 plays a role in purine and pyrimidine metabolism need

further investigation.

41

1.5. Rationale and objectives of the study

A common mechanism through which activated growth factor receptors control signaling

specificity is by recruiting adaptor proteins. In this thesis, I address the biological role of the

p66Shc adaptor isoform in connecting growth factor signaling to metabolim.

p52Shc is a well-characterized adaptor protein that acts downstream of receptor tyrosine kinases

to amplify signaling to the Ras-Erk and PI3K-Akt pathways. However, the Shc1 locus also

encodes a p66Shc isoform, identical to p52Shc with the exception of an N-terminal extension.

This isoform is only found in vertebrates and it regulates insulin sensitivity. In mammals, lack of

p66Shc confers resistance to oxidative stress-induced vascular damage, hyperglycemia-induced

endothelial dysfunction, diabetes and obesity. These observations raise the possibility that

p66Shc provides negative feedback regulation to insulin signaling, and here I test the hypothesis

that, in this capacity, p66Shc also suppresses glucose metabolism.

The objectives of this thesis are to:

1) Analyze the effect of p66Shc on cellular metabolism using a targeted mass spectrometry-

based metabolomics approach.

2) Determine metabolic pathways affected by p66Shc expression.

3) Explore potential molecular mechanisms by which p66Shc affects energy metabolism.

42

My data suggest that loss of p66Shc redirects carbon metabolism in favor of glycolysis in a

manner reminiscent of the Warburg effect. Specifically, silencing of p66Shc improves glucose

uptake, increases carbon flux towards lactate production, and redirects glucose carbon towards

anabolic pathways, including fatty acid biosynthesis, the pentose phosphate and hexosamine

pathways, resulting in increased cell size. Mechanistically, the heightened glycolytic metabolism

in p66Shc-deficient cells is partly rescued upon mTOR inhibition by rapamycin. These results

indicate that the Shc1 locus not only regulates mitogenic signaling, but also modulates anabolic

metabolism in an mTOR-dependent manner through the p66Shc isoform. The approach used in

this thesis could be of general utility in dissecting the roles of adaptor proteins regulating

metabolism in normal and disease states.

43

Chapter 2. Materials and Methods

44

2.1. Cell culture and treatments

HeLa cells and p66Shc knockout MEFs were cultured in Dulbecco’s Modified Eagle’s Medium

(Gibco) supplemented with 10% fetal bovine serum (HyClone), 4mM L-glutamine and 1mM

sodium pyruvate. Cells were maintained in a humidified atmosphere of 5% CO2 and 95% air at

37°C and media were changed every 2 days. Plastic tissue culture plates were supplied by Fisher.

For growth factor treatment, cells were serum-starved for 4 h and treated with a final

concentration of 100 ng/ml of EGF, IGF (Peprotech) or insulin (Novo Nordisk) or 10% serum

for 10 min. For amino acid stimulation, almost confluent cell cultures in 6-well plates were

rinsed once with amino acid-free RPMI (US Biological), incubated in amino acid-free RPMI for

50 min and stimulated with 10X amino acid mixture for 10 min before lysis. For rapamycin

treatment, rapamycin (Calbiochem) was dissolved in ethanol and cell were treated with a final

concentration of 20 nM for the times indicated in the figures. For Akt inhibition, cells were

treated with Akt Inhibitor IV (EMD Millipore) at a final concentration of 10 nM for the times

indicated in the experiments.

2.2. Freezing and thawing of cells

Cryopreservation was done by harvesting cells using trypsin-EDTA (Gibco-BRL) treatment

followed by centrifugation at 1200 rpm for 5 min. Cells were then resuspended in a medium

containing 10% FBS and 5-10% sterile dimethylsulfoxide (Sigma) to yield approximately l-2 X

l06 cells/ml. 1 ml aliquots of cell suspension were transferred to cryovials (Corning) and vials

were placed at -80°C overnight. Frozen cells were then placed in liquid nitrogen for long-term

storage. To thaw cells, vials of frozen cells were removed from liquid nitrogen and placed in a

37°C water bath for 2 min. The thawed cell suspension was quickly transferred to 15 ml tubes

(BD Falcon) containing fresh culture media supplemented with 10% FBS, centrifuged at 1200

45

rpm for 5 min, resuspended in 10 ml of culture media/10% FBS, transferred to 10 cm plates

(Corning) then incubated at 37°C in an atmosphere of 5% CO2.

2.3. Cell culture

Cells reaching confluence were incubated in a trypsin-EDTA solution at 37°C in a 5% CO2

incubator for 2-5 min until they lose anchorage from the tissue culture plates. Cells were

harvested by triturating with an appropriate volume of DMEM containing 10% FBS. Cells were

seeded onto fresh tissue culture dishes. Uniform spreading of cells was achieved by gently

shaking plates containing medium and cells in cross-shape direction.

2.4. Plasmid preparation and DNA constructs

Murine p66Shc cDNA (NCBI refseq number NP_001106802) was cloned in a gateway-

compatible pMX retroviral puromycin-resistant destination vector according to the

manufacture’s recommendation (Invitrogen). For p66Shc knockdown in HeLa cells, the

following hairpin sequence was cloned into a pMSCV retroviral puromycin-resistant vector

(with 3' LTR removed): sense 5'-cgg aat gag tct ctg tca tcg ct (tt)-3' and anti-sense 5'-ag cga tga

cag aga ctc att ccg (tt)-3'. Small scale and large scale plasmid preparations were done using the

Invitrogen plasmid preparation kits. These plasmid purification protocols are based on a

modified alkaline lysis procedure, followed by binding of plasmid DNA to an anion-exchange

resin under appropriate low-salt and pH conditions. RNA, proteins, dyes and low molecular

weight impurities were removed by a medium salt wash. Plasmid DNA was then eluted in a

high-salt buffer and then concentrated and desalted by isopropanol precipitation.

46

2.5. Cell transfection

Cells were seeded onto tissue culture plates at 40-50% confluence. 18-24 h later, when cells

reached 80-90% confluence, cells were transiently transfected using Lipofectamine 2000

transfection reagent (Invitrogen) or Fugene 6 (Roche). For Lipofectamine transfections, cells

were washed with PBS and media was replaced with serum-free medium (Opti-MEM media).

Plasmid DNA and Lipofectamine reagent were then mixed separately in antibiotic- and serum-

free medium for 5 min at room temperature then mixed together for 20 min with agitation at

room temperature. The ratio of µg of DNA to µL of Lipofectamine was 1:2. After 20 min of

incubation, the DNA: Lipofectamine complexes were added drop-wise to cells. Growth medium

was then replaced after incubation for 4-6 h at 37°C with fresh media containing 10% FBS. For

Fugene 6 transfections, media were replaced with fresh antibiotic free FBS-containing media.

Fugene 6 was added to antibiotic and serum-free media and incubated for 5 min. DNA was

added to the Fugene 6-media mixture and incubated for 20 min at room temperature. The ratio of

DNA: Fugene 6 used was 1:2. The mixture was added drop-wise to the plate of cells containing

media without antibiotics and distributed evenly by gently shaking the plate forth and back

several times. The cells were incubated at 37°C in a 5% CO2 incubator and harvested at the time

points indicated in each experiment. Amounts of DNA and Lipofectamine used were according

to the manufacture’s recommendation.

2.6. Retroviral production and cell infection

Vectors were packaged as recombinant retrovirus and pseudotyped – where the enveloped virus

particle gets assembled with a foreign viral glycoprotein - with VSV-G protein to enhance

infectivity. Either control shGFP or shp66Shc was co-transfected with pCMV-VSV-G into the

retroviral packaging cell line Platinum E (modified HEK293 cells) (Morita et al., 2000). Viral

47

supernatants from 10 cm plates were collected at 48 h posttransfection and filtered using 0.45

µm filters. Cells were infected with viral supernatant where media were supplemented with 10

μg/ml of hexadimethrine bromide (Sigma-Aldrich). 48 h post-infection, infected cells were

treated with puromycin (1 μg/ml) to generate cell lines stably expressing the constructs of

interest.

2.7. Metabolite extraction

The relative levels of metabolites were determined in HeLa cells and MEF using the protocol

indicated in (Abdel Rahman et al., 2013). Cells were seeded in 6-well plate in 6 replicates. After

24 h, the media were removed and the cells were washed 2X with warm PBS, then were placed

on dry ice. The metabolites were immediately extracted by adding 1 ml of dry ice-cold extraction

solution (40% acetonitrile, 40% methanol, and 20% water) containing internal standards

(500µg/ml and 300µg/ml of D7-Glucose and 13C915N-Tyrosine, respectively). Cells were

scraped and collected in 1.5 ml eppendorff tubes, and shaken for 1 h at 4°C/1400 rpm in a

Thermomixer (Eppendorf). Samples were centrifuged at 14000 rpm, for 10 min at 4°C, and

supernatants were transferred into fresh tubes to be evaporated to dryness in a CentreVap

concentrator at 40°C (Labconco). The dry extract samples were stored at -80°C until LC-MS

analysis. The dry metabolite extracts were reconstituted in 100 µL of water. Samples were

injected twice through the HPLC (Dionex Corporation) in gradient reversed phase column

Inertsil ODS-3, 4.6 mm internal diameter, 150 mm length, and 3 µM particle size for positive

and negative mode analysis. Automated washing procedure was used between samples. Eluted

metabolites were analyzed in MRM mode on electrospray ionization (ESI) triple-quadrupole

mass spectrometer (ABSciex4000Qtrap). Signals were normalized to both internal standard and

cell number. Metaboanalyst was used to analyze the data (Xia and Wishart, 2011). The LC-

48

MS/MS system does not resolve hexose and hexosamine isomers including glucose/galactose

and GlcNAc/GalNAc. To monitor trends in metabolic pathways, we referred to these isomers in

their glucose (Glc) forms.

2.8. Isotope labeling and kinetic profiling

Cells were seeded a day before the experiment in 6-well plates. Media of cell cultures in 6-well

dishes were replaced with unlabeled fresh DMEM containing 5% dialyzed FBS 2 h before

exposure to labeled glucose to allow equilibration to the new conditions. The media was then

replaced with DMEM/5% dialyzed FBS containing 25 mM [1,2-13

C2]-labeled glucose or 15

N2-

labeled glutamine (Cambridge Isotope Laboratories)for the times indicated in the figures

legends. Metabolites were immediately extracted with dry ice-cold extraction solution indicated

previously and analyzed using different MRM transitions that were developed for 1,2-13

C2-

glucose as detailed in Appendix 6.2.

2.9. [3H]-2-deoxy-D-glucose uptake assay

HeLa cells and MEF grown in 6-well plates were rinsed with HEPES-buffered saline [140 mM

NaCl, 20 mM HEPES, 5 mM KCl, 2.5 mM MgSO4, 1 mM CaCl2 (pH 7.4)]. [3H]-2-deoxy-D-

glucose (2-DG) uptake was performed for 3, 5, and 10 min in HEPES-buffered saline containing

2 mM unlabeled 2-DG and 0.2 μCi/ml [3H] 2-DG at room temperature. The reaction was

terminated by washing three times in ice-cold 0.9% NaCl (w/v). To quantify the radioactivity

incorporated, cells were lysed with 0.05 N NaOH and lysates were counted with scintillation

fluid in a β-counter. Nonspecific uptake was determined in the presence of cytochalasin B (10

μM) during the assay. The results are expressed as pmole 2-DG transported per min/mg of

49

protein. This experiment was done by Huogen Lu in Dr. George Fantus’ lab (Lunenfeld Research

Institute, University of Toronto).

2.10. Oxygen consumption rate measurement

Cells were seeded at 50,000 cells/well in XF-24 cell culture plates. The day after cell seeding,

cell culture media was replaced with RPMI and oxygen consumption was measured using an XF-

24 Flux Analyzer as described in (Birsoy et al., 2013). This experiment was done by Kıvanç

Birsoy in Dr. David Sabatini’s lab (Whitehead Institute, MIT).

2.11. Cell size determination

Cells were grown to 100% confluence in 10 cm plates, washed with 1X PBS, treated with 1 ml

of trypsin, and diluted 1:20 with counting solution (Isoton II Diluent, Beckman Coulter). Cell

diameters and volumes were determined using a particle size counter (Coulter Z2, Beckman

Coulter) with Coulter Z2 AccuComp software. For cell size determination using flow cytometry,

cells were washed with 1X PBS, treated with 1 ml of trypsin, suspended in 9 ml of DMEM/10%

FBS media, and centrifuged at 1200 rpm for 5 min. Media was discarded and cells were

resuspended in a mixture of 1 ml of 2% FBS/ 5 mM EDTA in PBS, filtered through the mesh cap

of BD Falcon polypropylene tubes. Cells were analyzed by flow cytometry for differences in

forward scatter (FSC correlates with cell size).

2.12. Cell lysis and immunoblotting

Cells growing in 6-well plates were rinsed once with 1X cold PBS and lysed on ice with 300 μl

of ice-cold Lysis Buffer (40 mM HEPES pH 7.5, 120 mM NaCl, 1 mM EDTA, 10 mM

pyrophosphate, 10 mM glycerophosphate, 10 mM NaF, 1 mM orthovanadate, and EDTA-free

protease inhibitors (Roche)) containing 0.5% NP-40. Cell suspensions were then transferred into

50

1.5 ml eppendorf tubes, centrifuged at 13,000 rpm for 10. 250 μl of cell supernatent was

transferred to a new tube and protein content was measured used Bradford reagent according to

the manufacturer’s instructions (Pierce). After normalization of protein concentration, 250 µL of

2X SDS sample buffer (2% SDS, 20% glycerol, 20 mM TrisCl pH 8.0, 4% β-mercaptoethanol,

2mM EDTA, and 0.02% bromophenol blue) was added and cell lysate was resolved by 8%-16%

SDS-PAGE gels.

2.13. Western blotting

For western blots, denatured whole cell lysates in sample buffer were electrophoresed at 100 V

for 2 h (8%-16% SDS-PAGE, Bio-Rad apparatus) and transferred to polyvinylidene difluoride

(PVDF) membranes for 2 h at 50 mAmp/gel using semi-dry transfer apparatus (Bio-Rad).

Membranes were blocked in 5% w/v milk in TBS buffer/0.1% tween (TBST) or 5% w/v BSA in

TBST for 1 h followed by overnight incubation with primary antibody in the blocking buffer.

Antibodies used for western blots in the course of this study include phospho-T389 S6K1 (CST,

#9234), phosho-T308 Akt (CST, #2965), phospho-S473 Akt (CST, #4058), phospho-T346

NDRG1 (CST, #3217), Akt (CST, #9272), S6K1 (CST, #2708) from Cell Signaling Technology

(1:1000 dilution); mouse (BD, #610878) and rabbit (BD, # 610082) Shc1 antibodies from BD

Biosciences (1:1000 dilution); FLAG M2 (#F3165, 1:1000 dilution) and tubulin (#T6793,

1:10,000 dilution) antibodies from Sigma-Aldrich. Horseradish peroxidase (HRP)-conjugated

secondary antibodies were used for western blotting (Santa Cruz Biotechnology, 1:2000 for goat

α-rabbit and 1:5000 for goat α-mouse) and blots were developed using chemiluminescent

substrate (Thermo Scientific).

51

2.14. Mass spectrometry analysis of the p66Shc protein-interactions

The p66Shc protein interaction network was determined following a protocol described in

(Zheng et al., 2013). Briefly, p66+ MEFs (expressing 3XFLAG-p66Shc) were washed with ice-

cold PBS and lysed in NP40 lysis buffer (50 mM HEPES pH 8, 150 mM NaCl, 1 mM EGTA,

0.5% NP40, 100 mM NaF, 2.5 mM MgCl2, 10 mM Na4P2O7, 1 mM DTT, 10% glycerol)

supplemented with protease and phosphatase inhibitors (50 mM β-glycerolphosphate, 10 µg/ml

aprotinin, 10 µg/ml leupeptin, 1 mM Na3VO4, 100 nM calyculin A, 1 mM PMSF

(phenylmethylsulphonyl fluoride)). The total cell lysates were centrifuged at 10,000 rpm for 15

min to pellet the nuclei and insoluble material. Nuclear-free lysates were pre-cleared by 1 h

incubation with protein A sepharose and normalized for total protein concentration using the

Bio-Rad protein assay. 3xFLAG-p66Shc was immunoprecipitated by incubating lysates with 5 µl

(bed volume) anti-Flag M2 antibody-conjugated agarose for 4 h at 4 °C. Beads were washed 3X

with lysis buffer, then 2X with 50 mM ammonium bicarbonate. After aspiration of washing

buffers, proteins on beads were treated with 0.5 µg trypsin (Promega) overnight at 37ºC. Beads

were centrifuged at 10,000 rpm for 2 min and trypsinized peptides were transferred to new 0.5

ml Eppendorf tubes and dried in a CentriVap at 40°C for 3 h (Labconco). Dried peptides were

kept at -80 ºC until LC-MS analysis. Dried tryptic samples were reconstituted with 3% formic

acid. Samples were loaded onto a 75 mm inner diameter (ID)/360 mm outer diameter (OD)

pulled tip packed with 3 mm ReproSil C18 and analysed on an TripleTOF 5600 mass

spectrometer (AB SCIEX) coupled to an Eksigent nano LC Ultra 1D plus pump with a flow rate

of 200 nl/min and a gradient of 2% to 35% acetonitrile over 90 min. A cycle time of 1.3 sec was

employed using a survey TOF scan of 250 msec at 30,000 resolution followed by selection of the

top 20 most intense peptides for MS/MS for 50 msec each with high sensitivity (at 18,000

52

resolution). Only peptides with a charge state above 11 were selected for MS/MS and dynamic

exclusion was set to 15 sec for all ions within 20 ppm.

2.15. RNA-seq

cDNA library was prepared using Illumina TruSeq RNA Sample Prep Kit v2 (Cat#RS-122-

2001). Briefly, 1 µg of high quality total RNA was isolated from p66Shc KO and p66+ MEFs to

generate the cDNA library according to the Illunima kit protocol. The generated barcoded cDNA

library has an average fragment size of 350-400 bp. This bar-coded library is quality checked

with Agilent Bioanalyzer and quantified with qPCR using KAPA SYBR FAST Universal 2x

qPCR Master Mix (Kapa Biosystem, Cat#KK4602). The quality checked libraries are then

loaded on a flowcell for cluster generation using Illumina c-Bot and TruSeq PE Cluster Kit v3

(Cat#: PE-401-3001). Sequencing was done on HiSeq2000 with TruSeq SBS Kit v3 (pair-ended

200 cycles, Cat#: FC401-3001). The real-time base call (.bcl) files are converted to fastq files

using CASAVA 1.8.2 (on CentOS 6.0 data storage and computation linux servers).

53

Chapter 3. p66Shc Inhibits Anabolic Metabolism

A version of this chapter appeared in the following article:

The Adaptor Protein p66Shc Inhibits mTOR-Dependent Anabolic Metabolism.

Science Signaling, 7, ra17 (2014).

Mohamed A. Soliman, Anas M. Abdel Rahman, Dudley W. Lamming, Kivanç Birsoy, Judy

Pawling, Maria E. Frigolet, Huogen Lu, I. George Fantus, Adrian Pasculescu, Yong Zheng,

David M. Sabatini, James W. Dennis, and Tony Pawson.

Experiments in Fig. 3.9 were done by Kivanç Birsoy

Experiments in Fig. 3.14 and 3.16 were done by Huogen Lu

54

3.1. Background

There are three isoforms of Shc1 with molecular masses of 46, 52, and 66 kDa. Shc1 adaptor

proteins mediate signal transduction, linking multiple tyrosine kinase growth factor receptors to

activation of the Ras-MAPK and PI3K-Akt pathway (Ravichandran, 2001). Like p52/p46shc,

p66Shc is tyrosine-phosphorylated following EGF stimulation, binds to activated EGFRs and

forms stable complexes with the adaptor protein Grb2 (Migliaccio et al., 1997). However, unlike

p52/46Shc, p66Shc is unable to transform mouse fibroblasts in vitro, and does not increase

MAPK activation following EGF stimulation (Migliaccio et al., 1997). In addition, it competes

with p52Shc for Grb2 binding, inhibiting signaling downstream of IGF-1 receptor in vascular

smooth muscle cells (Xi et al., 2008). p66Shc has a unique role amongst the other Shc1 isoforms

in promoting oxidative stress and pro-apoptotic signaling. p66Shc-deficient MEFs shows

enhanced resistance to apoptosis in response to hydrogen peroxide treatment (Migliaccio et al.,

1999). Knockout studies suggest a pivotal role for p66Shc in regulating insulin signaling and

glucose metabolism. Deletion of p66Shc in mice improves glucose tolerance and insulin

sensitivity at organism and tissue levels (Tomilov et al., 2011). Suggestive of an inhibitory role

in insulin signaling and glucose metabolism, deletion of p66Shc caused a marked increase in

basal glucose transport in skeletal muscle cell lines (Natalicchio et al., 2009). These observations

led me to investigate the role of p66Shc as a repressor of glucose metabolism in particular and

anabolic metabolism in general. To test this possibility, I generated HeLa cell lines stably

expressing shRNA against control or p66Shc and examined the change in glycolytic metabolism

in cells that are competent or deficient in p66Shc expression. To test whether p66Shc is

sufficient to alter glycolytic metabolism, I did “add-back” experiments where I performed

metabolomic analysis of p66Shc KO MEFs stably expressing GFP or 3xFLAG-p66Shc.

55

3.2. Loss of p66Shc enhances glycolytic metabolism

To elucidate the function of p66Shc in cellular metabolism, I performed a targeted metabolomic

analysis using liquid chromatography-tandem mass spectrometry (LC-MS/MS) in multiple

reaction monitoring (MRM) mode. Approximately 250 metabolites were measured in positive or

negative mode runs (Appendix 6.1). Metabolite spectral patterns were validated using standard

mixture of metabolites. To assess the role of p66Shc in cancer cell metabolism, HeLa cells were

stably transfected with either a short hairpin shRNA that specifically targets the isoform p66Shc

or a control hairpin (Fig. 3.1) (Kisielow et al., 2002). The profiles of metabolites extracted from

the two cell types were analyzed. Loss of p66Shc resulted in increased abundance of

intermediates of glucose metabolism (Fig. 3.1). Specifically, p66Shc deficiency was

accompanied by significant increases in glucose-6-phosphate (G6P), and downstream glycolysis

intermediates including fructose-1,6-bisphosphate (F1,6BP), phosphoenolpyruvate (PEP), and

pyruvate (Fig. 3.2). p66Shc-deficient HeLa cells produced more lactate than control cells,

consistent with the Warburg effect. The metabolic shift also redirects glucose-derived citrate

from the TCA cycle into lipid synthesis to generate biomass (Hatzivassiliou et al., 2005).

Consistent with this effect, p66Shc-deficient cells had higher citrate concentrations (Fig. 3.2).

These data suggest that depletion of p66Shc is sufficient to enhance anabolic metabolism in

HeLa cells.

3.3. Loss of p66Shc promotes glucose metabolism through the pentose phosphate and

hexosamine biosynthesis pathways

In addition to the role of these molecules in glycolysis, G6P and F6P are precursors of the

hexosamine biosynthesis and the pentose phosphate pathways, which are essential anabolic

56

pathways in proliferating cells. The hexosamine biosynthesis pathway provides UDP-GlcNAc as

a substrate for O-GlcNAcylation of cytosolic proteins, and N-linked and O-linked glycosylation

of proteins produced in the secretory pathway (Dennis et al., 2009). The pentose phosphate

pathway provides ribose for nucleic acid synthesis and NADPH to maintain the reductive

environment of the cell. In HeLa cells lacking p66Shc, a ~ 4-fold increase in N-

acetylglucosamine-6-phosphate (GlcNAcP), and a ~ 2-fold increase in UDP-GlcNAc abundance,

the major products of the hexosamine biosynthesis pathway, was observed (Fig. 3.3A). A 2-3

fold increase in UDP-GlcNAc enhances N-glycosylation of growth factor receptors, thereby

promoting positive feedback to signaling (Lau et al., 2007; Wellen et al., 2010). Concentrations

of ribose-5-phosphate (R5P) and xylulose-5-phosphate (X5P) were also increased, which may

involve the oxidative pentose phosphate pathway that contributes to redox balance (Fig. 3.3B).

We could not determine whether p66Shc loss enhances either or both of the oxidative and non-

oxidative branches of the pentose phosphate pathway as the mass spectrometry signals for 6-

phosphogluconolactone or 6-phosphogluconate, the main metabolites of the oxidative branch,

were too weak to be accurately measured.

3.4. Restoring p66Shc expression inhibits glycolytic metabolism

To test whether p66Shc was sufficient for enhanced glycolytic metabolism, immortalized p66Shc

KO MEFs were stably infected with either GFP or 3xFLAG-p66Shc retroviral vectors (hereafter

referred to as KO and p66+ MEFs, respectively) (Fig. 3.4). Principal component analysis (PCA)

was used to evaluate replicate consistency, and revealed a notable difference in the overall

metabolic profiles of the p66Shc KO and p66+ cells (Fig. 3.5). Consistent with the p66Shc-

mediated effects in HeLa cells, F6P and G6P were among the metabolites with the most

57

significant changes, and the abundance of glycolytic intermediates was generally lower in p66+

MEFs compared to the KO cells (Table 1). The abundance of glycolysis intermediates changed

to a similar extent in MEFs as in HeLa cells; we observed ~ 3 fold decrease in G6P

concentrations, and a concomitant decrease in downstream three-carbon glycolytic metabolites

including PEP and lactate in p66+ cells (Fig. 3.6).

