CHARACTERIZATION OF NIMA-RELATED 10 (NEK10): A ROLE IN

CHECKPOINT CONTROL

by

Larissa S Moniz

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Medical Biophysics University of Toronto

© Copyright by Larissa S Moniz 2010 Abstract

Characterization of NimA-related kinase 10 (NEK10): a role in checkpoint control

Larissa S Moniz, Doctor of Philosophy 2010. Department of Medical Biophysics, University of Toronto.

Deregulation of the is a hallmark of neoplastic transformation and plays a central role in both the initiation and progression of cancer. Members of the NimA-related kinase (NEK) family of are emerging as important players in regulation of the eukaryotic cell cycle during normal cell cycle progression and checkpoint activation in response to genotoxic stresses. The focus of this thesis is NEK10, a previously uncharacterized member of the NEK family. While little is known about the biology of

NEK10, recent cancer genomics studies have identified NEK10 as a candidate susceptibility at 3p24 in cancer.

Work herein describes a role for NEK10 in the cellular response to ultraviolet (UV) irradiation. NEK10 was required for the activation of ERK1/2 signaling upon UV irradiation, but not in response to mitogens, such as the epidermal growth factor. NEK10 interacted with

Raf and MEK and enhanced MEK activity through a novel mechanism involving MEK autoactivation. Significantly, appropriate maintenance of the G2/M checkpoint following

UV irradiation required NEK10 expression and ERK1/2 activation. In support of a conserved role for NEK10 in the cellular response to UV irradiation, nekl-4, the NEK10 C.elegans homologue, affected embryonic sensitivity to UV-irradiation.

In search of regulatory inputs into NEK10, using mass spectrometry, our laboratory identified 19 distinct sites of NEK10 phosphorylation. Characterization of a number of these

ii sites revealed a role for intermolecular autophosphorylation in achieving full NEK10 catalytic activity through activation loop phosphorylation on S684 and S688. Further, a C- terminal phosphorylation site on NEK10, S933, was found to be a 14-3-3 , and was essential for NEK10 cytoplasmic to nuclear translocation following UV irradiation.

Taken together, my studies have discovered a role for NEK10 in the engagement of the G2/M cell cycle checkpoint and provided a mechanistic insight into the relationship between NEK10 and the Raf/MEK/ERK cascade, and the control of NEK10 subcellular localization. This work will serve as a foundation for future studies aimed at understanding the molecular mechanism of NEK10 action and its function in development and tumourigenesis.

iii Acknowledgements

There are so many people who have contributed to my work and to my life over the past seven and a half years. Without them this thesis would look very different, and more importantly I would not have had such an enjoyable journey.

First and foremost on this list is my supervisor, Vuk Stambolic. Through your example, you have promoted curiosity, excitement, camaraderie, scientific excellence and most importantly, independent thought. These are qualities that I hope to emulate and share.

Thank you for your openness, guidance and patience. I definitely made the right choice.

Following a close second I want to thank all the members of my lab, both past and present: Tam, John, Claudia, Ben, Luke, Jason, Previn, Diana, Chris, Ryan, Vanessa and

Nasir. Thank you for “helpful discussions”, thank you for teaching me, thank you for troubleshooting with me and for listening to me complain, and thank you for making me happy to come to work everyday. An extra thank you to Tam, who has shared each step of this long and often daunting process. You have pushed me to be smarter and calmer, and have helped me to not fear but to embrace challenges.

Thank you to my committee members: Dan Durocher, Jim Woodgett and Wen-Chen

Yeh who have always been supportive, informative and challenging. I have worked hard to earn your respect. This is perhaps especially true for Jim. I have been in awe of you ever since I realized that you have the kinase domain sequence memorized.

Thank you to the other students in MBP who made the long days seem normal and the evenings fun. Thanks especially to Nick and Kellie who are my colleagues and friends.

Thank you for sharing wine and laughter, all mixed in with scientific ideas and secret projects.

iv Thank you to my non-science friends who have always been excited about my work and through their lives have let me discover excitement in so many other ways. Special thanks to Agnes, Pamela, Caroline, Carol, Catriona, Kirsten and Lilian.

Last but not least I want to thank my family both immediate and extended, in particular my parents, Ethel and Emano, and my siblings Celine, Francis, and Russ. My parents have instilled in me the importance of education and hard work. They have made it inevitable that I would be successful in whatever path I chose and by their unwavering love and support have made it easy to achieve it.

v Table of Contents

Abstract ...... ii

Acknowledgements ...... iv

Table of Contents ...... vi

List of Figures ...... xi

List of Abbreviations ...... xiii

Chapter 1: Introduction ...... 1

1.1 The cell cycle ...... 2

1.1.1 Cyclin-dependent kinases (CDKs) ...... 2

1.1.2 Cyclin/CDK complexes control cell cycle progression ...... 5

1.1.2.1 Progression through G1 ...... 5

1.1.2.2 Progression through ...... 6

1.1.3 Other cell cycle regulated kinase families: Plk, Aurora and NEK kinases ...... 9

1.1.3.1 Polo-like kinases (Plks) ...... 9

1.1.3.2 Aurora kinases ...... 11

1.1.4 Checkpoint signaling ...... 12

1.1.4.1 ATM and ATR signaling in the checkpoint response...... 13

1.1.4.2 Chk1/2 signaling in the checkpoint response ...... 15

1.1.4.3 p38 signaling in the checkpoint response ...... 17

vi 1.2. NimA-related kinases ...... 18

1.2.1 Never in mitosis A (NimA) ...... 18

1.2.2 Organization of NEK ...... 20

1.2.3 Mammalian NEK kinases ...... 23

1.2.3.1 NEK kinases in cell cycle progression ...... 23

1.2.3.2 NEK kinases in checkpoint controls ...... 25

1.2.3.3 NEK kinases in ciliagenesis ...... 26

1.3 NEK10 ...... 29

1.3.1 NEK10 ...... 29

1.3.2 NEK10 and carcinogenesis ...... 30

1.4 ERK1/2 signaling ...... 33

1.4.1 Organization of MAPK signaling modules...... 33

1.4.2 ERK1/2 signaling ...... 34

1.4.2.1 MAPKKK: Raf kinase ...... 36

1.4.2.2 MAPKK: MEK1/2 ...... 39

1.4.2.3 MAPK: ERK1/2 ...... 40

1.4.3 Molecular mechanisms of ERK signaling regulation ...... 43

1.4.3.1 Regulation by subcellular localization ...... 44

1.4.3.2 Regulation by scaffolds ...... 45

1.4.4 Physiological roles of ERK1/2 signaling ...... 46

1.4.4.1 ERK1/2 signaling in oncogenesis ...... 46

1.4.4.2 ERK1/2 signaling in cell cycle regulation ...... 47

1.4.4.3 ERK1/2 signaling in development and differentiation ...... 50

vii 1.4.4.4 ERK1/2 in the response to genotoxic stress ...... 51

1.5 Ph.D. Outline ...... 52

1.5.1 Study Rationale ...... 52

1.5.2 Thesis Aims ...... 52

Chapter 2: NEK10 enhances MEK autophosphorylation following UV irradiation. 53

2.1 Abstract ...... 54

2.2 Introduction ...... 55

2.3 Materials and Methods ...... 56

2.4 Results ...... 62

2.4.1 Cloning and characterization of NEK10, a novel NimA kinase ...... 62

2.4.2 NEK10 is a stimulus-specific modulator of ERK1/2 ...... 62

2.4.3 NEK10, Raf-1 and MEK1 form a ternary complex ...... 67

2.4.4 UV irradiation promotes MEK1 autophosphorylation ...... 71

2.4.5 NEK10 promotes MEK1 autophosphorylation by a mechanism involving Raf-1. .76

2.4.6 NEK10 participates in the maintenance of the G2/M checkpoint following

UV irradiation...... 78

2.4.7 NEK10 has a conserved role in the UV response...... 80

2.5 Discussion ...... 84

2.5.1 NEK10 promotes MEK autoactivation ...... 84

2.5.2 NEK10’s role in checkpoint control ...... 86

2.5.3 Characterization of C.elegans nekl-4 during the UV response ...... 87

viii Chapter 3: Characterization of NEK10 regulation by phosphorylation ...... 89

3.1 Abstract ...... 90

3.2 Introduction ...... 91

3.3 Materials and Methods ...... 93

3.4 Results ...... 97

3.4.1 Biochemical characterization of NEK10 catalytic activity ...... 97

3.4.2 NEK10 is a phosphoprotein ...... 99

3.4.3 NEK10 requires activation loop phosphorylation for full catalytic activity . . . . . 101

3.4.4 NEK10 localization, but not catalytic activity, is modulated by UV irradiation. . 104

3.4.5 NEK10 nuclear localization requires S933 phosphorylation...... 105

3.4.6 S933 is a 14-3-3 binding site ...... 105

3.5 Discussion ...... 110

3.5.1 Characterization of NEK10 catalytic activity ...... 110

3.5.2 NEK10 undergoes activation loop autophosphorylation ...... 111

3.5.3 NEK10 exhibits tyrosine kinase activity ...... 112

3.5.4 NEK10 subcellular localization is regulated by UV irradiation ...... 113

3.5.5 14-3-3 proteins regulate NEK10 subcellular localization ...... 114

Chapter 4: Future Directions ...... 116

4.1 Materials and methods ...... 118

4.2 NEK10 and myogenesis ...... 119

4.3 NEK10 and cell adhesion ...... 122

4.4 NEK10 interacts with tubulin in vivo ...... 126

ix 4.5 NEK10 in tumourigenesis...... 129

4.6 Concluding Remarks ...... 130

5.0 References ...... 131

x List of Figures

Figure 1.1: The major features of the cell cycle in mammalian cells ...... 3

Figure 1.2: The principal events in mitosis ...... 7

Figure 1.3: The main features of the DNA damage checkpoint ...... 14

Figure 1.4: Phylogenetic tree of NimA-related kinases ...... 21

Figure 1.5: Key structural features of NimA-related kinases ...... 22

Figure 1.6: Low NEK10 expression is associated with poor prognosis in breast cancer . .32

Figure 1.7: Canonical ERK signaling ...... 35

Figure 2.1: Cloning and characterization of NEK10, a novel NimA-related kinase ...... 63

Figure 2.2: NEK10 is a stimulus-specific modulator of ERK1/2 ...... 64

Figure 2.3: EsiRNA knockdown of NEK10 ...... 66

Figure 2.4: NEK10 is required for UV-induced MEK1/2 and ERK1/2 activation ...... 68

Figure 2.5: NEK10 interacts with Raf-1 ...... 69

Figure 2.6: NEK10 interacts with the Raf-1-MEK1 complex ...... 70

Figure 2.7: NEK10 and Raf-1 exhibit a multi-site interaction ...... 72

Figure 2.8: NEK10 does not phosphorylate Raf-1 or MEK1 in vitro ...... 74

Figure 2.9: NEK10 promotes MEK1 autophosphorylation following UV irradiation . . . .75

Figure 2.10: UV-induced MEK1 autophosphorylation requires Raf-1 ...... 77

Figure 2.11: ERK1/2 activation is required for maintenance of the G2/M checkpoint following UV irradiation ...... 79

Figure 2.12: NEK10 activation is required for maintenance of the G2/M checkpoint

xi following UV irradiation ...... 81

Figure 2.13: NEK10 homologue, nekl-4, mediates the UV response in C.elegans embryos. . .

...... 83

Figure 3.1: Characterization of NEK10 catalytic activity ...... 98

Figure 3.2: Characterization of NEK10 phosphorylation...... 100

Figure 3.3: NEK10 requires activation loop phosphorylation for full catalytic activity. . . .103

Figure 3.4: NEK10 undergoes nuclear translocation following UV irradiation...... 106

Figure 3.5: S933 phosphorylation is required for NEK10 nuclear localization...... 107

Figure 3.6: NEK10 binds 14-3-3 ...... 109

Figure 4.1: NEK10 expression and activity is induced during muscle differentiation. . . . .120

Figure 4.2: Schematic representation of myogenesis ...... 121

Figure 4.3: NEK10 does not effect MEK1 S298 phosphorylation...... 124

Figure 4.4: NEK10 activity in adherent and suspended cells...... 125

xii List of Abbreviations

A Alanine (Ala) aa amino acids

ATM Ataxia-telangiectasia mutated

ATP Adenosine triphosphate

ATR ATM and Rad3-related

BRCA Breast cancer 1, early onset

BSA Bovine serum albumin

CAK CDK activating kinase

Cdc2 Cell division cycle 2

Cdc25 Cell division cycle 25

CDK Cyclin-dependent kinase

CHAPS 3-[(3-Cholamidopropyl)dimethylammonio]-1-propanesulfonate

Chk1/2 Checkpoint kinase 1/2

D Aspartate (Asp)

DAPI 4',6-diamidino-2-phenylindole

DMEM Dulbecco’s modified Eagle’s medium

DNA Deoxyribonucleic acid

DSB Double strand break

E Glutamate (Glu)

EGF Epidermal growth factor

ERK Extracellular regulated kinases esiRNA Endoribonuclease-prepared small interfering RNA

xiii FACS Fluorescence-activated cell sorting

FBS Fetal bovine serum

FGF Fibroblast growth factor

FITC Fluorescein isothiocyanate

G0 Quiesence

G1 Gap period 1

G2 Gap period 2

GDP Guanosine diphosphate

GST Glutathione S-

GTP Guanosine triphosphate

HA Hemagglutinin

HEK293 Human embryonic kidney 293

HU Hydroxyurea

IR Ionizing radiation jck Junior cystic kidney

JNK Jun NH2 terminal kinases kat Kidney, anemia, testes

KSR Kinase suppressor of Ras

M Mitosis

Mad1/2 Mitotic arrest deficient 1/2

BubR1 Budding uninhibited by benzimidazoles related 1

MAPK Microtubule associated/mitogen activated protein kinase

MEF Mouse embryonic fibroblast

xiv MMC Mitomycin C

MT Microtubule

MTOC Microtubule organizing centre

Myt1 Membrane associated tyrosine/threonine 1

NEK NimA-related kinase nekl NimA-related kinase (NEK) like

NES Nuclear export sequence

NGF Nerve growth factor

NimA Never in mitosis A

NLS Nuclear localization sequence

PAK p21 activated kinase

PBS Phosphate buffered saline

PCR Polymerase chain reaction

PFA Paraformaldehyde

PIKK Phosphatidylinositol 3-kinase-related kinases

PKB

PKD Polycystic kidney disease

Plk Polo-like kinase

PP-1 Protein phosphatase-1 pRb Retinoblastoma protein

PVDF Polyvinylidene fluoride

RNA Ribonucleic acid

RNAi RNA interference

xv ROS Reactive oxygen species

RTK Receptor tyrosine kinase

S phase DNA synthesis phase

S Serine (Ser)

SAC Spindle assembly checkpoint

SCF-β-TrCP Skp Cullin F-box complex-β-transducin repeat-containing protein

SDS-PAGE Sodium dodecyl sulfate polyacrylamide gel electrophoresis ssDNA Single strand DNA

T Threonine (Thr)

UV Ultraviolet

Y Tyrosine (Tyr)

xvi CHAPTER 1

GENERAL INTRODUCTION

1 1.1 The cell cycle

The cell cycle is a series of ordered events, during which time DNA is replicated and

segregated between two daughter cells (reviewed in (242)). The accurate completion of cell

division is essential for normal organism development and deregulation of this program can result in loss of genomic stability and cell viability. The cell cycle can be divided into four phases: G1, S, G2 and M (Fig 1.1). During G1, the first gap period, the cell integrates

information from extracellular signals, the most important of which is the presence of

adequate growth factors, to determine if a round of cell division should occur. If nutrients are

lacking, the cell may enter a temporary state of quiescence, termed G0, or may undergo differentiation. If conditions are favorable, the cell will prepare to replicate its DNA. This decision occurs at the restriction point (R) in late G1 and once made, irreversibly commits the cell to completing cell division. Following DNA replication in S phase, cells pass through

a second gap period, G2, ensuring that replication is completed and preparing for division in mitosis (M). Mitosis is the final stage of the cell cycle and involves the appropriate and equal segregation of genetic material into two new daughter cells.

1.1.1 Cyclin-dependent kinases (Cdks)

Cell cycle progression is controlled by numerous regulatory proteins, which are highly conserved from yeast to mammals. The master regulators of the cell cycle are complexes composed of a regulatory component, a cyclin, and a catalytic component, a serine-threonine cyclin-dependent kinase (Cdk). Yeast have a single Cdk, cell division cycle

28 (cdc28) in Saccharomyces cerevisiae and Cdc2 in Schizosaccharomyces pombe, whereas

2 3 there are multiple Cdks in mammalian cells, which are sequentially activated at different

stages of the cell cycle (Fig 1.1). In mammalian cells, Cdk4 or Cdk6 complexed with

cyclinD, and Cdk2 bound to cyclinE, promote early and late G1 progression, respectively. S phase is controlled largely by cyclinA/Cdk2 complexes while Cdk1 (also known as cdc2 or

p34cdc2) in association with cyclinB and to a lesser extent cyclinA is activated during G2/M

and M. The activity of Cdks is tightly controlled, most notably by protein-protein interactions

and phosphorylation (reviewed in (276)).

In order to be activated, Cdks have to associate with cyclins, which are expressed in a

cell cycle dependent manner (31, 94, 275). Interaction with a cyclin controls Cdk activity in

two ways, by inducing a permissive conformation of the Cdk catalytic site and by providing

substrate specificity (174, 356). In the case of Cdk6, cyclin-binding is necessary and

sufficient for activation (357). For other Cdks, such as Cdk1 and Cdk2, interaction with a

cyclin only partially activates the kinase. These kinases also require phosphorylation by the

Cdk activating kinase (CAK) within the activation segment (T161 in Cdk1, T160 in Cdk2)

(102, 174, 349, 372).

Cdk activity is also subject to negative regulation throughout the cell cycle. The mitotic kinase, Cdk1, is inhibited by phosphorylation at T14 and Y15 within the ATP- binding loop (201). These sites are phosphorylated by the and membrane associated

tyrosine/threonine 1 (Myt1) kinases, whereas dephosphorylation and Cdk1 activation at the

G2/M transition is achieved by the dual-specificity phosphatases, Cdc25 A, B and C (121,

135, 201, 228, 279, 304)(reviewed in (347)).

The G1/S Cdks are predominantly regulated by their interactions with 2 families of

Cdk-inhibitors, the inhibitor of Cdk4 (INK4) and the Cdk interacting protein/kinase

4 inhibitory protein (Cip/Kip) families (32, 213)(reviewed in (306)). There are four INK4

proteins, p16INK4a, p15INK4b, p18INK4c, and p19INK4d that form stable complexes with

the early G1 Cdks, Cdk4 and Cdk6. This interaction inhibits Cdk4/6 activation by preventing

their association with cyclinD (154). The Cip/Kip family, composed of p21 (cip1/waf1), p27

(kip1) and p57 (kip2), display lower specificity and can inhibit all cyclin/Cdk complexes to some degree (439)(reviewed in (28)). However, in proliferating cells, Cip/Kip proteins are primarily inhibitory towards cyclinE/Cdk2 complexes. Interestingly, Cip/Kip proteins also promote G1 progression by facilitating the assembly and nuclear translocation of cyclinD/Cdk4/6 complexes (207). Moreover, as cells progress through G1, newly formed

cyclinD/Cdk4/6 complexes bind and sequester Cip/Kip proteins allowing downstream activation of cyclinE/Cdk2 (123, 308). In situations of cell cycle arrest, such as when cells undergo senescence or are arrested through contact inhibition or genotoxic stress, upregulation and post-translational modification of Cip/Kip proteins increases their ability to inhibit both cyclinE/Cdk2 and cyclinD/Cdk4/6 complexes (171, 294, 315, 378).

1.1.2 Cyclin/Cdk complexes control cell cycle progression

1.1.2.1 Progression through G1

Cyclin/Cdk complexes control the decision to pass R and to commit to a round of cell division. One of the key mechanistic events regulating this decision is hyperphosphorylation and inactivation of the retinoblastoma protein (pRB), mediated by cyclin/Cdk complexes

(reviewed in (124, 143)). Prior to R, unphosphorylated or hypophosphorylated pRB inhibits

5 transcription of required for S phase entry by at least two mechanisms. pRB binds to

the transactivation domain of the E2 transcription factor (E2F) preventing transcriptional

activation of target genes (103). In addition, in complex with E2F, pRB can bind directly to

promoters and actively repress transcription by a mechanism involving chromatin remodeling

via histone deacetylase 1 (HDAC1) (458)(reviewed in (243)). Once hyperphosphorylated,

pRB loses it’s ability to interact with E2F, allowing transcription of genes required for DNA replication, such as cyclinE, cyclinA, DNA polymerase subunits, and proliferating cell

nuclear antigen (PCNA) (70, 81, 82, 295, 385).