The lower concentration of citrate in p66+ cells was accompanied by a concomitant decrease in

the amounts of the fatty acid synthesis precursors acetyl-CoA and malonyl-CoA (Fig. 3.7).

Malonyl-CoA is a high energy and committed intermediate in the fatty acid biosynthesis

pathway, thus exclusively an anabolic metabolite. These results suggest that p66Shc inhibits de

novo lipid biosynthesis. In addition, the abundance of intermediates in the hexosamine

biosynthesis and pentose phosphate pathways was decreased in p66+ MEFs (Fig. 3.8). This

profile confirms that p66Shc deficiency in non-transformed cells enhanced glycolysis at the

expense of decreasing oxidative mitochondrial metabolism. Indeed, p66Shc deficient MEFs

displayed lower oxygen consumption (Fig. 3.9A) but improved energy utilization (AMP/ATP)

(Fig. 3.9B).

3.5. p66Shc expression inhibits amino acid biosynthesis and pyrimidine metabolism

The levels of amino acids were measured in p66Shc-competent and p66Shc-deficient cells.

Higher amounts of non-essential amino acids, including alanine, serine and aspartate, were

detected in p66Shc-deficient MEFs (Fig. 3.10). To determine whether the effect of p66Shc

signaling on the steady state abundance of nonessential amino acids reflects regulation of

metabolic flux through de novo synthesis, we measured the relative flux with a pulse of stable-

isotope labeled 15

N2-glutamine. Increased incorporation of labeled nitrogen into nonessential

58

amino acids was detected in p66Shc KO MEFs (Fig. 3.11). Increased flux of nitrogen from

extracellular glutamine into nonessential amino acids in p66Shc KO MEFs is consistent with

utilization catabolism of glutamine and reprogramming that favors anabolism. Pyrimidine

derivatives (dCTP and UTP) were among the top 10 metabolites whose abundance was

significantly decreased in p66+ MEFs (Table 3.1). Pyrimidine nucleotides are required as high

energy donors for phospholipid and glycoconjugate biosynthesis. p66+ MEFs showed lower

amounts of the pyrimidine synthesis intermediates orotate and dihydroorotate as compared to

knockout MEFs (Fig. 3.12). These results confirm that p66Shc inhibits anabolic metabolic

pathways, and demonstrate a prominent role for p66Shc in reprogramming cell metabolism

towards glucose catabolism and oxidative respiration.

3.6. p66Shc regulates redox homeostasis

By associating with cytochrome c, p66Shc promotes the generation of ROS (Giorgio et al., 2005;

Pinton et al., 2007). Consistent with these findings, a higher ratio of reduced to oxidized

glutathione (GSH/GSSG) was observed in p66Shc KO MEFs compared to p66+ MEFs (Fig.

3.13A). The ratio of NADH/NAD+ also reflects the redox state of the cell (Fisher-Wellman and

Neufer, 2012), and a ~ 3-fold increase in NADH/NAD+ indicates a more reducing environment

in p66Shc-deficient cells (Fig. 3.13B). NADH is produced in the second half of glycolysis

during the glyceraldehyde-3-phosphate dehydrogenase (GAPDH)-catalyzed step, and this may

contribute to the altered NADH/NAD+ ratio. The higher GSH/GSSG ratio in KO MEFs is

consistent with the reported lower amounts of ROS in p66Shc KO MEFs (Giorgio et al., 2005).

A previous report has shown that ROS contribute to insulin resistance and decrease glucose

metabolism (Houstis et al., 2006).

59

3.7. p66Shc is necessary and sufficient to alter glucose uptake and metabolism

To directly test whether p66Shc alters glucose uptake, we measured 3H-labeled 2-deoxy-D-

glucose (2-DG) uptake at different time points. The rate of 2-DG uptake in p66Shc-deficient

HeLa cells and in p66Shc KO MEFs was greater than in p66Shc-expressing cells (Fig. 3.14).

Consistent with the enhanced glucose consumption and higher glycolytic metabolism in p66Shc-

deficient cells, increased concentrations of extracellular lactate were observed in the media of

HeLa cells and MEFs lacking p66Shc (Fig. 3.15). However, western blots showed no apparent

difference in the abundance of glucose transporter 1 (Glut1) between p66Shc-competent and

p66Shc-deficient cells (Fig. 3.16).

To monitor the effect of p66Shc on glucose catabolism, the metabolites of 13

C-labeled glucose in

control and p66Shc-deficient HeLa cells were examined. The metabolism of [1,2-13

C2] glucose

generates M0, and M2 and M4 mass isotopomers corresponding to ion fragments that contain

zero, two or four labeled carbons, respectively (Fig. 3.17A and Appendix 6.2). Rapid transfer of

cells from unlabeled medium to identical medium containing [1,2-13

C2]-labeled glucose ensures

that metabolism is minimally disturbed (Munger et al., 2008). Consistent with enhanced

glycolytic metabolism in p66Shc-depleted cells, the amounts of the M2-labeled form of G6P and

downstream intermediates including M2 pyruvate were increased in p66Shc-depleted HeLa cells

(Fig. 3.17B). p66Shc silencing enhanced the amount of labeled glucose-derived citrate nearly 2

fold, consistent with the redirection of glucose-derived carbons towards lipid biosynthesis (Fig.

3.17B). Furthermore, the levels of M2-labeled form of hexosamine pathway intermediates were

also higher in p66Shc-deficient HeLa cells (Fig. 3.18). Conversely, re-expressing p66Shc in

p66Shc KO MEFs inhibited the heightened abundance of [1,2-13

C2] labeled glycolytic

60

metabolites and downstream anabolic intermediates, including pyruvate and citrate, observed in

p66Shc-deficient cells (Fig. 3.19). Additionally, p66Shc inhibited the de novo synthesis of amino

acids (Fig. 3.20). Collectively, our data suggest that p66Shc is necessary and sufficient to

reprogram glucose utilization into anabolic pathways.

3.8. Lack of p66Shc enhances glycolytic flux and anabolic metabolism

To study the dynamics of incorporation of isotope-labeled glucose into downstream anabolic

metabolites, kinetics flux profiling in p66Shc KO and p66+ MEFs using [1,2-

13C2] labeled

glucose was performed by measuring the relative amounts of labeled glycolytic intermediates

over 6 time points (0, 1, 2.5, 5, 10, 15 min). Consistent with the steady state data, p66Shc

inhibited labeled-glucose flux into glycolytic intermediates such as pyruvate in p66+ MEFs (Fig.

3.21A). In addition, the amounts of labeled citrate, UDP-GlcNAc and R5P were lower in cells

expressing p66Shc (Fig. 3.21, B-D, Appendix 6.2). UDP-GlcNAc is a high-energy donor

required for protein glycosylation and an example of an anabolic metabolite where the GlcNAc

portion is not catabolized back to glucose (Wellen et al., 2010). Overall, these results suggest

that increased glycolytic flux in p66Shc-deficient cells accounts for the observed increase in

metabolites at steady state, consistent with a role for p66Shc as an inhibitor of glucose

catabolism.

61

A B

Fig. 3.1. Effect of p66Shc on glycolytic metabolism.

(A) Abundance of the three Shc1 isoforms in HeLa cells stably expressing shRNAs targeting

GFP or p66Shc. (B) Summary of the changes in intracellular metabolite amounts associated with

p66Shc knockdown in HeLa cells. Glucose-6-phosphate (G6P); fructose-6-phosphate (F6P);

fructose-1,6-phosphate (F1,6BP); dihydroxyacetone phosphate (DHAP); glyceraldehyde 3-

phosphate (GADP); 1,3-bisphosphoglycerate (1,3BPG); 3-phosphoglycerate (3PG);

phosphoenolpyruvate (PEP); acetyl-CoA(ACoA); ribose-5-phosphate (R5P); xylulose 5-

phosphate (X5P); glucosamine-6-phosphate (GlcN6P); N-acetylglucosamine phosphate

(GlcNAcP); uridine diphosphate N-acetylglucosamine (UDP-GlcNAc).

Glucose transporter

62

Fig. 3.2. Lack of p66Shc enhances glycolytic metabolism.

Fold change of glycolytic intermediates in p66Shc-competent and p66Shc-deficient HeLa cells

as measured by LC-MS/MS. Error bars represent SD of at least three biological replicates (* p <

0.05, ** p < 0.01, *** p < 0.001). Glucose-6-phosphate (G6P); fructose-1,6-phosphate (F1,6BP);

phosphoenolpyruvate (PEP).

63

A

B

Fig. 3.3. p66Shc deficiency increases the levels of the pentose phosphate and the

hexosamine pathway intermediates.

Fold change of the hexosamine biosynthesis (A) and the pentose phosphate (B) pathway

intermediates in p66Shc-competent and p66Shc-deficient HeLa cells as measured by LC-

MS/MS. Error bars represent SD of at least three biological replicates (* p < 0.05, ** p < 0.01,

*** p < 0.001). Fructose-6-phosphate (F6P); glucosamine-6-phosphate (GlcN6P); uridine

diphosphate N-acetylglucosamine (UDP-GlcNAc); ribose-5-phosphate (R5P); xylulose 5-

phosphate (X5P).

64

Fig. 3.4. Levels of Shc1 isoforms in p66Shc KO and p66+ MEFs.

Abundance of the three Shc1 isoforms in p66Shc KO MEFs stably infected with GFP (KO) or

3xFLAG-p66Shc (p66+).

65

Fig. 3.5. Unsupervised principal component analysis for targeted metabolomics screen in

p66Shc KO and p66+ MEFs.

Each dot represents a biological replicate. Red dots indicate replicates for p66Shc KO MEFs, and

green dots indicate replicates for p66+ MEFs. The analysis demonstrates clear separation of

overall metabolomic profiles of the two genotypes with statistical significance 99.7% (PC1) and

0.2% (PC2) (p < 0.05). The metabolites with the most significant changes in p66+ MEFs

compared to KO MEFs are summarized in Table 3.1.

66

Fig. 3.6. p66Shc expression decreases the levels of glycolytic intermediates.

Fold change of glycolytic metabolites in p66Shc-deficient (white) and p66Shc-competent (black)

MEFs as measured by LC-MS/MS. Error bars represent SD of at least three biological replicates

(* p < 0.05, ** p < 0.01, *** p < 0.001). Glucose-6-phosphate (G6P); fructose-1,6-phosphate

(F1,6BP); phosphoenolpyruvate (PEP).

67

Fig. 3.7. p66Shc inhibits fatty acid biosynthesis

Fold change of acetyl-CoA (ACoA) and malonyl-CoA (MCoA) in p66Shc-deficient (white) and

p66Shc-competent (black) MEFs as measured by LC-MS/MS. Error bars represent SD of at least

three biological replicates (* p < 0.05, ** p < 0.01, *** p < 0.001).

68

A

B

Fig. 3.8. p66Shc expression decreases the levels of the pentose phosphate and the

hexosamine pathway intermediates.

Fold change of the hexosamine biosynthesis (A) and the pentose phosphate (B) pathways

metabolic intermediates in p66Shc-deficient (white) and p66Shc-competent (black) MEFs as

measured by LC-MS/MS. Error bars represent SD of at least three biological replicates (* p <

0.05, ** p < 0.01, *** p < 0.001). Fructose-6-phosphate (F6P); glucosamine-6-phosphate

(GlcN6P); uridine diphosphate N-acetylglucosamine (UDP-GlcNAc); ribose-5-phosphate (R5P);

xylulose 5-phosphate (X5P).

69

A B

Fig. 3.9. Deficiency of p66Shc inhibits oxygen consumption rate and lowers AMP/ATP

ratio.

(A) Oxygen consumption rate in p66Shc-deficient (white) and p66Shc-competent (black) MEFs

were measured by XF-24 Flux Analyzer. Error bars represent SD of three biological replicates (*

p < 0.05). (B) Relative amounts of AMP/ATP ratio in p66Shc KO and p66+ MEFs by LC-

MS/MS. Error bars represent SD of three biological replicates (* p < 0.05).

Oxygen consumption rate

KO p66+ KO p66+

70

Fig. 3.10. p66Shc expression inhibits the synthesis of nonessential amino acids

Fold change of the amino acids alanine, serine and aspartate in p66Shc-deficient (white) and

p66Shc-competent (black) MEFs as measured by LC-MS/MS. Error bars represent SD of at least

three biological replicates (*** p < 0.001).

71

Fig. 3.11. Tracing of 15

N-labeled-amino acids in p66Shc KO and p66+ cells.

Fold change of intracellular amounts of 15

N-labeled alanine, serine and aspartate in p66Shc KO

(white) and p66+ (black) MEFs grown in media containing

15N-glutamine for 2h. Error bars

represent SD of at least three biological replicates (*** p < 0.001)

72

Fig. 3.12. p66Shc inhibits de novo pyrimidine synthesis intermediates.

Fold change of de novo pyrimidine synthesis intermediates, N-carbamoyl aspartate, orotate and

uridine monophosphate (UMP) in p66Shc KO (white) and p66+ (black) MEFs. Error bars

represent SD of at least three biological replicates (* p < 0.05).

73

A B

Fig. 3.13. p66Shc regulates redox homeostasis.

Relative amounts of the ratio of reduced/oxidized glutathione (GSH/GSSG) (A) and

reduced/oxidized NAD+ (NADH/NAD

+) (B) in p66Shc KO (white) and p66

+ (black) MEFs. Error

bars represent SD of at least three biological replicates (*** p < 0.001).

74

A

B

Fig. 3.14. Lack of p66Shc enhances 2-DG uptake.

Tracing of 3H-labeled-2-deoxy glucose (2-DG) uptake over time in p66Shc-competent and

p66Shc-deficient HeLa cells stably expressing shRNAs targeting GFP and p66Shc (A) and

p66Shc KO and p66+

MEFs (B). Error bars represent SD of at least three biological (p < 0.05).

75

A

B

Fig. 3.15. p66Shc inhibits cellular lactate secretion.

Fold change of extracellular lactate in media from p66Shc-competent and p66Shc-deficient

HeLa (A) and MEFs (B). Error bars represent SD of at least three independently prepared

samples (*** p < 0.001).

76

A

B

Fig. 3.16. Abundance of Glut1 in p66Shc-deficient and p66Shc-competent cells.

p66Shc-competent and p66Shc-deficient HeLa and MEFs express similar levels of Glut1. (A)

HeLa cells stably transfected with indicated shRNAs were lysed and cell lysates were analyzed

for the indicated proteins by immunoblotting. (B) Lysates of p66Shc KO and p66+ MEFs were

examined as in (A).

77

A

B

Fig. 3.17. Isotope-tracing of 13

C-labeled glucose in p66Shc-deficient and p66Shc-competent

HeLa cells.

(A) Schematic diagram of 13

C-labeling patterns of metabolic products with [1,2-13

C2] labeled

glucose as a tracer. Blue circles indicate 13

C-labeled carbons. (B) Fold change of 13

C-labeled

glycolytic intermediates in p66Shc-competent (white) and p66Shc-deficient cells (black) HeLa

cells. Error bars represent SD of at least three biological replicates (* p < 0.05, *** p < 0.001).

Glucose-6-phosphate (G6P); fructose-6-phosphate (F6P); fructose-1,6-phosphate (F1,6BP); 3-

phosphoglycerate (3PG).

78

Fig. 3.18. Isotope-tracing of 13

C-labeled hexosamine pathway intermediates in p66Shc-

deficient and p66Shc-competent HeLa cells.

Fold change of 13

C-labeled hexosamine biosynthesis pathway intermediates in p66Shc-

competent (white) and p66Shc-deficient cells (black) HeLa cells. Error bars represent SD of at

least three biological replicates (** p < 0.01, *** p < 0.001). Fructose-6-phosphate (F6P);

glucosamine-6-phosphate (GlcN6P); uridine diphosphate N-acetylglucosamine (UDP-GlcNAc).

79

Fig. 3.19. Isotope-tracing of 13

C-labeled glycolytic intermediates in p66Shc KO and p66+

MEFs.

Fold change of 13

C-labeled glycolytic glycolytic intermediates in p66Shc KO (white) and p66+

(black) MEFs. Error bars represent SD of at least three biological replicates (** p < 0.01, *** p <

0.001). Glucose-6-phosphate (G6P); fructose-6-phosphate (F6P).

80

Fig. 3.20. Isotope-tracing of 13

C-labeled nonessential amino acids in p66Shc KO and p66+

MEFs.

Fold change of 13

C-labeled serine (A) and alanine (B) in p66Shc KO (white) and p66+ (black)

MEFs. Error bars represent SD of at least three biological replicates (*** p < 0.001).

M2 Alanine

KO p66+

0.0

0.5

1.0

1.5

Fo

ld c

han

ge

M2 Serine

KO p66+

0.0

0.5

1.0

1.5

Fo

ld c

han

ge

A B

*** ***

81

A B

C D

Fig. 3.21. Flux analysis of 13

C-labeled glucose in p66Shc KO and p66+ MEFs.

Relative amounts of 13

C-labeled pyruvate (A), citrate (B), and the hexosamine pathway

intermediates UDP-GlcNAc (C) and R5P (D) after incubating cells with [1,2-13

C2] labeled

glucose for the indicated time points. Data are represented as the ratio of 13

C-labeled metabolites

to an internal standard (D7-glucose). Error bars represent SD of at least three biological

replicates (p < 0.05). Uridine diphosphate N-acetylglucosamine (UDP-GlcNAc); ribose-5-

phosphate (R5P).

M4 UDP-GlcNAc

0 5 10 150.000

0.001

0.002

0.003

0.004

0.005KO

p66+

Time (min)

Are

a R

ati

o

KO

p66+

Time (min)

M4 UDP-GlcNAC M2 R5P

M2 Pyruvate M2 Citrate

82

Table 3.1: Fold change of the most significantly p66Shc-inhibited metabolites.

The 10 metabolites showing the most statistically significant changes in p66+

over

p66Shc KO MEFs. FDR: False discovery rate.

Metabolite Fold change p-value FDR

Dihydroxyacetone phosphate (DHAP) 0.23 2.44E-05 0.00024

Fructose 6-phosphate (F6P) 0.29 7.64E-05 0.00027

Glucose 6-phosphate (G6P) 0.36 8.11E-05 0.00027

Fructose 1,6-bisphosphate (F1,6BP) 0.29 0.000207 0.00045

Malonyl Co-enzyme A (MCoA) 0.30 0.000226 0.00045

Erythrose-4-phosphate (E4P) 0.26 0.000598 0.00091

NADH/NAD+ 0.34 0.000641 0.00091

Acetyl Co-enzyme A (ACoA) 0.29 0.001544 0.00193

Deoxycytidine triphosphate (dCTP) 0.41 0.00265 0.00294

Uridine 5-triphosphate (UTP) 0.45 0.003853 0.003853

83

Chapter 4. p66Shc Inhibits Signaling to The Metabolic Sensor

mTOR

A version of this chapter appeared in the following article:

The Adaptor Protein p66Shc Inhibits mTOR-Dependent Anabolic Metabolism.

Science Signaling, 7, ra17 (2014).

Mohamed A. Soliman, Anas M. Abdel Rahman, Dudley W. Lamming, Kivanç Birsoy, Judy

Pawling, Maria E. Frigolet, Huogen Lu, I. George Fantus, Adrian Pasculescu, Yong Zheng,

David M. Sabatini, James W. Dennis, and Tony Pawson.

84

4.1. Background

The evolutionarily conserved Ser/Thr kinase TOR integrates signals from growth factors and

nutrients to activate anabolic pathways driving cell growth. Chronic activation of mTORC1 in

Tsc2 KO MEFs enhances glucose uptake, lactate production and the biosynthesis of lipid

metabolites, and rapamycin treatment inhibits the heightened glycolytic phenotype in Tsc2 KO

MEFs (Duvel et al., 2010). The same study showed that mTORC1 increases the levels of the

pentose phosphate pathway intermediates, which provide the ribose sugar component for the

pyrimidine nucleotides (Duvel et al., 2010). In line with this study, recent metabolomic and

phosphoproteomic analyses described an essential role for mTORC1-S6K1 in de novo

pyrimidine biosynthesis (Ben-Sahra et al., 2013; Robitaille et al., 2013). mTORC2 also plays a

role in cell metabolism. In gliobastoma tumors, mTORC2 controls glycolytic metabolism by

regulating the cellular level of Myc. The enhancement of glucose metabolism correlated with

shorter survival of glioblastoma patients (Masui et al., 2013).

My data suggest that silencing of p66Shc stimulated glycolytic and lipid metabolism (Chapter

3). Given the central role of mTOR in coordinating growth signaling and cellular metabolism

(Shimobayashi and Hall, 2014), I hypothesized that deficiency of p66Shc enhances insulin

signaling to mTOR, hence increasing the utilization of glucose into branching anabolic pathways

to meet the growth demands of proliferating cells. This chapter addresses the effect of p66Shc

expression on mTORC1 and mTORC2 activation to provide insight into the molecular

mechanism by which p66Shc regulates cellular metabolism.

85

4.2. p66Shc inhibits growth factor signaling to the metabolic sensor mTOR

Several aspects of the metabolic profile of p66Shc KO MEFs resemble that of cells with chronic

mTOR activation; these include enhanced amounts of glycolytic and pentose phosphate pathway

intermediates (G6P, F6P, and R5P) (Duvel et al., 2010), increased amounts of de novo

pyrimidine metabolites (orotate) (Ben-Sahra et al., 2013; Robitaille et al., 2013) and increased

lipid biosynthesis precursors (acetyl-CoA and malonyl-CoA) (Duvel et al., 2010). To monitor

mTOR signaling, phosphorylation of Thr389 in the ribosomal protein S6 kinase (S6K1), a

substrate of mTORC1, Ser473 in Akt which is a direct target of mTORC2 (Sarbassov et al.,

2005) and Thr346 NDRG1, an indirect target of mTORC2 (Garcia-Martinez and Alessi, 2008),

were examined (Zoncu et al., 2011). Following serum stimulation, p66Shc-deficient HeLa cells

displayed markedly increased phosphorylation of both mTORC1 and mTORC2 targets compared

to control cells, despite having equal abundance of the p52Shc and p46Shc isoforms (Fig. 4.1).

Similar results were obtained with insulin stimulation (Fig. 4.2). We also found that amino acid

activation of the mTORC1 pathway was enhanced in p66Shc-deficient HeLa cells (Fig. 4.3). In

p66+ MEFs, p66Shc inhibited the activation of mTOR targets following stimulation with IGF-1

and insulin, but not with EGF (Fig. 4.4). This observation is consistent with previous reports

showing that p66Shc has little effect on EGF signaling (Migliaccio et al., 1997). The

phosphorylation of mTORC1 and mTORC2 targets was decreased in p66+ MEFs following

serum (Fig. 4.5) and amino acid stimulation (Fig. 4.6). In contrast to the inhibitory effects of

p66Shc expression on mTOR signaling in KO MEFs, mTOR activation was sustained in KO

MEFs stably overexpressing the p52Shc isoform (p52+ cells) (Fig. 4.7). In line with this,

elevated glycolytic metabolism was maintained in p52+ cells (Fig. 4.8). This confirms earlier

86

reports that the p66Shc and p52Shc isoforms have opposing effects attributable to the unique N-

terminal CH2 region of p66Shc (Migliaccio et al., 1997; Xi et al., 2008).

4.3. p66Shc expression causes a decrease in cell size

mTOR functions as a central regulator of cell growth. The inhibitory effect of p66Shc on mTOR

activation therefore predicted that p66Shc deficiency would lead to a larger cell size. Indeed,

stable re-expression of p66Shc in p66Shc KO MEFs caused a decrease in cell size as measured

by FACS forward scatter analysis (Fig. 4.9). These findings were confirmed using Coulter

Counter measurement (Fig. 4.10A). Conversely, HeLa cells depleted of p66Shc displayed an

increase in median cell size, consistent with redirection of glucose-derived carbon towards

biomass synthesis (Fig. 4.10B). Together, these results suggest that p66Shc acts as a negative

regulator of the nutrient-sensing mTOR signaling pathway leading to inhibition of cell growth

and anabolic metabolism.

4.4. Effects of p66Shc on glycolytic metabolism are mediated through mTOR

To determine which mTOR complex contributes to the metabolic phenotype observed in p66Shc

KO cells, we treated p66Shc KO and p66+ MEFs with rapamycin for 16 h, a time frame that

inhibits both mTORC1 and mTORC2 enzymes (Fig. 4.11). In line with reports linking mTORC1

and mTORC2 to glycolytic metabolism (Duvel et al., 2010; Hagiwara et al., 2012; Lamming et

al., 2014), we found that inhibition of both mTOR complexes with rapamycin reversed the

metabolic phenotype of p66Shc KO MEFs, notably diminishing the increase in the pentose

phosphate pathway (R5P) and the hexosamine biosynthesis pathway (UDP-GlcNAc) (Fig. 4.12).