During early G1, low Cdk activity maintains pRB in a hypophosphorylated and active

state. Following sustained growth factor stimulation, cyclinD expression is induced, leading

to formation of active cyclinD/Cdk4/6 complexes and phosphorylation of pRB on multiple

sites (185)(reviewed in (312)). In late G1, E2F-induced cyclinE, complexed with Cdk2,

further reinforces pRB phosphorylation, initiating transcription of genes required for S-phase

and irreversibly committing the cell to division (144)(reviewed in (143)).

1.1.2.2 Progression through mitosis

The function of mitosis is to faithfully partition the duplicated genome between two

daughter cells. Mitosis can be divided into 5 consecutive and morphologically distinct

phases: prophase, prometaphase, metaphase, anaphase, and telophase (Fig 1.2) (reviewed in

(285)). Early mitotic events include chromosome condensation and nuclear envelope breakdown (prophase). Concurrently, duplicated centrosomes migrate towards opposite ends of the cell and nucleate dynamic microtubules to form the mitotic spindle. Microtubules are

6 7 captured by kinetochores that have formed on mitotic (prometaphase), and once sister chromatids have undergone bipolar attachment they align in the centre of the cell at the metaphase plate (metaphase). At the metaphase-anaphase transition, cohesion between sister chromatids is lost and sister chromatids are pulled to opposite ends of the cell

(anaphase). Following this, the nuclear envelope begins to reform and chromosomes decondense (telophase). Finally, an actomyosin contractile ring is formed and generates two daughter cells during cytokinesis (reviewed in (285)).

To enter mitosis a cell requires cyclinB/Cdk1 activation. This occurs at the G2/M transition following dephosphorylation of T14 and Y15 (201). At this stage the balance is shifted towards dephosphorylation of T14/Y15 by inactivation of the Wee1 and Myt1 kinases and activation of the Cdc25 phosphatases through phosphorylation by a number of mitotic kinases including Cdk1 and polo-like kinase 1 () (17, 32, 247, 284, 403, 425)(reviewed

in (225)). Therefore, Cdk1 contributes to a positive feedback loop that enhances its own

activity. Activated Cdk1 initiates early mitotic events by phosphorylating a wide variety of

targets including, nuclear lamins, condensins, golgi-matrix components, kinesin-related

motors and microtubule binding proteins, leading to nuclear envelope breakdown,

chromosome condensation, golgi fragmentation, centrosome separation and mitotic spindle

assembly, respectively (11, 60, 72, 190, 239).

While Cdk1 activity stimulates mitotic entry, it has to be inhibited for a cell to exit

mitosis (281). Cdk1 contributes to it’s own inactivation and to mitotic exit by promoting

degradation of cyclinB via the E3 ubiquitin , anaphase promoting complex/cyclosome

(APC/C). At the metaphase-anaphase transition, mitotic kinases including Cdk1 and Plk1

phosphorylate and activate APC/C (129, 198, 199, 346). APC/C stimulates mitotic exit by

8 the timely degradation of mitotic regulators, including cyclinB, Plk1 and Aurora A, as well

as inhibitors of chromatid separation, such as securin (157, 365, 368, 463).

The key event in mitosis is the equal segregation of genetic material. To ensure that

this occurs, APC/C activity and chromatid separation is inhibited by the spindle assembly

checkpoint (SAC), until the point at which all chromatids have undergone bipolar

microtubule attachment (335, 336). The SAC is composed of a diffusible signal of kinetochore-interacting proteins, including Mad1/2, Bub1/2, BubRI and Aurora B. These proteins act as molecular sensors and detect unattached kinetochores and changes in microtubule-kinetochore tension that occur following bipolar attachment (reviewed in (282,

309)). The SAC is crucial for genomic stability and in animal cells elimination of this checkpoint results in aneuploidy and cell death (77).

1.1.3 Other cell cycle regulated kinase families: Plk, Aurora, and NEK kinases

While Cdks are considered the master regulators of cell cycle progression, in recent years, a number of other cell cycle regulated kinase families have been characterized. These families, including the Plks, Aurora kinases, and NimA-Related Kinases (NEK) encode

serine-threonine kinases that play important roles throughout the cell cycle and particularly during mitosis. Plk and Aurora kinases are discussed below. For information on NEK kinases see Chapter 1.2.

1.1.3.1 Polo-like kinases (Plks)

9 In vertebrates, there are 4 Plks (Plk1-4). The best characterized mammalian isoform is

Plk1, which is most closely related to the single Plk found in yeast and Drosophila melanogaster (Polo) and to Xenopus laevis Plx1. Plk1 is activated prior to mitotic entry via a molecular mechanism involving transcriptional regulation and activation loop phosphorylation (172)(reviewed in (12)). Phosphorylation is mediated by Aurora A during late G2 and likely by another, currently unidentified, kinase at later points of mitosis (244,

361).

Plk1 contributes to many aspects of mitosis, including regulation of mitotic entry, generation of the microtubule organizing center (MTOC) and cytokinesis (reviewed in (12)).

Plk1 promotes mitotic entry by directly phosphorylating Wee1, Myt1 and Cdc25C, leading to their inactivation and activation respectively. This contributes to the positive feedback loop that stimulates Cdk1 activity at G2/M (166, 205, 425). Consistently, depletion of Plx1, in

Xenopus oocytes, and Plk1, in human cells, causes a delay in early mitotic progression (322,

381). Plks also regulate formation of the mitotic spindle and one of the first phenotypes

associated with Polo kinases was the presence of monopolar spindles, seen in both D.

melanogaster polo mutants, and in cultured human cells with disrupted Plk1 (211, 384). This

can be attributed to the function of Plks in centriole duplication, where inhibition or

overexpression, of Plk4 in human cells and polo in Drosophila, results in decreased and

increased centriole numbers, respectively. Further, at mitotic entry, Plk1 promotes

microtubule recruitment to centrosomes during centrosome maturation (29, 139, 212). In

addition, Plks regulate cytokinesis, and in human cells chemical inhibition of Plk1 disrupts

cell division. This is mediated by Plk1-dependent activation of RhoA GTPase, which

participates in contractile ring formation (310). In addition to these functions, Plk activity is

10 implicated in maintaining sister chromatid cohesion, recovery from DNA damage induced

checkpoints and in asymmetric division during embryonic development (40, 244,

382)(reviewed in (12)).

1.1.3.2 Aurora kinases

Another family with a prominent function in mitosis is the Aurora family of kinases, in mammals consisting of 3 members, Aurora A, B and C (reviewed in (410)). As with Plks,

Aurora kinases are expressed in a cell cycle dependent manner. In addition, kinase activation requires phosphorylation and allosteric interaction with specific co-factors. The best- characterized co-factors are targeting protein for Xklp2 (TPX2) for Aurora A, and inner

centromere protein (INCENP) for Aurora B and C (95, 223). These interactions induce

conformational changes in Aurora kinases required for full activation (21, 364).

Aurora A is required for mitotic entry and for bipolar spindle formation. At the G2/M

transition, Aurora A directly phosphorylates and activates Plk1 and Cdc25B (90, 244, 361).

Together, these events promote cyclinB/Cdk1 activation and mitotic entry. Aurora A also

regulates the mitotic spindle by promoting centrosome maturation and separation, and in

Drosophila, C.elegans, and Xenopus model systems, inhibition of activity

results in monopolar spindle formation (27, 128, 142, 230).

Of the three mammalian Aurora kinases, Aurora B is most closely related to the

single Aurora homologue found in lower eukaryotes. Primary functions of Aurora B feature

during chromatid separation and cytokinesis. Aurora B regulates appropriate chromatid

segregation by multiple mechanisms (reviewed in (410)). For example, Aurora B activity

11 stabilizes bi-polar kinetochore-microtubule attachments and is required to maintain the SAC prior to the metaphase-anaphase transition (76, 149, 210). Aurora B also plays an essential role in cytokinesis, and depletion in C.elegans embryos and Drosophila S2 cells leads to impaired cytokinesis and polyploidy (126, 358). Aurora B may have an additional function during chromatin condensation either by phosphorylating histone H3 (S10) and/or by chromosomal targeting of the Condensin I complex, which drives condensation (125, 226).

The mechanism of Aurora B signaling for this event appears to differ between organisms and has not been fully defined in mammalian cells (reviewed in (410)).

Little is known about Aurora C, the third , whose expression is low in most adult tissues, with prominent expression only seen in the testis (189). In HeLa cells, ectopic expression of Aurora C can rescue multinucleation defects caused by Aurora B depletion. Taken together with the fact that it interacts with INCENP, Aurora C may have partially redundant functions with Aurora B in cells (223).

1.1.4 Checkpoint signaling

All eukaryote cells have multiple molecular mechanisms to identify and repair damaged DNA and preserve genomic integrity (reviewed in (353)). An important aspect of this process is activation of a checkpoint, or cell cycle arrest, to allow the cell time to repair damage. Cell cycle arrest can be triggered at G1/S, intra-S and G2/M by damage caused by endogenous sources, such as stalled replication forks, or by exogenous agents including ultraviolet (UV) radiation, ionizing radiation (IR), reactive oxygen species (ROS), and certain chemotherapeutic agents. Upon successful repair, the cell will re-enter the cell cycle.

12 However, if the damage is too severe, apoptosis may be initiated. Loss of checkpoint or

repair functions promotes propagation of damaged DNA and can have deleterious consequences on cell and organismal viability.

1.1.4.1 ATM and ATR signaling in the checkpoint response

The DNA damage checkpoint is controlled by phosphorylation cascades initiated by

two phosphatidylinositol 3-kinase-related kinases (PIKK), ataxia-telangiectasia mutated

(ATM) and ATM and Rad3-related (ATR) (Fig 1.3). ATM was originally identified as the

causal mutation in the neurodegenerative disease ataxia telangiectasia (A-T) (355)(reviewed

in (215)). Cells established from patients with A-T display hypersensitivity to ionizing

radiation and defects in G1/S and G2/M checkpoints (161, 184, 300). ATR is the human

homologue of the S.pombe gene rad3, mutants of which display checkpoint defects following

UV, IR, or inhibition of DNA replication by hydroxyurea (7, 25, 55). Mutations in ATR were

identified in patients with Seckel syndrome, which is characterized by dwarfism and mental

retardation (134). Fibroblasts derived from a patient with Seckel syndrome exhibit impaired

DNA damage response and hypersensitivity to UV irradiation and mitomycin C (MMC)

(290).

ATM is primarily activated by agents that cause double strand breaks, such as IR,

whereas ATR responds to accumulation of single strand DNA caused by stalled replication

forks, UV-induced damage or DNA crosslinks, and to a lesser extent by IR-induced damage

(44, 290, 424). While the exact molecular mechanisms differ, both kinases are activated

following recruitment to sites of DNA damage by multi-protein complexes, and

13 14 phosphorylate numerous substrates with overlapping specificity (254, 371)(reviewed in (54)).

The best-characterized downstream targets of ATM/ATR are the effector kinases, checkpoint

kinase 1 and 2 (Chk1 and Chk2), and the tumour suppressor protein p53 (44, 137, 173, 256,

398).

In mammalian cells, the p53 transcription factor mediates a cell’s decision to undergo

cell cycle arrest or to commit apoptosis following cellular stress (reviewed in (417)). In the absence of stress, p53 is downregulated by an interaction with the E3 ubiquitin ligase, murine

double minute 2 (Mdm2). In addition to promoting the ubiquitination and degradation of p53,

Mdm2-binding inhibits interaction of p53 with transcriptional co-activators and blocks transactivation (204, 273). Following DNA damage, ATM/ATR stabilizes p53 through direct and indirect mechanisms. ATM/ATR directly phosphorylate p53 at S15, inhibiting its interaction with Mdm2, and Mdm2 at S395, promoting its autodegradation (44, 380, 398).

ATM/ATR activation of other kinases such as Chk1/2 and Abelson murine leukemia viral

oncogene homolog 1 (Abl1) may also promote p53 stability through their phosphorylation of

p53 and Mdm2 (reviewed in (265)). One of the key targets of p53 is p21, which is rapidly induced following DNA damage and inhibits cyclinD/Cdk4/6 and cyclinE/Cdk2 complexes

(86, 227). In addition to p21, p53 regulates the expression of a number of anti-proliferative

genes and while p53 can contribute to all checkpoints, the G1/S cell cycle arrest is dependent

on functional p53 (184). Many human tumours and immortalized cell lines have

compromised p53 activity and G1/S arrest following damage. In these cases, the G2/M

checkpoint takes on increasing importance for maintaining genomic stability.

1.1.4.2 Chk1/2 signaling in the checkpoint response

15 Chk1 and Chk2 are non-homologous serine-threonine kinases that act downstream of

ATM/ATR (231, 255). Phosphorylation of two sites within the non-catalytic C-terminus by

ATM/ATR, increase Chk1 activity, likely by relieving autoinhibition (120, 186,

231)(reviewed in (51)). By comparison, Chk2 is phosphorylated within the non-catalytic N- terminus by ATM/ATR following DNA damage, promoting Chk2 dimerization, and facilitating autophosphorylation, both within the autoinhibitory loop and the activation loop

(216, 256, 437, 441)(reviewed in (5)). While Chk1 and 2 phosphorylate many of the same

downstream targets, they also have distinct functions, highlighted by the phenotypes of mice deficient for their genes. Chk1-null mice die early during embryogenesis, while Chk1- deficient embryonic stem (ES) cells and embryos display a defective G2/M arrest following

UV and IR (231, 387). In contrast, mice deficient for Chk2 are viable, whereas mouse

embryonic fibroblasts (MEFs) derived from them are radioresistant and display mild checkpoint defects in response to IR (155, 169, 386). These and other studies suggest that

Chk1 is the major checkpoint kinase, while Chk2 has a specific function in IR-induced

apoptosis and plays a supporting role in other circumstances (reviewed in (19)).

One of the key targets of Chk1/2 are the Cdk-activating phosphatases, Cdc25A, B and

C, which are inhibited through phosphorylation by Chk1/2. Cdc25A promotes cyclin/Cdk

activation at all stages of the cell cycle, whereas Cdc25B/C act predominantly during G2/M

to dephosphorylate and activate Cdk1 (177, 209, 268, 272)(reviewed in (80)). Following

genotoxic stress, Cdc25A is polyubiquitinated and degraded via the ubiquitin ligase Skp

Cullin F-box complex-β-transducin repeat-containing protein (SCF-β-TrCP). Interaction

between Cdc25A and SCF-β-TrCP requires phosphorylation of a DSG motif within Cdc25A

16 which occurs in a Chk1/2-dependent manner (42, 79). In mammalian cells, Chk1/2

phosphorylates Cdc25A, priming it for further DSG phosphorylation by NEK11 (79, 176,

260). In contrast to Cdc25A, Cdc25B/C are functionally inactivated by Chk1/2

phosphorylation, which creates 14-3-3 binding sites and leads to their cytoplasmic

sequestration, away from their nuclear target Cdk1 (62, 64, 307)(reviewed in (35)). In

addition, using a Xenopus egg/embryo system, Chk1 but not Chk2, was shown to inhibit

Cdc25A through phosphorylation at a C-terminal site (T504 in Xenopus Cdc25A), which

promoted 14-3-3 binding and inhibited Cdc25 interaction with cyclin/Cdk complexes (409).

Sites equivalent to T504 exist in mammalian Cdc25A, B and C indicating that phosphorylation of this residue may contribute to phosphatase-inactivation in all three proteins.

1.1.4.3 p38 signaling in the checkpoint response

The serine-threonine kinase, p38, has been implicated in checkpoint control parallel to Chk1/2 (reviewed in (331)). p38 kinases are members of the family of mitogen activated

protein kinases (MAPKs) and are part of the general stress response. They are activated by a

variety of stimuli including osmotic stress, heat shock, UV, IR and reactive oxygen species

(ROS) and, depending on the stimuli and cell context, can exert pro-survival, pro-apoptotic

or pro-differentiation effects (reviewed in (418)).

Following genotoxic stress, p38 plays a critical role in checkpoint initiation by

promoting inactivation of Cdc25B/C (41, 423)(reviewed in (331)). The p38-associated and -

activated kinase, MAPK activated protein kinase 2 (MK2, also known as Mapkap kinase 2),

17 phosphorylates Cdc25B and Cdc25C and, in a similar manner as Chk1/2 phosphorylation,

this leads to 14-3-3 binding and cell cycle arrest (41, 248). Consistent with an integral role in

checkpoint function, MK2-depleted U2OS cells (human osteosarcoma cells) are defective in

both G1/S and G2/M checkpoints after UV irradiation (248). Several lines of evidence

indicate that checkpoint control may be a conserved function of p38. In fission yeast, the

p38/MK2 homologues, Sty1/Srk1, are required for G2/M arrest following osmotic stress, via

a mechanism that involves Srk1 phosphorylation of Cdc25, 14-3-3 binding and cytoplasmic

sequestration (2).

The mechanism of p38 activation is not completely understood, but appears to differ

depending on the specific genotoxic stress involved. Raman et al. demonstrated that in HeLa

cells, thousand and one amino acid (Tao) kinases, which are directly phosphorylated by

ATM, were required for p38 activation in response to IR (327). Interestingly, this likely

reflects a stimulus-specific response, as in ATM-deficient and ATR-defective fibroblasts,

p38/MK2 activation was inhibited in response to doxorubicin, but not UV irradiation (330).

1.2. NimA-related kinases (NEKs)

1.2.1 Never in mitosis A (NimA)

The NEK family of serine-threonine kinases, which in mammals consists of 11 members, are the least characterized of the mitotic kinases. Never in mitosis A (NimA) is the

founding member of the NEK family and was identified in the filamentous fungus,

Aspergillus nidulans, in a screen for temperature sensitive mutations that blocked entry into

18 mitosis (278). At the non-permissive temperature, mutants of nimA arrest in late G2 with

duplicated spindle pole bodies, which are the fungal equivalent of centrosomes (293).

Conversely, overexpression of NimA caused premature entry into mitosis marked by chromosome condensation and aberrant mitotic spindle formation (278, 297).

NimA expression and activity is controlled in a cell cycle-dependent manner and by transcriptional, post-translational modification and proteolytic mechanisms, is restricted to a brief window during the G2-M transition and into mitosis (reviewed in (298)). In all eukaryotes studied to date, activation of p34cdc2 is sufficient to initiate mitosis. The sole exception, is A.nidulans, where there is also a requirement for NimA (296). Unlike in higher eukaryotes, mitosis in A.nidulans is closed, meaning that nuclear envelope breakdown does not occur. In cells mutant for nimA, active p34cdc2 remains in the suggesting that

NimA is required to transport active cyclinB/p34cdc2 complexes into the nucleus prior to

mitosis (436). Interestingly, NimA itself is activated downstream of p34cdc2, likely via direct

phosphorylation (451). Additional mitotic roles for NimA include induction of chromosome

condensation through phosphorylation of histone H3 and a potential function in regulating nuclear membrane fission during mitotic exit (67, 69).

The function of NimA in promoting cell cycle progression in A.nidulans raised the possibility that homologues of NimA existed in higher eukaryotes. Consistent with this, overexpression of NimA in S.pombe and in human HeLa cells induced chromosome condensation in the absence of other mitotic events such as microtubule spindle assembly or p34cdc2 activation (240, 289). Furthermore, overexpression of a catalytically-inactive NimA delayed entry into mitosis, suggesting that it may be interfering with the activity of an

endogenous kinase or sequestering a conserved substrate (240). Indeed, NimA-related

19 kinases (NEKs) have been identified in higher eukaryotes where the family expanded through evolution. While a single NimA homologue exists in yeast, 2, 4 and 11 homologues were identified in D.melanogaster, C.elegans and mammals, respectively (Fig 1.4). While they display a high degree of sequence similarity within their kinase domains, NEK family members share little homology outside this region, indicative of a divergence in function. In support of this notion, the only gene that can functionally complement the nimA mutation is nim-1 from the related fungus Neosporra crassa (320). Neither of the yeast homologues of nimA (fin1 in S.pombe; KIN3 in S.cerevesiae) or NEK2, the closest mammalian homologue is able to rescue the cell cycle defect incurred by defects in nimA (202, 288).

1.2.2 Organization of NEK proteins

NimA has an N-terminal catalytic domain and, in terms of substrate specificity,

exhibits specificity towards N-terminal hydrophobic residues and a phenylalanine at position

–3, relative to the target residue (FR/KR/KS/T) (241). NimA contains coiled-coiled domains, which are important for oligomerization, and PEST sequences, which mediate ubiquitin-

dependent proteolysis, a process that may be required for A.nidulans to exit mitosis (Fig 1.5)

(319). Despite low overall , the organizational features of NimA are

broadly conserved among mammalian NEK kinases. For instance, all NEK kinases, except

NEK10, contain N-terminal catalytic domains, whereas NEK4, 6 and 7 are the only family members that do not contain coiled-coiled motifs. Moreover, 6 of 11 mammalian NEK kinases contain putative PEST sequences.