87

Furthermore, inhibition of Akt, a kinase that is upstream of mTORC1 and downstream of

mTORC2 (Zoncu et al., 2011), significantly decreased the amounts of glycolytic metabolites in

p66Shc-deficient MEFs (Fig. 4.13).

88

Fig. 4.1. p66Shc inhibits mTORC1 and mTORC2 activation following serum stimulation.

HeLa cells stably transfected with indicated shRNAs were serum starved, treated with either

DMSO or Torin for 1 h, and then stimulated with 10% serum (10 min). Cell lysates were

analyzed for the indicated proteins and phosphorylation states by immunoblotting. Pre-treatment

of cells with the mTOR inhibitor Torin1 (Thoreen et al., 2009) abolished the phosphorylation of

mTOR targets in both cell types confirming the specificity of the signal.

89

Fig. 4.2. p66Shc inhibits insulin signaling to mTOR.

HeLa cells stably transfected with indicated shRNAs were serum starved, treated with either

DMSO or Torin for 1 h, and then stimulated with 100 nM Insulin (10 min). Cell lysates were

analyzed for the indicated proteins and phosphorylation states by immunoblotting. Pre-treatment

of cells with the mTOR inhibitor Torin1 (Thoreen et al., 2009) abolished the phosphorylation of

mTOR targets in both cell types confirming the specificity of the signal.

90

Fig. 4.3. p66Shc inhibits mTORC1 activation in response to amino acid stimulation.

HeLa cells stably transfected with shRNAs targeting the indicated genes were starved of amino

acids for 50 min and stimulated either with dialyzed serum or normal serum for 10 min.

Phosphorylation of S6K1 was analyzed by Western blotting.

91

Fig. 4.4. p66Shc expression inhibits mTOR activation in response to insulin and IGF1, but

not to EGF, stimulation.

p66Shc KO and p66+ MEFs were serum starved, then stimulated with 100 nM EGF, IGF, or

insulin (10 min). Cell lysates were analyzed for the indicated proteins and phosphorylation states

by immunoblotting.

92

Fig. 4.5. p66Shc expression inhibits mTOR activation in response to serum stimulation.

p66Shc KO and p66+ MEFs were serum starved for 4 h, then stimulated with 10% serum for 10

min. Cell lysates were analyzed for the indicated proteins and phosphorylation states by

immunoblotting. Pre-treatment of cells with the mTOR inhibitor Torin1 (Thoreen et al., 2009)

abolished the phosphorylation of mTOR targets in both cell types confirming the specificity of

the signal.

93

Fig. 4.6. p66Shc expression inhibits mTOR activation in response to amino acid

stimulation.

p66Shc KO and p66+ MEFs were starved of amino acids for 50 min and stimulated either with

dialyzed serum or normal serum for 10 min. Cell lysates were analyzed for the indicated proteins

and phosphorylation states by immunoblotting.

94

Fig. 4.7. Stable expression of p66Shc, but not p52Shc, in p66Shc KO cells inhibits the

mTOR pathway.

p66Shc KO, p66+ and p52

+ MEFs were serum starved, then stimulated with 10% serum for 10

min. Cell lysates were analyzed for the indicated proteins and phosphorylation states by

immunoblotting.

95

A B

Fig. 4.8. Stable expression of p66Shc, but not p52Shc, in p66Shc KO cells inhibits glycolytic

metabolism.

Fold change of G6P (A) and F6P (B) in p66Shc KO (white), p66+ (black) and p52

+ MEFs

(srtiped). Error bars represent SD of at least three independently prepared samples (** p < 0.01).

** **

G6P F6P

96

A B

Fig. 4.9. p66Shc expression decreases cell size.

Cell size distribution of p66Shc KO and p66+ MEFs using flow cytometry forward scattering

(A). Average cell size of KO and p66+ from three independent experiments is represented (B),

*** p < 0.001.

97

A

B

Fig. 4.10. p66Shc mediates cell growth.

(A) Cell size measurement of p66Shc KO and p66+

cells. Cell size was measured using coulter

counter. (B) Cell size measurement of p66Shc–deficient and p66Shc-competent HeLa cells.

Experiment was done as in (A).

98

Fig. 4.11. Effect of rapamycin on mTOR signaling in p66Shc KO and p66+ MEFs.

Rapamycin treatment for 16 h is sufficient to inhibit both mTORC1 and mTORC2 signaling in

p66Shc KO and p66+ MEFs. Cell lysates were analyzed for the indicated proteins and

phosphorylation states by immunoblotting.

99

A

Fig. 4.12. mTOR mediates the effects of p66Shc on glycolytic and pyrimidine metabolism.

p66Shc KO and p66+ MEFs were treated with vehicle or rapamycin for 16 h, and amounts of

glycolysis (A), pentose phosphate (B), hexosamine biosynthesis (C), and pyrimidine

biosynthesis (D) pathways metabolic intermediates were quantified by LC-MS/MS. Error bars

represent SD of three independently prepared samples (ANOVA analysis, * p < 0.05, ** p <

0.01, *** p < 0.001). Glucose-6-phosphate (G6P); fructose-6-phosphate (F6P); ribose-5-

phosphate (R5P); uridine diphosphate N-acetylglucosamine (UDP-GlcNAc); uridine

monophosphate (UMP); cytidine monophosphate (CMP).

B

C

D

100

Fig. 4.13. Effect of Akt inhibition on the abundance of glycolytic metabolites in p66Shc-

deficient and p66Shc-competent MEFs.

Fold change of 3PG (A), PEP (B), R5P (C), and citrate (D) in p66Shc KO MEFs (white), p66Shc

KO MEFs treated with Akt inhibitor (black), p66+ (grey) and p66

+ treated with Akt inhibitor

(stripped) MEFs. Error bars represent SD of at least three independently prepared samples (* p <

0.05, *** p < 0.001). 3-phosphoglycerate (3PG); phosphoenolpyruvate (PEP); ribose-5-

phosphate (R5P).

A B

C D

***

***

***

*

**

* *

***

***

101

Chapter 5. Discussion and Future Directions

102

Understanding the biological role of the adaptor protein p66Shc in regulating metabolic

homeostasis was the primary motivation in undertaking this study. Since the discovery of the

adaptor protein Shc1 in 1992 (Pelicci et al., 1992), extensive research has confirmed an essential

role for Shc1 in development, growth and diseases such as cancer (Ravichandran, 2001). The

most abundantly expressed Shc1 isoform, p52Shc, acts principally downstream of receptor

tyrosine kinases to amplify the Ras-MAPK (Rozakis-Adcock et al., 1992) and the PI3K-Akt

pathways signaling (Gu et al., 2000). However, my curiosity was piqued by the p66Shc isoform

(Migliaccio et al., 1997), identical to p52Shc1 but containing an N-terminal extension that has

been reported to posess unique properties affecting metabolism and oxidative stress (Luzi et al.,

2000). For example, in mammals, lack of p66Shc confers resistance to hyperglycemia-induced

endothelial dysfunction (Camici et al., 2007) and early atherogenesis in mice fed a high-fat diet

(Napoli et al., 2003). Indeed, deletion of p66Shc in mice improves glucose tolerance and insulin

sensitivity (Ranieri et al., 2010; Tomilov et al., 2011). These observations raised the possibility

that p66Shc evolved in mammals in response to variations in nutrient availability as a suppressor

for insulin signaling and metabolism, although a thorough study of the role of p66Shc in

metabolism using modern mass spectrometry-based metabolomics has not been done.

My results confirmed that p66Shc expression is associated with a dampening in the cellular

signaling response to insulin stimulation and a reduction in glycolytic metabolism. This

suggested that p66Shc mediates feedback inhibition of both growth factor signaling and, in turn,

glucose metabolism. The metabolomic analysis of p66Shc-competent and p66Shc-deficient cells

shows an increase in glucose-6-phosphate and downstream intermediates, including fructose-6-

phosphate and pyruvate, in cells lacking p66Shc. The kinetic flux analysis using 13

C-labeled

103

glucose shows faster carbon incorporation and higher steady state levels in p66Shc-deficient

cells. No statistically significant differences are found between p66Shc-deficient and wild-type

cells for other metabolites, including GTP, ADP, ATP, adenine, flavin adenine dinucleotide

(FAD), CoA, and glyoxalic acid. It should be noted that our analytical method targeted a specific

set of metabolites, including intermediates of glycolysis, glucosamine, pentose phosphate and

hexosamine pathways, amino acids, nucleic acids, fatty acids synthesis and oxidation, TCA and

urea cycles, and bile acid biosynthesis. Hence, I cannot exclude the possibility that p66Shc

affects other metabolic pathways that were not included in our targeted approach.

p66Shc silencing in both transformed and non-transformed immortalized cells improves glucose

uptake and enhances lactate production. Furthermore, p66Shc deficiency increases metabolite

abundance for fatty acid biosynthesis, the hexosamine pathway, pentose phosphate pathway and

increases cell size. This metabolic shift depends in part on the mTOR pathway, because

rapamycin treatment partly reversed the glycolytic shift caused by p66Shc loss. However, I could

not exclude the possibility that other signaling pathways might mediate p66Shc metabolic

effects, especially the Wnt signaling pathway. Wnt signaling can modulate glucose homeostasis

and insulin sensitivity (Yoon et al., 2010) and Wnt3A can induce aerobic glycolysis by

increasing the level of key glycolytic enzymes (Esen et al., 2013). This metabolic regulation is

mediated through mTORC2-Akt signaling (Esen et al., 2013). Given that p66Shc also signals to

mTORC2, it will be of interest to see if knocking down Wnt3a will inhibit the heightened

glycolytic metabolism observed in p66Shc-deficient cells.

104

My data suggest that depletion of p66Shc is sufficient to enhance the Warburg effect in cancer

cell lines, demonstrating that p66Shc may function as a potential tumor suppressor. Previous

reports showed that Shc1 signaling is essential for breast tumour progression, where tyrosine

phosphorylation sites Tyr239/240/317 on p52Shc are essential for ErbB2‐induced mammary

tumour outgrowth (Ahn et al., 2013; Ursini-Siegel et al., 2008; Ursini-Siegel and Muller, 2008).

However, p66Shc expression has been shown to inversely correlate to tumorigenesis (Stevenson

and Frackelton, 1998). This is consistent with my results in which p66Shc loss enhances mTOR

signaling and anabolic metabolism to support cell growth. Clinical studies report that both an

increase in tyrosine phosphorylated Shc and a decrease in the expression of p66Shc correlated

with breast tumor recurrence (Davol et al., 2003). Unlike p52Shc, the abundance of p66Shc is

substantially decreased in ErbB2-overexpressing breast cancer cell lines (Stevenson and

Frackelton, 1998). This antagonistic effect of p66Shc on RTK signaling might arise from a

competition between p66Shc and p52Shc for common targets downstream of activated receptors

(Migliaccio et al., 1997; Okada et al., 1997).

5.1. p66Shc signaling to mTOR: an open question

My results show that p66Shc inhibits insulin signaling to the mTOR pathway. Identifying

upstream and downstream regulators of p66Shc could provide valuable insight into how p66Shc

regulates mTOR activation and the associated changes in glycolytic metabolism. A limitation of

studying the functional interaction between p66Shc and mTOR is the difficulty of preserving the

p66Shc protein complexes. Using Co-IP and MS analysis, interactions between p66Shc and

mTOR, Raptor and Rictor were undetectable. The Scansite Motif bioinformatics tool analysis of

the unique CH2 region of the p66Shc isoform suggests putative binding of GSK3 to the SPSASS

105

(SxxxS) motif. This is reasonable as GSK3 inhibits mTOR signaling via phosphorylation of the

TSC2 tumor suppressor (Inoki et al., 2006). In addition, GSK3 regulates glucose transport in

several cell types, reducing glucose uptake by almost two-fold in an mTOR-mediated manner

(Buller et al., 2008). GSK3-mediated inhibition of glucose uptake is restored in TSC2 null cells

expressing wild-type TSC2, but not by a TSC2 construct with mutated GSK3 phosphorylation

sites (Buller et al., 2008). However, I was unable to detect an interaction between GSK3 and

p66Shc using co-immunoprecipitation, LC-MS or BioID (proximity-dependent biotin

identification). As well, attempts using milder cell lysis conditions failed. In future work,

chemical cross-linking might capture this putative transient interaction. If an interaction can be

confirmed, then understanding the functional role of p66Shc-GSK3 interaction in regulating

mTOR signaling and glycolytic metabolism will be essential.

Importantly, the contribution of the p66Shc PTB and SH2 domains and the three conserved Tyr

residues in the CH1 region to the observed glycolytic phenotype warrants further investigation.

Despite the fact that all Shc1 isoforms share the same PTB and SH2 domains, it is possible that

these domains and conserved Tyr residues may not bind the same proteins at the same time for

the three isoforms. Domain binding to specific partners will also depend on the local molecular

environment surrounding each isoform. Due to the high percentage of Pro and Gly, it is predicted

that the 110 amino acid peptide at the N-terminus of p66Shc is highly unstructured. It is possible

that this region can form a weak intramolecular interaction with the PTB domain, the SH2

domain or the CH1 region, blocking access of substrates to p66Shc binding surfaces. It will be

interesting to test whether expressing a p66Shc mutant with either a non-functional PTB

(R285Q), a non-phosphotyrosine binding SH2 domain (R507K) or with all three CH1 region

106

phosphotyrosine residues substituted by phenylalanine (3Y to 3F) would be able to rescue the

glycolytic phenotype in p66Shc KO MEFs.

Inhibition of PI3K is a potential mechanism by which p66Shc might suppress mTOR activation.

As PI3K acts upstream of mTORC1 and mTORC2, inhibition of PI3K would affect both mTOR

complexes. It has been reported that p66Shc-deficient vascular smooth muscle cells display

enhanced IGF-I-stimulated PI3K activity, as measured by an increase in the levels of PIP3 (Xi et

al., 2010b). Mechanistically, silencing of p66Shc enhanced IGF-I-stimulated association between

Grb2 and p85, the regulatory unit of PI3K, which led to PI3K-Akt activation (Xi et al., 2010b).

Membrane fractionation studies showed that IGF-I-stimulated membrane recruitment of Akt was

inhibited by overexpression of p66Shc and enhanced by knockdown of p66Shc (Xi et al.,

2010b).

5.2. Competition between Shc1 isoforms in regulating PI3K-mTOR signaling

In contrast to the inhibitory effects of p66Shc, mTOR pathway activation was sustained in

p66Shc KO MEFs stably overexpressing the p52Shc isoform. These effects are particularly

striking since the p52Shc isoform is identical to p66Shc in all but the N-terminal CH2 region.

Signaling downstream of p66Shc and p52Shc to the mTOR pathway might vary depending on

the cell type and the relative abundance of each protein. My proteomic analysis shows that

p66Shc has common binding proteins to those known for p52Shc (Appendix 6.4) (Zheng et al.,

2013). Thus, it is probable that p66Shc competes with p52Shc for binding partners in a dynamic

manner that disrupts downstream signaling to mTOR. Future studies could include the mass-

107

spectrometry based analysis of the time-dependent recruitment of p52Shc versus p66Shc

signaling complexes by receptors, which could reveal the molecular basis for p66Shc action on

signaling and anabolic metabolism. I cannot exclude the possibility that additional independent

pathways could be signaling to p66Shc and p52Shc independently, allowing cells to fine tune

mTOR activity in response to nutrient and environmental cues. Given the complexity of insulin-

Shc-mTOR signaling, and the feedback loops among the components of mTOR pathway, a better

understanding of the physiological consequences of altering the p66Shc-p52Shc signaling

balance will require generating mouse models with inducible overexpression or deletion of

p66Shc.

5.3. Regulation of receptor and glucose transporter glycosylation by p66Shc

UDP-GlcNAc and hexosamine pathway intermediates were among the metabolites showing the

largest increases in p66Shc knockout cells, and may contribute to the Warburg-like phenotype in

cells lacking p66Shc. UDP-GlcNAc is an essential substrate for the Golgi N-glycan branching

pathway and consequently for galectin binding, which enhances the residency of cytokine

receptors at the cell surface (Lau et al., 2007; Partridge et al., 2004). More recently, regulation by

N-glycan branching has been extended to surface residency of glucose transporters, as reported

for Glut2 in β-cells (Ohtsubo et al., 2005), Glut1 in tumor cells (Kitagawa et al., 1995), and

Glut4 in cells simulated by insulin or UDP-GlcNAc (Haga et al., 2011).

The N-glycans are remodeled in the Golgi by the branching N-acetylglucosaminyltransferases

enzymes encoded by the genes Mgat1-5. Previous work showed that Mgat5-deficient mice

108

display hypoglycemia, lower body weight and fail to gain weight on a high fat diet (Cheung et

al., 2007). Since p66Shc knockout mice are also insulin sensitive (Tomilov et al., 2011) and

show resistance to high fat diet (Napoli et al., 2003), it is possible that p66Shc functionally

interacts with the HBP and N-glycan branching to affect the level of glycosylation of nutrient

transporters. We expect that silencing HBP and N-glycan branching enzymes, such as Mgat5,

would suppress the p66Shc-mediated metabolic phenotype. Conversely, forced expression of N-

glycan branching enzymes would be expected to oppose p66Shc-mediated suppression of

glycolytic metabolism.

5.4. p66Shc and fatty acid signaling to mTOR

Higher levels of malonyl-CoA were observed in p66Shc-deficient MEFs. Malonyl Co-A is a

strictly anabolic metabolite that supplies and commits the 2-carbon building blocks to fatty acid

chain biosynthesis. Although the scope of the project did not allow for investigating lipid

biosynthesis, it would be important to examine the effect of p66Shc expression on the abundance

of various lipids using a non-targeted metabolomics approach. Earlier reports have suggested

that mTOR binds lipids, in particular phosphatidic acid, and that this binding increases mTOR

kinase activity (Fang et al., 2001). In addition, one of the enzymes responsible for the production

of phosphatidic acid, phospholipase D, acts as an upstream regulator in the mTOR pathway

(Fang et al., 2003). Hence, high intracellular levels of particular lipid species might activate

mTOR directly, bypassing the normal requirement of growth factor and amino acid signaling.

This raises the possibility that p66Shc could be playing an indirect role in regulating mTOR

activity through its effect on modulating lipid metabolism.

109

5.5. Genes regulated through p66Shc expression

RNA-seq analysis showed that p66Shc-competent and p66Shc-deficient cells have comparable

expression levels of genes encoding metabolic enzymes and glucose transporters (Appendix

6.3). These results are surprising given that mTOR has been shown to regulate the transcription

of metabolic genes in a HIF1- and SREBP-dependent manner (Duvel et al., 2010). The

transcriptome of Tsc1 and Tsc2 KO MEFs revealed that mTORC1 promotes the transcription of

genes involved in glycolysis, the pentose phosphate pathway and de novo lipogenesis (Duvel et

al., 2010).

Regulation by allostery and posttranslational modifications to metabolic enzymes and

transporters, rather than transcriptional regulation, may account for much of the observed

p66Shc-depdendent metabolic reprogramming. However, the stable cell lines used in our

experiments have adapted to p66Shc expression or loss, and the associated changes in signaling

and metabolic flux may be reinforced by other changes in gene expression and posttranslational

modifications. Importantly, transcripts showing a change in abundance were mainly clustered in

the category of secretory proteins. The list of genes showing differential expression in p66Shc-

competent cells included transcripts that have been reported to be involved in regulating insulin

signaling, such as Wnt4 (Heller et al., 2011). Additional studies are required to assess their

contribution to the p66Shc metabolic phenotype.

110

5.6. Negative regulation of insulin signaling by adaptor proteins

In addition to p66Shc, other adaptor proteins, including IRS1, Grb10 and Grb14, facilitate

negative feedback inhibition of the insulin/IGF-1 signaling pathway. Major regulatory

mechanisms include phosphorylation and O-GlcNAcylation of Ser/Thr residues of the insulin

receptor substrate 1 (IRS1) (D'Alessandris et al., 2004; Harrington et al., 2004; Shah et al.,

2004). Repression of IRS1 gene expression and direct phosphorylation of IRS1 by S6K1, where

IRS1 becomes limiting for signal transmission from the insulin receptor to the PI3K-mTOR

pathway, drives constitutive activation of the mTORC1-S6K1 axis (Harrington et al., 2004; Shah

et al., 2004). In addition to p66Shc and IRS1, other adaptor proteins including the family of

growth factor receptor-bound proteins Grb10 and Grb14, and SH2B2 have been strongly linked

to the inhibition of insulin signaling.

Grb10 and Grb14 have emerged as major negative regulators of insulin/IGF-1 action (Holt and

Siddle, 2005). Various studies have provided links between these adaptors, the pathways they are

involved in and metabolic diseases such as diabetes. For example, phosphoproteomic analysis of

mTOR targets has revealed that Grb10 suppresses insulin sensitivity through feedback inhibition

of insulin-PI3K-mTOR signaling (Hsu et al., 2011; Yu et al., 2011). These reports showed that

overexpression of Grb10 and Grb14 inhibited insulin-stimulated tyrosine phosphorylation of

substrates, including IRS1, and downstream PI3K-Akt-mTORC1, while deficiency of Grb10

through gene silencing potentiated insulin-induced substrate phosphorylation (Hsu et al., 2011;

Kasus-Jacobi et al., 1998; Yu et al., 2011). Consistent with the cell culture data, disruption of the

maternal allele of Grb10 in mice improved insulin sensitivity and glucose tolerance, and

enhanced insulin-induced Akt activation (Smith et al., 2007; Wang et al., 2007). Similarly,

111

Grb14 knockout mice showed increased insulin sensitivity, glucose tolerance and activation of

PI3K-Akt in skeletal muscles and liver (Cooney et al., 2004). Deletion of the Sh2b2 gene in mice

led to an increase in insulin sensitivity and insulin-stimulated glucose transport in adipocytes

(Minami et al., 2003).

Further evidence of the link between adaptor proteins and metabolic diseases was provided from

human genome-wide association studies. Genome sequencing revealed an association between

single nucleotide polymorphisms (SNPs) at the GRB10 (Di Paola et al., 2006; Rampersaud et al.,

2007) and the GRB14 (Kooner et al., 2011; Morris et al., 2012) gene loci and incidence of type 2

diabetes, and with an increase in fasting blood glucose and insulin levels (Manning et al., 2012;

Scott et al., 2012). Overall, the role of these adaptor proteins as negative regulators of

physiological insulin/IGF1 signaling represents an exciting avenue for further investigation.

Much remains to be done to identify new binding partners and functions for p66Shc, Grb10 and

Grb14, to identify transcriptional factors that regulate their expression, and to assess the

implications of their altered signaling in human metabolic diseases.

5.7. Summary

Deficiency of the adaptor protein p66Shc improves glucose tolerance of mice. My results

indicate that the Shc1 proteins not only regulate mitogenic signaling, but also modify anabolic

metabolism in an mTOR-dependent manner. Using high-throughput metabolomic approaches, I

show that p66Shc silencing enhanced glucose metabolism, increased the abundance of

intermediates in various biosynthetic pathways and increased cell size. These changes in

112

metabolism required the mTOR pathway, which couples energy, nutrient, and growth factor

(including insulin) signals to processes that mediate cellular growth and metabolism. To my

knowledge, this is the first study that directly assesses the role of an adaptor-type protein in

regulating energy metabolism using a metabolomic approach. My data points to the prospect that

monitoring or modulating p66Shc abundance could be important in the management of

pathological conditions in which metabolic signaling is dysregulated, such as cancer and

diabetes.