20 21 22 Outside regions of homology, certain NEK kinases contain unique protein domains that point to the acquisition of novel functions compared to the ancestral NimA. NEK8 and

NEK9 feature regulator of chromosome condensation (RCC1) repeats. These repeats are also

found in RCC1, a guanine nucleotide exchange factor (GEF) for the small GTPase, Ras-

related nuclear protein (Ran). While the function of RCC1 repeats has not been defined in

NEK8, in NEK9 this domain acts as a negative regulator of NEK9 catalytic activity and can

interact with Ran. However, the consequence of this interaction for either NEK9 or Ran

function is not clear (340). Additional unique domains in NEK family members include a

predicted DEAD-box helicase-like domain in NEK5 and a cluster of armadillo repeats in

NEK10 (288).

1.2.3 Mammalian NEK kinases

1.2.3.1 NEK kinases in cell cycle progression

Based on sequence homology within the kinase domain, NEK2 is the closest

mammalian NimA homologue. Unlike NimA however, NEK2 is not essential for mitotic

entry, but regulates centrosome separation during mitosis. NEK2 localizes to centrosomes

during interphase and early mitosis and interacts and phosphorylates several centrosomal

proteins including cNap-1, Rootletin and β-catenin (14, 15, 110, 111). It has been proposed

that the balance between phosphorylation and dephosphorylation of target substrates by

NEK2 and its’ interacting partner, protein phosphatase-1 (PP-1), controls centrosome

separation (150, 263). In support of a role for NEK2 in centrosome regulation,

23 overexpression of NEK2 in U2OS cells results in premature centrosome splitting during

interphase, whereas expression of catalytically inactive NEK2 inhibits centrosome

segregation and increases the rate of spindle defects and multinucleated cells (98, 111).

Consistently, knockdown of NEK2 substrates cNap-1, Rootletin and β-catenin in human cells also causes defects in centrosome separation and spindle formation (14, 15).

In addition to centrosomes, there is preliminary evidence that NEK2 localizes to condensed chromatin during mitosis, and the midbody and kinetochores of dividing NIH3T3 cells (138). NEK2 interacts with certain centromeric proteins, including Hec1 and Mad1 and, in HeLa cells, NEK2 phosphorylation of Hec1 is required for proper chromosome alignment at metaphase (85, 238). Further, knockdown of NEK2 in HeLa cells leads to displacement of the centromeric protein Mad2 from kinetochores and impaired chromosome segregation

(238). Taken together, these studies indicate that NEK2 plays multiple roles in coordinating cell division.

NEK6, 7 and 9 also participate in proper mitotic progression. NEK6 and NEK7 are highly related and are composed almost entirely of catalytic domains, which share 87% identity (181). They were originally identified based on their ability to phosphorylate p70 S6 kinase in vitro (23). However, this kinase is likely not a NEK6/7 target in vivo (234). Instead, there is increasing evidence that NEK6 and NEK7 are required for mitotic progression acting downstream of NEK9 (24, 292). All three kinases display elevated catalytic activity during mitosis, at which time NEK9 interacts with and activates both NEK6 and NEK7 (238).

During interphase, NEK6 and NEK7 are repressed by a unique autoinhibitory mechanism, which is relieved by NEK9 in mitosis (334). Activation of NEK6 and NEK7 results from

24 interaction with NEK9’s non-catalytic C-terminal tail, which relieves autoinhibition, and

direct NEK9-mediated phosphorylation within NEK6/7 activation loops (334, 340).

In HeLa cells, knockdown of either NEK6 or NEK7 leads to apoptosis following

mitotic arrest (292). NEK6 or NEK7 deficiency in HeLa cells results in fragile spindle

formation during mitosis and a prolonged activation of the SAC (292). If the SAC is

inhibited, NEK6- or NEK7-depleted cells will continue to progress through mitosis but will

then arrest during cytokinesis. Therefore, NEK6 and NEK7 function at multiple points during

mitosis to regulate microtubule organization at the spindle poles, mitotic spindle and central

spindle (292). Consistently, inhibition of NEK9, resulting from microinjection of α-NEK9

antibodies, impairs spindle assembly and chromosome alignment during metaphase (340).

Interestingly, despite their strong homology, NEK6 and NEK7 are not redundant and cannot

compensate for the loss of the other kinase (292). This may be explained by their differential

tissue distribution and subcellular localization (101, 292).

In addition to a mitotic role, NEK9 may also function during interphase, when it

interacts with the facilitates chromatin transcription complex (FACT), which modulates

chromatin structure (52). Finally, HeLa cells depleted for NEK9 exhibit slower progression through G1 and S, implicating NEK9 in control of transcription and/or DNA replication (52).

1.2.3.2 NEK kinases in checkpoint control

In addition to roles in normal cell cycle progression, many members of the NEK family have been implicated in checkpoint control and the DNA damage response. Thus far, the molecular mechanism by which NEK11 contributes to checkpoint function has been best

25 characterized. Meliexetian et al. demonstrated that in response to IR, NEK11 was activated

through Chk1-mediated phosphorylation (260). It has been previously established that Chk1

also phosphorylates Cdc25A on S76 (131). This event primes Cdc25A for further

phosphorylation within the DSG motif, which is required for SCF β-TrCP-mediated

ubiquitination and degradation of Cdc25A (176). Significantly, NEK11 has been identified as

the Cdc25A DSG-motif kinase (260). Consistently, HeLa cells depleted for NEK11 display

elevated levels of Cdc25A protein and fail to undergo IR-induced G2/M arrest (260).

In addition to NEK11, NEK1 and NEK2 also participate in IR-induced checkpoints.

Following this insult, NEK1-/- cells are defective in the G1/S and G2/M checkpoints and in

their ability to repair DNA, leading to accumulation of double strand breaks (313). Moreover,

in HK2 and HeLa cells, NEK1 expression and catalytic activity are elevated in response to IR

(50). Conversely, IR inhibits NEK2 activity in a PP-1-dependent manner, which is essential for the radiation-induced inhibition of centrosome splitting (266). In addition to IR, the catalytic activity of NEK1, NEK2, NEK6 and NEK11 is sensitive to genotoxic stresses such as UV and etoposide (104, 218, 286, 313). Thus, various NEK kinases participate in the cellular response to genotoxic stress and can act as positive and negative regulators of damage-induced checkpoints.

1.2.3.3 NEK kinases in ciliagenesis

Cilia are organelles that protrude from the cell surface and are structurally and functionally similar to flagella (reviewed in (93)). A is classified as a motile or non- motile/primary cilium and is composed of a cylindrically organized, microtubule-based

26 cytoskeleton, called an . Motile cilia function to move fluid and debris, and are

found on various cell types including the efferent ductules of the testies, the ependymal

lining of the brain and epithelial cells of the respiratory tract and oviduct (93). Until recently considered vestigial structures, primary cilia are present in most cells. They are generated

during interphase from the mother centriole, which serves as the (reviewed in

(68))(375). The cilia dissemble prior to mitosis, allowing the centriole to nucleate the mitotic

spindle (406)(reviewed in (324)). Both during development and in adulthood, the primary

cilium performs key roles in coordinating extracellular conditions with cellular responses.

Indeed, mutations in ciliary proteins are the basis of a number of human genetic disorders

termed ciliopathies, including retinal degeneration, polycystic kidney, liver and pancreatic

diseases and abnormalities in neural tube closure (reviewed in (26)).

A link between NEK kinases and ciliagenesis originates from identification of loss-

of-function mutations in NEK1 and NEK8 as the causal events in independent mouse models

of polycystic kidney disease (PKD) (232, 408). Two mouse models of PKD feature loss-of

function mutations in NEK1. The kidney, anemia, testes (kat) allele results from an internal

deletion (bp 791-2105), whereas the kat2J allele contains a insertion at position 966,

which generates a premature stop codon (408). The junior cystic kidney (jck) mouse contains

a glycine 488 to valine (G488V) mutation in NEK8, which alters the subcellular localization

of NEK8. Immunohistochemical staining of mouse kidneys revealed that NEK8 is localized

to the apical cytoplasm of duct collecting cells in wildtype mice but exhibits diffuse cytoplasmic staining in jck mice (232). Further supporting a causal role for NEK8 in PKD, zebrafish embryos depleted for NEK8 with antisense morpholinos develop pronephric cysts

(232).

27 While the molecular mechanism by which NEK1 and NEK8 cause PKD is not fully

understood, both proteins have been shown to regulate ciliagenesis. In primary kidney epithelial cells derived from wildtype mice, NEK8 localizes to primary cilia (370). This

localization is lost in cells derived from jck mice and correlates with increased cilia length.

Conversely, IMCD3 cells, derived from the inner medullary collecting duct of the murine

kidney, overexpressing wildtype NEK1 do not contain primary cilia (430). Supporting the

importance of NEK1 kinase activity in ciliagenesis, a catalytically inactive mutant of NEK1

localizes to primary cilia, but fails to affect cilia formation.

Appropriate regulation of cilia is tightly linked to cell cycle progression. The primary

cilium is nucleated by the centriole, which needs to be liberated for proper spindle formation

to occur during mitosis (406)(reviewed in (324)). Indeed, cell cycle proteins contribute to

cilia assembly. For instance, mutations in Fa2, a NEK protein in Chlamydomonas, cause

defects in both flagellar detachment and progression through the G2/M transition (246). A

further example is Aurora A, which regulates centrosome maturation and mitotic entry

during the cell cycle (27, 90). This kinase is also activated during cilium disassembly and is

necessary for cilia-resorption in the human retinal epithelial pigment cell line (hTERT-RPE1)

(321).

There may be a correlation between an expansion of NEK kinases in the genomes of

higher eukaryotes, and the need to coordinate primary cilia with cell cycle progression (323).

For example, and yeast are non-ciliated and contain a single

NimA/NEK protein. Drosophila and C.elegans have 2 and 4 NEK kinases, respectively.

While these organisms contain ciliated cells, these are terminally differentiated and do not

coordinate cilia function with the cell cycle. In contrast, organisms such mammals,

28 Chlamydomonas and Tetrahymena, contain proliferating ciliated cells and feature an expansion in the number of NEK kinases they contain, with 11, 10 and 35 members, respectively. In these organisms, NEKs have been implicated in regulating cilia and/or centrioles/centrosomes (reviewed in (323)). For example, in mammals, in addition to NEK1 and NEK8, NEK2, 6, 7 and 9 associate with centrioles and/or the mitotic spindle. Future studies aimed at this largely unexplored aspect of the biology of NEK kinases will shed further light on their relationship with basal bodies, cilia and centrioles (reviewed in (323)).

1.3 NEK10

1.3.1 NEK10

NEK10 is a previously uncharacterized member of the NEK family and is evolutionarily its most divergent member. NEK10 is located on human chromosome 3p24 and encodes a 1125 amino acid protein. Unlike other NEK and NimA-related kinases, which feature N-terminal catalytic domains, NEK10’s kinase domain is centrally located. In addition, NEK10 contains coiled-coiled domains flanking the catalytic domain, a putative

PEST sequence in the C-terminus and 4 N-terminal armadillo repeats (Fig 1.5).

Armadillo repeats are related to the HEAT (Huntington, Elongation Factor 3,

PR65/A, TOR) domain and together they are found in less than 10 human kinases (249).

Each armadillo repeat is composed of approximately 40 amino acids, which form three α- helices (162). It has been shown that multiple contiguous armadillo repeats are required to create a functional domain (156). Armadillo repeats have been implicated in coordinating

29 protein-protein interactions. For example, β-catenin’s armadillo repeats mediate interactions

with several partner proteins including axin, cadherin and adenamatous polyposis coli (APC)

(reviewed in (440)). Another well-characterized interaction is between the armadillo repeats

of importin-α, which is part of the nuclear import machinery and the nuclear localization sequences of target proteins (151). The significance of armadillo repeats in NEK10 is currently unknown.

Judged by sequence homology within the catalytic domain, NEK10 is most similar to the mitotic kinases NEK6 and NEK7. Its closest homologue, however, is the uncharacterized

C.elegans protein nekl-4, which also contains armadillo repeats and a centrally located catalytic domain. NEK10 and nekl-4 share 72% homology within the kinase domain and

34% overall homology. By comparison, NEK10 shares only 54% homology within the

kinase domain with NEK6/7. nekl-4 expression as detected by in situ hybridization is

ubiquitous during C.elegans embryogenesis, but is restricted during larval and adult stages to

a limited number of cells, likely either muscle or nerve cells near the pharynx

(http://nematode.lab.nig.ac.jp/). In large-scale, loss of function screens, knockdown of nekl-4 has not been associated with any of the common phenotypes examined, such as sterility, lethality or defects in movement or egglaying, and its physiological function is currently unknown (http://www.wormbase.org/).

1.3.2 NEK10 and carcinogenesis

Growing evidence indicates that NEK10 may participate in oncogenesis. Whole genome sequencing of primary tumours and immortalized human cancer cell lines uncovered

30 more than a 1000 somatic mutations within the coding sequences of the 518 predicted human protein kinases (66, 136). Mutations in NEK10 were found in primary tumours (ovarian mucinous carcinoma (A66K), large cell carcinoma (R878M)) and cultured cell lines

(metastatic melanoma (E379K), and lung neuroendocrine carcinoma (P1115L)) (136). None of the identified mutations map to the catalytic domain of NEK10, and their effect on protein function is currently unknown. Somatic mutations fall into two classes, driver mutations, which confer a growth advantage and are positively selected, and passenger mutations, which occur by chance and are not subject to selection. Based on the frequency of mutations,

NEK10 was defined as one of 120 kinases predicted to contain a driver mutation (136). Of

note, NEK10 mutations were found with the same frequency (4/33) as the mutations of B-

Raf and liver kinase B1 (LKB1), kinases previously implicated in tumourigenesis (66).

A further suggestion of potential NEK10 involvement in cancer came from a genome

wide association study (GWAS) involving over 37,000 breast cancer samples and 40,000

controls, which identified a strong breast cancer susceptibility locus within 3p24 (p value =

4.1 x 10-23). Importantly, the sub-region of 3p24 identified by this GWAS contains only two genes, NEK10 and solute carrier family 4, sodium bicarbonate co-transporter, member 7

(SLC4A7).

Prompted by the reports of NEK10 mutations in human cancers, our lab explored

NEK10 expression profiles in a series of breast cancers. Mining the NKI295 breast cancer

dataset, consisting of 295 breast cancer patients, we found that low NEK10 expression was

significantly associated with >20% reduction in 5-year and overall disease-free survival (p =

0.0014) (Fig 1.6C) (411). When cross-referenced with key pathology parameters, low

NEK10 expression was associated with a number of characteristics indicative of poor

31 32 prognosis including triple negative status and poorly differentiated tumours (Fig 1.6A, 1.6B).

Consistent with this data, low NEK10 expression was associated with higher tumour grade in an independent cohort of 88 breast cancers isolated and analyzed at Princess Margaret

Hospital/Ontario Cancer Institute (in collaboration with Dr. Wey-Liang Leong, Surgeon-

Scientist, PMH). Finally, compared to nonmalignant control tissue, reduced NEK10 expression was found in high-grade serous ovarian carcinoma and in a series of nasopharyngeal carcinomas (362, 401). Taken together, these studies strongly implicate

NEK10 expression as a possible risk factor for disease severity in several human cancers.

1.4 ERK1/2 signaling

Work described in Chapter 2 links NEK10 to regulation of the ERK1/2 pathway. This

section will provide a brief overview of ERK1/2 signaling, with emphasis on its function in mammalian systems.

1.4.1 Organization of MAPK signaling modules

Mitogen activated protein kinase (MAPK) pathways are highly conserved kinase

cascades. They are activated by a wide variety of extracellular stimuli and are critical for

numerous cellular processes such as proliferation, differentiation, migration, and apoptosis

(reviewed in (48)). The three best-characterized MAPK subfamilies are extracellular

regulated kinases (ERK1/2) (also known as MAPK1/2), Jun NH2 terminal kinases

33 (JNK1/2/3, also known as stress activated protein kinases (SAPK1/2/3)) and p38α/β/γ/δ

kinases.

MAPK cascades are three-tiered signaling modules consisting of a MAPK, a MAPK

kinase (MAPKK) and a MAPKK kinase (MAPKKK) (reviewed in (203)). MAPKKK are

serine-threonine kinases that are activated in response to extracellular stimuli by phosphorylation and/or interactions with small GTP binding proteins such as Ras or Rho.

MAPKKK phosphorylate and activate the dual specificity kinases, MAPKK, which in turn

activate the MAPK by phosphorylation of a conserved TXY motif within the activation loop.

MAPKs are proline-directed serine-threonine kinases and phosphorylate substrates on S/T

residues that are directly followed by a proline residue.

ERK1/2 were the first identified MAPK subfamily and are strongly activated by

growth factors, serum and phorbol esters (57, 342). In addition, ERK1/2 show lower levels of

activation to a wide variety of stimuli including cytokines, G-protein coupled receptor

activation and genotoxic stress (99, 229, 233). JNK1/2 and p38 MAPK families are most

strongly activated by cellular stresses including UV irradiation, genotoxic stress and in the

case of p38, osmotic stress (141, 152)(reviewed in (182)).

1.4.2 ERK1/2 signaling

The canonical ERK1/2 signaling module is composed of Raf (MAPKKK), MEK

(MAPKK) and ERK (MAPK) kinases which are activated downstream of the small GTPase

Ras (Fig 1.7) (reviewed in (259)). The ERK cascade is one of the most extensively studied signaling pathways and plays a central role in organismal development and homeostasis and

34 35 when deregulated, in carcinogenesis. Activating mutations in Ras and Raf are frequently

found in tumours and ERK signaling is best characterized in the context of growth factor stimulation and control of cell proliferation.

1.4.2.1 MAPKKK: Raf kinase

Raf is the primary MAPKKK involved in ERK signaling and constitutes the first step

in the kinase cascade. There are three Raf isoforms in mammals, A-Raf, B-Raf and C-Raf

(also known as Raf-1) (165). B-Raf is most closely related to the single Raf isoform present

in lower eukaryotes such as D.melanogaster (DRaf) and C.elegans (lin-45) (457). The three

kinases have unique and overlapping functions. A-Raf knockout mice are viable, but exhibit

neurological and intestinal defects after birth (318). Deletion of either B-Raf or Raf-1 in mice

results in embryonic lethality around midgestation (164, 267, 434). Interestingly, in response

to growth factors, both A-Raf- and Raf-1-deficient fibroblasts exhibit normal ERK activation

whereas ERK activation is reduced, but not abolished in cells derived from B-Raf-deficient

embryos (164, 267, 318, 434).

Raf-1 was originally identified as the cellular homologue of the viral oncogene v-raf,

and is the best characterized of the three isoforms (328). More recently, interest in B-Raf has

been sparked by the discovery that activating mutations in B-Raf frequently occur in human

tumours (65). In contrast, little in known about the physiological role of A-Raf, although it is

predicted that it’s mechanism of activation is similar to Raf-1 (252)(reviewed in(429)). Raf-1

activation following growth factor stimulation is complex and involves protein-protein

interactions, changes in subcellular localization and multiple phosphorylation events. In

36 unstimulated cells, Raf-1 is cytoplasmic and is held in an autoinhibited state by phosphorylation and 14-3-3 binding at S259 and S621 (220, 407). Following growth factor stimulation, Raf-1 is recruited to the plasma membrane by an interaction with the effector domain of the activated GTPase Ras (220, 274). This promotes dephosphorylation of inhibitory phospho-sites and phosphorylation of positive regulatory sites including S338,

Y341 and T491 and S494 within the activation loop (71, 220)(reviewed in (74)). Candidate upstream kinases that contribute to phosphorylation of the activating sites include protein

kinase B (PKB, also known as Akt) for S259, p21-activated kinase (Pak) for S338 and Src

and Janus kinase (JAK) kinases for Y341 (96, 191, 251, 462).

In comparison to Raf-1, activation of B-Raf is simpler in response to extracellular

stimuli, requiring only Ras binding and activation loop phosphorylation (252). Raf-1 contains

five activating phosphorylation sites, while in B-Raf only three appear to play similar

functions. This includes S728 (equivalent to S621 in Raf-1), which mediates 14-3-3 binding,

and activation loop sites T598 and S601 (equivalent to T491 and S494 in Raf-1) (reviewed in

(429)). In Raf-1 phosphorylation of Y341 and S338 are also required for activation. In

contrast, B-Raf exists in a “primed” state and in B-Raf these residues are replaced by an aspartic acid (D448) and are constitutively phosphorylated (S445), respectively (253). The fact that B-Raf exists in a primed state may contribute to the fact that it, but not Raf-1 or A-

Raf, is found mutated in human tumours. Raf-1 and B-Raf are also subject to feedback phosphorylation by ERK (16, 83, 337). In the case of Raf-1, the significance of this event is not fully understood whereas for B-Raf, phosphorylation acts to negatively regulate its activity.

37 Defining the specific roles of B-Raf and Raf-1 in ERK activation is further complicated by the ability of B-Raf and Raf-1 to homo- and heterodimerize. The importance

of Raf dimerization for ERK activation was highlighted by the discovery of oncogenic mutants of B-Raf, which have reduced catalytic activity but still activate ERK (420). This seemingly contradictory property can be explained by the fact that these B-Raf mutants bind and stimulate Raf-1 catalytic activity (117, 420). Systematic evaluation of Raf dimers revealed that B-Raf/Raf-1 heterodimers display higher catalytic activity compared to homodimers/monomers of B-Raf, which themselves have higher activity than the homodimers/monomers of Raf-1 (117, 348). Interestingly, heterodimers composed of one active and one inactive Raf kinase exhibit similar activity to heterodimers composed of two active kinases (348). This suggests that enhanced catalytic activity is a result of the physical interaction and is not caused by intermolecular phosphorylation. Initially, the fact that growth factor-stimulated ERK activation is compromised in B-Raf- but not Raf-1-null fibroblasts was interpreted to indicate that B-Raf was the predominant kinase for MEK (164, 267, 434).

Nevertheless, subsequent studies demonstrated that loss of B-Raf also impaired Raf-1 activity, suggesting that the effect on ERK phosphorylation likely arises from a combined decrease in B-Raf and Raf-1 activity within the cell (117).

MEK1/2 are the only well-defined Raf substrates. However there is accumulating evidence that Raf-1 has kinase-independent functions in apoptosis. Raf-1 deficient embryos display increased apoptosis, while Raf1-/- cells are hypersensitive to pro-apoptotic stimuli

such as etoposide and α-Fas antibody (164, 267). Hypersensitivity to apoptosis in the

absence of Raf-1 is independent of MEK, as it can be rescued by expression of a catalytically impaired mutant of Raf-1 (Raf-1 Y340F Y341F), but appears dependent on the ability of

38 Raf-1 to interact with, and inhibit, other pro-apoptotic kinases, including apoptosis signal-

regulating kinase 1 (ASK1) and MST2 (164, 291, 443)(reviewed in (116)).

In addition to Raf kinases, other MAPKKKs that activate MEK in a cell-type and

stimulus-specific manner exist. For instance, the serine-threonine kinase, mouse sarcoma

(Mos), acts as the MAPKKK in germ cells where it is necessary for MEK/ERK activation during oocyte maturation (316, 414)(reviewed in (350)). Oocyte maturation occurs during

ovulation and is the process by which oocytes arrested at the first meiotic prophase reenter

meiosis in response to hormones such as progesterone (reviewed in (350)). These cells subsequently arrest at the second meiotic metaphase (metaphase II) as mature unfertilized

oocytes. Mos and MAPK are essential components of the cytostatic factor (CSF), which is responsible for metaphase II arrest (140, 414)(reviewed in (350)). Another stimulus-specific

MAPKKK is tumour progression locus 2 (Tpl-2, also known as MAP3K8 and Cot),

implicated in innate immunity. Tpl-2 links toll-like receptor (TLR) activation to tumour

necrosis factor (TNF) production by regulating ERK1/2. This is best exemplified by the lack

of ERK1/2 activation and TNF-α production in LPS-treated macrophages isolated from Tpl-

2 knockout mice (87). Interestingly, in MEFs, Tpl-2 is also implicated in activating ERK1/2 and JNK1/2 in response to TNF-α (63).

1.4.2.2 MAPKK: MEK1/2

In the ERK1/2 cascade, the dual specificity kinases, MEK1/2 are the MAPKK which lie immediately downstream of the MAPKKK. MEK1 and MEK2 are highly homologous, displaying 80% identity, and perform both overlapping and non-redundant roles within the

39 cell. MEK2-null mice are viable and fertile and cells derived from them exhibit normal ERK

activation (22). By contrast, mice null for MEK1 are embryonic lethal. Interestingly, MEFs

derived from these mice display normal ERK2 activation, indicating that this function of

MEK1 may be compensated for by MEK2 (127). Consistent with distinct physiological roles,

MEK1 may be regulated independently from MEK2. MEK1, but not MEK2, is subject to positive (S298 by Pak1) and negative (T292 by ERK) regulatory phosphorylation and in certain circumstances may be preferentially activated. For instance, in serum-stimulated

NIH3T3 cells MEK1, but not MEK2, forms a complex with Ras and Raf-1 (175).

Generally, MEK1/2 are activated by phosphorylation of S218 and S222 (in MEK1) within the activation loop by the upstream MAPKKK, most commonly Raf kinases (459). An additional level of regulation can be provided by other phosphorylation events, in a context- dependent manner. For instance, in adherent cells, MEK1 activation is promoted by phosphorylation on S298 by Pak1, which can be inhibited through feedback phosphorylation on MEK1 T292 by ERK1/2 (109). Phosphorylation of S298 has also been shown to promote

MEK1 autophosphorylation in the context of cell adhesion (303). ERK1/2 are the only established MEK1/2 substrates, a testament to the high degree of substrate specificity of

MEK1/2.

1.4.2.3 MAPK: ERK1/2

ERK1/2 (also known as p42/p44MAPK and MAPK1/2) are serine-threonine kinases originally identified as growth-factor stimulated kinases, which phosphorylate microtubule- associated protein-2 (MAP-2) and myelin basic protein (MBP) (159, 341). Since their

40 discovery, many lines of investigation have revealed their essential functions in numerous cellular processes. Despite the high degree of homology between ERK1 and ERK2 (85% amino acid identity), knockout mice display distinct phenotypes. ERK1-deficient mice are viable but display defective thymocyte maturation (299). In particular, ERK1 is critically required for positive selection of thymocytes, as well as for thymocyte proliferation (299). In

contrast, ERK2-null mice are embryonic lethal at gastrulation (E6.5), due to a defect in

mesoderm differentiation (450). Immunohistochemical analysis of ERK2-/- embryos does not reveal any activation specific-ERK staining, indicating that ERK2 but not ERK1 is active during early embryogenesis (450). Interestingly, ERK2-/- embryos (E6.5 and E7.5) and ES cells display normal proliferation, suggesting that at this early stage in development, ERK

signaling is dispensable for control of proliferation.

ERK1/2 are activated by phosphorylation of T202/Y204 (in ERK1) within the TEY motif by the dual specificity MAPKKs, MEK1/2 (43, 460). Phosphorylation of these sites is necessary and sufficient for ERK activation and dephosphorylation of either site by serine- threonine, tyrosine or dual specificity phosphatases is a powerful means of downregulation.

The dual specificity MAPK phosphatases (MKPs) are the largest family of phosphatases that act primarily on MAPKs. There are 10 MKPs, which display varying degrees of specificity towards the ERK, p38 and JNK MAPKs (reviewed in (187)). Many of the MKPs are present at low levels in unstimulated cells and are transcriptionally induced upon appropriate extracellular stimuli including serum, growth factors and cytokines (36, 92). Significantly, expression and protein stability of several MKPs is MAPK dependent (reviewed in (187,

305)). For example, MKP-1 (also known as DUSP-1) protein is undetectable in quiescent

CCL39 cells, a Chinese hamster fibroblast cell line, but is induced within 30 minutes of

41 serum stimulation in an ERK-dependent manner (36, 37). The increase in MKP1 protein

levels is mediated by increased , and direct phosphorylation by ERK. This

indicates that ERK signaling can stimulate feedback loops and negatively regulate its own activation.

Unlike Raf and MEK, ERK1/2 have numerous downstream targets, both in the cytoplasm and in the nucleus. These include MAPK activated protein kinases (MAPKAPs)

such as ribosomal S6 kinases (RSKs), mitogen and stress activated protein kinases (MSKs)

and MAPK integrating kinases (MNKs), cytoskeletal proteins, tau and paxillin, and nuclear

substrates, including the transcription factors, E-26 like protein 1 (Elk-1) and

myelocytomatosis viral oncogene homolog (Myc), and FBJ murine osteosarcoma viral

oncogene homolog (fos) and jun, which comprise the activator protein 1 (AP-1) transcription

factor (reviewed in (454)). ERK1/2 are proline-directed kinases and phosphorylate

serine/threonine residues that are followed by a proline residue (133). Based on the looseness of this consensus sequence, ERK1/2 can theoretically phosphorylate 1000’s of proteins.

However, additional specificity is provided by the presence of docking motifs on both

ERK1/2 and putative targets.

ERK kinases recognize at least two motifs: the docking (D) motif and the docking site

for ERK and FXFP (DEF, also known as FXF) motif (170, 449). The D motif is recognized

by most MAPKs with varying specificity. In contrast, the DEF motif is specifically targeted

by ERK1/2, but not by related MAPKs, p38 or JNK (170, 448). The D and DEF motifs,

either alone or in combination, are found in many proteins that interact with ERK (170). For

example, the transcription factor elk-1 contains a D motif and a DEF motif. The former

mediates interactions with ERK and JNK1/2, whereas the latter binds ERK alone (170). In

42 comparison kinase suppressor of ras (KSR), which interacts with ERK but not JNK or p38

contains a DEF motif alone (170). The presence of two ERK interaction motifs in elk-1 is consistent with the stronger affinity that ERK has for elk-1 compared to KSR which contains one interaction motif (170).

In addition to docking motifs present on substrates, ERK1/2 itself contains protein sequences that facilitate protein-protein interactions. The C-terminal common docking (CD)

site, also present on other MAPKs, mediates interaction with D-domains, whereas the ERK docking site (ED), only found on ERK1/2, recognizes the DEF motif (75, 219, 391, 392).

The modular system of recognition motifs allows MAPKs to achieve substrate specificity and

explains how different MAPK families can have partially overlapping substrate specificities.

1.4.3 Molecular mechanisms of ERK signaling regulation

ERKs are activated by a wide array of extracellular stimuli, which promote various

biological outcomes in a cell-type-, tissue- and stimulus-specific manner. For instance,

sustained ERK activation stimulates proliferation in fibroblasts, whereas it is associated with

neurite outgrowth and differentiation in PC12 cells (rat pheochromocytoma cells) (280, 404).

Even within the same cell type, ERK activation can produce different phenotypes depending

on the strength and duration of the signal. In PC12 cells, nerve growth factor (NGF) and

epidermal growth factor (EGF) stimulation leads to sustained and transient ERK activation,

respectively. While transient ERK activation drives proliferation, sustained is responsible for

differentiation (404, 405). The wide spectrum of ERK targets necessitates their tight

regulation.

43 1.4.3.1 Regulation by subcellular localization

In most unstimulated cells, ERK is cytoplasmic. As much as half of ERK activity is associated with microtubules where it influences cytoskeleton dynamics (160, 333). ERK is also retained in the cytoplasm by association with a number of cytoplasmic proteins, including MEK1/2 (113). Following stimulation by certain extracellular stimuli, such as growth factors and phorbol esters, active ERK can translocate to the nucleus via both passive diffusion and active transport (3, 49, 221). The importance of ERK nuclear translocation is highlighted by the fact that it is required for a number of ERK functions such as stimulation of NIH3T3 cell proliferation and PC12 cell differentiation (39, 339).

Nuclear translocation facilitates access of ERK to its nuclear substrates. For example, in primary human foreskin fibroblasts (HFFs), ERK nuclear translocation is induced by phorbol 12-myristate 13-acetate (PMA) but not EGF (431). RSK, a cytoplasmic target of

ERK1/2, is phosphorylated to equal levels in PMA- and EGF-treated cells. However in

PMA-, but not EGF-treated cells, this is accompanied by the accumulation of c-fos, a nuclear

ERK1/2 target, and cyclinD1, which is a transcriptional target of c-fos. Interestingly, the difference in nuclear localization is independent of the amplitude and duration of ERK1/2 activation, which is similar in response to either stimuli (431). Therefore, differences in subcellular localization provide flexibility to generate distinct phenotypic responses to different extracellular stimuli.

44 1.4.3.2 Regulation by scaffolds

Scaffold proteins act as docking platforms that serve to colocalize the components of

a particular signaling cascade. The first identified MAPK scaffold was Ste5 in budding yeast

(53, 200). Since then, numerous proteins that fulfill similar functions have been

characterized. MAPK scaffolds interact with multiple components of the MAPK cascades and modulate the strength, duration and location of activation (reviewed in (73)).

Furthermore, they can provide signal specificity by promoting activation in response to particular stimuli and access to specific substrates (45).

KSR is the best characterized ERK1/2 scaffold and was concurrently identified in

D.melanogaster and C.elegans in screens for modifiers of Ras signaling (196, 383, 396). In cultured cells, KSR constitutively binds MEK and ERK, and following mitogenic stimulation, Raf (397, 438, 455). KSR promotes ERK activation by bringing MEK/ERK into

close proximity to Raf, thereby facilitating cascade activation (344). In addition, KSR binding may directly stimulate Raf catalytic activity (326). In cells derived from KSR1-null

mice, the amplitude and duration of ERK activation following growth factor stimulation is

decreased (197). Furthermore, while these cells proliferate normally, their tumourigenic

potential in response to oncogenic Ras expression is lowered (197).

Other scaffolds with capacity to provide stimulus and substrate specificity to the ERK

cascade exist. Ectopic expression of one of these, MAPK organizer-1 (MORG-1), in NIH3T3

cells, promotes ERK1 activation in response to PMA and lysophosphatidic acid, but not

EGF, whereas in PC12 cells, another scaffold, regulator of G-protein signaling-12 (RGS-12),

45 is required for NGF-, but not fibroblast growth factor (FGF)-induced neurite outgrowth (416,

433). Finally, through direct interaction with ERK cascade components, scaffold proteins can direct ERK activation to specific subcellular compartments, such as MEK partner 1 (MP-1)

to endosomes, similar expression to fgf genes (Sef) to the Golgi, and Paxillin to focal

adhesions (167, 395, 402).

1.4.4 Physiological roles of ERK1/2 signaling

ERK1/2 are activated by numerous extracellular signals, which produce discrete

outcomes in a cell- and stimulus-specific manner. Traditionally, ERK1/2 signaling has been

associated with promoting proliferation and cell transformation. Nevertheless, ERK1/2

activation also contributes to cell differentiation, migration, embryonic development and the

cellular response to genotoxic and inflammatory stresses.

1.4.4.1 ERK1/2 signaling in oncogenesis

Deregulated ERK1/2 signaling can contribute to cell transformation and

tumourigenesis. Activation of ERK1/2 signaling is found in many tumours due to the

activating mutations in its upstream regulators (receptor tyrosine kinases (RTKs), Ras, and

Raf) or downregulation of its negative regulators (Sprouty, neurofibromatosis type 1 (NF1))

(reviewed in (183)). ERK activation likely contributes to tumourigenesis through multiple mechanisms, promoting proliferation, survival, motility and angiogenesis. Curiously, mutations in MEK and ERK have thus far not been identified in human cancers.

46 Activating Ras mutations are found in approximately 30% of human tumours with the highest incidence in pancreatic, colon and lung cancers (reviewed in (33, 183)). Ras acts via several downstream effectors that can contribute to transformation. While ERK activation is not necessary for Ras-mediated NIH3T3 transformation, constitutive activation of Raf or

MEK is sufficient to promote cell transformation and tumourigenesis in mouse models (250,

328, 377).

Even though Raf-1 was initially identified as a viral oncogene, mutations in Raf-1 or

A-Raf have not been found in human tumours (328). Nevertheless, activating mutations in B-

Raf are found in approximately 7% of human tumours, most commonly in melanoma, colorectal, ovarian and papillary thyroid carcinomas (65). The V600E mutation accounts for

80% of B-Raf mutations found in cancer and encodes a constitutively active, Ras- independent protein kinase (65). Other mutations in B-Raf have been identified at much lower frequencies and include at least three mutants that display decreased catalytic activity, yet retain the ability to activate the ERK pathway through a Raf-1-dependent mechanism (65,

420).

1.4.4.2 ERK1/2 signaling in cell cycle regulation

Early work on ERK1/2 linked them to activation by a variety of growth signals. A close correlation was also found between sustained ERK1/2 activity and DNA synthesis

(179, 262). Further work has established that entry into S-phase requires sustained ERK signaling through till late G1 (427, 444). During G1, one of the key events downstream of

ERK1/2 is cyclinD1 induction (427). The cyclinD1 promoter contains an AP-1 binding site

47 and sustained ERK likely induces cyclinD1 expression by promoting stabilization of fos and jun family members, which form the AP-1 transcription factor (18). This is consistent with

the ability of c-fos to act as a molecular sensor for ERK signal duration (280). c-fos is

unstable in conditions of transient ERK activation (280). However, following prolonged

activation, ERK phosphorylates the C-terminus of c-fos. This stabilizes c-fos and exposes a

DEF domain allowing further ERK-mediated phosphorylation. The hyperphosphorylated

form of c-fos is active and can promote expression from AP-1 target genes and cellular transformation (280).

In addition to fos and jun, ERK targets the transcription factor c-Myc to promote G1 progression. ERK stabilizes c-Myc by direct phosphorylation at S62, leading to an upregulation of c-Myc targets, including cyclinD2, Cdk4, and Cdc25A (359)(reviewed in

(261)). Finally, continuous ERK activation is required through G1 to repress expression of up

to 100 antiproliferative genes (444). The importance of this inhibition is demonstrated in

NIH3T3 cells where ectopic expression of a number of these genes, including JunD and

Gadd45α, inhibits serum-stimulated proliferation (444).

Mammalian cells are only sensitive to growth factor stimulation prior to passing R.

However, there is accumulating evidence that ERK is also important at later stages of the cell

cycle. After the G1/S transition, inhibition of ERK activity can delay entry into mitosis and

increase its duration, indicating that ERK signaling participates in both the G2/M transition

and mitotic progression (338)(reviewed in (47)). ERK signaling may impact these processes

by promoting activation and nuclear localization of the cyclinB1/Cdk1 complex.

CyclinB1/Cdk1 is primarily cytoplasmic in interphase and accumulates in the nucleus prior

to mitosis (84). In Xenopus oocytes, ERK directly phosphorylates the cytoplasmic retention

48 sequence (CRS) of cyclin B1, which in conjunction with phosphorylation by Plx, increases

its nuclear localization at the G2/M transition (419). ERK activity also regulates proteins that

control Cdk1 (p34cdc2) activation. During Xenopus oocyte maturation, RSK, phosphorylates and inhibits Myt1, a p34cdc2 inhibitory kinase (301). Additionally, in both Xenopus oocytes

and human cells, ERK activates the Cdk-phosphatase, Cdc25C, by direct phosphorylation on

T48 (422). Finally, localization of active ERK on the spindle and midbody during mitosis

suggests that it may also influence chromosome dynamics during mitosis (432)(reviewed in

(47)).

Somewhat paradoxically, while ERK1/2 activation promotes proliferation and

transformation in many cultured cell lines, its hyperactivation can lead to differentiation or

cell cycle arrest in both primary and immortalized cells (352, 363)(reviewed in (47)). ERK-

mediated cell cycle arrest is accompanied by elevated levels of the Cdk inhibitors p21 and

p16INK4a and can be overcome by co-expression of the adenoviral oncogene E1A or

ablation of a tumour suppressor such as p53 or p16(INK4A) (363, 461). This suggests that there is selective pressure to mutate p53 or p16INK4a during tumourigenesis to avoid cell cycle arrest. Indeed, Ras mutations are often found in combination with p53 or p16INK4a mutations in human tumours (363). ERK hyperactivation may also promote cell cycle arrest through a Cdc25A dependent mechanism. In Xenopus oocytes, Cdc25A degradation is induced by expression of constitutively active MEK, by a mechanism involving both ERK and p90RSK activity (168). Taken together, these studies suggest that in the presence of deregulated mitogenic stimulation, cells may induce senescence as a tumour suppression mechanism to prevent inappropriate cell growth and proliferation.

49 1.4.4.3 ERK1/2 signaling in development and differentiation

Early characterization of ERK signaling in Drosophila and C.elegans revealed their roles in eye and vulval development, respectively (30, 208). In addition to defining the

hierarchal order of the ERK cascade, these studies highlighted a function for ERK signaling

in organism development and cellular differentiation. This has been substantiated by studies

examining the activation pattern of ERK during embryogenesis in zebrafish, Xenopus,

Drosophila, and mice, where it is largely FGF-dependent (59, 61, 114, 115). Despite the high

level of proliferation that is occurring in the embryo, active ERK displays restricted spatio-

temporal localization, suggesting that it is predominantly involved in tissue patterning during

embryogenesis (59). Whereas ERK1-/- mice are viable, ERK2-/- mice are early embryonic

lethal (E6.5) due to a defect in mesoderm differentiation. Interestingly ERK2-null embryos display normal proliferation, indicating that at this stage ERK signaling is required for

differentiation but not proliferation. Other transgenic mouse models have demonstrated critical functions of ERK signaling in placental angiogenesis, and bone and brain development during embryogenesis, and in adipogenesis, learning, and memory in adult organisms (34, 148, 257, 258)(reviewed in (351)).

Developmental functions of ERK signaling are further highlighted by a class of developmental syndromes termed RASopathies caused by germline mutations in several constituents of the ERK pathway, such as Raf-1 germline mutations associated with Noonan

and Leopard syndromes (302). RASopathies display partially overlapping phenotypes

including facial and cardiac abnormalities, neurocognitive delays and malignancy (reviewed

50 in (399)). Finally, a genetic link has been discovered between ERK1 and autism.

Approximately 1% of all cases of autism are associated with a duplication or deletion in chromosome 16p11.2 a locus that encodes ERK1 (206, 428).