113

Chapter 6. Appendix

114

6.1. LC-MS/MS transitions for the metabolites measured in this study

Name RT Q1 Q3 DP CE HMDB

(-)-Riboflavin 2.2 377 63 31 75 HMDB00244

1,4-diaminobutane 2.4 87 45 41 21 HMDB01414

1-Methylhistamine 3.3 126 109 30 20 HMDB00898

2`-Deoxyadenosine 6.2 252 136 40 20 HMDB00101

2`-Deoxycytidine 2.8 228 112 111 53 HMDB00014

2`-Deoxyuridine 2.8 229 113 50 20 HMDB00012

2'-Deoxy-D-ribose 3.9 134 117 31 15 HMDB03224

Tryptophanol 4.5 162 144 61 19 HMDB03447

3-OH-anthranilate 6.3 154 136 30 18 HMDB01476

4-guanidinobutyrate 2.5 146 87 76 25 HMDB03464

Acetoacetate 1.9 85 70 40 20 HMDB00060

Acetylcholine 2.5 147 87 25 21 HMDB00895

Adenosine 6.1 268 136 40 30 HMDB00050

(A)symetric dimethyl arginine 1.9 203 70 50 40 HMDB01539

Allantoin 6.5 159 116 50 11 HMDB00462

Aminoisobutyrate 2.4 104 86 40 16 HMDB01906

Anthranilate 9.4 138 120 25 18 HMDB01123

Argininosuccinate 3.2 291 70 50 54 HMDB00052

Betaine 10.3 118 58 40 41 HMDB00043

Carnitine 2.5 163 85 25 29 METPA0048

Carnosine 2.1 227 110 40 33 HMDB00033

Choline 2.8 105 60 50 27 HMDB00097

Cobalamin 18.1 678 147 60 52 HMDB02174

Creatine 3.2 132 90 50 17 HMDB00064

Creatinine 2.3 114 44 25 28 HMDB00562

Cysteamine 2.2 78 61 40 16 HMDB02991

Cysteine 3.2 122 76 25 20 METPA0075

Cytidine 2.8 244 112 25 17 HMDB00089

Cytosine 2.3 112 95 40 26 HMDB00630

D-alpha-Aminobutyrate 3.3 104 58 30 17 HMDB00452

Dimethylglycine 3.3 104 58 30 20 HMDB00092

DL-2-Aminoadipate 3.8 162 98 46 25 HMDB00510

DL-Homocystine 3.4 136 90 50 20 HMDB00575

Dopamine 3.9 154 137 50 16 HMDB00073

Epinephrine 11.5 184 166 27 13 HMDB00068

gamma-Aminobutyrate 2.4 104 87 41 15 HMDB00112

gamma-Aminoisobutyrate 2.7 104 86 40 16 HMDB00112

Glucosamine 2.2 180 162 50 10 HMDB01514

Gluconolactone 5.8 179 133 46 15 HMDB00150

115

Glycerol 3.6 93 57 30 10 HMDB00131

Glycine 3.1 76 30 20 21 HMDB00123

Guanidinoacetate 3.7 118 72 25 18 HMDB00128

Guanine 4.5 152 135 40 30 HMDB00132

Guanosine 6.2 284 152 40 25 HMDB00133

Histamine 2.3 112 95 40 26 HMDB00870

HO-Tryptophan 7.5 221 204 40 18 HMDB03447

Hydroxybutyrate 4.6 105 87 70 10 HMDB00008

Inosine 6 269 137 1 15 HMDB00195

Isobutyrylglycine 1.9 146 72 50 22 HMDB00730

Kynurenine 3.1 209 146 115 45 HMDB00684

L-Alanine 3.2 90 44 25 17 HMDB00161

L-alpha-Aminoadipate 3.9 163 73 41 37 HMDB00510

L-alpha-Aminobutyrate 3.3 104 58 30 17 HMDB00452

L-Arabinose 10.1 151 107 51 35 HMDB00646

L-Arginine 2.8 175 70 25 32 HMDB00517

L-Asparagine 3.2 133 74 30 23 HMDB00168

L-Aspartate 6.3 134 74 25 21 HMDB00191

L-Canavanine 2.1 177 76 56 35 HMDB02706

L-Carnitine 2.5 162 103 86 25 HMDB00062

L-Citrulline 3.2 176 70 41 27 HMDB00904

L-Cystathyonine 17.3 223 149 31 29 HMDB00099

L-Cystine 2.1 241 109 46 37 HMDB00192

L-Glutamate 3.8 148 84 25 23 HMDB00148

L-Glutamine 3.2 147 84 25 25 HMDB00641

L-Histidine 2.1 156 110 25 21 HMDB00177

L-Homoserine 3.2 120 74 50 40 HMDB00719

Lipoamide 7.5 206 189 41 15 HMDB00962

L-Isoleucine 4.8 132 86 50 18 HMDB00172

L-Leucine 5.6 132 86 50 18 HMDB00687

L-Lysine 2.1 147 84 25 25 HMDB00182

L-Methionine 4.6 150 61 40 31 HMDB00696

N-Monomethyl-L-arginine 3.2 189 70 60 40 NA

L-Phenylalanine 6.7 166 120 50 19 HMDB00159

L-Proline 3.5 116 70 50 20 HMDB00162

L-Sarcosine 2.9 91 73 46 17 HMDB00271

L-Serine 3.2 106 60 25 18 HMDB00187

L-Threonine 3.2 120 74 50 20 HMDB00167

L-Tryptophan 7.5 205 188 25 16 HMDB00929

L-Tyrosine 5.6 182 136 25 19 HMDB00158

L-Valine 3.7 118 72 25 18 HMDB00883

Melatonin 5.4 233 174 35 18 HMDB01389

Metanephrine 3.4 198 181 40 13 HMDB04063

116

N-Carbamoyl-Beta-Alanine 3.4 133 115 40 12 HMDB00026

Niacinamide 6 123 80 30 30 HMDB01406

Normetanephrine 11.5 184 166 27 13 HMDB00819

Hydroxy proline 3.3 132 86 50 18 HMDB00725

Ornithine 2 133 70 16 31 HMDB00214

Phosphocholine 4.2 184 125 66 20 HMDB01565

phosphoethanoloamine 3.2 142 44 50 20 HMDB00224

Phosphoserine 7.8 186 130 56 22 HMDB00272

Protoporphyrin IX 0.2 563 129 251 27 HMDB00241

Pyridoxal 3 168 150 51 79 HMDB01545

Pyridoxamine 11.7 169 123 21 23 HMDB01431

Pyridoxine 3 170 152 40 17 HMDB00239

S-(5`-Adenosyl)-L-homocysteine 5.4 385 136 60 25 HMDB00939

Serotonin 3.2 177 160 50 32 HMDB00259

Spermidine 1.9 146 72 46 18 HMDB01257

Spermine 1.9 203 129 40 25 HMDB01256

Taurine 3.3 126 108 50 20 HMDB00251

Thiamine 2.8 265 122 46 18 HMDB00235

Thiamine monophosphate 2.3 345 122 66 29 HMDB02666

Thymidine 6.8 243 127 30 35 HMDB00273

Thymine 6.5 127 110 40 16 HMDB00262

Trans-4-hydroxy-L-Proline 3.3 132 68 66 23 HMDB00725

Trimethylamine-N-oxide 2.9 76 58 25 35 HMDB00925

Uracil 5.3 113 70 71 29 HMDB00300

Uridine 5.9 245 113 96 25 HMDB00296

Xanthine 5.9 153 110 40 21 HMDB00292

Xanthosine 14.6 285 63 41 27 HMDB00299

2,3-Dihydroxybenzoate 8.8 153 109 -45 -30 HMDB00397

2,3-Pyridinedicarboxylate 5 166 122 -110 -54 HMDB00232

2-Aminoadipate 6.4 160 116 -50 -18 HMDB00510

2'-Deoxy cytidine 5'-

monophosphate (dCMP)

7.6 306 79 -60 -54 HMDB01202

2'-Deoxy cytidine diphosphate

(dCDP)

8.2 386 79 -30 -60 HMDB01245

2'-Deoxy thymidine

monophosphate (dTMP)

7 321 195 -60 -18 HMDB01227

2'-Deoxy uridine monophosphate

(dUMP)

7.5 307 79 -60 -50 HMDB01409

2'-Deoxyadenosine 8.6 250 134 -60 -24 HMDB00101

2'-Deoxycytidine 5-triphosphate

(dCTP)

8.5 466 159 -60 -30 HMDB00998

2'-dexyuridine 5-triphosphate

(dUTP)

11 467 159 -40 -40 HMDB01191

2-Oxobutyrate 8.1 101 57 -30 -10 HMDB00005

3-Indoleacetate 8.1 174 130 -25 -14 HMDB00197

117

3-Phosphoglycerate 8.3 185 97 -40 -22 HMDB00807

4-Aminobenzoate 7.8 136 92 -40 -14 HMDB01392

4-Hydroxy-3-

methoxyphenylglycolaldehyde

9.5 183 150 -40 -30 HMDB04061

4-Hydroxybenzoate 7.9 137 93 -20 -20 HMDB00500

4-Hydroxyphenylpyruvate 3.3 179 107 -110 -54 HMDB00707

4-Pyridoxate 8.8 182 138 -40 -30 HMDB00017

5-hydroxyindole-3-acetate 7.2 190 146 -50 -16 HMDB00763

5-Methyl THF 9.5 485 329 -70 -30 HMDB01396

Acetoacetate 8.1 101 57 -40 -15 HMDB00060

Acetyl Co-enzyme A (ACoA) 8.7 808 159 -150 -88 HMDB01206

Adenine 5.5 134 107 -65 -24 HMDB00034

Adenosine 5'-monophosphate

(AMP)

8.3 346 79 -30 -66 HMDB00045

Adenosine diphosphate (ADP) 8.6 426 159 -75 -36 HMDB01341

Adenosine triphosphate (ATP) 8.6 506 159 -100 -44 HMDB00538

Adenylosuccinate 8.3 462 79 -60 -48 HMDB00536

Adipate 9.1 145 101 -110 -54 HMDB00448

α-ketoglutarate 8.2 145 101 -20 -12 HMDB00208

Ascorbate 8.1 175 115 -40 -17 HMDB00044

Bilirubin 11.1 583 285 -30 -36 HMDB00054

Biotin 4.8 243 200 -60 -18 HMDB00030

Butyrate 7.9 87 45 -30 -10 HMDB00039

Chenodeoxycholate 13.4 391 374 -80 -50 HMDB00518

Cholate 4.3 407 343 -100 -20 HMDB00619

Aconitate 8.3 173 85 -25 -16 HMDB00072

Citrate 8.7 191 111 -25 -18 HMDB00094

Co-enzyme A (CoA) 8.7 766 159 -135 -86 HMDB01423

Creatine phosphate 8 210 79 -35 -24 HMDB01511

Cyclic adenosine monophosphate

(cAMP)

8.6 328 134 -80 -42 HMDB11616

Cyclic guanosine monophosphate

(cGMP)

8.8 344 150 -40 -34 HMDB11616

Cytidine 5-diphosphate (CDP) 8.5 402 158 -60 -30 HMDB01546

Cytidine 5-triphosphate (CTP) 8.5 482 159 -60 -40 HMDB00082

Cytidine monophosphate (CMP) 7.5 322 139 -70 -32 HMDB00095

D7-Glucose (Internal Standard) 3.4 186 124 -55 -12 NA

D-Arabino-1,4-lactone 7.7 147 59 -55 -18 METPA0132

Dihydrofolate 9.5 442 176 -50 -40 HMDB01056

Dihydroxyacetone phosphate

(DHAP)

7.4 169 97 -40 -15 HMDB01473

Isocitrate 8.7 191 111 -35 -20 HMDB00193

D-Maltose 3.2 341 161 -40 -12 HMDB00163

D-Pantothenate 7.5 218 88 -30 -16 HMDB00210

118

D-Rib(ul)ose-5-Phosphate (R5P) 7.5 229 79 -15 -58 HMDB01548

D-Xylose 3.4 149 89 -50 -8 HMDB00098

Erythrose-4-Phosphate (E4P) 7.3 199 97 -40 -16 HMDB01321

Folate 8 440 311 -80 -30 HMDB00121

Folinate 13.4 472 315 -40 -40 HMDB01562

Fructose 1,6-bisphosphate (Fru 1,6

DP)

8.3 339 241 -30 -22 HMDB01058

Fructose 6-phosphate (Fru-6P) 6.9 259 169 -50 -18 HMDB00124

Fumarate 8.1 115 71 -30 -10 HMDB00134

Geranyl pyrophosphate 7.3 313 79 -35 -37 METPA0034

Glucoronate 8.6 193 113 -60 -22 HMDB00127

Glucosamine (GlcN) 3.5 178 145 -30 -12 HMDB01514

Glucosamine 6-Phosphate (GlcNP) 4.1 258 97 -75 -18 HMDB01254

Glucose (Glc) 3.3 179 119 -60 -12 HMDB00122

Glucose 6-Phosphate 6.9 259 199 -55 -16 HMDB01401

Glyceraldehyde (GA) 3.4 89 59 -35 -10 HMDB01051

Glycerate 2-P 8.3 185 79 -40 -30 HMDB00362

Glycerol-3-Phosphate (G3P) 7.2 171 79 -50 -22 HMDB00126

Glycochenodeoxycholate 10 448 74 -80 -58 HMDB00637

Glycocholate 13.9 464 74 -30 -60 HMDB00138

Glyoxylate 7.4 73 45 -45 -10 HMDB00119

Guanosine diphosphate (GDP) 8.7 442 159 -75 -38 HMDB01201

Guanosine diphosphate- Fucose

(GDP-Fuc)

8 588 159 -100 -60 HMDB01095

Guanosine diphosphate-Mannose

(GDP-Man)

8.7 604 159 -120 -70 HMDB01163

Guanosine monophosphate (GMP) 7.5 362 79 -70 -60 HMDB01397

Guanosine triphosphate (GTP) 8.8 522 159 -95 -48 HMDB01273

Hippurate 3.5 178 134 -50 -16 HMDB00714

Homocystine 8.9 267 132 -30 -22 HMDB00575

Homogentisate 6.2 167 123 -50 -23 HMDB00130

Homovanillate 5.7 181 137 -50 -12 HMDB00118

Hypoxanthine 4.8 135 92 -30 -22 HMDB00157

Inosine 5'-monophosphate (IMP) 7.7 347 79 -65 -100 HMDB00175

Inositol 7.2 179 81 -50 -15 HMDB00211

Itaconate 8.2 129 85 -40 -12 HMDB02092

Kynurenate 8.1 188 144 -30 -20 HMDB00715

L-(-)-Sorbose 3.3 179 89 -40 -12 HMDB01266

Lactose (L) 3.3 341 161 -50 -11 HMDB00186

L-Dihydroorotate 5.7 157 113 -35 -12 HMDB03349

L-Fucose 3.7 163 103 -55 -10 HMDB00174

L-Lactate 7.41 89 71 -55 -16 HMDB00190

L-Malate 9 133 115 -40 -14 HMDB00156

Maleate 7.4 115 71 -110 -54 HMDB00176

119

Malonate 7.3 103 59 -40 -15 HMDB00691

Malonyl Co-enzyme A (MCoA) 8.7 852 159 -130 -96 HMDB01175

Melibiose 3.2 341 89 -50 -28 HMDB00048

Methylmalonate 2.7 117 73 -110 -54 HMDB00202

Mevalonate 7.4 147 59 -40 -19 HMDB00227

myo-inositol 3.2 179 87 -65 -26 HMDB00211

N-acetylglucosamine (GlcNAc) 3.5 220 119 -60 -10 HMDB00215

N-acetylglucosamine phosphate

(GlcNAcP)

7 300 199 -75 -20 HMDB02817

N-Acetylglutamate 8.1 188 143 -35 -18 HMDB01138

NADH 8.1 664 408 -130 -46 HMDB01487

NADPH 11.6 372 134 -60 -30 HMDB00221

Neopteron 5.8 252 192 -50 -22 HMDB00845

Nicotinamide adenine dinucleotide

(NAD+)

6.9 662 540 -90 -28 HMDB00902

Nicotinamide adenine dinucleotide

phosphate (NADP)

9 371 309 -40 -10 HMDB00217

Nicotinate ribonucleotide 9.9 334 290 -50 -13 HMDB01132

Nicotinate 7.8 122 78 -35 -18 HMDB01488

hydroxyphenylpyruvate 3.3 179 107 -40 -12 HMDB00205

o-Nitrophenol 4.9 138 108 -50 -22 HMDB01232

o-Phosphoryl-ethanol-amine 4.7 140 79 -110 -54 HMDB00224

Orotate 8.8 155 111 -50 -22 HMDB00226

Orotidine 5'-monophosphate

(OMP)

3.6 367 323 -50 -17 HMDB00218

Oxaloacetate 8.7 133 87 -30 -12 HMDB00223

Oxidized glutathione (GSSG) 7.8 611 306 -60 -60 HMDB03337

Palmitate 6.5 255 69 -75 -80 HMDB00220

Pantothenate 7.5 218 88 -55 -22 HMDB00210

Phenylpyruvate 9.5 163 91 -25 -14 HMDB00205

Phosphoenolpyruvate 8.5 167 79 -40 -31 HMDB00263

Phosphotyrosine 11.1 260 79 -50 -30 HMDB06049

Phytate 11.6 329 79 -40 -98 HMDB03502

PPA 9.5 163 91 -40 -15 HMDB04110

Propionate 4 73 55 -30 -20 HMDB00237

Prostaglandin E2 14.3 351 315 -40 -20 HMDB01220

Pyridoxal-5-Phosphate 7.4 246 97 -60 -20 HMDB01491

Pyruvate 7.7 87 43 -35 -10 HMDB00243

Quinolinate 9.1 166 122 -30 -13 HMDB00232

Reduced glutathione (GSH) 7 306 143 -40 -32 HMDB00125

Salicylurate 12.3 194 150 -60 -20 HMDB00840

Shikimate 8.9 173 93 -25 -20 HMDB03070

Sialate 7 308 170 -50 -22 HMDB00230

Sorbitol 3.4 181 89 -60 -20 HMDB00247

120

Succinate 8 117 73 -30 -16 HMDB00254

Succinyl Co-enzyme A (SCoA) 8.8 866 159 -135 -90 HMDB01022

Taurochenodeoxycholate 14.3 498 80 -90 -88 HMDB00951

Taurocholate 10.5 514 80 -50 -90 HMDB00036

Thiamine pyrophosphate 3.6 424 382 -10 -24 HMDB01372

Trehalose 3.2 341 59 -80 -52 HMDB00975

UDP-GlcNAc 8 606 159 -110 -66 HMDB00290

UDP-glucuronate (UDP-G) 8.5 579 403 -50 -28 HMDB00935

Urate 6.2 167 124 -60 -22 HMDB00289

Uridine 5'-monophosphate (UMP) 7.9 323 79 -65 -66 HMDB00288

Uridine 5-triphosphate (UTP) 8.6 483 159 -75 -45 HMDB00285

Uridine diphosphate (UDP) 8.6 403 159 -90 -36 HMDB00295

Uridine diphosphate-glucose 8.1 565 323 -85 -30 HMDB00286

Xanthosine 5-monophosphate

(XMP)

7.4 363 779 -40 -60 METPA1379

Xanthurenate 8.3 204 160 -30 -33 HMDB00881

Xyl(ul)ose-5P 7.5 229 97 -30 -10 HMDB00868

121

6.2. LC-MS/MS transitions for 1,2-13

C2 Glucose intermediates.

DP: de-clustering potential, EP: entrance potential, CE: Collision energy, CXP: collision exit

potential

Q1 Q3 RT ID DP CE CXP

87 43 7.41

Pyruvate -35 -14 -3

89 45 [1,2]13

C-Pyruvate [M2] -35 -14 -3

89 43

7.41

Lactate -55 -16 -1

90 44 [1]13

C-lactate [M1] -55 -16 -1

91 45 [1,2]13

C-lactate [M2] -55 -16 -1

191 111 8.6

Citrate -30 -18 -7

193 112 [1,2]13

C-Citrate [M2] -30 -18 -7

259 97 7.33 G6P -40 -28 -5

261 79 7.33

[1,2]13

C-G6P [M2] -40 -28 -5

261 97 [1,2]13

C-G6P [M2] -40 -28 -5

259 79 8.26 D-Fructose-6P -55 -72 -11

261 79 8.26 [1,2]13

C-Fructose-6P [M2] -55 -72 -11

300 79 7.7

GlcNAcP -45 -76 -15

302 79 [1,2]13

C-GlcNAcP [M2] -60 -76 -15

606 159

8.5

UDP-GlcNAc -110 -66 -1

608 159 UDP-[1,2]13

C-GlcNAc [M2] -110 -66 -1

610 159 UDP-[1,2]13

C-GlcNAc [M4] -110 -66 -1

808 159

8.5

Acetyl Co-enzyme A (ACoA) -150 -88 -9

809 159 [2]13

C-Acetyl Co-enzyme A (ACoA) [M1] -150 -88 -9

810 159 [1,2]13

C-Acetyl Co-enzyme A (ACoA) [M2] -150 -88 -9

122

6.3. List of gene differentially regulated by p66Shc expression

Gene Gene full Name p66Shc KO -

p66+(average log 10)

p-value SD

Lrp2 low density lipoprotein receptor-

related protein 2

-1.915792811 0.00053 0.265341548

Lingo4 leucine rich repeat and Ig domain

containing 4

-1.811543741 0.00053 0.08564836

Fam19a2 family with sequence similarity 19,

member A2

-1.635935145 0.00053 0.094915819

Fras1 Fraser syndrome 1 homolog (human) -1.632833069 0.00053 0.088349326

Epha7 Eph receptor A7 -1.617034208 0.00053 0.265811188

2810405K02Rik RIKEN cDNA 2810405K02 gene -1.545497146 0.00053 0.104072241

Pcdhgb8 protocadherin gamma subfamily B, 8 -1.505347374 0.00053 0.134934987

Nol4 nucleolar protein 4 -1.491293872 0.00053 0.147634015

Elavl2 ELAV (embryonic lethal, abnormal

vision, Drosophila)-like 2 (Hu antigen

B)

-1.475961437 0.00053 0.172756757

Slc38a4 solute carrier family 38, member 4 -1.450549921 0.00053 0.082508643

Slc16a7 solute carrier family 16

(monocarboxylic acid transporters),

member 7

-1.422033949 0.00053 0.180917562

Fam132b family with sequence similarity 132,

member B

-1.408791619 0.00053 0.067851145

Fndc3c1 fibronectin type III domain containing

3C1

-1.317312614 0.00053 0.082532254

Cpa6 carboxypeptidase A6 -1.311506175 0.00053 0.39594187

Lrrtm1 leucine rich repeat transmembrane

neuronal 1

-1.30232534 0.00053 0.261829654

Tmem178 transmembrane protein 178 -1.229814213 0.00053 0.252731011

Pcdh9 protocadherin 9 -1.209521591 0.00053 0.06012591

Rnf128 ring finger protein 128 -1.200148584 0.00053 0.295306529

Mfap3l microfibrillar-associated protein 3-like -1.184275636 0.00053 0.471271669

123

Nr3c2 nuclear receptor subfamily 3, group C,

member 2

-1.179205418 0.00053 0.245215328

Epha3 Eph receptor A3 -1.141945788 0.00053 0.18094104

Pde3b phosphodiesterase 3B, cGMP-

inhibited

-1.126724114 0.00053 0.058435767

Armcx4 armadillo repeat containing, X-linked

4

-1.118182406 0.00053 0.009426341

Pcsk5 proprotein convertase subtilisin/kexin

type 5

-1.110683672 0.00053 0.122133709

Mycn v-myc myelocytomatosis viral related

oncogene, neuroblastoma derived

(avian)

-1.109619722 0.00053 0.085221891

Fibin fin bud initiation factor homolog

(zebrafish)

-1.075052764 0.00053 0.053954081

Mei4 meiosis-specific, MEI4 homolog (S.

cerevisiae)

-1.027144733 0.00054 0.404083607

Slc26a7 solute carrier family 26, member 7 -1.018616084 0.00053 0.265390359

Pmaip1 phorbol-12-myristate-13-acetate-

induced protein 1

-0.991734978 0.00053 0.180923943

Sfrp2 secreted frizzled-related protein 2 -0.974846673 0.00053 0.172783959

Tll1 tolloid-like -0.957443259 0.00053 0.175013394

Tmem151b transmembrane protein 151B -0.952357544 0.00053 0.148567314

Tnfrsf21 tumor necrosis factor receptor

superfamily, member 21

-0.92950432 0.00053 0.067073319

Lef1 lymphoid enhancer binding factor 1 -0.929191993 0.00054 0.311455724

Hunk hormonally upregulated Neu-

associated kinase

-0.925378588 0.00054 0.488512276

Tdrkh tudor and KH domain containing

protein

-0.924298141 0.00053 0.016466993

Plcl1 phospholipase C-like 1 -0.922570232 0.00053 0.050416995

Pkia protein kinase inhibitor, alpha -0.917178399 0.00053 0.145255127

Tm6sf2 transmembrane 6 superfamily member

2

-0.909526811 0.00054 0.265433652

Mtap7d3 MAP7 domain containing 3 -0.905721395 0.00054 0.265381819

124

Zfp583 zinc finger protein 583 -0.89964319 0.00054 0.231461609

Fam38b family with sequence similarity 38,

member B

-0.886354776 0.00054 0.393893458

Sort1 sortilin 1 -0.883128964 0.00053 0.036959691

Hist1h2bg histone cluster 1, H2bg -0.870775698 0.00054 0.08807413

Snca synuclein, alpha -0.850845519 0.00054 0.101000827

Dok5 docking protein 5 -0.829479412 0.00054 0.098944847

Rell2 RELT-like 2 -0.824736054 0.00054 0.261803243

Gabre gamma-aminobutyric acid (GABA) A

receptor, subunit epsilon

-0.824104256 0.00054 0.134844879

Hist1h4k histone cluster 1, H4k -0.821945916 0.00054 0.296458294

Gm10406 predicted gene 10406 -0.808209064 0.00054 0.180904117

Cacna1b calcium channel, voltage-dependent, N

type, alpha 1B subunit

-0.805835164 0.00054 0.143616773

Hoxb9 homeobox B9 -0.797828438 0.00054 0.198343907

Asxl3 additional sex combs like 3

(Drosophila)