1.4.4.4 ERK1/2 in the response to genotoxic stress

The JNK and p38 kinases are strongly activated by cellular stress and are considered the predominant MAPKs involved in the stress response. However, ERK activation in response to these stimuli can also be detected, albeit to a lesser extent than in response to

mitogens. In particular, ERK1/2 activation has been demonstrated following UV, IR,

etoposide, adriamycin and MMC exposure, as well as following breast cancer 1, early onset

(BRCA1) overexpression, implicating ERK1/2 in the cellular response to genotoxic stress

(329, 390, 446).

ERK1/2 activation following genotoxic insults has been linked with appropriate cell

cycle arrest. Yan et al. demonstrated that BRCA1 overexpression in MCF7 cells, which led to a G2/M cell cycle arrest, was marked by Cdc25C degradation, and Wee1 and ERK1/2 activation (446). Notably, in this system, checkpoint activation and Cdc25C and Wee1

regulation, required ERK1/2 activation (446). ERK1/2 activation was also required for

appropriate hydroxyurea-induced, S-phase arrest, in MCF-7, NIH3T3 and HCT116 cells,

where inhibition of ERK1/2 was associated with decreased induction of γH2AX foci and

ATR nuclear localization (435). In addition, MCF-7 cells treated with a MEK1/2 specific

inhibitor, U0126, had defects in the G2/M cell cycle arrest after IR (445). Interestingly, this may be a MEK2 specific response, as in IR-treated HeLa cells, expression of dominant

51 negative MEK2, but not MEK1, inhibited initiation and recovery of cells from G2/M arrest

(1).

1.5 Ph.D. Outline

1.5.1 Study Rationale

Deregulation of the cell cycle is a hallmark of neoplastic transformation and

participates in both the initiation and progression of cancer. Several kinase families play an

indispensable role in governing the mammalian cell cycle including Cdk, Aurora, Polo and

NEK kinases. The focus of this thesis is the biology of NEK10, an unexplored member of the

NEK family.

In Chapter 2, I discuss the importance of NEK10 for G2/M arrest and MEK autoactivation in response to UV irradiation. In Chapter 3, I examine the mechanism of

NEK10 regulation by phosphorylation and subcellular localization. Given that loss of NEK10 has been correlated with poor prognostic characteristics in breast cancer, these studies will

provide insight into the molecular function of NEK10 and its role in tumourigenesis.

1.5.2 Thesis Aims

1. Investigate the cellular role of NEK10 in response to UV irradiation

2. Investigate the role of phosphorylation in regulating NEK10

52 CHAPTER 2

NEK10 ENHANCES MEK AUTOPHOSPHORYLATION FOLLOWING UV

IRRADIATION

A version of this chapter is in revision with MCB

53 2.1 Abstract

Appropriate cell cycle checkpoint control is essential for the maintenance of cell and organismal homeostasis. Members of the NimA-related kinase (NEK) family of serine/threonine protein kinases have been implicated in regulation of various aspects of the cell cycle. We report the cloning and characterization of NEK10, a novel member of the

NEK family and demonstrate a role for NEK10 in the cellular UV response. NEK10 was required for the activation of ERK1/2 signaling upon UV irradiation, but not in response to mitogens, such as epidermal growth factor. NEK10 physically associated with Raf-1 and

MEK1 in a Raf-1-dependent manner, and the formation of this complex was necessary for

NEK10-mediated MEK1 activation. NEK10 did not affect the kinase activity of Raf-1 but instead promoted autophosphorylation-dependent activation of MEK1. The appropriate maintenance of the G2/M checkpoint following UV irradiation required NEK10 expression and ERK1/2 activation. Consistently, the NEK10 C.elegans homologue, nekl-4, modulated

UV-irradiation sensitivity during embryogenesis. Taken together, our results uncover a conserved role for NEK10 in the cellular response to UV irradiation.

54 2.2 Introduction

The NEK kinases are a family of cell cycle regulated, serine-threonine kinases. The

founding member of the family, the Aspergillus nidulans’ Never in mitosis A (NimA) is

essential for mitotic entry (293). Based on the amino acid homology within their respective

catalytic domains, 11 mammalian NEK kinases have been identified (249) and several have

been shown to play diverse roles both during mitosis and at the other phases of the cell cycle.

Further, NEK family members have been implicated in checkpoint control and the DNA

damage response.

Within the NEK family, the molecular mechanism of NEK11 impact on checkpoint

control is the best characterized. The Cdk1-activating phosphatase, Cdc25A, undergoes

ubiquitin-mediated degradation following genotoxic stress, and this event promotes cell cycle

arrest (42). NEK11 directly phosphorylates Cdc25A, enhancing its interaction with the E3 ubiquitin ligase, SCF β-TrCP (260). Significantly, HeLa cells depleted for NEK11 display elevated levels of Cdc25A protein and fail to undergo IR-induced G2/M arrest (260).

In addition to NEK11, other NEK kinases are regulated by genotoxic stresses. In

HeLa cells, NEK2 is inactivated following ionizing radiation, which appears to be essential for radiation-induced inhibition of centrosome splitting (266). Conversely, NEK1 expression

and catalytic activity are elevated in HK2 and HeLa cells treated with IR, whereas

kat2J/NEK1-/- cells are deficient in their ability to repair DNA following this genotoxic stress

(50, 313). Finally, the catalytic activity of NEK1, NEK2, NEK6 and NEK11 appears

sensitive to genotoxic stresses such as ultraviolet (UV) radiation, IR, and etoposide (104,

55 218, 286, 313). Thus, various NEK kinases participate in the cellular response to genotoxic stress and can act as positive and negative regulators of various damage-induced checkpoints.

Many cellular stresses, including UV irradiation, lead to activation of MAPK families, JNK/SAPK, p38 and ERK1/2. While UV-induced JNK activation mediates activation of the AP-1 transcription factor and the cell’s survival response, p38 is required for

engagement of the G2/M checkpoint (41, 423, 456). The physiological relevance and the mechanism of ERK1/2 activation in response to UV irradiation are less well characterized.

Nevertheless, activation of ERK1/2 is emerging as an important aspect of G2/M checkpoint control in a cell type- and stimulus-specific manner. For instance, ERK1/2 activation by IR and etoposide in MCF7 and NIH3T3 cells is required for G2/M arrest (390, 445).

Here, we describe the cloning of human NEK10, a novel member of the NEK family and a recently identified candidate susceptibility gene in cancer (4, 66, 136). Our results demonstrate a role for NEK10 in the maintenance of the G2/M checkpoint following UV irradiation. Mechanistically, NEK10 was found to act as a positive regulator of ERK1/2

signaling in response to UV irradiation, but not mitogenic stimuli, by forming a complex with Raf-1 and MEK1 and enhancing MEK autoactivation. Importantly, our data indicate that checkpoint regulation may be a conserved function of NEK10 between C.elegans and

mammalian cells.

2.3 Materials and Methods

All materials were obtained from Sigma unless otherwise indicated. UV-C irradiation

(254nm) was performed using a UV stratalinker 2400 (Stratagene, La Jolla, CA, USA).

56 Plasmids

NEK10 cDNA was isolated by PCR from a skeletal muscle cDNA library (Clontech

HL5505u) based on the longest predicted NEK10 transcript and was confirmed by sequencing (249). The resulting cDNA was subcloned into the EcoRI and Kpn1 sites of

3xFLAG-CMV-7.1. Deletion mutants of NEK10 were generated by standard recombinant

DNA procedures (details available upon request). Catalytically inactive NEK10 (KD) was generated by site-directed mutagenesis of lysine 548 to arginine. pEBG-Raf-1 was provided by J.Woodgett. Catalytically inactive Raf-1 (KD) was generated by site-directed mutagenesis of lysine 375 to tryptophan. pMCL HA-MEK1, pMCL HA-MEK1 K97M (KD) by M.Cobb, pcDNA HA-MEK1, MEK1∆270-307, Pak1 K299R (KD) by A.Catling.

Cell Culture and Transfection

All cells were cultured in DMEM/10% FBS unless otherwise indicated. HEK293 cells were transfected using the calcium phosphate method. For knockdown using esiRNA, cells were transfected with Dharmafect #1 (Dharmacon) according to manufacturer’s instructions.

Briefly, 1ug of esiRNA was transfected per 35mm plate on Day 1 and Day 2. Media was

changed 24 h following transfection. For FACS analysis, cells were UV irradiated on Day 3

and gathered 24 h later. For Semi-Quantitative (Semi-Q) RT PCR RNA was extracted on

Day 3. For preparation of whole cell lysates, cells were lysed on Day 4.

57 Cell Lysis and Immunoprecipitation

Cells were lysed in CHAPS lysis buffer (40 mM Hepes, pH 7.5, 0.3% CHAPS, 120 mM

NaCl, 1 mM EDTA, 50 mM NaF, 1 mM Na2VO3, 20 mM ß-glycerophosphate, 1mM DTT

and protease inhibitors). Immunoprecipitations were performed with Anti-Flag M2-Agarose

or Glutathione Sepharose 4 Fast Flow (GE Healthcare) for 2 h at 4°C, washed 4 times with

CHAPS lysis buffer containing 220mM NaCl, and eluted by boiling in sample buffer.

Western Blot Analysis and Antibodies

Whole cell lysates or immunoprecipitates were resolved by SDS–PAGE and blotted to PVDF membranes. Proteins were probed using appropriate primary antibodies from the following sources: α-ERK1/2 (#9102), α-pERK1/2 (T202/Y204) (#9101), α-pMEK1/2 (S217/221)

(#9121) α-pS298 MEK1 (#9128), α-pJNK1/2 (T183/T185) (#9251), α-pRaf-1 (S338)

(#9427) from Cell Signaling, α-Raf-1 (C20) (#sc-227), Vinculin from Abcam, α-Flag M2 from Sigma, α-GST from GE Healthcare, and α-HA (12CA5) was harvested from supernatants of the corresponding hybridoma.

NEK10 in vitro Kinase Assay

Flag-NEK10 immunoprecipitates were washed 2 times with CHAPS Lysis buffer + 0.4M

LiCl and 2 times with NEK10 kinase assay buffer (KAB) (50mM MOPS pH7.4, 10mM

MgCl2, 10MnCl2, 2mM EGTA, 20mM ß-glycerophosphate, 1mM DTT). Kinase assays were performed in a 40ul reaction containing 30ul KAB supplemented with 5uM ATP and 5uCi γ-

32P ATP and either recombinant Histone H3, or GST-Raf-1 or HA-MEK

immunoprecipitated and eluted from HEK293 cells. Reactions were carried out for 30min at

58 30°C and stopped with SDS-loading buffer. Reactions were separated by SDS-PAGE and

detected by autoradiography.

esiRNA Preparation

esiRNA was produced as described in Kittler et al. 2005 (192). Briefly, using Flag-NEK10 cDNA as a template, 1kb fragments of NEK10 were PCR amplified with primers containing

T7 promoter sequences. The PCR primer sequences for esiNEK10#2 were: sense: 5’- cgtaatacgactcactatagggacttgaagctcctctgg-3’, antisense: 5’- cgtaatacgactcactatagggggatgataaagctgct-3’ for esiNEK10#3 were: sense: 5’-cgtaatacgactcactatagggtatgcaattttgg-3’, antisense: 5’- cgtaatacgactcactatagggtgtgtgcgtcttcgttc-3’. As a control esiRNA was generated against

dsRed.

PCR product was used as a template for in vitro transcription reaction using MEGAscript kit

(Ambion). Resulting RNA was denatured, annealed and digested with GST-RNaseIII

(provided by L.Pelletier) at 26°C for 5 h, followed by 37°C for 1 h. Digested RNA was

purified using Q-Sepharose.

Semi-Quantitative PCR

For RNA preparation, cells were harvested, and total RNA extracted using RNeasy Mini kit

(Qiagen). The first-strand cDNA was prepared with 0.5ug of RNA using qScript cDNA

SuperMix (Quanta Biosciences). In preliminary experiments, the relative amounts of cDNA

and the range of PCR cycles that permit the linear amplification of NEK10 and ß-Actin were

determined. The PCR primer sequences for NEK10 were: sense: 5’-

59 ATGAGGGATCCATGTTATCAGGAAATAC-3’, antisense: 5’-

TGGGGCTCTGCACAAAGTA-3’, and those for ß-Actin were: sense 5’-

GCCAACCGCGAGAAGATGACC-3’, antisense: 5’-

CTCCTTAATGTCACGCACGATTTC-3’

The PCR conditions for NEK10 were 94°C for 2 min, followed by 30 cycles of 94°C for 30s,

55°C for 30s and 72°C for 1 min. PCR conditions for ß-Actin were the same except the

amount of cDNA used as template was 10 times lower than for NEK10. Using ImageJ, the

relative expression of the NEK10 mRNA was evaluated by calculating the band intensity

ratio of NEK10/ ß-Actin.

FACS analysis

Cell cycle distributions were determined by the flow cytometric analysis of propidium

iodide–labeled cells. Cells were collected, fixed in 70% ethanol, and stored at –20°C overnight. They were then washed with PBS and incubated with RNaseA and propidium iodide. For quantification of mitotic cells, cells were collected, fixed in 70% ethanol, and stored at –20°C overnight. They were washed with PBS and incubated with α-pS10 H3

antibody (1:1000) (provided by P.Cheung) for 1 hr at room temperature. Cells were washed

three times and incubated with α-mouse Alexa Fluor 488 (1:200) (Molecular Probes) for 30

min at room temperature. Cells were washed one time and incubated with RNaseA and

propidium iodide. Cell cycle analysis was done by FACScan flow cytometry (BD

Biosciences, San Jose, CA).

Cloning of nekl-4 RNAi feeding vector

60 nekl-4 sequences corresponding to nucleotides #1-1000 (nekl4 RNAi #1) and #1019-2050

(nekl4 RNAi#2) were amplified from C.elegans cDNA, cloned into an RNAi vector L4440

(provided by W.B. Derry) and sequenced.

RNAi by Feeding in C.elegans

Gene expression was inhibited using the feeding method (400). Bacteria expressing double- stranded RNA (dsRNA) to nekl-4 were grown on nematode growth media (NGM) plates

supplemented with 25ug/ml Carbenicillin and 2.5 mM isopropyl--d-thiogalactopyranoside to induce dsRNA production. Bacteria expressing an empty RNAi vector (L4440) were used as

controls.

61 2.4 Results

2.4.1 Cloning and characterization of NEK10, a novel NimA kinase

A cDNA for NEK10, a previously uncharacterized member of the NEK family, encoding an 1125aa protein was isolated by PCR from a human skeletal muscle cDNA library. Amino acid sequence alignment within the kinase domain revealed that NEK10 is the most divergent member of the NEK family sharing 54% homology within this region with its

closest mammalian relatives NEK6 and 7. Unlike other NEK kinases that feature N-terminal

kinase domains, NEK10 contains a centrally located catalytic domain (Fig 2.1A). In addition,

NEK10 contains 4 N-terminal armadillo repeats, coiled-coiled repeats flanking the catalytic

domain, and a putative PEST sequence near the C-terminus. Judged by RT-PCR, NEK10

appears to be ubiquitously expressed at low levels in most mouse tissues with prominent

expression in the mammary gland, lung, spleen and kidney (Fig 2.1B).

2.4.2 NEK10 is a stimulus-specific modulator of ERK1/2

In attempting to define a cellular role for NEK10 we observed that ectopic expression of NEK10 in HEK293 cells resulted in an increase in activation-specific phosphorylation of

ERK1/2 (Fig 2.2A). Curiously, under conditions of lower NEK10 expression, which did not significantly enhance ERK1/2 activation in non-stimulated cells, NEK10 expression did not affect ERK1/2 phosphorylation elicited by mitogenic stimuli, such as the epidermal growth factor (EGF) (Fig 2.2B) but did enhance ERK1/2 phosphorylation in response to UV

62 63 64 irradiation (Fig 2.2B). The effect of NEK10 expression on ERK1/2 activation post-UV was dependent on MEK1/2 activity, as it was sensitive to the pretreatment of cells with the

MEK1/2 inhibitor U0126 (Fig 2.2B). Two other MAPK sub-families, the JNK and p38 kinases, are known to be robustly activated by stress stimuli including UV irradiation

(reviewed in (343)). Nevertheless, NEK10 expression did not enhance activation-specific phosphorylation of JNK or p38 following UV irradiation (Fig 2.2B).

Mitogenic stimulation leads to ERK1/2 activation by phosphorylation within its activation loop (T202/Y204 in ERK1) by the dual specificity kinases MEK1/2, which are themselves activated by phosphorylation (S217/S221 in MEK1) by the serine-threonine kinase Raf-1 (reviewed in (343)). We investigated if NEK10 affected activation of Raf-1 and

MEK1/2. Judged by activation-specific MEK1 phosphorylation, cells ectopically expressing

NEK10 displayed enhanced activation of MEK1 following UV irradiation (Fig 2.2C).

Significantly, although Raf-1 was activated following UV irradiation, as measured by S338 phosphorylation, NEK10 did not enhance Raf-1 activation (Fig 2.2D).

To examine the function of endogenous NEK10 in human cells, esiRNA against two distinct regions of the human NEK10 cDNA were generated (Fig 2.3A, top panel). HEK293 cells were transfected with NEK10 esiRNA or dsRed esiRNA as control, and the efficiency of knockdown determined by RT-PCR. While both of the NEK10 esiRNA pools successfully decreased endogenous NEK10 transcript, esiNEK10#2 consistently produced greater knockdown, resulting in up to 70% decrease compared to controls (Fig 2.3B). Both NEK10 esiRNAs efficiently reduced the expression of transfected Flag-NEK10 protein, further demonstrating their specificity towards NEK10 mRNA and protein (Fig 2.3C).

65 66 The requirement for NEK10 in ERK1/2 activation was examined in HEK293 cells depleted for endogenous NEK10. While NEK10 knockdown impaired ERK1/2 activation following both high (250J/m2) and low (20J/m2) doses of UV irradiation (Fig 2.4A, 2.4C), it had no effect on ERK1/2 activation in response to EGF stimulation (Fig 2.4B, 2.4C) or on the activation of JNK or p38 following UV irradiation (Fig 2.4B). Consistent with the effect of ectopic NEK10 expression on MEK1/2 and Raf-1 phosphorylation, UV-irradiated

NEK10-depleted cells displayed impaired MEK1/2 but not Raf-1 phosphorylation (Fig 2.4B).

Thus, NEK10 appears to specifically mediate ERK1/2 activation in response to UV irradiation.

2.4.3 NEK10, Raf-1 and MEK1 form a ternary complex

To further probe the mechanism of ERK1/2 activation by NEK10, its ability to directly interact with components of the ERK1/2 signaling cascade was examined. While an interaction between co-transfected Flag-NEK10 and HA-H-Ras could not be detected, Flag-

NEK10 co-precipitated both with GST-Raf-1 from transfected HEK293 cells (Fig 2.5A) and endogenous Raf-1 (Fig 2.5B). Possible association of NEK10 with MEK1 was also tested. A weak interaction between the two proteins was significantly enhanced by co-expression with

GST-Raf-1 suggesting that Raf-1 may bridge the NEK10-MEK1 association (Fig 2.6A). To explore this further, we used a previously characterized mutant MEK1∆270-307, which lacks its proline-rich repeat and is unable to interact with Raf-1 (Fig 2.6C left) (46). Upon co- expression, NEK10 associated with wildtype, but not ∆270-307 MEK1, in a Raf-1-dependent manner, further demonstrating that the interaction between NEK10 and MEK1 requires Raf-1

67 68 69 70 binding to MEK1 (Fig 2.6C right). NEK10 does not appear to be required for association between Raf-1 and MEK1, as complex formation was unaffected by co-expression of

NEK10. Interestingly, the ternary complex containing NEK10, Raf-1 and MEK1 formed in the absence of specific cell stimulation and was not modulated upon UV irradiation, suggesting that the complex is present under cycling, non-stimulated conditions (Fig 2.6B).

To map the interaction between Raf-1 and NEK10, we designed a series of NEK10

deletion mutants and tested their ability to interact with Raf-1 and MEK1 (Fig 2.7A). All of

the generated mutants retained the ability to interact with Raf-1 and MEK1, except the

NEK10 protein composed of amino acids 1-478 (NEK10 Nt) encompassing the N-terminal

third of the protein, which includes the armadillo repeats (Fig 2.7A). A predicted coiled-

coiled domain (aa 784-817) was common to all of the remaining NEK10 proteins, raising the

possibility that it mediates the Raf-1/MEK1/NEK10 interaction. Nevertheless, NEK10 protein lacking this region (NEK10 ΔCtΔcc) retained the ability to interact with Raf-1 and

MEK1 (Fig 2.7B), implying that NEK10 and Raf-1 interact through multiple sites.