-0.791955705 0.00054 0.054836257

Pcdhb2 protocadherin beta 2 -0.787980234 0.00054 0.160613469

Snx10 sorting nexin 10 -0.784827431 0.00054 0.100231686

Pcdhb3 protocadherin beta 3 -0.779873929 0.00054 0.114090528

4921528I01Rik RIKEN cDNA 4921528I01 gene -0.777509771 0.00054 0.192939047

Wnt2b wingless related MMTV integration

site 2b

-0.755864987 0.00054 0.0903634

Peg10 paternally expressed 10 -0.744421658 0.00054 0.10331376

Fndc5 fibronectin type III domain containing

5

-0.744227916 0.00054 0.074817703

Ntng2 netrin G2 -0.741595235 0.00054 0.121176342

Klf12 Kruppel-like factor 12 -0.738830144 0.00054 0.041258524

Npy1r neuropeptide Y receptor Y1 -0.736048621 0.00054 0.125603467

Gpr137c G protein-coupled receptor 137C -0.73215916 0.00059 0.534608914

8430408G22Rik RIKEN cDNA 8430408G22 gene -0.724087278 0.00055 0.222465861

125

Bhlhe22 basic helix-loop-helix family, member

e22

-0.718249387 0.00054 0.150545533

Nlrx1 NLR family member X1 -0.713254896 0.00054 0.030703712

Elovl4 elongation of very long chain fatty

acids (FEN1/Elo2, SUR4/Elo3, yeast)-

like 4

-0.711290071 0.00055 0.223211263

Chdh choline dehydrogenase -0.705962742 0.00056 0.361115226

Sema6a sema domain, transmembrane domain

(TM), and cytoplasmic domain,

(semaphorin) 6A

-0.703651997 0.00054 0.059363453

Cthrc1 collagen triple helix repeat containing

1

-0.701838551 0.00055 0.173849535

Enox1 ecto-NOX disulfide-thiol exchanger 1 -0.699117306 0.00054 0.114088181

Fbxl7 F-box and leucine-rich repeat protein 7 -0.698448854 0.00054 0.044721972

Snrpn small nuclear ribonucleoprotein N -0.696287165 0.00055 0.123683742

1700010I14Rik RIKEN cDNA 1700010I14 gene -0.687747012 0.00056 0.2527405

Tmem169 transmembrane protein 169 -0.686456758 0.00057 0.395917183

Palm3 paralemmin 3 -0.684145211 0.00059 0.50659648

Tnfaip8l1 tumor necrosis factor, alpha-induced

protein 8-like 1

-0.683349349 0.00055 0.029609797

Cd200 CD200 antigen -0.681445884 0.00055 0.046093045

Pcdh8 protocadherin 8 -0.680995051 0.00056 0.27531672

Dclk2 doublecortin-like kinase 2 -0.676081914 0.00055 0.134788259

Igf2 insulin-like growth factor 2 -0.674359709 0.00055 0.036664029

Fut4 fucosyltransferase 4 -0.672636529 0.00057 0.404091035

F2rl1 coagulation factor II (thrombin)

receptor-like 1

-0.664956186 0.00055 0.016152716

Tmem200a transmembrane protein 200A -0.663179322 0.00055 0.036468114

Tbx2 T-box 2 0.663640332 0.00055 0.021998753

Gm1661 predicted gene 1661 0.665248537 0.00055 0.108950542

Diras2 DIRAS family, GTP-binding RAS-

like 2

0.667441057 0.00056 0.246125103

126

Rab3b RAB3B, member RAS oncogene

family

0.667673045 0.00055 0.087214478

Col28a1 collagen, type XXVIII, alpha 1 0.668464877 0.00055 0.103303578

Agt angiotensinogen (serpin peptidase

inhibitor, clade A, member 8)

0.668787912 0.00055 0.059258563

Slc22a23 solute carrier family 22, member 23 0.670209459 0.00055 0.036892269

Il13ra2 interleukin 13 receptor, alpha 2 0.670898039 0.00055 0.077902592

Syt5 synaptotagmin V 0.673105204 0.00055 0.134876587

Cfh complement component factor h 0.673403902 0.00055 0.014587192

Cdh26 cadherin-like 26 0.674748099 0.00055 0.171407963

1700034H15Rik RIKEN cDNA 1700034H15 gene 0.676046005 0.00056 0.339792831

Lgals9 lectin, galactose binding, soluble 9 0.678298366 0.00055 0.051080452

Gm14393 predicted gene 14393 0.680908099 0.00058 0.403850786

Adh7 alcohol dehydrogenase 7 (class IV),

mu or sigma polypeptide

0.68327306 0.00055 0.152894719

Cygb cytoglobin 0.683903613 0.00055 0.157736829

4-Sep septin 4 0.684369152 0.00061 0.376028174

Nsg1 neuron specific gene family member 1 0.68443605 0.00055 0.027139644

Tmem117 transmembrane protein 117 0.6868859 0.00055 0.123397066

Siglecg sialic acid binding Ig-like lectin G 0.687365882 0.00055 0.072380418

Cbln3 cerebellin 3 precursor protein 0.688270313 0.00055 0.066008871

Islr immunoglobulin superfamily

containing leucine-rich repeat

0.692047166 0.00054 0.019871733

Adamtsl1 ADAMTS-like 1 0.695033204 0.00054 0.104145348

Lama3 laminin, alpha 3 0.69513384 0.00054 0.108472544

Acy3 aspartoacylase (aminoacylase) 3 0.696801941 0.00054 0.086061403

Atf3 activating transcription factor 3 0.69773903 0.00054 0.024900306

Rsph1 radial spoke head 1 homolog

(Chlamydomonas)

0.701830671 0.00055 0.207766755

Phactr1 phosphatase and actin regulator 1 0.703918435 0.00054 0.055588964

127

9930023K05Rik RIKEN cDNA 9930023K05 gene 0.704413698 0.00058 0.462960081

Atp8b1 ATPase, class I, type 8B, member 1 0.704855288 0.00054 0.066155661

Timp3 tissue inhibitor of metalloproteinase 3 0.705980282 0.00054 0.009519119

Ticam2 toll-like receptor adaptor molecule 2 0.706980197 0.00055 0.216094637

Nhsl2 NHS-like 2 0.708152301 0.00055 0.187820581

Tcf7 transcription factor 7, T cell specific 0.710676527 0.00054 0.079743011

Asgr1 asialoglycoprotein receptor 1 0.710908571 0.00054 0.102195858

Gem GTP binding protein (gene

overexpressed in skeletal muscle)

0.711129665 0.00054 0.074775012

Acss1 acyl-CoA synthetase short-chain

family member 1

0.713136814 0.00056 0.332812454

Adrbk2 adrenergic receptor kinase, beta 2 0.71389249 0.00054 0.017370486

Mapkapk3 mitogen-activated protein kinase-

activated protein kinase 3

0.717198672 0.00054 0.11169346

Map3k5 mitogen-activated protein kinase

kinase kinase 5

0.722403796 0.00054 0.060276991

Islr2 immunoglobulin superfamily

containing leucine-rich repeat 2

0.723526789 0.00054 0.137727559

Ggt7 gamma-glutamyltransferase 7 0.724810409 0.00054 0.121828837

Zfp296 zinc finger protein 296 0.726254956 0.00059 0.440252288

Sh2d1b1 SH2 domain protein 1B1 0.726373223 0.00054 0.086700286

B3gnt8 UDP-GlcNAc:betaGal beta-1,3-N-

acetylglucosaminyltransferase 8

0.729573708 0.00057 0.303349205

Tnni3 troponin I, cardiac 3 0.731254977 0.00058 0.514112037

9130019O22Rik RIKEN cDNA 9130019O22 gene 0.732389495 0.00054 0.098879186

Ccbe1 collagen and calcium binding EGF

domains 1

0.733942913 0.00054 0.137946031

Gstt1 glutathione S-transferase, theta 1 0.734574564 0.00054 0.095911655

Rtp4 receptor transporter protein 4 0.737249757 0.00054 0.050153752

Abi3 ABI gene family, member 3 0.739334553 0.00054 0.101934592

Aldh1a3 aldehyde dehydrogenase family 1,

subfamily A3

0.742734742 0.00054 0.085524474

128

Rasl11a RAS-like, family 11, member A 0.743417171 0.00054 0.069964518

Epb4.1l4a erythrocyte protein band 4.1-like 4a 0.743720766 0.00054 0.067778459

Ica1 islet cell autoantigen 1 0.746449159 0.00058 0.532783978

Hyal3 hyaluronoglucosaminidase 3 0.746670131 0.00058 0.532544673

Ifi27l2a interferon, alpha-inducible protein 27

like 2A

0.748805838 0.00054 0.128921448

Matn4 matrilin 4 0.751296432 0.00055 0.309099514

Afap1l2 actin filament associated protein 1-like

2

0.754643517 0.00054 0.040200803

Abcc3 ATP-binding cassette, sub-family C

(CFTR/MRP), member 3

0.755485412 0.00055 0.246203752

Deptor DEP domain containing MTOR-

interacting protein

0.758137134 0.00054 0.116223933

Smoc1 SPARC related modular calcium

binding 1

0.759456057 0.00056 0.432661364

Dcn decorin 0.759916709 0.00054 0.063105273

Cldn1 claudin 1 0.761833728 0.00054 0.04408684

Traf1 TNF receptor-associated factor 1 0.764048952 0.00054 0.065079109

Cfb complement factor B 0.76455062 0.00054 0.088308557

Vtcn1 V-set domain containing T cell

activation inhibitor 1

0.765821423 0.00055 0.267925708

Acta1 actin, alpha 1, skeletal muscle 0.765964795 0.00054 0.043835532

Gm12216 predicted gene 12216 0.76644448 0.00054 0.135900469

1700003F12Rik RIKEN cDNA 1700003F12 gene 0.768055495 0.00054 0.089105998

Saa3 serum amyloid A 3 0.773580491 0.00054 0.173053452

Cdhr1 cadherin-related family member 1 0.774388432 0.00054 0.045177575

Ptgfr prostaglandin F receptor 0.775513945 0.00054 0.090611927

Susd2 sushi domain containing 2 0.775594375 0.00054 0.097255044

Cdkn1c cyclin-dependent kinase inhibitor 1C

(P57)

0.776172936 0.00054 0.06379327

Mndal myeloid nuclear differentiation antigen

like

0.779296909 0.00055 0.379651909

129

Clec11a C-type lectin domain family 11,

member a

0.780058561 0.00054 0.055581965

Apol10b apolipoprotein L 10B 0.782834626 0.00054 0.137041661

Cyp26b1 cytochrome P450, family 26,

subfamily b, polypeptide 1

0.782836703 0.00054 0.117528419

Kctd14 potassium channel tetramerisation

domain containing 14

0.784766049 0.00056 0.360430947

AI428936 expressed sequence AI428936 0.787294696 0.00054 0.099279262

Adam33 a disintegrin and metallopeptidase

domain 33

0.788839092 0.00054 0.085605242

Aox1 aldehyde oxidase 1 0.790952518 0.00054 0.028503273

Blnk B cell linker 0.790963356 0.00054 0.301575058

Ecscr endothelial cell surface expressed

chemotaxis and apoptosis regulator

0.793488682 0.00054 0.035874271

Rarres2 retinoic acid receptor responder

(tazarotene induced) 2

0.793997302 0.00054 0.137749917

Epas1 endothelial PAS domain protein 1 0.794441813 0.00054 0.093825829

Fam179a family with sequence similarity 179,

member A

0.797152443 0.00054 0.085500996

Sncg synuclein, gamma 0.797621076 0.00054 0.089113939

Angptl7 angiopoietin-like 7 0.79815192 0.00054 0.044478057

Eln elastin 0.798869667 0.00054 0.018790951

Gprin3 GPRIN family member 3 0.800688631 0.00054 0.128589472

Rhox5 reproductive homeobox 5 0.801127842 0.00055 0.265045427

Pcdhga2 protocadherin gamma subfamily A, 2 0.802342483 0.00055 0.4339809

Ppl periplakin 0.806596571 0.00054 0.017650785

Pdzrn4 PDZ domain containing RING finger

4

0.816216272 0.00054 0.03802449

Bdh1 3-hydroxybutyrate dehydrogenase,

type 1

0.818013017 0.00054 0.13315328

Gbp6 guanylate binding protein 6 0.820104623 0.00054 0.059587685

Auts2 autism susceptibility candidate 2 0.823419711 0.00054 0.017035484

130

Peg3 paternally expressed 3 0.824601124 0.00054 0.095374777

Slfn2 schlafen 2 0.828174735 0.00054 0.180890243

Espnl espin-like 0.829729488 0.00054 0.187549219

Raet1d retinoic acid early transcript delta 0.831300419 0.00054 0.084477615

Col10a1 collagen, type X, alpha 1 0.834622555 0.00054 0.043812899

Usp18 ubiquitin specific peptidase 18 0.841243699 0.00054 0.33200192

Rab40b Rab40b, member RAS oncogene

family

0.845386805 0.00054 0.216744132

Ddit4l DNA-damage-inducible transcript 4-

like

0.846972376 0.00054 0.038527775

Panx1 pannexin 1 0.847590486 0.00054 0.159315477

Isg15 ISG15 ubiquitin-like modifier 0.856119102 0.00054 0.114701267

Lxn latexin 0.85697897 0.00054 0.065921894

Cercam cerebral endothelial cell adhesion

molecule

0.860725077 0.00054 0.094089887

Ankrd6 ankyrin repeat domain 6 0.861965839 0.00054 0.148669382

Naalad2 N-acetylated alpha-linked acidic

dipeptidase 2

0.863873157 0.00054 0.043166595

Ceacam1 carcinoembryonic antigen-related cell

adhesion molecule 1

0.868657678 0.00054 0.09876767

Nptx1 neuronal pentraxin 1 0.868733414 0.00055 0.432899662

Tnfrsf9 tumor necrosis factor receptor

superfamily, member 9

0.869749964 0.00054 0.0822082

Apol6 apolipoprotein L 6 0.870822486 0.00054 0.096801623

Thsd4 thrombospondin, type I, domain

containing 4

0.874631254 0.00054 0.375134475

Pcdhb13 protocadherin beta 13 0.877924773 0.00054 0.085463636

Acsbg1 acyl-CoA synthetase bubblegum

family member 1

0.878723314 0.00054 0.30885885

Serpina3n serine (or cysteine) peptidase inhibitor,

clade A, member 3N

0.879110484 0.00053 0.041690783

Aoc3 amine oxidase, copper containing 3 0.881136113 0.00053 0.018828905

131

5430407P10Rik RIKEN cDNA 5430407P10 gene 0.881624059 0.00053 0.072724759

Serpina3m serine (or cysteine) peptidase inhibitor,

clade A, member 3M

0.883754442 0.00053 0.060574541

Bst1 bone marrow stromal cell antigen 1 0.886785141 0.00056 0.540574601

Stac2 SH3 and cysteine rich domain 2 0.892804709 0.00054 0.147549381

Gjb3 gap junction protein, beta 3 0.893450437 0.00055 0.431657445

Nap1l3 nucleosome assembly protein 1-like 3 0.893947626 0.00054 0.139544026

Kng2 kininogen 2 0.895607628 0.00053 0.074813929

Il33 interleukin 33 0.896838083 0.00054 0.144757383

Ngfr nerve growth factor receptor (TNFR

superfamily, member 16)

0.898674402 0.00053 0.085524089

Sorl1 sortilin-related receptor, LDLR class

A repeats-containing

0.899089977 0.00054 0.152335302

H2-T10 histocompatibility 2, T region locus 10 0.902292843 0.00054 0.380676837

Insc inscuteable homolog (Drosophila) 0.90417236 0.00053 0.112156662

Spink2 serine peptidase inhibitor, Kazal type

2

0.905591442 0.00054 0.14676946

Clstn3 calsyntenin 3 0.908201748 0.00053 0.089411025

Gpr4 G protein-coupled receptor 4 0.909466117 0.00054 0.216203667

Cxcl17 chemokine (C-X-C motif) ligand 17 0.909828619 0.00055 0.537563025

Abcg1 ATP-binding cassette, sub-family G

(WHITE), member 1

0.91071001 0.00054 0.274435175

Plcd4 phospholipase C, delta 4 0.913182402 0.00054 0.171352782

Ifit1 interferon-induced protein with

tetratricopeptide repeats 1

0.918782586 0.00053 0.175317001

Cd74 CD74 antigen (invariant polypeptide

of major histocompatibility complex,

class II antigen-associated)

0.926534365 0.00053 0.095906798

Rhpn2 rhophilin, Rho GTPase binding

protein 2

0.927138782 0.00053 0.08551548

Dgkg diacylglycerol kinase, gamma 0.927520067 0.00053 0.100662246

Adh1 alcohol dehydrogenase 1 (class I) 0.928183835 0.00054 0.359441834

132

Edar ectodysplasin-A receptor 0.928254958 0.00053 0.072900312

Il2rb interleukin 2 receptor, beta chain 0.928805896 0.00054 0.267954993

Aim1 absent in melanoma 1 0.929351782 0.00053 0.057430612

Prex2 phosphatidylinositol-3,4,5-

trisphosphate-dependent Rac exchange

factor 2

0.93719502 0.00053 0.08997535

Ptk2b PTK2 protein tyrosine kinase 2 beta 0.938135831 0.00053 0.027228824

Ldhd lactate dehydrogenase D 0.93871423 0.00054 0.439647844

1700024P16Rik RIKEN cDNA 1700024P16 gene 0.938819604 0.00054 0.360716271

Gstt3 glutathione S-transferase, theta 3 0.948569948 0.00053 0.024634112

Sim2 single-minded homolog 2

(Drosophila)

0.953991468 0.00053 0.143385491

Efhd1 EF hand domain containing 1 0.956252223 0.00053 0.103256292

Oasl2 2'-5' oligoadenylate synthetase-like 2 0.959954864 0.00053 0.023324485

Neurl3 neuralized homolog 3 homolog

(Drosophila)

0.964136118 0.00053 0.134849684

Kng1 kininogen 1 0.977131516 0.00053 0.049991385

Sp7 Sp7 transcription factor 7 0.978273933 0.00054 0.309076084

Parm1 prostate androgen-regulated mucin-

like protein 1

0.980508708 0.00053 0.058662293

Foxf2 forkhead box F2 0.981038713 0.00053 0.059632285

C1qtnf5 C1q and tumor necrosis factor related

protein 5

0.982903387 0.00054 0.290011907

Spib Spi-B transcription factor (Spi-1/PU.1

related)

0.983279881 0.00053 0.157542051

Fgd4 FYVE, RhoGEF and PH domain

containing 4

0.984260836 0.00053 0.139318882

D14Ertd668e DNA segment, Chr 14, ERATO Doi

668, expressed

0.988864035 0.00053 0.127634024

Dll1 delta-like 1 (Drosophila) 0.992361465 0.00054 0.432629062

Rnase1 ribonuclease, RNase A family, 1

(pancreatic)

0.993085648 0.00054 0.285009166

Acox2 acyl-Coenzyme A oxidase 2, branched 0.993184757 0.00053 0.094325878

133

chain

Cybrd1 cytochrome b reductase 1 0.994028526 0.00053 0.225943878

Sult1a1 sulfotransferase family 1A, phenol-

preferring, member 1

0.995373222 0.00053 0.12120022

Tfap2a transcription factor AP-2, alpha 0.997539572 0.00053 0.135302675

Pdgfb platelet derived growth factor, B

polypeptide

0.998420792 0.00053 0.163272181

Hck hemopoietic cell kinase 0.999070587 0.00053 0.250495191

C2cd4a C2 calcium-dependent domain

containing 4A

0.999639773 0.00053 0.208161313

Pcdhgb1 protocadherin gamma subfamily B, 1 1.005347499 0.00053 0.080214417

9030224M15Rik RIKEN cDNA 9030224M15 gene 1.013271323 0.00054 0.359742465

Mustn1 musculoskeletal, embryonic nuclear

protein 1

1.015777183 0.00053 0.127381615

Tbxa2r thromboxane A2 receptor 1.015799552 0.00053 0.044372957

Rapsn receptor-associated protein of the

synapse

1.017472248 0.00053 0.106670772

Mgst1 microsomal glutathione S-transferase

1

1.018475319 0.00053 0.0203985

Gas6 growth arrest specific 6 1.018597236 0.00053 0.012972127

Ptprq protein tyrosine phosphatase, receptor

type, Q

1.023762028 0.00053 0.09260246

Mrvi1 MRV integration site 1 1.024551664 0.00053 0.047280673

Selenbp1 selenium binding protein 1 1.027124543 0.00053 0.136774659

H28 histocompatibility 28 1.027624929 0.00053 0.114131969

Pcdhga6 protocadherin gamma subfamily A, 6 1.027721531 0.00053 0.273393221

Serpina3f serine (or cysteine) peptidase inhibitor,

clade A, member 3F

1.031536232 0.00053 0.178581046

Ccl27a chemokine (C-C motif) ligand 27A 1.03329066 0.00054 0.484304318

Iigp1 interferon inducible GTPase 1 1.036102758 0.00053 0.065017448

Apol9b apolipoprotein L 9b 1.039598534 0.00053 0.153952589

Mmp13 matrix metallopeptidase 13 1.040888988 0.00053 0.083124011

134

Fmo4 flavin containing monooxygenase 4 1.046495119 0.00053 0.216815067

Steap4 STEAP family member 4 1.052428065 0.00053 0.002584317

Cln3 ceroid lipofuscinosis, neuronal 3,

juvenile (Batten, Spielmeyer-Vogt

disease)

1.058161466 0.00053 0.309026893

C1qtnf3 C1q and tumor necrosis factor related

protein 3

1.062870795 0.00053 0.085885934

Dio2 deiodinase, iodothyronine, type II 1.07779171 0.00053 0.209791746

Tmprss6 transmembrane serine protease 6 1.079375769 0.00053 0.033600766

Xdh xanthine dehydrogenase 1.085912708 0.00053 0.079133761

Pcdhgb6 protocadherin gamma subfamily B, 6 1.088637837 0.00053 0.059579985

Alox5ap arachidonate 5-lipoxygenase

activating protein

1.09044588 0.00053 0.166524406

Notum notum pectinacetylesterase homolog

(Drosophila)

1.10405039 0.00053 0.094076015

Lbp lipopolysaccharide binding protein 1.113722934 0.00053 0.03739157

Clec3b C-type lectin domain family 3,

member b

1.116350535 0.00053 0.076252528

Enpp2 ectonucleotide

pyrophosphatase/phosphodiesterase 2

1.117753237 0.00053 0.347134136

Acpp acid phosphatase, prostate 1.120588722 0.00053 0.137820192

Rragb Ras-related GTP binding B 1.125262389 0.00053 0.309078515

Gdnf glial cell line derived neurotrophic

factor

1.132545993 0.00053 0.109641032

Hp haptoglobin 1.13791102 0.00053 0.43130165

Lcn2 lipocalin 2 1.139064261 0.00053 0.08005039

Hspa12b heat shock protein 12B 1.14116058 0.00053 0.085648791

Syt13 synaptotagmin XIII 1.143486055 0.00053 0.027679077

Slc16a2 solute carrier family 16

(monocarboxylic acid transporters),

member 2

1.159110478 0.00053 0.398843774

Sod3 superoxide dismutase 3, extracellular 1.167418756 0.00053 0.056023357

Ugt1a6b UDP glucuronosyltransferase 1 1.170555987 0.00053 0.128208789

135

family, polypeptide A6B

Slc22a18 solute carrier family 22 (organic cation

transporter), member 18

1.174411334 0.00053 0.145295089

Entpd3 ectonucleoside triphosphate

diphosphohydrolase 3

1.180120689 0.00053 0.1037217

Cxcl10 chemokine (C-X-C motif) ligand 10 1.193671012 0.00053 0.140252908

Zfp941 zinc finger protein 941 1.194086034 0.00053 0.085421735

Elfn1 leucine rich repeat and fibronectin

type III, extracellular 1

1.194848088 0.00053 0.020379081

Apol9a apolipoprotein L 9a 1.197174371 0.00053 0.246164347

Gpbar1 G protein-coupled bile acid receptor 1 1.198285099 0.00053 0.135408239

Cda cytidine deaminase 1.205054123 0.00053 0.237088505

Fam107a family with sequence similarity 107,

member A

1.205425335 0.00053 0.039519945

Nr1h3 nuclear receptor subfamily 1, group H,

member 3

1.206406134 0.00053 0.091558614

Plcg2 phospholipase C, gamma 2 1.206711745 0.00053 0.398864245

Trim30a tripartite motif-containing 30A 1.208785913 0.00053 0.144143522

Csmd1 CUB and Sushi multiple domains 1 1.210273167 0.00053 0.057302114

Psca prostate stem cell antigen 1.212806007 0.00053 0.091731027

Psmb9 proteasome (prosome, macropain)

subunit, beta type 9 (large

multifunctional peptidase 2)

1.213436685 0.00053 0.1345473

Pcsk6 proprotein convertase subtilisin/kexin

type 6

1.216391075 0.00053 0.032624856

Prelp proline arginine-rich end leucine-rich

repeat

1.216786271 0.00053 0.051151785

A4galt alpha 1,4-galactosyltransferase 1.22310685 0.00053 0.061975306

Vnn1 vanin 1 1.22704685 0.00053 0.016872923

Fxyd1 FXYD domain-containing ion

transport regulator 1

1.231112772 0.00053 0.523075014

Trpv2 transient receptor potential cation

channel, subfamily V, member 2

1.248179355 0.00053 0.216250134

136

Renbp renin binding protein 1.248781807 0.00053 0.017383204

Grhl1 grainyhead-like 1 (Drosophila) 1.253278296 0.00053 0.270603819

Msl3l2 male-specific lethal 3-like 2

(Drosophila)