2.4.4 UV irradiation promotes MEK1 autophosphorylation

The ability of NEK10 to enhance MEK but not Raf-1 activation following UV irradiation (Fig 2.4B), combined with the lack of any apparent effect of NEK10 on Raf-

1/MEK1 association (Fig 2.6C), prompted us to further explore the mechanism(s) by which

NEK10 impacts MEK activity. We first probed the possibility that Raf-1 or MEK were

targets of direct phosphorylation by NEK10, by performing NEK10 in vitro kinase assays

using affinity-purified GST-Raf-1 or HA-MEK1 as substrates. While Flag-NEK10

71 72 immunoprecipitated from HEK293 cells readily autophosphorylated and phosphorylated a

generic serine/threonine kinase substrate, Histone H3, it failed to phosphorylate both Raf-1

and MEK1 in vitro (Fig 2.8).

We reasoned that NEK10 might enhance MEK activation by an unidentified kinase or

phosphatase. An alternative explanation was that NEK10 promotes MEK

autophosphorylation following UV irradiation. To explore this latter possibility, we

examined the role of autophosphorylation in MEK activation following UV irradiation using

two MEK inhibitors with different modes of action. U0126 is a direct, non-ATP-competitive

inhibitor of MEK catalytic activity, whereas PD98059 prevents phosphorylation of MEK by

upstream activators, such as Raf-1, by binding and sequestering inactive forms of MEK (9,

100). In response to EGF stimulation, MEK phosphorylation was inhibited by pretreatment

with PD98059, but was unaffected by incubation with U0126 (Fig 2.9A). Interestingly, both

MEK inhibitors impaired MEK phosphorylation following UV irradiation (Fig 2.9A). This result indicates that MEK activation proceeds by distinct mechanisms following UV irradiation and EGF stimulation. In particular, UV-induced MEK activation requires MEK catalytic activity.

Consistent with this, both wildtype MEK1 and MEK1 KD were comparably phosphorylated at S217/S222 upon EGF stimulation, whereas phosphorylation of MEK1 KD at these sites following UV irradiation was reduced (Fig 2.9B). We next assessed the effect of NEK10 on MEK1 phosphorylation. While NEK10 expression did not affect EGF- stimulated MEK1 phosphorylation, it enhanced UV-induced phosphorylation of wildtype but not catalytically inactive MEK1 (MEK1 KD) (Fig 2.9B). Significantly, NEK10-induced

MEK1 phosphorylation was also sensitive to U0126 (Fig 2.9C). Taken together, these data

73 74 75 reveal a distinct mechanism of MEK activation following UV irradiation that requires MEK

autocatalytic activity. NEK10 specifically enhances MEK autoactivation in response to UV, but does not promote MEK activation in response to mitogens.

2.4.5 NEK10 promotes MEK1 autophosphorylation by a mechanism involving Raf-1

MEK1 autoactivation has been implicated as a non-canonical means of stimulating

ERK1/2 signaling, specifically in the context of cell attachment (303). In rat embryo

fibroblasts, Pak1 phosphorylation of MEK1 at S298 led to MEK1 autophosphorylation at

S217/S222 (303). We examined the requirement for S298 phosphorylation in UV-induced

MEK autophosphorylation using a catalytically inactive, dominant-negative mutant of Pak1

(Pak1 KD). While MEK1 phosphorylation at S298 was abolished by expression of Pak1 KD, there was no inhibition of S217/S222 phosphorylation by Pak1 KD following UV irradiation, indicating that MEK autoactivation in this context does not require S298 phosphorylation

(Fig 2.10A).

Autoactivation of MEK upon UV irradiation raised the possibility that Raf-1 may be dispensable for MEK activation in response to this stimulus. However, MEK1∆270-307,

which does not bind to Raf-1 (Fig 2.6C), was not phosphorylated upon UV irradiation or the

concomitant NEK10 overexpression (Fig 2.10B), indicating that Raf-1 is required for MEK

activation following UV irradiation. Interestingly, Raf-1 catalytic activity was not necessary

for this effect, as ectopic expression of catalytically inactive Raf-1 (Raf-1 KD) enhanced UV-

induced MEK1 phosphorylation (Fig 2.10C). Significantly, and consistent with a distinct

mechanism of MEK activation in response to mitogenic stimuli, Raf-1 KD appeared to act in

76 77 a dominant-negative manner and reduced MEK1 activation in response to EGF (Fig 2.10C).

Taken together, these data suggest that Raf-1 plays a non-catalytic role in MEK autoactivation following UV irradiation.

2.4.6 NEK10 participates in the maintenance of the G2/M checkpoint following UV irradiation

In response to genotoxic stress, such as ionizing radiation and the intercalating agent etoposide, ERK1/2 activity is required for appropriate checkpoint function, and inhibition of

ERK1/2 by pretreatment with MEK1/2 inhibitors or expression of dominant negative MEK1 led to a reduction in G2/M arrest (390, 445). While ERK1/2 activation by UV irradiation has been well documented, a role for ERK1/2 signaling in regulating the G2/M checkpoint following UV irradiation has not been established (264, 317, 325). We explored the cell cycle distribution of HEK293 cells following UV irradiation by propidium iodide (PI) staining and flow cytometry. Twenty-four hours after UV-irradiation, a 25% increase in the

G2/M population could be detected, indicative of a G2/M checkpoint arrest (Fig 2.11).

Pretreatment of cells with the MEK1/2 inhibitor U0126 markedly attenuated the G2/M arrest in HEK293 cells, indicating that ERK1/2 is required for the response to UV irradiation (Fig

2.11).

NEK10’s ability to promote UV-stimulated ERK1/2 activation prompted us to further investigate its function in the UV response. Cell cycle distribution of HEK293 cells depleted of NEK10 by esiRNA was profiled under basal conditions or following UV-irradiation.

NEK10 depletion did not affect the cell cycle profile of unsynchronized cycling HEK293

78 79 cells, indicating that NEK10 is not essential for normal cell cycle progression in this system

(Fig 2.12A, 2.12B). Nevertheless, following UV irradiation, cells depleted for NEK10 by two independent esiRNA pools, consistently displayed a decreased G2/M population compared to control cells (Fig 2.12A, 2.12B). While this is indicative of an impaired G2/M checkpoint, a similar result would be also observed if NEK10-depleted cells were arresting earlier in the cell cycle at the G1/S transition. To distinguish between these possibilities, we measured the proportion of mitotic, cycling UV-irradiated cells by staining for phospho-Serine 10 Histone

H3. In control cells, UV irradiation led to a greater than 3-fold reduction in the proportion of mitotic cells, whereas in NEK10-depleted cells only a modest decrease in the mitotic cell population could be detected (Fig 2.12C). Based on the fact that cells transfected with esiNEK10#2, which displays greater potency in NEK10 knockdown (Fig 2.3B), consistently displayed a more severe G2/M arrest defect compared to cells transfected with esiNEK10#3

(Fig 2.12A, 2.12C), NEK10 may exert a dose-dependent effect on the UV-induced G2/M arrest.

2.4.7 NEK10 has a conserved role in the UV response

To further explore the physiological function of NEK10 in the UV response we exploited C.elegans as a model organism. There are 4 NEK-like kinases in C.elegans, nekl-1,

-2, -3 and -4, which correspond to mammalian NEK9, 8, 7 and 10 respectively. NEK10’s

closest homologue is the C.elegans protein nekl-4 (previously designated D1044.8), which

also contains armadillo repeats and a centrally located kinase domain. Remarkably, NEK10

and nekl-4 share 72% homology within the kinase domain and 34% overall homology (Fig

80 81 2.13A). By comparison, NEK10 shares only 54% homology within the kinase domain with its closest mammalian homologues NEK6 and NEK7.

The role of nekl-4 in the UV response was explored by measuring the radiation sensitivity of nekl-4-depleted embryos. N2 strain worms (F0) were fed on bacteria containing two independent RNAi against nekl-4 or empty vector as control, and the F1 adults were used in subsequent experiments. Consistent with several large-scale RNAi screens (132, 180, 245,

374), no obvious defects in worm viability, development or progeny survival were detected upon nekl-4 knockdown using our RNAi constructs.

C.elegans embryos display differential sensitivity to UV irradiation at various stages of development (145). We therefore irradiated embryos at three stages: pachytene nuclei stage, fertilization/early embryogenesis and gastrulation/early morphogenesis and determined the survival rate of control or nekl-4-depleted embryos by measuring the rate of egg hatching. Twenty-four hours after laying, non-irradiated embryos, both control and nekl-4- depleted, displayed close to a 100% survival as measured by hatching rate (Fig 2.13B).

Following UV irradiation of embryos in either pachytene nuclei stage or fertilization/early embryogenesis, no significant differences in the hatching rates could be detected between control and nekl-4-depleted eggs (Fig 2.13C). However, when embryos were UV irradiated during gastrulation/early morphogenesis, an approximately 30% decrease in survival of nekl-

4-depleted embryos compared to controls was observed (Fig 2.13D). Thus, nekl-4 appears to mediate radiation sensitivity in C.elegans embryos, further supporting the notion that NEK10 has a conserved role in the UV response.

82 83 2.5 Discussion

In this study we demonstrate a function for a previously uncharacterized member of

the NEK family, NEK10, in promoting MEK/ERK activation and G2/M arrest in response to

UV irradiation. Our results indicate that NEK10 is a stimulus-specific modulator of ERK1/2 signaling, as its expression enhanced MEK and ERK activation in response to UV irradiation, but not to mitogenic stimuli such as EGF (Fig 2.2B). Consistent with this, NEK10 depletion led to impaired MEK/ERK activation in response to UV irradiation but not to EGF stimulation (Fig 2.4B, 2.4C). Our results indicate that the specificity of the ERK signaling response to UV can be attributed to the ability of NEK10 to promote a non-canonical mechanism of MEK activation following UV irradiation. Interestingly, this mechanism requires MEK catalytic activity (Fig 2.9A, 2.9B) and Raf-1 binding (Fig2.10B), but not Raf-1 kinase activity (Fig 2.10C).

2.5.1 NEK10 promotes MEK autoactivation

ERK1/2 has traditionally been associated with mitogenic stimulation and regulation of cell proliferation, but is also activated by a diverse range of other stimuli, including cytokines and various stresses, such as UV and IR, as well as hypoxia (269, 325, 390). One of the means of achieving control in ERK1/2 signaling, is the differential interaction of

ERK1/2 cascade components with various scaffolding proteins (reviewed in (195)). These interactions have been found to modulate the stimulus specificity, amplitude and duration of

84 pathway activation, as well as to impact subcellular localization, access to substrates and

cellular outcome (45).

NEK10 promoted MEK autoactivation in response to UV, but not following EGF

stimulation (Fig 2.9C), suggesting a distinct mechanism of MEK activation from the

canonical Ras/Raf/MEK cascade. This is reminiscent of the ability of p38α to undergo

MAPKK-independent autophosphorylation in response to specific stimuli. In HEK293 cells,

p38α autoactivation occurred following stimulation with TNF or peroxynitrite, but not

anisomycin or sorbitol (122). Significantly, autophosphorylation was stimulated by binding

to TAB1 (transforming growth factor-β-activated protein kinase 1 (TAK1)-binding protein 1)

(122).

MEK autoactivation has been previously described in the context of cell adhesion,

where it was induced by Pak1-mediated S298 phosphorylation (303). Significantly, UV-

induced MEK autophosphorylation is independent of S298 phosphorylation, indicating that it

proceeds by a distinct mechanism (Fig 2.10A). Our data are consistent with the formation of

a MEK autoactivation-competent complex consisting of NEK10, Raf-1 and MEK.

Interestingly, while NEK10 association with Raf-1 and MEK1 was needed for enhanced

MEK1 activation in response to UV (Fig 2.10B), the formation of the ternary complex was not sensitive to UV (Fig 2.6B). Further, Raf-1 protein, but not its catalytic activity, was required for MEK1 activation following UV (Fig 2.10C). The ability of Raf-1 to bridge the association between NEK10 and MEK (Fig 2.6A) points towards a scaffolding, non-catalytic role for Raf-1 in the formation of the MEK autoactivation complex.

The fact that signaling by the NEK10/Raf/MEK module depends on MEK

autocatalysis, points towards the existence of distinct spatial and temporal controls for MEK

85 activation in response to specific stimuli. In the case of UV, NEK10 participation in this complex may promote association with additional regulators, facilitate a permissive change in MEK conformation leading to its autoactivation and/or govern access to specific substrates. Uncoupling of MEK activation from RTKs, Ras, and their other effectors allows for discrete control of MEK/ERK signaling targets leading to specific outcomes, such as engagement of the G2/M checkpoint in response to UV (Fig 2.11, 2.12). Whether MEK activation is dependent on autocatalysis and NEK10 following other non-mitogenic stimuli, such as genotoxic agents, cellular stressors or cytokines, remains to be determined.

2.5.2 NEK10’s role in checkpoint control

ERK signaling has previously been implicated in regulation of cell cycle checkpoints in response to ionizing radiation and etoposide in MCF-7 and NIH3T3 cells, respectively

(390, 445, 446). Our results revealed that ERK1/2 activation was required for engagement of the G2/M checkpoint following UV-irradiation of HEK293 cells (Fig 2.11). A similar phenotype was also observed in UV-irradiated cells depleted for NEK10 by esiRNA (Fig

2.12B). Interestingly, we did not observe any impact of NEK10 knockdown on normal cell cycle progression, indicating that unlike it’s closest homologues NEK6/7, NEK10 is not required for mitotic progression (Fig 2.12B) (188, 452, 453). Taken together, our results suggest that NEK10-mediated ERK1/2 activation participates in the UV-induced G2/M arrest. To further define NEK10’s role in the response to genotoxic stress, future studies may explore the ability of NEK10 to impact checkpoints at earlier stages of the cell cycle, both in response to genotoxic agents, or endogenous replication stress. Furthermore, the possibility

86 that NEK10 is required for appropriate DNA repair in addition to checkpoint control, akin to

other kinases including ATR and NEK1 should be pursued (50)(reviewed in (106)).

2.5.3 Characterization of C.elegans nekl-4 during the UV response

Examination of the C.elegans NEK10 homolog, nekl-4, further supports the notion

that NEK10 kinases have a conserved function in the UV response. Namely, C.elegans

embryos depleted for nekl-4 displayed increased lethality in response to UV irradiation, as

measured by the rate of egg hatching (Fig 2.13D). This phenotype is reminiscent of the

increased radiation sensitivity observed in worms depleted for genes involved in checkpoint

functions and DNA repair, including the xeroderma pigmentosum complementation group A

(XPA, formerly known as rad-3), suppressor of mek-null (smk-1, formerly known as rad-2)

and C.elegans Fanconi anemia complementation group D2 (FANCD2) (217) (145).

Interestingly, the hatching defect was only apparent when nekl-4 depleted embryos were

irradiated during gastrulation/early morphogenesis (0-5h post laying) when the embryo is

undergoing the majority of its cell divisions, but not during earlier stages of development

(Fig 2.13C, 2.13D) (10). This is similar to the phenotypes of mutants of rad-2 and xpa, which are involved in nucleotide excision repair, who displayed increased radiation sensitivity as they passed through embryogenesis (13, 145).

The original C.elegans radiation sensitive (rad) mutants were identified using embryonic lethality as a measure of UV sensitivity, and some of them were later shown to

play roles in checkpoint signaling and DNA repair (146, 217). However, such experiments do

not directly test checkpoint function, as C.elegans embryos do not undergo cell cycle arrest

87 (178). Therefore, as an extension of our work, the role of nekl-4 in cells in the adult germline, which undergo cell death or cell cycle arrest following DNA damage, can be investigated (119). These responses are spatially separated within the gonad and produce morphological changes, which can be monitored microscopically (118). In the mitotic region of the gonad, UV-irradiated germ cells undergo cell cycle arrest, marked by a decrease in mitotic figures and reduced density of enlarged nuclei. In contrast, cells within the pachytene region of the gonad undergo apoptosis following irradiation. This can be measured by scoring the number of germ cell corpses in the germline of live animals by Nomarski microscopy. Compared to wildtype worms, checkpoint mutants exhibit decreased cell death, whereas worms defective in DNA repair have higher levels of cell death (118).

Our work in mammalian cells demonstrates an interaction between NEK10 and

Raf/MEK/ERK, which may contribute to checkpoint activation. Suggestive of a conserved role for MEK/ERK in checkpoint function in lower eukaryotic systems, Drosophila embryos lacking MEK exhibit faster cell division and fail to arrest following ionizing radiation, indicating an involvement in both intrinsic and extrinsic checkpoint control (271).

Interestingly this function was independent of Raf (271). In C.elegans, mutations in lin-

45/mek-2/mpk-1 (Raf/MEK/ERK) are not associated with increased sensitivity to genotoxic stress, however strong phenotypes of null mutations, including sterility and lethality, may preclude identification of a milder phenotype. Therefore, the radiosensitivity of mutants containing alleles that weakly inactivate ERK1/2 signaling should be studied either alone or in combination with nekl-4 knockdown.

88 CHAPTER 3

CHARACTERIZATION OF NEK10 REGULATION BY PHOSPHORYLATION

89 3.1 Abstract

Phosphorylation is a primary means of transmitting information from extracellular stimuli to defined cellular outputs. The activity of protein kinases is tightly controlled by multiple mechanisms, including post-translational modification and . In search of regulatory inputs into NEK10, using mass spectrometry, our laboratory identified nineteen distinct sites of NEK10 phosphorylation. While the majority of the detected phosphorylation events depended on NEK10 catalytic activity, likely due to prominent autophosphorylation, several phospho-sites were also found on the catalytically inactive

NEK10, indicative of phosphorylation by other kinases. In addition to the NEK10 activation loop phosphorylation sites, which were obligate for catalytic activity, a C-terminal NEK10 phosphorylation site, predicted to be a target of an upstream kinase(s) was characterized.

This phospho-site was essential for NEK10 cytoplasmic to nuclear translocation following

UV irradiation. Moreover, phosphorylation of this site facilitated interaction of NEK10 with

14-3-3, revealing a possible role in control of NEK10 subcellular localization. Taken together, this work defines a role for phosphorylation in NEK10 regulation.

90 3.2 Introduction

Protein kinases constitute one of the most common protein domains in the eukaryotic

proteome, representing 2% of all encoded proteins, and participate in control of most aspects of cell physiology and function (345). Consistent with their potent biological activity, kinase catalytic activity is strictly controlled, temporally and spatially.

The kinase domain is a well-conserved structure composed of two lobes, a smaller N- terminal lobe and a larger C-terminal lobe (reviewed in (163)). The N lobe is composed of a five-stranded β-sheet and one α-helix (αC) whereas the C lobe is predominantly α-helical.

The C lobe bears the catalytic loop, which contains residues that are directly involved in catalysis of phosphate transfer from ATP to a serine, threonine or tyrosine residue on the substrate protein. ATP binds in the cleft between the N and C lobes and is coordinated by interactions with the glycine-rich phosphate-binding (P) loop. The kinase substrate is correctly positioned by the centrally located activation loop, which is phosphorylated in most active kinases. These motifs (αC helix and catalytic, P and activation loops) have to be appropriately spatially coordinated for a kinase to achieve full activity. The transition from an inactive to an active kinase conformation is controlled by several mechanisms, including phosphorylation and allosteric regulation.

Activation loop phosphorylation is the most common means of regulating kinase activity (reviewed in (287)). In kinases that require activation loop phosphorylation for activation, the unphosphorylated activation loop collapses into the catalytic core, blocking the interaction of the kinase with both ATP and substrate (reviewed in (163)). Following phosphorylation, the activation loop undergoes a large conformational change, moving away

91 from the catalytic core, allowing the kinase domain to bind and catalyze phosphate transfer to

substrates. Certain kinases require additional phosphorylation events to achieve full activity.

For example, while PKB is partially activated by activation loop phosphorylation (T308 in

PKB1), phosphorylation within the hydrophobic motif (S473 in PKB1), located outside of

the kinase domain is required to stabilize the active conformation (8, 108, 447).

Allosteric regulation of kinases can influence catalytic activity either alone or in

combination with phosphorylation events. A well-studied example of this is the relationship

between cyclins and Cdks, in which cyclin-binding induces an active conformation within

the αC helix of a Cdk (174). The conformation of the αC helix is directly related to the ability of a kinase to bind ATP. This is mediated by the formation of an ion pair between a conserved glutamate (E51 in Cdk2) within the αC helix and the ATP-coordinating lysine

(K33 in Cdk2). In the absence of cyclin-binding, the E51-K33 interaction is disrupted in

Cdks, preventing nucleotide binding and catalysis (174). Certain Cdks, such as Cdk2, also require activation loop phosphorylation for full activity, whereas others, including Cdk5 and

Cdk6, can be activated by cyclin binding alone (349, 357, 393). Kinase activity can also be modulated by autoinhibitory mechanisms. For example, the presence of autoinhibitory or pseudosubstrate sequences in the non-catalytic regions of proteins, such as in Src and calcium/-dependent protein kinase (CaMK) kinases respectively, limits kinase

activity until activating stimuli cause a change in conformation that relieves the inhibition

(130, 270, 366).

Protein phosphorylation represents a major mediator of signal transduction.

Phosphorylation can influence protein conformation and enzymatic activity, as well as create binding sites for domains that bind phospho-residues. For example, SH2 domains, FHA

92 domains and 14-3-3 proteins target motifs that contain phospho-tyrosine, phospho-threonine

and phospho-serine/phospho-threonine residues, respectively (89, 194, 283). Thus,

phosphorylation can induce protein-protein interaction and in the case of modular proteins,

nucleate multiprotein complexes. Coupling a dynamic event such as

phosphorylation/dephosphorylation with protein interaction provides flexibility and increases

the efficiency with which information can be transmitted through the cell.

Our previous work described a role for NEK10 in mediating the cellular UV

response. NEK10 was found to participate in activation of MEK/ERK and the establishment

of the G2/M checkpoint following UV irradiation. The mechanism of NEK10 regulation

following irradiation remains unknown. To address this, the role of phosphorylation in

regulation of NEK10 activity, protein interactions and subcellular localization was explored.

3.3 Materials and Methods

For additional materials and methods see Chapter 2.3.

Plasmids

myc-14-3-3η was provided by R.Rottapel. For list of additional plasmids see Chapter 2.3.

Cell Culture and Transfection

HEK293 cells were cultured and transfected as in Chapter 2.3. HeLa cells were transfected with Effectene (Qiagen) according to manufacturer’s instructions.

93 Cell Lysis, Immunoprecipitation and Kinase Assays

Cell lysis, immunoprecipitation and kinase assays were performed as in Chapter 2.3.

Western Blot Analysis and Antibodies

Whole cell lysates or immunoprecipitates were resolved by SDS–PAGE and blotted to PVDF membranes. Proteins were probed using appropriate primary antibodies from the following sources: α-GFP (sc-8334) and α-pTyr (PY99) (sc-7020) from Santa Cruz, α-Flag M2-FITC from Sigma and α-HA (12CA5) and α-Myc (9E10) were harvested from supernatants of the corresponding hybridoma. Phosphospecific antibodies against NEK10 (α-pNEK10), phosphorylated on S684 and S688 were raised by immunizing rabbits with the peptide

CEQENpSKLTpSVVG linked to keyhole limpet hemocyanin (KLH) (Antibodies Inc.).

Phosphospecific antibodies were purified from rabbit serum by two-step affinity chromatography on a sulfolink coupling gel column on sulfolink coupling gel column

(Pierce) with covalently coupled dephosphopeptide (first step) and phosphopeptide (second step) immunogen. For a list of additional antibodies see Chapter 2.3.

[32P]orthophosphate labeling in cells

Transfected HEK293 cells were incubated overnight with 1mCi of [32P]orthophosphate in

phosphate-free DMEM + 10% dialyzed/heat inactivated FBS. After treatments, cells were

washed twice with ice cold PBS(-/-) and lysed with CHAPS lysis buffer. After lysates were

clarified, Flag-NEK10 was immunoprecipitated with Anti-Flag M2-Agarose.

Immunoprecipitates were washed four times with lysis buffer. Immunoprecipitates were

separated by SDS-PAGE, transferred to PVDF membrane and exposed by autoradiography.

94 Phosphoamino acid Analysis

Immunoprecipitates of radiolabelled proteins (from orthophosphate labeling in cells or in vitro kinase assays) were separated by SDS-PAGE and transferred to PVDF membrane. The protein band of interest was identified and cut out of the PVDF membrane. The piece of membrane was soaked in 0.5% polyvinyl pyrollidone (PVP-360) in 100mM acetic acid, at

37C for 30 min, washed twice in water and then incubated with 10ug of TPCK-treated trypsin in fresh 50mM ammonium bicarbonate pH8.0, at 37C overnight. The next morning reactions were spiked with 10ug of trypsin and incubated, at 37C for 3 additional hours. At the end of the reaction the supernatant was removed and reserved. The membrane was washed one time with water and the wash liquid pooled with reserved supernatant. Pooled liquid was lyophilized in a centrifugal vacuum concentrator (Speedvac). Once dried, peptides were hydrolyzed in 5.7M HCl, at 110C for 1h, and subsequently lyophilized. Samples were resuspended in pH 1.9 buffer (2.5% formic acid (88%), 7.8% glacial acetic acid), which contained 15 parts buffer to 1 part cold phosphoamino acid standards (1.0mg/ml of each phosphoserine (P-S), phosphothreonine (P-T) and phosphotyrosine (P-Y) in water).

Phosphoamino acids were separated on cellulose-TLC plates in the first dimension from cathode (-) to anode (+) at 1.5kV for 20 min in buffer pH 1.9. The plate was rotated 90 degrees counterclockwise and electrophoresed in the second dimensionn from cathode (-) to anode (+) at 1.3kV for 16 min in buffer pH 3.5 (5% glacial acetic acid, 0.5% pyridine). The location of the phosphoamino acids was mapped by ninhydrin staining, and the presence of

32P-labelled amino acids detected by autoradiography.

95 Immunofluorescence

For immunofluorescence experiments cells were grown on glass coverslips. Cells were fixed for 10 min in 2% PFA in PBS and permeabilised for 2 min in 0.2% Triton X-100 in water, at room temperature. Staining was performed at 37C. Samples were blocked for 30 min in blocking buffer (1.5% BSA/1.5% goat serum in PBS), and incubated for 1h with fitc- conjugated antibodies in blocking buffer and washed. Samples were mounted with

Mowiol/DAPI and analysed by inverted fluorescence microscopy (Zeiss Axiovert 200M).

96 3.4 Results

3.4.1 Biochemical characterization of NEK10 catalytic activity

To gain insight into NEK10 catalytic properties we developed a NEK10 in vitro kinase assay. The assay conditions were based on those previously established for NEK10’s closest mammalian relatives NEK6 and NEK7 (23). Flag-NEK10 immunoprecipitated from transiently transfected HEK293 cells was used as the kinase source. To generate a catalytically-inactive kinase (NEK10 KD), the conserved ATP-binding lysine (K548) was mutated to arginine (R). NEK10 immunopurified from exponentially growing cells was

catalytically active, as measured by autophosphorylation, as well as phosphorylation of

various histones, including 1, 2 and 3. NEK10 however failed to phosphorylate β-casein,

another generic kinase substrate (Fig 3.1A).

Phosphoamino acid analysis of NEK10, autophosphorylated in vitro, revealed

phosphorylated serine, threonine and tyrosine residues, with the majority of phosphate

incorporated into serine (Fig 3.1B). Based on sequence homology, NEK10 is predicted to be

a serine-threonine kinase. Surprisingly, its ability to autophosphorylate on tyrosine suggested

that it might possess dual-specificity activity, at least towards self. Consistent with this,

immunoprecipitates of wildtype NEK10 contained a single band, which corresponded to

NEK10, which was reactive with the phospho-tyrosine specific antibody (PY99: α-pTyr).

However, catalytically inactive NEK10 was not detected with this antibody indicating that

catalytic activity is required for NEK10 tyrosine phosphorylation in vivo (Fig 3.1C).

Furthermore, the phospho-tyrosine signal was increased when wildtype NEK10

97 98 immunoprecipitates were subjected to an in vitro kinase assay (Fig 3.1C), further supporting

the notion that NEK10 possesses dual serine/threonine and tyrosine kinase activity.

3.4.2 NEK10 is a phosphoprotein

To examine if phosphorylation was required for NEK10 activity, kinase assays were performed on NEK10 immunoprecipitates that had been pretreated with calf intestinal phosphatase (CIP) predicted to robustly dephosphorylate NEK10. CIP treatment abolished

NEK10 catalytic activity illustrating the importance of phosphorylation for NEK10 kinase activity (Fig 3.2A).

To explore NEK10 phosphorylation status in vivo, Flag-NEK10 was immunoprecipitated from transiently transfected HEK293 cells metabolically labeled with

[32P]orthophosphate. While both wildtype and NEK10 KD were phosphorylated in vivo, wildtype NEK10 displayed significantly higher levels of phosphate incorporation. This implies that the majority of NEK10 phosphorylation may depend on its autocatalytic activity

(Fig 3.2B upper). Interestingly, phosphoamino acid analysis of orthophosphate labeled

NEK10 showed that NEK10 was phosphorylated on serine and tyrosine but not threonine in vivo (Fig 3.2B lower).

To comprehensively identify sites of NEK10 phosphorylation, liquid chromatography tandem mass spectrometry (LC/MS/MS) was performed on wild-type and catalytically inactive Flag-NEK10 immunopurified from HEK293 cells (Protana). Mass spectrometric data were analyzed with the bioinformatics database system Mascot (Matrix Science,

London) and every phosphorylation site identified was confirmed manually. Wildtype

99 100 NEK10 displayed eighteen distinct sites of phosphorylation (Fig 3.2C). Of these, four were also found on NEK10 KD suggesting that the majority of NEK10 phospho-sites arise either as a result of autophosphorylation or are dependent on NEK10 catalytic activity.

Interestingly, one additional site of phosphorylation, S846, appeared to be unique to NEK10

KD (Fig 3.2C).

Consistent with previous results (Fig 3.2B), wildtype NEK10 was phosphorylated on

serine and tyrosine but not threonine, whereas NEK10 KD was exclusively phosphorylated

on serine (Fig 3.2C). Using Scansite (http://scansite.mit.edu/), which searches for motifs that

are likely to be phosphorylated by specific kinases or bind to domains such as SH2, 14-3-3 or

PDZ domains, we analysed NEK10 protein sequence to determine if any of the identified phosphorylation sites were putative targets of known protein kinases or part of phospho- binding motifs. pS933 was predicted to be within a 14-3-3 binding sequence, as well as a target of CaMK2G and PKB kinases, while pS352 and pS797 displayed limited homology to sites targeted by (CK2) and ATM, respectively. None of the other sites

displayed significant homology to the established kinase consensus sequences.

3.4.3 NEK10 requires activation loop phosphorylation for full catalytic activity

Two phosphorylation sites identified exclusively on wildtype but not KD NEK10,

S684 and S688, are located within the activation loop of the catalytic domain.

Phosphorylation at equivalent residues in other kinases, including NEK6, is important for

their activation. NEK6 phosphorylation at S206 (equivalent to NEK10 S688) by Nek9 and to

101 a lesser degree S202 (equivalent to NEK10 S684) is required for NEK6 to achieve full

catalytic activity (24).

To explore the importance of activation loop phosphorylation for NEK10 catalytic

activity, S684 and S688 were mutated to alanine (A), a non-phosphorylatable residue,

individually, or in combination. Mutation of either S684 or S688 strongly impaired NEK10

catalytic activity, although the impact was more pronounced following mutation of S688 (Fig

3.3A). Replacement of S684, but not S688, with the phospho-mimetic residue, aspartic acid

(D), restored full NEK10 catalytic activity (Fig 3.3A). Interestingly, the double mutant

S684D/S688D was also impaired for catalytic activity. These results indicate that both S684 and S688 are required for full catalytic activity. However, S688 may play a primary role in

NEK10 activation, as was the case for the equivalent residue in NEK6 (24).

The involvement of S684/S688 phosphorylation in control of NEK10 catalytic activity prompted us to develop a polyclonal antibody against a phosphorylated peptide encompassing S684 and S688 (α-pNEK10). Wildtype NEK10, immunoprecipitated from cycling HEK293 cells, was detected by α-pNEK10, but the signal was markedly reduced

when either S684 or S688 were mutated to an alanine, and completely abolished in the

S684A/S688A double mutant (Fig 3.3B). This indicates that α-pNEK10 specifically

recognizes NEK10 phosphorylated on S684 and S688.

Mass spectrometry analysis did not identify S684 and S688 as phospho-sites on

NEK10 KD. Consistently, α -pNEK10 did not react with Flag-NEK10 KD

immunoprecipitated from HEK293 cells (Fig 3.3B). The lack of NEK10 KD activation loop phosphorylation suggested that this might occur as a result of autophosphorylation.

Activation loop autophosphorylation can occur by an intramolecular (in cis) mechanism, as

102 103 in glycogen synthase kinase 3 α/β (GSK3α/β) and dual-specificity tyrosine-phosphorylation- regulated kinase (DYRK) or by an intermolecular (in trans) mechanism, as in Chk2, NEK2 and EGF receptor (EGFR) (6, 56, 158, 237, 332). In support of NEK10 autophosphorylation occurring in trans, expression of increasing levels of wildtype, but not KD GFP-NEK10, led to phosphorylation of the activation loop of co-expressed Flag-NEK10 KD in vivo (Fig

3.3C). Dimerization is a common feature of kinases that undergo intermolecular autophosphorylation (311). Indeed, HA-NEK10 co-precipitated with both wildtype and KD

Flag-NEK10, consistent with NEK10 dimerization (Fig 3.3D). Taken together, our results are consistent with a mechanism of activation loop phosphorylation that involves NEK10 intermolecular autophosphorylation.

3.4.4 NEK10 localization, but not catalytic activity, is modulated by UV irradiation

Considering that my previous work has implicated NEK10 in mediating the UV response, I was interested in examining how NEK10 may be regulated by UV irradiation.

Significantly, certain NEK family members exhibit sensitivity to a variety of genotoxic stresses. In response to stimuli such as IR, UV and etoposide, NEK1, NEK2, NEK6 and

NEK11 catalytic activity is modulated (218, 266, 286, 313). For instance, IR activates NEK1 and NEK11, whereas NEK2 and NEK6 are inhibited under this condition (218, 260, 266,

313). In addition, following IR, NEK1 relocalizes from the cytoplasm to nuclear foci within

10 minutes of IR treatment, and this persists for up to 8 hours (313).

We first examined if UV irradiation affected NEK10 catalytic activity. Flag-NEK10 was immunoprecipitated from UV-irradiated HEK293 cells and its kinase activity assayed

104 using Histone H3 as substrate. At both early (30 minutes) and late (3 hours) time points

following UV irradiation, NEK10 activity was not modulated (Fig 3.4A). Consistently,

activation loop phosphorylation of NEK10 at S684/S688 was unaffected by UV (Fig 3.4B).

We next explored NEK10 subcellular localization following UV irradiation in HeLa cells, using indirect immunofluorescence. In cycling cells, ectopically expressed NEK10 displayed predominantly cytoplasmic localization (Fig 3.4C). However, 3h post-UV irradiation, prominent NEK10 relocalization to the nucleus was observed (Fig 3.4C).

3.4.5 NEK10 nuclear localization requires S933 phosphorylation

During the characterization of NEK10 localization, the subcellular distribution of a phosphorylation mutant of NEK10, NEK10 S933A, was examined. S933 was identified by mass spectrometry as being phosphorylated on both wildtype and kinase dead NEK10.

Significantly, NEK10 S933A, displayed impaired nuclear localization following UV

irradiation when compared to wildtype NEK10 (Fig 3.5). Even in the absence of UV, NEK10

S933A exhibited reduced nuclear localization compared to wildtype protein, suggesting that in cycling cells NEK10 may constantly shuttle between the cytoplasm and the nucleus.

3.4.6 S933 is a 14-3-3 binding site

NEK10 S933 is located within a consensus binding site for 14-3-3 proteins, which are known to regulate localization, stability and activity of their target proteins (reviewed in

(413)). Consistent with this, Flag-NEK10 and myc-14-3-3η co-precipitated from transfected

105 106 107 HEK293 cells (Fig 3.6A). Intriguingly, this interaction was not modulated by UV irradiation.

Moreover, NEK10 S933A retained the ability to interact with 14-3-3 (Fig 3.6A).

In addition to S933, NEK10 contains other predicted 14-3-3 binding sites and may engage in multi-site interactions with 14-3-3. To probe for this, a truncated NEK10 protein, composed of the C-terminal 339aa of NEK10 (Flag NEK10 Ct), and including S933, was utilized. Similar to the full-length protein, Flag-NEK10 Ct co-precipitated with 14-3-3η in a

UV-independent manner. However, S933 mutation in Flag-NEK10 Ct abolished its interaction with 14-3-3η (Fig 3.6B), demonstrating that NEK10 S933 is a bona fide 14-3-3 binding site.

In addition to impacting protein subcellular localization, 14-3-3 binding has also been shown to modulate protein-protein interactions and kinase catalytic activity. Judging by the comparable activity of NEK10 S933A to wildtype NEK10, 14-3-3 binding to this site does not appear to affect the catalytic function of NEK10 (Fig 3.6C). My previous work demonstrated that NEK10 interacts with Raf-1 (Fig 2.5) and that the NEK10 C-terminus is sufficient to mediate this interaction (Fig 2.7A). Raf-1 is a well-characterized 14-3-3 binding protein and it’s interactions with a number of proteins including A20 and ζ are bridged by a 14-3-3 dimer (412, 415). Nevertheless, 14-3-3 binding to S933 was dispensable for the NEK10/Raf-1 interaction, as endogenous Raf-1 co-precipitated to equal degrees with Flag NEK10 Ct and Flag NEK10 Ct S933A (Fig 3.6D).

108 109 3.5 Discussion

Work described in this chapter examined the role of phosphorylation in regulation of

NEK10 catalytic activity and subcellular localization. This data indicates that NEK10 is fully

active in cycling cells. Furthermore, that NEK10 catalytic activity requires activation loop

phosphorylation, which likely occurs as a result of autophosphorylation (Fig 3.3.A, 3.3B).

While no effect of UV irradiation on NEK10 catalytic activity was observed (Fig 3.4A),

NEK10 underwent nuclear translocation following irradiation (Fig 3.4C). Significantly, this

was dependent on phosphorylation of S933, a residue within one of the predicted NEK10 14-

3-3 binding sites (Fig 3.5, 3.6B).

3.5.1 Characterization of NEK10 catalytic activity

NEK10 displayed robust autophosphorylation and had a preference for histones compared to β-casein as a generic substrate (Fig 3.1A). This may indicate NEK10’s preference for basic over acidic substrates. NEK kinases can be grouped according to their general substrate specificity, with NimA, NEK1, 2 and 3 exhibiting robust phosphorylation

of β-casein, and NEK6, 9 and 10 with no activity towards this substrate (112, 147, 222, 340,

389). The substrate specificity of NEK10 is consistent with its greater homology to NEK6

compared to NEK1/2/3. Analysis of NEK6 substrate specificity demonstrated that it had a

preference for an aromatic residue at the –4 and +1 position, relative to the phospho-target

site, and did not tolerate a proline at +1 (F/YXXXpS/TF/Y) (234). It would be informative

110 to determine if NEK10 shared similar substrate specificity to NEK6, using a method such as

oriented peptide library screening (373).

3.5.2 NEK10 undergoes activation loop autophosphorylation

Eighteen sites of serine and tyrosine phosphorylation were detected within wildtype

NEK10 (Fig 3.2C). Five serine phosphorylation sites were identified within NEK10 KD,

indicating that the majority of NEK10 phosphorylation is dependent on NEK10 catalytic activity. While some of these sites are undoubtedly sites of autophosphorylation, others might be targeted by unidentified upstream kinases. In such a scenario, limited NEK10 autophosphorylation may serve to prime NEK10 for additional phosphorylation events or may alter NEK10 protein conformation allowing it to interact with upstream kinases.

Two likely sites of NEK10 autophosphorylation are S684 and S688 within its

activation loop. Similar to other kinases, phosphorylation of these sites was required for

activity and mutation of these residues rendered NEK10 inactive (Fig 3.3A). Several pieces

of evidence implicate these residues as targets of intermolecular autophosphorylation. First,

S684 and S688 were identified as phosphorylation sites on wildtype, but not NEK10 KD, by

both mass spectrometry and immunoblotting with a phosphospecific antibody (Fig 3.2C,

3.3B). Second, in vivo, NEK10 expression induces activation loop phosphorylation of

NEK10 KD, indicating that phosphorylation may occur in trans (Fig 3.3.C). Finally, NEK10

was found to dimerize with both wildtype and catalytically inactive NEK10 (Fig 3.3.D), a

feature associated with kinases that autophosphorylate in trans. For example, NEK2 and

NEK9 are activated by autophosphorylation and are able to dimerize by a mechanism that

111 requires their coiled-coil motifs (reviewed (323)). In contrast, NEK6 and NEK7 do not

dimerize and are activated by NEK9-mediated activation loop phosphorylation (24). This

suggests that for NEK kinases, the ability to dimerize may be predictive of an activation mechanism by autophosphorylation.

3.5.3 NEK10 exhibits tyrosine kinase activity

An interesting feature of NEK10 phosphorylation is the apparent absence of tyrosine phosphorylation on catalytically inactive NEK10 and the ability of wildtype NEK10 to autophosphorylate on tyrosine in vitro (Fig 3.1B,C, 3.2C). Despite sequence homology to serine-threonine kinases, tyrosine autophosphorylation has been noted for certain kinases, including GSK3, DYRK and CK2 (56, 78, 236, 237). Autophosphorylation of serine- threonine kinases on tyrosine may be kinetically- and sterically-favored compared to tyrosine phosphorylation of exogenous substrates, especially when the reaction occurs in cis, as is the case for GSK3 and DYRK. Thus, to further explore NEK10 as a bona fide dual specificity , it should be systematically tested for its ability to phosphorylate exogenous substrate(s) on S/T and Y residues (224).