1.25986456 0.00053 0.137146309

Arhgap6 Rho GTPase activating protein 6 1.266639625 0.00053 0.180941556

Ccl5 chemokine (C-C motif) ligand 5 1.300214635 0.00053 0.131616389

H60b histocompatibility 60b 1.301586423 0.00053 0.311046603

H2-Q4 histocompatibility 2, Q region locus 4 1.303535724 0.00053 0.084559364

C3 complement component 3 1.322856021 0.00053 0.063090714

AI607873 expressed sequence AI607873 1.346856052 0.00053 0.137647255

Cyp2j9 cytochrome P450, family 2, subfamily

j, polypeptide 9

1.359477206 0.00053 0.086144262

Tnfsf10 tumor necrosis factor (ligand)

superfamily, member 10

1.375656394 0.00053 0.185884401

C130074G19Rik RIKEN cDNA C130074G19 gene 1.377624185 0.00053 0.13747556

Serping1 serine (or cysteine) peptidase inhibitor,

clade G, member 1

1.386331531 0.00053 0.095017679

Nfe2l3 nuclear factor, erythroid derived 2,

like 3

1.389071601 0.00053 0.032589724

Gda guanine deaminase 1.400864825 0.00053 0.026537201

Crabp1 cellular retinoic acid binding protein I 1.404406104 0.00053 0.242432756

Htatip2 HIV-1 tat interactive protein 2,

homolog (human)

1.407270603 0.00053 0.057078139

Plb1 phospholipase B1 1.44324694 0.00053 0.269786455

Oas2 2'-5' oligoadenylate synthetase 2 1.443352045 0.00053 0.137723246

Inmt indolethylamine N-methyltransferase 1.460243408 0.00053 0.310783802

Pik3r5 phosphoinositide-3-kinase, regulatory

subunit 5, p101

1.467004844 0.00053 0.08535067

Pdzk1ip1 PDZK1 interacting protein 1 1.474467624 0.00053 0.042997429

Mmp3 matrix metallopeptidase 3 1.479116108 0.00053 0.058432453

Tril TLR4 interactor with leucine-rich

repeats

1.50993966 0.00053 0.137914316

137

Tubg2 tubulin, gamma 2 1.527010062 0.00053 0.309441049

Pdpn podoplanin 1.555027038 0.00053 0.4399842

Sp140 Sp140 nuclear body protein 1.570560246 0.00053 0.250312212

Apod apolipoprotein D 1.61198472 0.00053 0.221408291

Chi3l1 chitinase 3-like 1 1.636684153 0.00053 0.583086161

Unc93b1 unc-93 homolog B1 (C. elegans) 1.66968256 0.00053 0.360200982

H2-DMb1 histocompatibility 2, class II, locus

Mb1

1.67565171 0.00053 0.359347716

Slc15a2 solute carrier family 15 (H+/peptide

transporter), member 2

1.700204294 0.00053 0.385075496

Wnt4 wingless-related MMTV integration

site 4

1.713137287 0.00053 0.360869216

Fmo2 flavin containing monooxygenase 2 1.815536475 0.00053 0.308692687

Sparcl1 SPARC-like 1 1.816996676 0.00053 0.063385442

Cyp2f2 cytochrome P450, family 2, subfamily

f, polypeptide 2

1.83370284 0.00053 0.078208358

Klhl29 kelch-like 29 (Drosophila) 1.88351542 0.00053 0.138028824

Eml2 echinoderm microtubule associated

protein like 2

1.970482521 0.00053 0.027241317

Slpi secretory leukocyte peptidase inhibitor 2.08642332 0.00053 0.441915633

Lgals6 lectin, galactose binding, soluble 6 5.671892314 0.00053 0.198517766

138

6.4. List of identified p66Shc-interacting proteins

Identified protein No. of peptides p66+

MEF –

serum starvation

No. of peptides

p66+

MEF – serum

stimulation / 10 min

Shc1 900 920

Sgk269 22 23

FAM59A 4 27

Beta2-adaptin 5 23

Alpha1-adaptin 4 22

Grb2 0 12

Lrrk1 0 9

Anks1 1 7

PTPN12 1 5

Gab1 0 2

PP2C-beta 8 14

PP1A 2 2

139

References

Abdel Rahman, A.M., Ryczko, M., Pawling, J., and Dennis, J.W. (2013). Probing the

hexosamine biosynthetic pathway in human tumor cells by multitargeted tandem mass

spectrometry. ACS chemical biology 8, 2053-2062.

Ahn, R., Sabourin, V., Ha, J.R., Cory, S., Maric, G., Im, Y.K., Hardy, W.R., Zhao, H., Park, M.,

Hallett, M., et al. (2013). The ShcA PTB domain functions as a biological sensor of

phosphotyrosine signaling during breast cancer progression. Cancer research 73, 4521-4532.

Andronesi, O.C., Kim, G.S., Gerstner, E., Batchelor, T., Tzika, A.A., Fantin, V.R., Vander

Heiden, M.G., and Sorensen, A.G. (2012). Detection of 2-hydroxyglutarate in IDH-mutated

glioma patients by in vivo spectral-editing and 2D correlation magnetic resonance spectroscopy.

Sci Transl Med 4, 116ra114.

Bar-Peled, L., Chantranupong, L., Cherniack, A.D., Chen, W.W., Ottina, K.A., Grabiner, B.C.,

Spear, E.D., Carter, S.L., Meyerson, M., and Sabatini, D.M. (2013). A Tumor suppressor

complex with GAP activity for the Rag GTPases that signal amino acid sufficiency to mTORC1.

Science 340, 1100-1106.

Bar-Peled, L., Schweitzer, L.D., Zoncu, R., and Sabatini, D.M. (2012). Ragulator is a GEF for

the rag GTPases that signal amino acid levels to mTORC1. Cell 150, 1196-1208.

Batzer, A.G., Rotin, D., Urena, J.M., Skolnik, E.Y., and Schlessinger, J. (1994). Hierarchy of

binding sites for Grb2 and Shc on the epidermal growth factor receptor. Molecular and cellular

biology 14, 5192-5201.

Ben-Sahra, I., Howell, J.J., Asara, J.M., and Manning, B.D. (2013). Stimulation of de Novo

Pyrimidine Synthesis by Growth Signaling Through mTOR and S6K1. Science 339, 1323-1328.

Bensaad, K., Tsuruta, A., Selak, M.A., Vidal, M.N., Nakano, K., Bartrons, R., Gottlieb, E., and

Vousden, K.H. (2006). TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell 126,

107-120.

Bertout, J.A., Patel, S.A., and Simon, M.C. (2008). The impact of O2 availability on human

cancer. Nature reviews Cancer 8, 967-975.

Berwick, D.C., Hers, I., Heesom, K.J., Moule, S.K., and Tavare, J.M. (2002). The identification

of ATP-citrate lyase as a protein kinase B (Akt) substrate in primary adipocytes. The Journal of

biological chemistry 277, 33895-33900.

140

Bhattacharyya, R.P., Remenyi, A., Yeh, B.J., and Lim, W.A. (2006). Domains, motifs, and

scaffolds: the role of modular interactions in the evolution and wiring of cell signaling circuits.

Annual review of biochemistry 75, 655-680.

Bieging, K.T., and Attardi, L.D. (2012). Deconstructing p53 transcriptional networks in tumor

suppression. Trends in cell biology 22, 97-106.

Birsoy, K., Wang, T., Possemato, R., Yilmaz, O.H., Koch, C.E., Chen, W.W., Hutchins, A.W.,

Gultekin, Y., Peterson, T.R., Carette, J.E., et al. (2013). MCT1-mediated transport of a toxic

molecule is an effective strategy for targeting glycolytic tumors. Nature genetics 45, 104-108.

Brown, E.J., Albers, M.W., Shin, T.B., Ichikawa, K., Keith, C.T., Lane, W.S., and Schreiber,

S.L. (1994). A mammalian protein targeted by G1-arresting rapamycin-receptor complex. Nature

369, 756-758.

Brugarolas, J., Lei, K., Hurley, R.L., Manning, B.D., Reiling, J.H., Hafen, E., Witters, L.A.,

Ellisen, L.W., and Kaelin, W.G., Jr. (2004). Regulation of mTOR function in response to

hypoxia by REDD1 and the TSC1/TSC2 tumor suppressor complex. Genes & development 18,

2893-2904.

Budanov, A.V., and Karin, M. (2008). p53 target genes sestrin1 and sestrin2 connect genotoxic

stress and mTOR signaling. Cell 134, 451-460.

Buller, C.L., Loberg, R.D., Fan, M.H., Zhu, Q., Park, J.L., Vesely, E., Inoki, K., Guan, K.L., and

Brosius, F.C., 3rd (2008). A GSK-3/TSC2/mTOR pathway regulates glucose uptake and GLUT1

glucose transporter expression. American journal of physiology Cell physiology 295, C836-843.

Cafferkey, R., Young, P.R., McLaughlin, M.M., Bergsma, D.J., Koltin, Y., Sathe, G.M.,

Faucette, L., Eng, W.K., Johnson, R.K., and Livi, G.P. (1993). Dominant missense mutations in

a novel yeast protein related to mammalian phosphatidylinositol 3-kinase and VPS34 abrogate

rapamycin cytotoxicity. Molecular and cellular biology 13, 6012-6023.

Cairns, R.A., Harris, I.S., and Mak, T.W. (2011). Regulation of cancer cell metabolism. Nature

reviews Cancer 11, 85-95.

Cam, H., Easton, J.B., High, A., and Houghton, P.J. (2010). mTORC1 signaling under hypoxic

conditions is controlled by ATM-dependent phosphorylation of HIF-1alpha. Molecular cell 40,

509-520.

Camici, G.G., Schiavoni, M., Francia, P., Bachschmid, M., Martin-Padura, I., Hersberger, M.,

Tanner, F.C., Pelicci, P., Volpe, M., Anversa, P., et al. (2007). Genetic deletion of p66(Shc)

adaptor protein prevents hyperglycemia-induced endothelial dysfunction and oxidative stress.

ProcNatlAcadSciUSA 104, 5217-5222.

141

Castilho, R.M., Squarize, C.H., Chodosh, L.A., Williams, B.O., and Gutkind, J.S. (2009). mTOR

mediates Wnt-induced epidermal stem cell exhaustion and aging. Cell stem cell 5, 279-289.

Chandel, N.S., McClintock, D.S., Feliciano, C.E., Wood, T.M., Melendez, J.A., Rodriguez,

A.M., and Schumacker, P.T. (2000). Reactive oxygen species generated at mitochondrial

complex III stabilize hypoxia-inducible factor-1alpha during hypoxia: a mechanism of O2

sensing. The Journal of biological chemistry 275, 25130-25138.

Chen, J., Zheng, X.F., Brown, E.J., and Schreiber, S.L. (1995). Identification of an 11-kDa

FKBP12-rapamycin-binding domain within the 289-kDa FKBP12-rapamycin-associated protein

and characterization of a critical serine residue. Proceedings of the National Academy of

Sciences of the United States of America 92, 4947-4951.

Cheung, P., Pawling, J., Partridge, E.A., Sukhu, B., Grynpas, M., and Dennis, J.W. (2007).

Metabolic homeostasis and tissue renewal are dependent on beta1,6GlcNAc-branched N-

glycans. Glycobiology 17, 828-837.

Choi, C., Ganji, S.K., DeBerardinis, R.J., Hatanpaa, K.J., Rakheja, D., Kovacs, Z., Yang, X.L.,

Mashimo, T., Raisanen, J.M., Marin-Valencia, I., et al. (2012). 2-hydroxyglutarate detection by

magnetic resonance spectroscopy in IDH-mutated patients with gliomas. Nat Med 18, 624-629.

Chowdhury, R., Yeoh, K.K., Tian, Y.M., Hillringhaus, L., Bagg, E.A., Rose, N.R., Leung, I.K.,

Li, X.S., Woon, E.C., Yang, M., et al. (2011). The oncometabolite 2-hydroxyglutarate inhibits

histone lysine demethylases. EMBO Rep 12, 463-469.

Christofk, H.R., Vander Heiden, M.G., Harris, M.H., Ramanathan, A., Gerszten, R.E., Wei, R.,

Fleming, M.D., Schreiber, S.L., and Cantley, L.C. (2008a). The M2 splice isoform of pyruvate

kinase is important for cancer metabolism and tumour growth. Nature 452, 230-233.

Christofk, H.R., Vander Heiden, M.G., Wu, N., Asara, J.M., and Cantley, L.C. (2008b). Pyruvate

kinase M2 is a phosphotyrosine-binding protein. Nature 452, 181-186.

Cooney, G.J., Lyons, R.J., Crew, A.J., Jensen, T.E., Molero, J.C., Mitchell, C.J., Biden, T.J.,

Ormandy, C.J., James, D.E., and Daly, R.J. (2004). Improved glucose homeostasis and enhanced

insulin signalling in Grb14-deficient mice. The EMBO journal 23, 582-593.

Corradetti, M.N., Inoki, K., Bardeesy, N., DePinho, R.A., and Guan, K.L. (2004). Regulation of

the TSC pathway by LKB1: evidence of a molecular link between tuberous sclerosis complex

and Peutz-Jeghers syndrome. Genes & development 18, 1533-1538.

D'Alessandris, C., Andreozzi, F., Federici, M., Cardellini, M., Brunetti, A., Ranalli, M., Del

Guerra, S., Lauro, D., Del Prato, S., Marchetti, P., et al. (2004). Increased O-glycosylation of

insulin signaling proteins results in their impaired activation and enhanced susceptibility to

142

apoptosis in pancreatic beta-cells. FASEB journal : official publication of the Federation of

American Societies for Experimental Biology 18, 959-961.

Dai, N., Christiansen, J., Nielsen, F.C., and Avruch, J. (2013). mTOR complex 2 phosphorylates

IMP1 cotranslationally to promote IGF2 production and the proliferation of mouse embryonic

fibroblasts. Genes & development 27, 301-312.

Dang, L., White, D.W., Gross, S., Bennett, B.D., Bittinger, M.A., Driggers, E.M., Fantin, V.R.,

Jang, H.G., Jin, S., Keenan, M.C., et al. (2009). Cancer-associated IDH1 mutations produce 2-

hydroxyglutarate. Nature 462, 739-744.

David, C.J., Chen, M., Assanah, M., Canoll, P., and Manley, J.L. (2010). HnRNP proteins

controlled by c-Myc deregulate pyruvate kinase mRNA splicing in cancer. Nature 463, 364-368.

Davol, P.A., Bagdasaryan, R., Elfenbein, G.J., Maizel, A.L., and Frackelton, A.R., Jr. (2003).

Shc proteins are strong, independent prognostic markers for both node-negative and node-

positive primary breast cancer. Cancer research 63, 6772-6783.

DeBerardinis, R.J., and Cheng, T. (2010). Q's next: the diverse functions of glutamine in

metabolism, cell biology and cancer. Oncogene 29, 313-324.

DeBerardinis, R.J., Mancuso, A., Daikhin, E., Nissim, I., Yudkoff, M., Wehrli, S., and

Thompson, C.B. (2007). Beyond aerobic glycolysis: transformed cells can engage in glutamine

metabolism that exceeds the requirement for protein and nucleotide synthesis. Proceedings of the

National Academy of Sciences of the United States of America 104, 19345-19350.

Demetriades, C., Doumpas, N., and Teleman, A.A. (2014). Regulation of TORC1 in Response to

Amino Acid Starvation via Lysosomal Recruitment of TSC2. Cell 156, 786-799.

Dengjel, J., Akimov, V., Olsen, J.V., Bunkenborg, J., Mann, M., Blagoev, B., and Andersen, J.S.

(2007). Quantitative proteomic assessment of very early cellular signaling events. Nature

biotechnology 25, 566-568.

Dennis, J.W., Nabi, I.R., and Demetriou, M. (2009). Metabolism, cell surface organization, and

disease. Cell 139, 1229-1241.

Deprez, J., Vertommen, D., Alessi, D.R., Hue, L., and Rider, M.H. (1997). Phosphorylation and

activation of heart 6-phosphofructo-2-kinase by protein kinase B and other protein kinases of the

insulin signaling cascades. The Journal of biological chemistry 272, 17269-17275.

DeYoung, M.P., Horak, P., Sofer, A., Sgroi, D., and Ellisen, L.W. (2008). Hypoxia regulates

TSC1/2-mTOR signaling and tumor suppression through REDD1-mediated 14-3-3 shuttling.

Genes & development 22, 239-251.

143

Di Paola, R., Ciociola, E., Boonyasrisawat, W., Nolan, D., Duffy, J., Miscio, G., Cisternino, C.,

Fini, G., Tassi, V., Doria, A., et al. (2006). Association of hGrb10 genetic variations with type 2

diabetes in Caucasian subjects. Diabetes care 29, 1181-1183.

Dibble, C.C., Elis, W., Menon, S., Qin, W., Klekota, J., Asara, J.M., Finan, P.M., Kwiatkowski,

D.J., Murphy, L.O., and Manning, B.D. (2012). TBC1D7 is a third subunit of the TSC1-TSC2

complex upstream of mTORC1. Molecular cell 47, 535-546.

Dibble, C.C., and Manning, B.D. (2013). Signal integration by mTORC1 coordinates nutrient

input with biosynthetic output. Nature cell biology 15, 555-564.

Duran, A., Amanchy, R., Linares, J.F., Joshi, J., Abu-Baker, S., Porollo, A., Hansen, M., Moscat,

J., and Diaz-Meco, M.T. (2011). p62 is a key regulator of nutrient sensing in the mTORC1

pathway. Molecular cell 44, 134-146.

Duran, R.V., Oppliger, W., Robitaille, A.M., Heiserich, L., Skendaj, R., Gottlieb, E., and Hall,

M.N. (2012). Glutaminolysis activates Rag-mTORC1 signaling. Molecular cell 47, 349-358.

Duvel, K., Yecies, J.L., Menon, S., Raman, P., Lipovsky, A.I., Souza, A.L., Triantafellow, E.,

Ma, Q., Gorski, R., Cleaver, S., et al. (2010). Activation of a metabolic gene regulatory network

downstream of mTOR complex 1. MolCell 39, 171-183.

Efeyan, A., Zoncu, R., Chang, S., Gumper, I., Snitkin, H., Wolfson, R.L., Kirak, O., Sabatini,

D.D., and Sabatini, D.M. (2013). Regulation of mTORC1 by the Rag GTPases is necessary for

neonatal autophagy and survival. Nature 493, 679-683.

Elstrom, R.L., Bauer, D.E., Buzzai, M., Karnauskas, R., Harris, M.H., Plas, D.R., Zhuang, H.,

Cinalli, R.M., Alavi, A., Rudin, C.M., et al. (2004). Akt stimulates aerobic glycolysis in cancer

cells. Cancer research 64, 3892-3899.

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.

Esen, E., Chen, J., Karner, C.M., Okunade, A.L., Patterson, B.W., and Long, F. (2013). WNT-

LRP5 signaling induces Warburg effect through mTORC2 activation during osteoblast

differentiation. Cell Metab 17, 745-755.

Fan, J., Kamphorst, J.J., Mathew, R., Chung, M.K., White, E., Shlomi, T., and Rabinowitz, J.D.

(2013). Glutamine-driven oxidative phosphorylation is a major ATP source in transformed

mammalian cells in both normoxia and hypoxia. Molecular systems biology 9, 712.

144

Fang, Y., Park, I.H., Wu, A.L., Du, G., Huang, P., Frohman, M.A., Walker, S.J., Brown, H.A.,

and Chen, J. (2003). PLD1 regulates mTOR signaling and mediates Cdc42 activation of S6K1.

Current biology : CB 13, 2037-2044.

Fang, Y., Vilella-Bach, M., Bachmann, R., Flanigan, A., and Chen, J. (2001). Phosphatidic acid-

mediated mitogenic activation of mTOR signaling. Science 294, 1942-1945.

Feng, Z., Hu, W., de Stanchina, E., Teresky, A.K., Jin, S., Lowe, S., and Levine, A.J. (2007).

The regulation of AMPK beta1, TSC2, and PTEN expression by p53: stress, cell and tissue

specificity, and the role of these gene products in modulating the IGF-1-AKT-mTOR pathways.

Cancer research 67, 3043-3053.

Ferrari, S., Bandi, H.R., Hofsteenge, J., Bussian, B.M., and Thomas, G. (1991). Mitogen-

activated 70K S6 kinase. Identification of in vitro 40 S ribosomal S6 phosphorylation sites. The

Journal of biological chemistry 266, 22770-22775.

Figueroa, M.E., Abdel-Wahab, O., Lu, C., Ward, P.S., Patel, J., Shih, A., Li, Y., Bhagwat, N.,

Vasanthakumar, A., Fernandez, H.F., et al. (2010). Leukemic IDH1 and IDH2 mutations result

in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic

differentiation. Cancer Cell 18, 553-567.

Findlay, G.M., Yan, L., Procter, J., Mieulet, V., and Lamb, R.F. (2007). A MAP4 kinase related

to Ste20 is a nutrient-sensitive regulator of mTOR signalling. The Biochemical journal 403, 13-

20.

Fisher-Wellman, K.H., and Neufer, P.D. (2012). Linking mitochondrial bioenergetics to insulin

resistance via redox biology. Trends EndocrinolMetab 23, 142-153.

Frias, M.A., Thoreen, C.C., Jaffe, J.D., Schroder, W., Sculley, T., Carr, S.A., and Sabatini, D.M.

(2006). mSin1 is necessary for Akt/PKB phosphorylation, and its isoforms define three distinct

mTORC2s. Current biology : CB 16, 1865-1870.

Ganley, I.G., Lam du, H., Wang, J., Ding, X., Chen, S., and Jiang, X. (2009).

ULK1.ATG13.FIP200 complex mediates mTOR signaling and is essential for autophagy. The

Journal of biological chemistry 284, 12297-12305.

Gao, P., Tchernyshyov, I., Chang, T.C., Lee, Y.S., Kita, K., Ochi, T., Zeller, K.I., De Marzo,

A.M., Van Eyk, J.E., Mendell, J.T., et al. (2009). c-Myc suppression of miR-23a/b enhances

mitochondrial glutaminase expression and glutamine metabolism. Nature 458, 762-765.

Garami, A., Zwartkruis, F.J., Nobukuni, T., Joaquin, M., Roccio, M., Stocker, H., Kozma, S.C.,

Hafen, E., Bos, J.L., and Thomas, G. (2003). Insulin activation of Rheb, a mediator of

mTOR/S6K/4E-BP signaling, is inhibited by TSC1 and 2. Molecular cell 11, 1457-1466.

145

Garcia-Martinez, J.M., and Alessi, D.R. (2008). mTOR complex 2 (mTORC2) controls

hydrophobic motif phosphorylation and activation of serum- and glucocorticoid-induced protein

kinase 1 (SGK1). The Biochemical journal 416, 375-385.

Gatenby, R.A., and Gillies, R.J. (2004). Why do cancers have high aerobic glycolysis? Nature

reviews Cancer 4, 891-899.

Gingras, A.C., Gygi, S.P., Raught, B., Polakiewicz, R.D., Abraham, R.T., Hoekstra, M.F.,

Aebersold, R., and Sonenberg, N. (1999). Regulation of 4E-BP1 phosphorylation: a novel two-

step mechanism. Genes Dev 13, 1422-1437.

Giorgio, M., Berry, A., Berniakovich, I., Poletaeva, I., Trinei, M., Stendardo, M., Hagopian, K.,

Ramsey, J.J., Cortopassi, G., Migliaccio, E., et al. (2012). The p66Shc knocked out mice are

short lived under natural condition. Aging Cell 11, 162-168.

Giorgio, M., Migliaccio, E., Orsini, F., Paolucci, D., Moroni, M., Contursi, C., Pelliccia, G.,

Luzi, L., Minucci, S., Marcaccio, M., et al. (2005). Electron transfer between cytochrome c and

p66Shc generates reactive oxygen species that trigger mitochondrial apoptosis. Cell 122, 221-

233.

Gordan, J.D., Thompson, C.B., and Simon, M.C. (2007). HIF and c-Myc: sibling rivals for

control of cancer cell metabolism and proliferation. Cancer Cell 12, 108-113.

Gottlieb, E., and Tomlinson, I.P. (2005). Mitochondrial tumour suppressors: a genetic and

biochemical update. Nature reviews Cancer 5, 857-866.

Gottlob, K., Majewski, N., Kennedy, S., Kandel, E., Robey, R.B., and Hay, N. (2001). Inhibition

of early apoptotic events by Akt/PKB is dependent on the first committed step of glycolysis and

mitochondrial hexokinase. Genes & development 15, 1406-1418.

Graiani, G., Lagrasta, C., Migliaccio, E., Spillmann, F., Meloni, M., Madeddu, P., Quaini, F.,

Padura, I.M., Lanfrancone, L., Pelicci, P., et al. (2005). Genetic deletion of the p66Shc adaptor

protein protects from angiotensin II-induced myocardial damage. Hypertension 46, 433-440.

Greer, S.N., Metcalf, J.L., Wang, Y., and Ohh, M. (2012). The updated biology of hypoxia-

inducible factor. The EMBO journal 31, 2448-2460.

Gu, H., Maeda, H., Moon, J.J., Lord, J.D., Yoakim, M., Nelson, B.H., and Neel, B.G. (2000).

New role for Shc in activation of the phosphatidylinositol 3-kinase/Akt pathway. MolCell Biol

20, 7109-7120.

Guertin, D.A., Stevens, D.M., Thoreen, C.C., Burds, A.A., Kalaany, N.Y., Moffat, J., Brown,

M., Fitzgerald, K.J., and Sabatini, D.M. (2006). Ablation in mice of the mTORC components

146

raptor, rictor, or mLST8 reveals that mTORC2 is required for signaling to Akt-FOXO and

PKCalpha, but not S6K1. Developmental cell 11, 859-871.

Gulati, P., Gaspers, L.D., Dann, S.G., Joaquin, M., Nobukuni, T., Natt, F., Kozma, S.C.,

Thomas, A.P., and Thomas, G. (2008). Amino acids activate mTOR complex 1 via Ca2+/CaM

signaling to hVps34. Cell Metab 7, 456-465.

Gwinn, D.M., Shackelford, D.B., Egan, D.F., Mihaylova, M.M., Mery, A., Vasquez, D.S., Turk,

B.E., and Shaw, R.J. (2008). AMPK phosphorylation of raptor mediates a metabolic checkpoint.

Molecular cell 30, 214-226.

Haga, Y., Ishii, K., and Suzuki, T. (2011). N-glycosylation is critical for the stability and

intracellular trafficking of glucose transporter GLUT4. The Journal of biological chemistry 286,

31320-31327.

Haghighat, A., Mader, S., Pause, A., and Sonenberg, N. (1995). Repression of cap-dependent

translation by 4E-binding protein 1: competition with p220 for binding to eukaryotic initiation

factor-4E. The EMBO journal 14, 5701-5709.

Hagiwara, A., Cornu, M., Cybulski, N., Polak, P., Betz, C., Trapani, F., Terracciano, L., Heim,

M.H., Ruegg, M.A., and Hall, M.N. (2012). Hepatic mTORC2 activates glycolysis and

lipogenesis through Akt, glucokinase, and SREBP1c. Cell Metab 15, 725-738.

Han, J.M., Jeong, S.J., Park, M.C., Kim, G., Kwon, N.H., Kim, H.K., Ha, S.H., Ryu, S.H., and

Kim, S. (2012). Leucyl-tRNA synthetase is an intracellular leucine sensor for the mTORC1-

signaling pathway. Cell 149, 410-424.

Hanahan, D., and Weinberg, R.A. (2011). Hallmarks of cancer: the next generation. Cell 144,

646-674.

Hara, K., Yonezawa, K., Kozlowski, M.T., Sugimoto, T., Andrabi, K., Weng, Q.P., Kasuga, M.,

Nishimoto, I., and Avruch, J. (1997). Regulation of eIF-4E BP1 phosphorylation by mTOR. The

Journal of biological chemistry 272, 26457-26463.

Hara, K., Yonezawa, K., Weng, Q.P., Kozlowski, M.T., Belham, C., and Avruch, J. (1998).

Amino acid sufficiency and mTOR regulate p70 S6 kinase and eIF-4E BP1 through a common

effector mechanism. The Journal of biological chemistry 273, 14484-14494.

Hardy, W.R., Li, L., Wang, Z., Sedy, J., Fawcett, J., Frank, E., Kucera, J., and Pawson, T.

(2007). Combinatorial ShcA docking interactions support diversity in tissue morphogenesis.

Science 317, 251-256.

147

Harrington, L.S., Findlay, G.M., Gray, A., Tolkacheva, T., Wigfield, S., Rebholz, H., Barnett, J.,

Leslie, N.R., Cheng, S., Shepherd, P.R., et al. (2004). The TSC1-2 tumor suppressor controls

insulin-PI3K signaling via regulation of IRS proteins. The Journal of cell biology 166, 213-223.

Hatzivassiliou, G., Zhao, F., Bauer, D.E., Andreadis, C., Shaw, A.N., Dhanak, D., Hingorani,

S.R., Tuveson, D.A., and Thompson, C.B. (2005). ATP citrate lyase inhibition can suppress

tumor cell growth. Cancer Cell 8, 311-321.

Heller, C., Kuhn, M.C., Mulders-Opgenoorth, B., Schott, M., Willenberg, H.S., Scherbaum,

W.A., and Schinner, S. (2011). Exendin-4 upregulates the expression of Wnt-4, a novel regulator

of pancreatic beta-cell proliferation. American journal of physiology Endocrinology and

metabolism 301, E864-872.

Hemmings, B.A., and Restuccia, D.F. (2012). PI3K-PKB/Akt pathway. Cold Spring Harbor

perspectives in biology 4, a011189.

Hitosugi, T., Kang, S., Vander Heiden, M.G., Chung, T.W., Elf, S., Lythgoe, K., Dong, S.,

Lonial, S., Wang, X., Chen, G.Z., et al. (2009). Tyrosine phosphorylation inhibits PKM2 to

promote the Warburg effect and tumor growth. Science signaling 2, ra73.

Holt, L.J., and Siddle, K. (2005). Grb10 and Grb14: enigmatic regulators of insulin action--and

more? The Biochemical journal 388, 393-406.

Hosokawa, N., Hara, T., Kaizuka, T., Kishi, C., Takamura, A., Miura, Y., Iemura, S., Natsume,

T., Takehana, K., Yamada, N., et al. (2009). Nutrient-dependent mTORC1 association with the

ULK1-Atg13-FIP200 complex required for autophagy. Molecular biology of the cell 20, 1981-

1991.

Houstis, N., Rosen, E.D., and Lander, E.S. (2006). Reactive oxygen species have a causal role in

multiple forms of insulin resistance. Nature 440, 944-948.

Hsu, P.P., Kang, S.A., Rameseder, J., Zhang, Y., Ottina, K.A., Lim, D., Peterson, T.R., Choi, Y.,

Gray, N.S., Yaffe, M.B., et al. (2011). The mTOR-regulated phosphoproteome reveals a

mechanism of mTORC1-mediated inhibition of growth factor signaling. Science 332, 1317-

1322.

Huang, J., Dibble, C.C., Matsuzaki, M., and Manning, B.D. (2008). The TSC1-TSC2 complex is

required for proper activation of mTOR complex 2. Molecular and cellular biology 28, 4104-

4115.

Hudson, C.C., Liu, M., Chiang, G.G., Otterness, D.M., Loomis, D.C., Kaper, F., Giaccia, A.J.,

and Abraham, R.T. (2002). Regulation of hypoxia-inducible factor 1alpha expression and

function by the mammalian target of rapamycin. Molecular and cellular biology 22, 7004-7014.

148

Ikenoue, T., Inoki, K., Yang, Q., Zhou, X., and Guan, K.L. (2008). Essential function of TORC2

in PKC and Akt turn motif phosphorylation, maturation and signalling. The EMBO journal 27,

1919-1931.

Inoki, K., Li, Y., Zhu, T., Wu, J., and Guan, K.L. (2002). TSC2 is phosphorylated and inhibited

by Akt and suppresses mTOR signalling. Nature cell biology 4, 648-657.

Inoki, K., Ouyang, H., Zhu, T., Lindvall, C., Wang, Y., Zhang, X., Yang, Q., Bennett, C.,

Harada, Y., Stankunas, K., et al. (2006). TSC2 integrates Wnt and energy signals via a

coordinated phosphorylation by AMPK and GSK3 to regulate cell growth. Cell 126, 955-968.

Inoki, K., Zhu, T., and Guan, K.L. (2003). TSC2 mediates cellular energy response to control

cell growth and survival. Cell 115, 577-590.

Isaacs, J.S., Jung, Y.J., Mole, D.R., Lee, S., Torres-Cabala, C., Chung, Y.L., Merino, M., Trepel,

J., Zbar, B., Toro, J., et al. (2005). HIF overexpression correlates with biallelic loss of fumarate

hydratase in renal cancer: novel role of fumarate in regulation of HIF stability. Cancer Cell 8,

143-153.

Jacinto, E., Facchinetti, V., Liu, D., Soto, N., Wei, S., Jung, S.Y., Huang, Q., Qin, J., and Su, B.

(2006). SIN1/MIP1 maintains rictor-mTOR complex integrity and regulates Akt phosphorylation

and substrate specificity. Cell 127, 125-137.

Jacinto, E., Loewith, R., Schmidt, A., Lin, S., Ruegg, M.A., Hall, A., and Hall, M.N. (2004).

Mammalian TOR complex 2 controls the actin cytoskeleton and is rapamycin insensitive. Nature

cell biology 6, 1122-1128.

Jiang, P., Du, W., Wang, X., Mancuso, A., Gao, X., Wu, M., and Yang, X. (2011). p53 regulates

biosynthesis through direct inactivation of glucose-6-phosphate dehydrogenase. Nature cell

biology 13, 310-316.

Jin, J., and Pawson, T. (2012). Modular evolution of phosphorylation-based signalling systems.

Philosophical transactions of the Royal Society of London Series B, Biological sciences 367,

2540-2555.

Jones, R.G., Plas, D.R., Kubek, S., Buzzai, M., Mu, J., Xu, Y., Birnbaum, M.J., and Thompson,

C.B. (2005). AMP-activated protein kinase induces a p53-dependent metabolic checkpoint.

Molecular cell 18, 283-293.

Jung, C.H., Jun, C.B., Ro, S.H., Kim, Y.M., Otto, N.M., Cao, J., Kundu, M., and Kim, D.H.

(2009). ULK-Atg13-FIP200 complexes mediate mTOR signaling to the autophagy machinery.

Molecular biology of the cell 20, 1992-2003.

149

Kasus-Jacobi, A., Perdereau, D., Auzan, C., Clauser, E., Van Obberghen, E., Mauvais-Jarvis, F.,

Girard, J., and Burnol, A.F. (1998). Identification of the rat adapter Grb14 as an inhibitor of

insulin actions. The Journal of biological chemistry 273, 26026-26035.

Khanday, F.A., Santhanam, L., Kasuno, K., Yamamori, T., Naqvi, A., Dericco, J., Bugayenko,

A., Mattagajasingh, I., Disanza, A., Scita, G., et al. (2006). Sos-mediated activation of rac1 by

p66shc. The Journal of cell biology 172, 817-822.

Khatri, S., Yepiskoposyan, H., Gallo, C.A., Tandon, P., and Plas, D.R. (2010). FOXO3a

regulates glycolysis via transcriptional control of tumor suppressor TSC1. The Journal of

biological chemistry 285, 15960-15965.

Kim, C.S., Jung, S.B., Naqvi, A., Hoffman, T.A., DeRicco, J., Yamamori, T., Cole, M.P., Jeon,

B.H., and Irani, K. (2008a). p53 impairs endothelium-dependent vasomotor function through

transcriptional upregulation of p66shc. Circulation research 103, 1441-1450.

Kim, D.H., Sarbassov, D.D., Ali, S.M., King, J.E., Latek, R.R., Erdjument-Bromage, H., Tempst,

P., and Sabatini, D.M. (2002). mTOR interacts with raptor to form a nutrient-sensitive complex

that signals to the cell growth machinery. Cell 110, 163-175.

Kim, D.H., Sarbassov, D.D., Ali, S.M., Latek, R.R., Guntur, K.V., Erdjument-Bromage, H.,

Tempst, P., and Sabatini, D.M. (2003). GbetaL, a positive regulator of the rapamycin-sensitive

pathway required for the nutrient-sensitive interaction between raptor and mTOR. Molecular cell

11, 895-904.

Kim, E., Goraksha-Hicks, P., Li, L., Neufeld, T.P., and Guan, K.L. (2008b). Regulation of

TORC1 by Rag GTPases in nutrient response. Nature cell biology 10, 935-945.

Kim, S., Kim, S.F., Maag, D., Maxwell, M.J., Resnick, A.C., Juluri, K.R., Chakraborty, A.,

Koldobskiy, M.A., Cha, S.H., Barrow, R., et al. (2011). Amino acid signaling to mTOR

mediated by inositol polyphosphate multikinase. Cell Metab 13, 215-221.

Kisielow, M., Kleiner, S., Nagasawa, M., Faisal, A., and Nagamine, Y. (2002). Isoform-specific

knockdown and expression of adaptor protein ShcA using small interfering RNA. BiochemJ 363,

1-5.

Kitagawa, T., Tsuruhara, Y., Hayashi, M., Endo, T., and Stanbridge, E.J. (1995). A tumor-

associated glycosylation change in the glucose transporter GLUT1 controlled by tumor

suppressor function in human cell hybrids. Journal of cell science 108 ( Pt 12), 3735-3743.

Kondoh, H., Lleonart, M.E., Gil, J., Wang, J., Degan, P., Peters, G., Martinez, D., Carnero, A.,

and Beach, D. (2005). Glycolytic enzymes can modulate cellular life span. Cancer research 65,

177-185.

150

Kooner, J.S., Saleheen, D., Sim, X., Sehmi, J., Zhang, W., Frossard, P., Been, L.F., Chia, K.S.,

Dimas, A.S., Hassanali, N., et al. (2011). Genome-wide association study in individuals of South

Asian ancestry identifies six new type 2 diabetes susceptibility loci. Nature genetics 43, 984-989.

Kovacina, K.S., Park, G.Y., Bae, S.S., Guzzetta, A.W., Schaefer, E., Birnbaum, M.J., and Roth,

R.A. (2003). Identification of a proline-rich Akt substrate as a 14-3-3 binding partner. The

Journal of biological chemistry 278, 10189-10194.

Kunz, J., Henriquez, R., Schneider, U., Deuter-Reinhard, M., Movva, N.R., and Hall, M.N.

(1993). Target of rapamycin in yeast, TOR2, is an essential phosphatidylinositol kinase homolog

required for G1 progression. Cell 73, 585-596.

Lai, K.M., Olivier, J.P., Gish, G.D., Henkemeyer, M., McGlade, J., and Pawson, T. (1995). A

Drosophila shc gene product is implicated in signaling by the DER receptor tyrosine kinase.

Molecular and cellular biology 15, 4810-4818.

Lai, K.M., and Pawson, T. (2000). The ShcA phosphotyrosine docking protein sensitizes

cardiovascular signaling in the mouse embryo. Genes & development 14, 1132-1145.

Lamming, D.W., Demirkan, G., Boylan, J.M., Mihaylova, M.M., Peng, T., Ferreira, J., Neretti,

N., Salomon, A., Sabatini, D.M., and Gruppuso, P.A. (2014). Hepatic signaling by the

mechanistic target of rapamycin complex 2 (mTORC2). FASEB journal : official publication of

the Federation of American Societies for Experimental Biology 28, 300-315.

Laplante, M., and Sabatini, D.M. (2012). mTOR signaling in growth control and disease. Cell

149, 274-293.

Lau, K.S., Partridge, E.A., Grigorian, A., Silvescu, C.I., Reinhold, V.N., Demetriou, M., and

Dennis, J.W. (2007). Complex N-glycan number and degree of branching cooperate to regulate

cell proliferation and differentiation. Cell 129, 123-134.

Lebiedzinska, M., Duszynski, J., Rizzuto, R., Pinton, P., and Wieckowski, M.R. (2009). Age-

related changes in levels of p66Shc and serine 36-phosphorylated p66Shc in organs and mouse

tissues. ArchBiochemBiophys.

Li, S.C., Lai, K.M., Gish, G.D., Parris, W.E., van der Geer, P., Forman-Kay, J., and Pawson, T.

(1996). Characterization of the phosphotyrosine-binding domain of the Drosophila Shc protein.

The Journal of biological chemistry 271, 31855-31862.

Li, T., Kon, N., Jiang, L., Tan, M., Ludwig, T., Zhao, Y., Baer, R., and Gu, W. (2012). Tumor

suppression in the absence of p53-mediated cell-cycle arrest, apoptosis, and senescence. Cell

149, 1269-1283.

151

Lim, W.A., and Pawson, T. (2010). Phosphotyrosine signaling: evolving a new cellular

communication system. Cell 142, 661-667.

Long, X., Lin, Y., Ortiz-Vega, S., Yonezawa, K., and Avruch, J. (2005). Rheb binds and

regulates the mTOR kinase. Current biology : CB 15, 702-713.

Lu, C., Ward, P.S., Kapoor, G.S., Rohle, D., Turcan, S., Abdel-Wahab, O., Edwards, C.R.,

Khanin, R., Figueroa, M.E., Melnick, A., et al. (2012). IDH mutation impairs histone

demethylation and results in a block to cell differentiation. Nature 483, 474-478.

Luo, W., Hu, H., Chang, R., Zhong, J., Knabel, M., O'Meally, R., Cole, R.N., Pandey, A., and

Semenza, G.L. (2011). Pyruvate kinase M2 is a PHD3-stimulated coactivator for hypoxia-

inducible factor 1. Cell 145, 732-744.

Luzi, L., Confalonieri, S., Di Fiore, P.P., and Pelicci, P.G. (2000). Evolution of Shc functions

from nematode to human. CurrOpinGenetDev 10, 668-674.

Ma, L., Chen, Z., Erdjument-Bromage, H., Tempst, P., and Pandolfi, P.P. (2005).

Phosphorylation and functional inactivation of TSC2 by Erk implications for tuberous sclerosis

and cancer pathogenesis. Cell 121, 179-193.

Ma, X.M., and Blenis, J. (2009). Molecular mechanisms of mTOR-mediated translational

control. Nature reviews Molecular cell biology 10, 307-318.

Ma, X.M., Yoon, S.O., Richardson, C.J., Julich, K., and Blenis, J. (2008). SKAR links pre-

mRNA splicing to mTOR/S6K1-mediated enhanced translation efficiency of spliced mRNAs.

Cell 133, 303-313.

Manning, A.K., Hivert, M.F., Scott, R.A., Grimsby, J.L., Bouatia-Naji, N., Chen, H., Rybin, D.,

Liu, C.T., Bielak, L.F., Prokopenko, I., et al. (2012). A genome-wide approach accounting for

body mass index identifies genetic variants influencing fasting glycemic traits and insulin

resistance. Nature genetics 44, 659-669.

Manning, B.D., Tee, A.R., Logsdon, M.N., Blenis, J., and Cantley, L.C. (2002). Identification of

the tuberous sclerosis complex-2 tumor suppressor gene product tuberin as a target of the

phosphoinositide 3-kinase/akt pathway. Molecular cell 10, 151-162.

Mardis, E.R., Ding, L., Dooling, D.J., Larson, D.E., McLellan, M.D., Chen, K., Koboldt, D.C.,

Fulton, R.S., Delehaunty, K.D., McGrath, S.D., et al. (2009). Recurring mutations found by

sequencing an acute myeloid leukemia genome. N Engl J Med 361, 1058-1066.

152

Masui, K., Tanaka, K., Akhavan, D., Babic, I., Gini, B., Matsutani, T., Iwanami, A., Liu, F.,

Villa, G.R., Gu, Y., et al. (2013). mTOR complex 2 controls glycolytic metabolism in

glioblastoma through FoxO acetylation and upregulation of c-Myc. Cell Metab 18, 726-739.

Matoba, S., Kang, J.G., Patino, W.D., Wragg, A., Boehm, M., Gavrilova, O., Hurley, P.J., Bunz,

F., and Hwang, P.M. (2006). p53 regulates mitochondrial respiration. Science 312, 1650-1653.

Maxwell, P.H., Wiesener, M.S., Chang, G.W., Clifford, S.C., Vaux, E.C., Cockman, M.E.,

Wykoff, C.C., Pugh, C.W., Maher, E.R., and Ratcliffe, P.J. (1999). The tumour suppressor

protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature 399,

271-275.

Mayer, C., Zhao, J., Yuan, X., and Grummt, I. (2004). mTOR-dependent activation of the

transcription factor TIF-IA links rRNA synthesis to nutrient availability. Genes & development

18, 423-434.

Mazurek, S., Boschek, C.B., Hugo, F., and Eigenbrodt, E. (2005). Pyruvate kinase type M2 and

its role in tumor growth and spreading. Seminars in cancer biology 15, 300-308.

Menon, S., Dibble, C.C., Talbott, G., Hoxhaj, G., Valvezan, A.J., Takahashi, H., Cantley, L.C.,

and Manning, B.D. (2014). Spatial Control of the TSC Complex Integrates Insulin and Nutrient

Regulation of mTORC1 at the Lysosome. Cell 156, 771-785.

Migliaccio, E., Giorgio, M., Mele, S., Pelicci, G., Reboldi, P., Pandolfi, P.P., Lanfrancone, L.,

and Pelicci, P.G. (1999). The p66shc adaptor protein controls oxidative stress response and life

span in mammals. Nature 402, 309-313.

Migliaccio, E., Mele, S., Salcini, A.E., Pelicci, G., Lai, K.M., Superti-Furga, G., Pawson, T., Di

Fiore, P.P., Lanfrancone, L., and Pelicci, P.G. (1997). Opposite effects of the p52shc/p46shc and

p66shc splicing isoforms on the EGF receptor-MAP kinase-fos signalling pathway. EMBO J 16,

706-716.

Minami, A., Iseki, M., Kishi, K., Wang, M., Ogura, M., Furukawa, N., Hayashi, S., Yamada, M.,

Obata, T., Takeshita, Y., et al. (2003). Increased insulin sensitivity and hypoinsulinemia in APS

knockout mice. Diabetes 52, 2657-2665.

Morita, S., Kojima, T., and Kitamura, T. (2000). Plat-E: an efficient and stable system for

transient packaging of retroviruses. Gene therapy 7, 1063-1066.

Morris, A.P., Voight, B.F., Teslovich, T.M., Ferreira, T., Segre, A.V., Steinthorsdottir, V.,

Strawbridge, R.J., Khan, H., Grallert, H., Mahajan, A., et al. (2012). Large-scale association

analysis provides insights into the genetic architecture and pathophysiology of type 2 diabetes.

Nature genetics 44, 981-990.

153

Munger, J., Bennett, B.D., Parikh, A., Feng, X.J., McArdle, J., Rabitz, H.A., Shenk, T., and

Rabinowitz, J.D. (2008). Systems-level metabolic flux profiling identifies fatty acid synthesis as

a target for antiviral therapy. NatBiotechnol 26, 1179-1186.

Nakae, J., Kitamura, T., Silver, D.L., and Accili, D. (2001). The forkhead transcription factor

Foxo1 (Fkhr) confers insulin sensitivity onto glucose-6-phosphatase expression. The Journal of

clinical investigation 108, 1359-1367.

Napoli, C., Martin-Padura, I., de Nigris, F., Giorgio, M., Mansueto, G., Somma, P., Condorelli,

M., Sica, G., De Rosa, G., and Pelicci, P. (2003). Deletion of the p66Shc longevity gene reduces

systemic and tissue oxidative stress, vascular cell apoptosis, and early atherogenesis in mice fed

a high-fat diet. Proceedings of the National Academy of Sciences of the United States of

America 100, 2112-2116.

Natalicchio, A., De Stefano, F., Perrini, S., Laviola, L., Cignarelli, A., Caccioppoli, C.,

Quagliara, A., Melchiorre, M., Leonardini, A., Conserva, A., et al. (2009). Involvement of the

p66Shc protein in glucose transport regulation in skeletal muscle myoblasts. American journal of

physiology Endocrinology and metabolism 296, E228-237.

Natalicchio, A., Laviola, L., De Tullio, C., Renna, L.A., Montrone, C., Perrini, S., Valenti, G.,

Procino, G., Svelto, M., and Giorgino, F. (2004). Role of the p66Shc isoform in insulin-like

growth factor I receptor signaling through MEK/Erk and regulation of actin cytoskeleton in rat

myoblasts. The Journal of biological chemistry 279, 43900-43909.

Nemoto, S., and Finkel, T. (2002). Redox regulation of forkhead proteins through a p66shc-

dependent signaling pathway. Science 295, 2450-2452.

Nicklin, P., Bergman, P., Zhang, B., Triantafellow, E., Wang, H., Nyfeler, B., Yang, H., Hild,

M., Kung, C., Wilson, C., et al. (2009). Bidirectional transport of amino acids regulates mTOR

and autophagy. Cell 136, 521-534.

Nobukuni, T., Joaquin, M., Roccio, M., Dann, S.G., Kim, S.Y., Gulati, P., Byfield, M.P., Backer,

J.M., Natt, F., Bos, J.L., et al. (2005). Amino acids mediate mTOR/raptor signaling through

activation of class 3 phosphatidylinositol 3OH-kinase. Proceedings of the National Academy of

Sciences of the United States of America 102, 14238-14243.

O'Connell, B.C., Cheung, A.F., Simkevich, C.P., Tam, W., Ren, X., Mateyak, M.K., and Sedivy,

J.M. (2003). A large scale genetic analysis of c-Myc-regulated gene expression patterns. The

Journal of biological chemistry 278, 12563-12573.

Oh, W.J., Wu, C.C., Kim, S.J., Facchinetti, V., Julien, L.A., Finlan, M., Roux, P.P., Su, B., and

Jacinto, E. (2010). mTORC2 can associate with ribosomes to promote cotranslational

phosphorylation and stability of nascent Akt polypeptide. The EMBO journal 29, 3939-3951.

154

Ohtsubo, K., Takamatsu, S., Minowa, M.T., Yoshida, A., Takeuchi, M., and Marth, J.D. (2005).

Dietary and genetic control of glucose transporter 2 glycosylation promotes insulin secretion in

suppressing diabetes. Cell 123, 1307-1321.

Okada, S., Kao, A.W., Ceresa, B.P., Blaikie, P., Margolis, B., and Pessin, J.E. (1997). The 66-

kDa Shc isoform is a negative regulator of the epidermal growth factor-stimulated mitogen-

activated protein kinase pathway. The Journal of biological chemistry 272, 28042-28049.

Orsini, F., Migliaccio, E., Moroni, M., Contursi, C., Raker, V.A., Piccini, D., Martin-Padura, I.,

Pelliccia, G., Trinei, M., Bono, M., et al. (2004). The life span determinant p66Shc localizes to

mitochondria where it associates with mitochondrial heat shock protein 70 and regulates trans-

membrane potential. The Journal of biological chemistry 279, 25689-25695.

Osthus, R.C., Shim, H., Kim, S., Li, Q., Reddy, R., Mukherjee, M., Xu, Y., Wonsey, D., Lee,

L.A., and Dang, C.V. (2000). Deregulation of glucose transporter 1 and glycolytic gene

expression by c-Myc. The Journal of biological chemistry 275, 21797-21800.

Panchaud, N., Peli-Gulli, M.P., and De Virgilio, C. (2013). Amino acid deprivation inhibits

TORC1 through a GTPase-activating protein complex for the Rag family GTPase Gtr1. Science

signaling 6, ra42.

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.

Partridge, E.A., Le Roy, C., Di Guglielmo, G.M., Pawling, J., Cheung, P., Granovsky, M., Nabi,

I.R., Wrana, J.L., and Dennis, J.W. (2004). Regulation of cytokine receptors by Golgi N-glycan

processing and endocytosis. Science 306, 120-124.

Pawson, T. (2007). Dynamic control of signaling by modular adaptor proteins. Current opinion

in cell biology 19, 112-116.

Pawson, T., and Scott, J.D. (1997). Signaling through scaffold, anchoring, and adaptor proteins.

Science 278, 2075-2080.

Pearce, L.R., Huang, X., Boudeau, J., Pawlowski, R., Wullschleger, S., Deak, M., Ibrahim, A.F.,

Gourlay, R., Magnuson, M.A., and Alessi, D.R. (2007). Identification of Protor as a novel

Rictor-binding component of mTOR complex-2. The Biochemical journal 405, 513-522.

Pelicci, G., Lanfrancone, L., Grignani, F., McGlade, J., Cavallo, F., Forni, G., Nicoletti, I.,

Pawson, T., and Pelicci, P.G. (1992). A novel transforming protein (SHC) with an SH2 domain

is implicated in mitogenic signal transduction. Cell 70, 93-104.

155

Pena-Llopis, S., Vega-Rubin-de-Celis, S., Schwartz, J.C., Wolff, N.C., Tran, T.A., Zou, L., Xie,

X.J., Corey, D.R., and Brugarolas, J. (2011). Regulation of TFEB and V-ATPases by mTORC1.

The EMBO journal 30, 3242-3258.

Peterson, T.R., Laplante, M., Thoreen, C.C., Sancak, Y., Kang, S.A., Kuehl, W.M., Gray, N.S.,

and Sabatini, D.M. (2009). DEPTOR is an mTOR inhibitor frequently overexpressed in multiple

myeloma cells and required for their survival. Cell 137, 873-886.

Pinton, P., Rimessi, A., Marchi, S., Orsini, F., Migliaccio, E., Giorgio, M., Contursi, C., Minucci,

S., Mantovani, F., Wieckowski, M.R., et al. (2007). Protein kinase C beta and prolyl isomerase 1

regulate mitochondrial effects of the life-span determinant p66Shc. Science 315, 659-663.

Pinton, P., and Rizzuto, R. (2008). p66Shc, oxidative stress and aging: importing a lifespan

determinant into mitochondria. Cell Cycle 7, 304-308.

Porstmann, T., Santos, C.R., Griffiths, B., Cully, M., Wu, M., Leevers, S., Griffiths, J.R., Chung,

Y.L., and Schulze, A. (2008). SREBP activity is regulated by mTORC1 and contributes to Akt-

dependent cell growth. Cell Metab 8, 224-236.

Potter, C.J., Pedraza, L.G., and Xu, T. (2002). Akt regulates growth by directly phosphorylating

Tsc2. Nature cell biology 4, 658-665.

Rabinowitz, J.D., and White, E. (2010). Autophagy and metabolism. Science 330, 1344-1348.

Rampersaud, E., Damcott, C.M., Fu, M., Shen, H., McArdle, P., Shi, X., Shelton, J., Yin, J.,

Chang, Y.P., Ott, S.H., et al. (2007). Identification of novel candidate genes for type 2 diabetes

from a genome-wide association scan in the Old Order Amish: evidence for replication from

diabetes-related quantitative traits and from independent populations. Diabetes 56, 3053-3062.

Ranieri, S.C., Fusco, S., Panieri, E., Labate, V., Mele, M., Tesori, V., Ferrara, A.M., Maulucci,

G., De, S.M., Martorana, G.E., et al. (2010). Mammalian life-span determinant p66shcA

mediates obesity-induced insulin resistance. ProcNatlAcadSciUSA 107, 13420-13425.

Ravichandran, K.S. (2001). Signaling via Shc family adapter proteins. Oncogene 20, 6322-6330.

Reiling, J.H., and Hafen, E. (2004). The hypoxia-induced paralogs Scylla and Charybdis inhibit

growth by down-regulating S6K activity upstream of TSC in Drosophila. Genes & development

18, 2879-2892.

Robey, R.B., and Hay, N. (2009). Is Akt the "Warburg kinase"?-Akt-energy metabolism

interactions and oncogenesis. Seminars in cancer biology 19, 25-31.

156

Robitaille, A.M., Christen, S., Shimobayashi, M., Cornu, M., Fava, L.L., Moes, S., Prescianotto-

Baschong, C., Sauer, U., Jenoe, P., and Hall, M.N. (2013). Quantitative phosphoproteomics

reveal mTORC1 activates de novo pyrimidine synthesis. Science 339, 1320-1323.

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.

Roux, P.P., Ballif, B.A., Anjum, R., Gygi, S.P., and Blenis, J. (2004). Tumor-promoting phorbol

esters and activated Ras inactivate the tuberous sclerosis tumor suppressor complex via p90

ribosomal S6 kinase. Proceedings of the National Academy of Sciences of the United States of

America 101, 13489-13494.

Rozakis-Adcock, M., McGlade, J., Mbamalu, G., Pelicci, G., Daly, R., Li, W., Batzer, A.,

Thomas, S., Brugge, J., Pelicci, P.G., et al. (1992). Association of the Shc and Grb2/Sem5 SH2-

containing proteins is implicated in activation of the Ras pathway by tyrosine kinases. Nature

360, 689-692.

Ruvinsky, I., Sharon, N., Lerer, T., Cohen, H., Stolovich-Rain, M., Nir, T., Dor, Y., Zisman, P.,

and Meyuhas, O. (2005). Ribosomal protein S6 phosphorylation is a determinant of cell size and

glucose homeostasis. Genes & development 19, 2199-2211.

Sabatini, D.M., Erdjument-Bromage, H., Lui, M., Tempst, P., and Snyder, S.H. (1994). RAFT1:

a mammalian protein that binds to FKBP12 in a rapamycin-dependent fashion and is

homologous to yeast TORs. Cell 78, 35-43.

Sabers, C.J., Martin, M.M., Brunn, G.J., Williams, J.M., Dumont, F.J., Wiederrecht, G., and

Abraham, R.T. (1995). Isolation of a protein target of the FKBP12-rapamycin complex in

mammalian cells. The Journal of biological chemistry 270, 815-822.

Sancak, Y., Bar-Peled, L., Zoncu, R., Markhard, A.L., Nada, S., and Sabatini, D.M. (2010).

Ragulator-Rag complex targets mTORC1 to the lysosomal surface and is necessary for its

activation by amino acids. Cell 141, 290-303.

Sancak, Y., Peterson, T.R., Shaul, Y.D., Lindquist, R.A., Thoreen, C.C., Bar-Peled, L., and

Sabatini, D.M. (2008). The Rag GTPases bind raptor and mediate amino acid signaling to

mTORC1. Science 320, 1496-1501.

Sancak, Y., Thoreen, C.C., Peterson, T.R., Lindquist, R.A., Kang, S.A., Spooner, E., Carr, S.A.,

and Sabatini, D.M. (2007). PRAS40 is an insulin-regulated inhibitor of the mTORC1 protein

kinase. Molecular cell 25, 903-915.

Sarbassov, D.D., Ali, S.M., Kim, D.H., Guertin, D.A., Latek, R.R., Erdjument-Bromage, H.,

Tempst, P., and Sabatini, D.M. (2004). Rictor, a novel binding partner of mTOR, defines a

157

rapamycin-insensitive and raptor-independent pathway that regulates the cytoskeleton. Current

biology : CB 14, 1296-1302.

Sarbassov, D.D., Ali, S.M., Sengupta, S., Sheen, J.H., Hsu, P.P., Bagley, A.F., Markhard, A.L.,

and Sabatini, D.M. (2006). Prolonged rapamycin treatment inhibits mTORC2 assembly and

Akt/PKB. MolCell 22, 159-168.

Sarbassov, D.D., Guertin, D.A., Ali, S.M., and Sabatini, D.M. (2005). Phosphorylation and

regulation of Akt/PKB by the rictor-mTOR complex. Science 307, 1098-1101.

Saucedo, L.J., Gao, X., Chiarelli, D.A., Li, L., Pan, D., and Edgar, B.A. (2003). Rheb promotes

cell growth as a component of the insulin/TOR signalling network. Nature cell biology 5, 566-

571.

Schwartzenberg-Bar-Yoseph, F., Armoni, M., and Karnieli, E. (2004). The tumor suppressor p53

down-regulates glucose transporters GLUT1 and GLUT4 gene expression. Cancer research 64,

2627-2633.

Scott, J.D., and Pawson, T. (2009). Cell signaling in space and time: where proteins come

together and when they're apart. Science 326, 1220-1224.

Scott, R.A., Lagou, V., Welch, R.P., Wheeler, E., Montasser, M.E., Luan, J., Magi, R.,

Strawbridge, R.J., Rehnberg, E., Gustafsson, S., et al. (2012). Large-scale association analyses

identify new loci influencing glycemic traits and provide insight into the underlying biological

pathways. Nature genetics 44, 991-1005.

Selak, M.A., Armour, S.M., MacKenzie, E.D., Boulahbel, H., Watson, D.G., Mansfield, K.D.,

Pan, Y., Simon, M.C., Thompson, C.B., and Gottlieb, E. (2005). Succinate links TCA cycle

dysfunction to oncogenesis by inhibiting HIF-alpha prolyl hydroxylase. Cancer Cell 7, 77-85.

Semenza, G.L. (2004). Hydroxylation of HIF-1: oxygen sensing at the molecular level.

Physiology (Bethesda) 19, 176-182.

Settembre, C., Zoncu, R., Medina, D.L., Vetrini, F., Erdin, S., Huynh, T., Ferron, M., Karsenty,

G., Vellard, M.C., Facchinetti, V., et al. (2012). A lysosome-to-nucleus signalling mechanism

senses and regulates the lysosome via mTOR and TFEB. The EMBO journal 31, 1095-1108.

Shah, O.J., Wang, Z., and Hunter, T. (2004). Inappropriate activation of the

TSC/Rheb/mTOR/S6K cassette induces IRS1/2 depletion, insulin resistance, and cell survival

deficiencies. Current biology : CB 14, 1650-1656.

Shimobayashi, M., and Hall, M.N. (2014). Making new contacts: the mTOR network in

metabolism and signalling crosstalk. Nature reviews Molecular cell biology 15, 155-162.

158

Smith, E.M., Finn, S.G., Tee, A.R., Browne, G.J., and Proud, C.G. (2005a). The tuberous

sclerosis protein TSC2 is not required for the regulation of the mammalian target of rapamycin

by amino acids and certain cellular stresses. The Journal of biological chemistry 280, 18717-

18727.

Smith, F.M., Holt, L.J., Garfield, A.S., Charalambous, M., Koumanov, F., Perry, M., Bazzani,

R., Sheardown, S.A., Hegarty, B.D., Lyons, R.J., et al. (2007). Mice with a disruption of the

imprinted Grb10 gene exhibit altered body composition, glucose homeostasis, and insulin

signaling during postnatal life. MolCell Biol 27, 5871-5886.

Smith, W.W., Norton, D.D., Gorospe, M., Jiang, H., Nemoto, S., Holbrook, N.J., Finkel, T., and

Kusiak, J.W. (2005b). Phosphorylation of p66Shc and forkhead proteins mediates Abeta toxicity.

The Journal of cell biology 169, 331-339.

Stevenson, L.E., and Frackelton, A.R., Jr. (1998). Constitutively tyrosine phosphorylated p52

Shc in breast cancer cells: correlation with ErbB2 and p66 Shc expression. Breast Cancer

ResTreat 49, 119-128.

Stocker, H., Radimerski, T., Schindelholz, B., Wittwer, F., Belawat, P., Daram, P., Breuer, S.,

Thomas, G., and Hafen, E. (2003). Rheb is an essential regulator of S6K in controlling cell

growth in Drosophila. Nature cell biology 5, 559-565.

Tee, A.R., Fingar, D.C., Manning, B.D., Kwiatkowski, D.J., Cantley, L.C., and Blenis, J. (2002).

Tuberous sclerosis complex-1 and -2 gene products function together to inhibit mammalian

target of rapamycin (mTOR)-mediated downstream signaling. Proceedings of the National

Academy of Sciences of the United States of America 99, 13571-13576.

Tee, A.R., Manning, B.D., Roux, P.P., Cantley, L.C., and Blenis, J. (2003). Tuberous sclerosis

complex gene products, Tuberin and Hamartin, control mTOR signaling by acting as a GTPase-

activating protein complex toward Rheb. Current biology : CB 13, 1259-1268.

Thoreen, C.C., Kang, S.A., Chang, J.W., Liu, Q., Zhang, J., Gao, Y., Reichling, L.J., Sim, T.,

Sabatini, D.M., and Gray, N.S. (2009). An ATP-competitive mammalian target of rapamycin

inhibitor reveals rapamycin-resistant functions of mTORC1. JBiolChem 284, 8023-8032.

Tomilov, A.A., Ramsey, J.J., Hagopian, K., Giorgio, M., Kim, K.M., Lam, A., Migliaccio, E.,

Lloyd, K.C., Berniakovich, I., Prolla, T.A., et al. (2011). The Shc locus regulates insulin

signaling and adiposity in mammals. Aging Cell 10, 55-65.

Trinei, M., Giorgio, M., Cicalese, A., Barozzi, S., Ventura, A., Migliaccio, E., Milia, E., Padura,

I.M., Raker, V.A., Maccarana, M., et al. (2002). A p53-p66Shc signalling pathway controls

intracellular redox status, levels of oxidation-damaged DNA and oxidative stress-induced

apoptosis. Oncogene 21, 3872-3878.

159

Tsun, Z.Y., Bar-Peled, L., Chantranupong, L., Zoncu, R., Wang, T., Kim, C., Spooner, E., and

Sabatini, D.M. (2013). The folliculin tumor suppressor is a GAP for the RagC/D GTPases that

signal amino acid levels to mTORC1. Molecular cell 52, 495-505.

Turcan, S., Rohle, D., Goenka, A., Walsh, L.A., Fang, F., Yilmaz, E., Campos, C., Fabius, A.W.,

Lu, C., Ward, P.S., et al. (2012). IDH1 mutation is sufficient to establish the glioma

hypermethylator phenotype. Nature 483, 479-483.

Urano, J., Tabancay, A.P., Yang, W., and Tamanoi, F. (2000). The Saccharomyces cerevisiae

Rheb G-protein is involved in regulating canavanine resistance and arginine uptake. The Journal

of biological chemistry 275, 11198-11206.

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.

Ursini-Siegel, J., and Muller, W.J. (2008). The ShcA adaptor protein is a critical regulator of

breast cancer progression. Cell Cycle 7, 1936-1943.

Vander Heiden, M.G., Cantley, L.C., and Thompson, C.B. (2009). Understanding the Warburg

effect: the metabolic requirements of cell proliferation. Science 324, 1029-1033.

Vanhaesebroeck, B., Stephens, L., and Hawkins, P. (2012). PI3K signalling: the path to

discovery and understanding. Nature reviews Molecular cell biology 13, 195-203.

Vousden, K.H., and Ryan, K.M. (2009). p53 and metabolism. Nature reviews Cancer 9, 691-700.

Wang, L., Balas, B., Christ-Roberts, C.Y., Kim, R.Y., Ramos, F.J., Kikani, C.K., Li, C., Deng,

C., Reyna, S., Musi, N., et al. (2007). Peripheral disruption of the Grb10 gene enhances insulin

signaling and sensitivity in vivo. MolCell Biol 27, 6497-6505.

Wang, X., Campbell, L.E., Miller, C.M., and Proud, C.G. (1998). Amino acid availability

regulates p70 S6 kinase and multiple translation factors. The Biochemical journal 334 ( Pt 1),

261-267.

Wang, X., Li, W., Williams, M., Terada, N., Alessi, D.R., and Proud, C.G. (2001). Regulation of

elongation factor 2 kinase by p90(RSK1) and p70 S6 kinase. The EMBO journal 20, 4370-4379.

Warburg, O. (1956). On the origin of cancer cells. Science 123, 309-314.

Ward, P.S., Patel, J., Wise, D.R., Abdel-Wahab, O., Bennett, B.D., Coller, H.A., Cross, J.R.,

Fantin, V.R., Hedvat, C.V., Perl, A.E., et al. (2010). The common feature of leukemia-associated

160

IDH1 and IDH2 mutations is a neomorphic enzyme activity converting alpha-ketoglutarate to 2-

hydroxyglutarate. Cancer Cell 17, 225-234.

Ward, P.S., and Thompson, C.B. (2012). Metabolic reprogramming: a cancer hallmark even

warburg did not anticipate. Cancer Cell 21, 297-308.

Watanabe, T., Nobusawa, S., Kleihues, P., and Ohgaki, H. (2009). IDH1 mutations are early

events in the development of astrocytomas and oligodendrogliomas. Am J Pathol 174, 1149-

1153.

Weinberg, F., Hamanaka, R., Wheaton, W.W., Weinberg, S., Joseph, J., Lopez, M.,

Kalyanaraman, B., Mutlu, G.M., Budinger, G.R., and Chandel, N.S. (2010). Mitochondrial

metabolism and ROS generation are essential for Kras-mediated tumorigenicity. Proceedings of

the National Academy of Sciences of the United States of America 107, 8788-8793.

Wellen, K.E., Lu, C., Mancuso, A., Lemons, J.M., Ryczko, M., Dennis, J.W., Rabinowitz, J.D.,

Coller, H.A., and Thompson, C.B. (2010). The hexosamine biosynthetic pathway couples growth

factor-induced glutamine uptake to glucose metabolism. Genes Dev 24, 2784-2799.

Wills, M.K., and Jones, N. (2012). Teaching an old dogma new tricks: twenty years of Shc

adaptor signalling. The Biochemical journal 447, 1-16.

Wilson, K.F., Wu, W.J., and Cerione, R.A. (2000). Cdc42 stimulates RNA splicing via the S6

kinase and a novel S6 kinase target, the nuclear cap-binding complex. The Journal of biological

chemistry 275, 37307-37310.

Wise, D.R., DeBerardinis, R.J., Mancuso, A., Sayed, N., Zhang, X.Y., Pfeiffer, H.K., Nissim, I.,

Daikhin, E., Yudkoff, M., McMahon, S.B., et al. (2008). Myc regulates a transcriptional program

that stimulates mitochondrial glutaminolysis and leads to glutamine addiction. Proceedings of the

National Academy of Sciences of the United States of America 105, 18782-18787.

Wullschleger, S., Loewith, R., Oppliger, W., and Hall, M.N. (2005). Molecular organization of

target of rapamycin complex 2. The Journal of biological chemistry 280, 30697-30704.

Xi, G., Shen, X., and Clemmons, D.R. (2008). p66shc negatively regulates insulin-like growth

factor I signal transduction via inhibition of p52shc binding to Src homology 2 domain-

containing protein tyrosine phosphatase substrate-1 leading to impaired growth factor receptor-

bound protein-2 membrane recruitment. MolEndocrinol 22, 2162-2175.

Xi, G., Shen, X., and Clemmons, D.R. (2010a). p66shc inhibits insulin-like growth factor-I

signaling via direct binding to Src through its polyproline and Src homology 2 domains, resulting

in impairment of Src kinase activation. JBiolChem 285, 6937-6951.

161

Xi, G., Shen, X., Radhakrishnan, Y., Maile, L., and Clemmons, D. (2010b). Hyperglycemia-

Induced p66shc Inhibits Insulin-Like Growth Factor I-Dependent Cell Survival via Impairment

of Src Kinase-Mediated Phosphoinositide-3 Kinase/AKT Activation in Vascular Smooth Muscle

Cells. Endocrinology.

Xia, J., and Wishart, D.S. (2011). Web-based inference of biological patterns, functions and

pathways from metabolomic data using MetaboAnalyst. NatProtoc 6, 743-760.

Xu, W., Yang, H., Liu, Y., Yang, Y., Wang, P., Kim, S.H., Ito, S., Yang, C., Xiao, M.T., Liu,

L.X., et al. (2011). Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of alpha-

ketoglutarate-dependent dioxygenases. Cancer Cell 19, 17-30.

Yan, H., Parsons, D.W., Jin, G., McLendon, R., Rasheed, B.A., Yuan, W., Kos, I., Batinic-

Haberle, I., Jones, S., Riggins, G.J., et al. (2009). IDH1 and IDH2 mutations in gliomas. N Engl

J Med 360, 765-773.

Yan, L., Mieulet, V., Burgess, D., Findlay, G.M., Sully, K., Procter, J., Goris, J., Janssens, V.,

Morrice, N.A., and Lamb, R.F. (2010). PP2A T61 epsilon is an inhibitor of MAP4K3 in nutrient

signaling to mTOR. Molecular cell 37, 633-642.

Yang, C.P., and Horwitz, S.B. (2000). Taxol mediates serine phosphorylation of the 66-kDa Shc

isoform. Cancer research 60, 5171-5178.

Yang, Q., Inoki, K., Ikenoue, T., and Guan, K.L. (2006). Identification of Sin1 as an essential

TORC2 component required for complex formation and kinase activity. Genes & development

20, 2820-2832.

Yang, W., Xia, Y., Ji, H., Zheng, Y., Liang, J., Huang, W., Gao, X., Aldape, K., and Lu, Z.

(2011). Nuclear PKM2 regulates beta-catenin transactivation upon EGFR activation. Nature 480,

118-122.

Yip, C.K., Murata, K., Walz, T., Sabatini, D.M., and Kang, S.A. (2010). Structure of the human

mTOR complex I and its implications for rapamycin inhibition. Molecular cell 38, 768-774.

Yoon, J.C., Ng, A., Kim, B.H., Bianco, A., Xavier, R.J., and Elledge, S.J. (2010). Wnt signaling

regulates mitochondrial physiology and insulin sensitivity. Genes & development 24, 1507-1518.

Yu, Y., Yoon, S.O., Poulogiannis, G., Yang, Q., Ma, X.M., Villen, J., Kubica, N., Hoffman,

G.R., Cantley, L.C., Gygi, S.P., et al. (2011). Phosphoproteomic analysis identifies Grb10 as an

mTORC1 substrate that negatively regulates insulin signaling. Science 332, 1322-1326.

162

Yuneva, M., Zamboni, N., Oefner, P., Sachidanandam, R., and Lazebnik, Y. (2007). Deficiency

in glutamine but not glucose induces MYC-dependent apoptosis in human cells. The Journal of

cell biology 178, 93-105.

Zhang, Y., Gao, X., Saucedo, L.J., Ru, B., Edgar, B.A., and Pan, D. (2003). Rheb is a direct

target of the tuberous sclerosis tumour suppressor proteins. Nature cell biology 5, 578-581.

Zhao, S., Lin, Y., Xu, W., Jiang, W., Zha, Z., Wang, P., Yu, W., Li, Z., Gong, L., Peng, Y., et al.

(2009). Glioma-derived mutations in IDH1 dominantly inhibit IDH1 catalytic activity and induce

HIF-1alpha. Science 324, 261-265.

Zheng, Y., Zhang, C., Croucher, D.R., Soliman, M.A., St-Denis, N., Pasculescu, A., Taylor, L.,

Tate, S.A., Hardy, W.R., Colwill, K., et al. (2013). Temporal regulation of EGF signalling

networks by the scaffold protein Shc1. Nature 499, 166-171.

Zinzalla, V., Stracka, D., Oppliger, W., and Hall, M.N. (2011). Activation of mTORC2 by

association with the ribosome. Cell 144, 757-768.

Zoncu, R., Efeyan, A., and Sabatini, D.M. (2011). mTOR: from growth signal integration to

cancer, diabetes and ageing. NatRevMolCell Biol 12, 21-35.