Of note, a member of the NEK family, NEK1, was originally identified as a dual- specificity kinase (222). NEK1 expression in bacteria was accompanied by tyrosine phosphorylation of bacterial proteins, and in vitro NEK1 displayed weak activity towards a generic acidic tyrosine kinase substrate, poly(Glu:Tyr). However, the physiological relevance of NEK1’s tyrosine kinase activity is unclear.

112 3.5.4 NEK10 subcellular localization is regulated by UV irradiation

Following up on our studies with NEK10 and the UV response, we were interested in

determining if NEK10 was regulated by UV irradiation. While we did not observe any

change in catalytic activity (Fig 3.4A) or binding to Raf-1 (Fig 2.6B), we detected a

cytoplasmic to nuclear translocation of NEK10 following UV irradiation (Fig 3.5). The physiological relevance of such behaviour is currently unknown, however, UV-dependent relocalization may serve as a means to control access of NEK10 to distinct pools of substrates, or interacting regulatory proteins in a stimulus-dependent manner. Future work

should explore if localization impacts NEK10 in checkpoint control. For example, NEK10- null MEFs, which are being generated in our laboratory, could be reconstituted with wildtype or S993A NEK10 and the fidelity of the G2/M arrest measured following UV irradiation.

Parallels can be drawn between the control of NEK10 subcellular localization and that of NEK1. The latter exhibits predominantly cytoplasmic localization in cycling cells, but translocates to the nucleus following IR, where it is found at discrete foci in the nucleus and colocalizes with γH2AX (313). NEK1 shuttling between the cytoplasm and nucleus likely depends on the presence of functional nuclear localization sequences (NLS) and nuclear export sequences (NES) (153). Based on these studies, it would be of interest to explore

NEK10 subcellular localization following other genotoxic stresses and to further examine

subnuclear localization of NEK10. Using publicly available prediction tools

(http://www.cbs.dtu.dk/services/NetNES/; http://cubic.bioc.columbia.edu/predictNLS/), 3

putative NES, but no NLS, were identified within NEK10. Examining the localization of

NEK10 NES point mutants or NEK10 truncation mutants, possibly in the context of

113 pharmacological modifiers of nuclear transport, such as leptomycin B, may identify regions of NEK10 that regulate its subcellular localization.

3.5.5 14-3-3 proteins regulate NEK10 subcellular localization

My data reveal that phosphorylation of NEK10 S933, a residue within a site of 14-3-3 binding, participates in control of NEK10 subcellular localization (Fig 3.6B). The 14-3-3 family are small, acidic, dimeric proteins that bind to phosphoserine/phosphothreonine residues within the consensus sequences: RSX(pS/sT)XP or RXFX(pS/pT)XP (442). 14-3-3 binding can have multiple consequences for the target protein. In the case of Raf-1, 14-3-3 binding influences catalytic activity and protein-complex formation. In non-stimulated cells,

14-3-3 binding to phosphorylated S259 inhibits Raf-1 by promoting an inactive conformation. In contrast, following mitogenic stimulation, 14-3-3 binding to S621 promotes

Raf-1 activation (407). This may occur through two mechanisms; stabilizing an active conformation and promoting Raf-1 dimerization (117, 348, 426). Significantly, 14-3-3 binding to S933 does not contribute to similar events in NEK10, and mutation of S933 to alanine did not have any effect on NEK10 catalytic activity or interaction with Raf-1, a known interacting partner (Fig 3.6C, 3.6D).

14-3-3 proteins also act as major modulators of subcellular localization. 14-3-3 binding is predominantly associated with cytoplasmic sequestration as in the case of

Cdc25B/C and members of the foxo family of transcription factors (38, 62). However, 14-3-3 binding to other proteins, including telomerase catalytic subunit (TERT) and Chk1, promotes nuclear localization (88, 360). It has been demonstrated that 14-3-3 binding often functions

114 by masking a subcellular targeting sequence. An example of this is TERT, where 14-3-3 binding obscures a nearby NES (360). While NEK10 does not contain any obvious NES or

NLS motifs in close proximity to S933, it is possible that cryptic localization signals exist within NEK10, and are altered by 14-3-3 binding. Alternatively, NEK10 S933 phosphorylation and subsequent 14-3-3 binding may affect its association with proteins that govern its subcellular localization.

The sequence surrounding NEK10 S933 constitutes a consensus target sequence for the AGC kinases and the calcium/calmodulin-dependent kinase (CaMK) families (Scansite).

Using a targeted approach, it may be possible to identify the S933 kinase. The ability of candidate kinases to phosphorylate NEK10 in a S933-dependent manner can be explored using recombinant kinases in in vitro kinase assays. The in vivo relevance of promising candidates may determined by examining the effect that inhibition of these kinases, by RNAi or treatment with chemical inhibitors has on UV-induced NEK10 translocation. The latter approach may be utilized to more broadly identify pathways that regulate NEK10, either dependent or independent of S933 status, using NEK10 subcellular localization as a readout.

The work described here provides the first insight into the importance of phosphorylation for regulation of NEK10. Identification of upstream regulators, as well as

NEK10 substrates and substrate recognition sequences will be instrumental in defining the molecular mechanisms of action of this kinase.

Attributions

J.Ho purified α-pNEK10 antibodies from rabbit serum. Mass spectral analysis on NEK10 was performed by Protana.

115 CHAPTER 4

FUTURE DIRECTIONS

116 The work described in Chapters 2 and 3 constitutes the first characterization of the

serine-threonine kinase NEK10. When I began work on NEK10, only a handful of

publications focused on mammalian NEK kinases. Since then, NEKs have been more

extensively studied and shown to impact cell cycle progression, ciliagenesis and the response to genotoxic stresses. The experiments described in Chapters 2 and 3 link NEK10 to G2/M

arrest following UV irradiation. Analogous to the reaction of NEK1 to IR, I found that

NEK10 translocates to the nucleus following UV irradiation. Intriguingly, NEK10 was found

to promote UV-induced ERK activation. NEK10 interacted with Raf and MEK and enhanced

MEK activity through a novel mechanism involving MEK autoactivation. Whether a

relationship between other NEKs and the ERK signaling pathway exists is a topic for future

study.

Work described in Chapter 3 provides a foundation for further biochemical

characterization of NEK10. We have demonstrated that NEK10 is a multiply phosphorylated

protein and that the majority of the phosphorylation sites, including S684 and S688 within

the activation loop, are likely targets of autophosphorylation. Of significance, five

phosphorylation sites that are independent of NEK10-catalytic activity were identified, and

represent candidate targets for unidentified upstream kinases. I have shown that one of these

sites, S933, is a 14-3-3 binding site and mediates NEK10 nuclear translocation following UV

irradiation. Further study of S933 and the other 4 non-autophosphorylation sites likely

provides the best opportunity to identify upstream regulators of NEK10.

In addition to the work presented in Chapters 2 and 3, several other lines of evidence

will inform future work on NEK10. Namely, a potential role for NEK10 in control of

117 myogenesis and cell adhesion, preliminary characterization of NEK10-interacting partners,

and possible means of NEK10 involvement in tumourigenesis will be discussed below.

4.1 Materials and Methods

Cell Culture and Transfection

HEK293 cells were cultured and transfected as in Chapter 2.3. C2C12 cells were cultured in growth media (gm) (DMEM/10% FBS). To induce differentiation, culture medium was changed to differentiation media (dm) (DMEM/2% horse serum) when cells reached confluency. For experiments involving non-adherent and re-plated cells, HEK293 cells were serum-starved overnight. To place cells in suspension, cells were detached by trypsinization

(5 min) and collected in DMEM containing 1mg/ml soybean trypsin inhibitor and incubated, with rotation, at 37C for 90 min. Cells were then maintained in suspension or re-plated onto

un-coated or fibronectin-coated plates for 20 min before lysis.

Cell Lysis, Immunoprecipitation and Kinase Assays

Cell lysis, immunoprecipitation and kinase assays were performed as in Chapter 2.3.

Western Blot Analysis and Antibodies

α-NEK10 (sc-100434) was from Santa Cruz, α-MHC ascites fluid from Developmental

Studies Hybridoma Bank (DSHB) at the University of Iowa. For list of additional antibodies

see Chapter 2.3.

118 4.2 NEK10 and myogenesis

In the course of characterizing a commercially available antibody against NEK10 (α-

NEK10), an immunoblot of mouse tissue lysates demonstrated prominent expression of

NEK10 in skeletal muscle (Fig 4.1A). Immunoprecipitated endogenous NEK10, using the

same antibody, from a mouse myoblast cell line (C2C12) cultured in growth media (gm)

displayed no autophosphorylation activity (Fig 4.1.B). Strikingly, NEK10

immunoprecipitates, from cells grown for 48-96h in differentiation media (dm), readily

autophosphorylated (Fig 4.1B). This coincides with increased NEK10 immunoreactivity in

lysates from differentiating C2C12 cells (Fig 4.1C). These results raise the possibility that

NEK10 is involved in muscle differentiation.

In cell culture, early stages of myogenesis can be studied using a myoblast cell line

such as the C2C12 cell line. In the absence of growth factor stimulation, these progenitor cells differentiate and fuse into multinucleated skeletal muscle cells called myotubes

(reviewed in (193)). Myotubes can further develop into striated muscle cells called

myofibrils, which contain the functional contractile apparatus (Fig 4.2). Commitment to

differentiation in this system requires irreversible cell cycle exit during G1. This is

accompanied by downregulation of proliferative signaling and is marked by decreased

expression of proteins that promote cell cycle progression, such as cyclinD1 and cyclinA

(421)(reviewed in (193)). There is a concurrent increase in the expression and activity of

proteins that inhibit cell cycle progression, including p27 and pRB, which also contribute to

the maintenance of the cell cycle arrest following differentiation (58, 105). Myogenesis is

under the control of the muscle specific, MyoD, family of basic helix loop helix (bHLH)

119 120 121 transcription factors, which contribute to cell cycle withdrawal and stimulate expression of

muscle-specific structural genes, such as myosin heavy chain (MHC) and α-actin

(376)(reviewed in (193, 214, 354, 388)).

The induction of NEK10 reactivity during C2C12 cell differentiation warrants exploration of the impact of its depletion on myoblast differentiation, as well as the status of abovementioned muscle-specific factors in NEK10-depleted cells. NEK10’s involvement in initiating or maintaining cell cycle arrest during differentiation should also be further explored. Work described in Chapter 2 outlines a role for NEK10 in promoting the G2/M checkpoint following UV irradiation, raising an intriguing possibility that cell cycle arrest may be a conserved role for NEK10. In proliferating cells, NEK10 may act at the level of checkpoint control, whereas it may promote cell cycle arrest or withdrawal in the context of cell differentiation. While it is clear that checkpoint arrest and cell cycle withdrawal during differentiation are distinctly regulated, they have partially overlapping molecular signatures

(reviewed in (314)). Identification of substrates or proteins that interact with NEK10 during myocyte differentiation would provide a greater understanding of NEK10’s role in myogenesis, and may be instrumental in uncovering the mechanism of NEK10-mediated cell cycle arrest following genotoxic stress

4.3 NEK10 and cell adhesion

In Chapter 2, NEK10 was shown to enhance MEK autoactivation following UV

irradiation. MEK autophosphorylation has also been described in the context of cell adhesion

and integrin engagement (303). In adherent cells, Pak phosphorylates MEK1 on S298 and

122 this is abolished when cells are brought into suspension (369). S298 phosphorylation increases MEK1 activation by enhancing MEK1’s interaction with its upstream activator

Raf-1 (109). In addition, S298 phosphorylation of MEK1 may stimulate MEK1 autophosphorylation following adhesion (303). Intrigued by these findings, I pursued

NEK10’s effect on MEK activity in the context of cell adhesion, but found that ectopic expression of NEK10 failed to stimulate MEK1 S298 phosphorylation in either adherent cells or cells re-plated onto fibronectin, indicating that NEK10 promotes MEK autophosphorylation by a mechanism independent of Pak and S298 phosphorylation (Fig

4.3).

In the course of these experiments, I also examined the effect of adherence on NEK10

catalytic activity. Endogenous NEK10 was immunoprecipitated from differentiated C2C12 myotubes and subjected to in vitro kinase assays. NEK10 kinase activity decreased when myotubes were placed in suspension and this could be recovered upon re-plating (Fig 4.4A).

Interestingly, Flag-NEK10 immunoprecipitated from suspended HEK293 cells exhibited a downward mobility shift in SDS-PAGE, indicative of a lower degree of phosphorylation.

Significantly, this mobility shift was abolished upon re-plating (Fig 4.4B), raising several interesting questions regarding possible regulation of NEK10 by cell adhesion.

In tissues, cells are in constant contact with other cells and the extracellular matrix. In

2-D culture, most non-transformed cells require continuous interaction with solid support and loss of adhesion often results in cell cycle arrest, or a form of apoptosis termed anoikis (107,

379). Integrins are one of the key proteins involved in mediating the flow of information

between the extra- and intracellular compartments (reviewed in (277)). In response to cell adhesion these transmembrane glycoproteins recruit and activate kinases cascades,

123 124 125 stimulating downstream signaling (reviewed in (277)). Integrin signaling is critically involved in cell cycle progression, cell shape, tissue organization and cell motility.

To extend on our preliminary observations, NEK10’s regulation by cell adhesion and

integrins should be explored. Initially, further insight into the effect of adhesion on NEK10

catalytic activity and/or phosphorylation status will be needed. Identification of adhesion-

dependent NEK10 phosphorylation sites would focus the pursuit of upstream regulators of its

activity upon adhesion. Candidate approaches aimed at proteins with established roles in

mediating signals downstream of cell adhesion/integrins would also be informative. Using

chemical inhibitors or RNAi strategies, the effect of inhibition of proteins such as Rac and

Rho GTPases and Pak kinases on NEK10 activity/phosphorylation status should be

determined. In parallel, contribution of NEK10 to cell adhesion and to the assembly and

activation of adhesion-dependent signaling complexes should be examined. The outlined

basic experimental strategy will inform subsequent steps in understanding NEK10’s function

in cell adhesion.

4.4 NEK10 interacts with tubulin in vivo

To identify interacting partners of NEK10, the protein content of NEK10 immunoprecipitates from transiently transfected HEK293 cells was analysed by mass spectrometry, in collaboration with the laboratory of A.C.Gingras (The Samuel Lunenfeld

Research Institute of Mount Sinai Hospital). α- and β-tubulin were isolated as significant binding partners for NEK10. Tubulins are commonly found in mass spectrometry-based interaction screens, and are often considered false-positive hits. However, the number of

126 NEK10-interacting tubulin peptides was considerably higher than what is considered

background. For example, 155 NEK10 peptide events, which covered 39% of the protein

were identified. By comparison, the top 4 NEK10-interacting proteins, isoforms of β-tubulin,

were each represented by over 250 peptide events providing 60% coverage. In total, over

1000 peptide events corresponding to α- and β-tubulin interacted with NEK10, and were not

detected in control immunoprecipitations. The number and variety of identified tubulin

peptides provides convincing evidence that tubulin is a true NEK10 interactor.

Significantly, NEK kinases have a role in regulating microtubule (MT) dynamics at

the centrosome, cilia and the mitotic spindle. NEK2 activity promotes centrosome separation

during early mitosis and deregulation of NEK2 increases the rate of spindle defects (111).

Inhibition of either NEK6 or NEK7 leads to fragile spindle formation, activation of the SAC

and defects in cytokinesis (292). Interestingly, while both of these kinases are diffusely

distributed throughout the cell cycle they also display distinct patterns of subcellular localization. NEK6 localizes to MT-based structures during mitosis and cytokinesis,

including the mitotic spindle and the central body respectively, while a small fraction of

NEK7 is associated with interphase centrosomes and mitotic spindle poles (292). In vitro,

both NEK6 and NEK7 interact with and phosphorylate MT preparations (292).

Future work should explore the biochemical properties and the physiological

relevance of NEK10 interaction with tubulin. Based on work with NEK6 and NEK7, it would

be relevant to confirm a physical interaction between NEK10 and tubulin and/or MT

preparations, as well as examine the possibility that NEK10 directly phosphorylates tubulin

or a MT-associated protein. The effect of MT stability on NEK10 catalytic activity should

127 also be determined, for instance in cells treated with microtubule stabilizing or

depolymerization agents, paclitaxel and colcemid, respectively.

Moreover, a detailed characterization of the subcellular localization of NEK10 should

be performed. In particular, localization of NEK10 to tubulin-based structures including

MTs, centrosomes and cilia during interphase, and centrosomes, mitotic spindles and the central body during mitosis should be examined. Based on the results of these experiments,

further studies could probe the effect of NEK10 on MT organization, stability and dynamics

using in vitro MT preparations and immunofluorescence-based assays in cultured cells. In case that NEK10 localizes to centrosomes, a role in regulating centrosomal numbers and duplication, separation and maturation should be examined. Finally, possible NEK10 functions in normal spindle assembly, chromosome segregation and cytokinesis should be explored.

In proliferating cells the centrosome nucleates MTs. In most differentiated cells however, MTs are organized by noncentrosomal mechanisms. In these systems MTs often

display linear, as opposed to radial, organization and increased stability (reviewed in (20)).

Considering our observations that NEK10 may participate in myogenesis, it may be

worthwhile to explore a role for NEK10 in noncentrosomal MT organization during myocyte

differentiation. Terminal differentiation of muscle cells involves extensive changes in the MT

architecture, from a centrosome-bound radial array into parallel fibres that are longitudinally-

organized. This may be triggered by a redistribution of centrosomal proteins from the

pericentriolar region to the nuclear periphery (394). Significantly, in differentiated myotubes,

nuclei gain the ability to nucleate MT assembly while the potential of centrosomes to do so

decreases (97). MT organization has been implicated in the process of myofibril generation

128 and in maintenance of cell shape, however the mechanism of this is largely unexplored. It is

intriguing to speculate that NEK10 may play a role in the reorganization, maintenance or

stability of the MT network during differentiation.

4.5 NEK10 in tumourigenesis

Our lab has generated transgenic mouse lines carrying conditionally targeted alleles

of NEK10 which upon Cre-recombinase expression are expected to lead to a complete loss of

NEK10 function (P.Dutt and V.Stambolic, unpublished). These mice will provide invaluable

tools to further explore the physiological roles of NEK10. The course of examination will depend on the viability and/or the phenotype of the mice lacking NEK10. Nevertheless, broad applications for this animal model can be anticipated. MEFs generated from NEK10 knockout mice will provide a powerful model system to study the cellular functions of

NEK10 and may be used in place of NEK10 knockdown in previously outlined experiments, as they will feature complete loss of NEK10 and be devoid of potential off-target effects associated with siRNA methods.

The NEK10 knockout mice will also be instrumental in exploring a role for NEK10 in oncogenesis. As outlined in section 1.3.2, there is accumulating evidence that NEK10 may be implicated in breast cancer. Thus, it would be highly informative to conditionally delete

NEK10 in the developing mammary gland of the mouse and observe if this event promotes tumourigenesis. Considering that NEK10-deletion alone may not be sufficient to generate mammary tumours, NEK10 deletion can be combined with overexpression of ErbB2, a major transforming oncogene, in the same mammary cells (reviewed in (91)). For example,

129 MMTV-NIC mice (W.J. Muller), which express the Cre-recombinase and ErbB2/Neu in the same mammary epithelial cells, can be crossed to NEK10 flox/flox mice (367). MMTV-NIC

mice develop tumours with 100% penetrance at approximately five months of age. Based

upon our data pertaining to the reduced survival of patients with low NEK10-tumour expression it may be predicted that MMTV-NIC;NEK10-/- mice would exhibit reduced tumour latency and/or increased tumour burden. The histopathological properties of tumours

should also be examined to determine the impact of NEK10 on the tumour spectrum,

possibly associated with the effect of NEK10 on genetic stability. Finally the impact of

NEK10 on the metastatic properties of tumours should be examined by monitoring tumour

spread to the lungs and liver.

4.6 Concluding Remarks

Impaired checkpoint function can have critical consequences for the cell, leading to

genomic instability, transformation and loss of viability. Work described in this thesis

provides the first insight into NEK10 regulation and its role in checkpoint control. These

studies will serve as a foundation for future studies aimed at understanding the molecular

mechanism of NEK10 action and its function in organismal development and

tumourigenesis.

130 5.0 References

1. Abbott, D. W., and J. T. Holt. 1999. Mitogen-activated protein kinase kinase 2 activation is essential for progression through the G2/M checkpoint arrest in cells exposed to ionizing radiation. J Biol Chem 274:2732-42. 2. Abraham, R. T. 2005. MAPKAP kinase-2: Three's company at the G(2) checkpoint. Molecular Cell 17:163-164. 3. Adachi, M., M. Fukuda, and E. Nishida. 1999. Two co-existing mechanisms for nuclear import of MAP kinase: passive diffusion of a monomer and active transport of a dimer. Embo Journal 18:5347-5358. 4. Ahmed, S., G. Thomas, M. Ghoussaini, C. S. Healey, M. K. Humphreys, R. Platte, J. Morrison, M. Maranian, K. A. Pooley, R. Luben, D. Eccles, D. G. Evans, O. Fletcher, N. Johnson, I. Dos Santos Silva, J. Peto, M. R. Stratton, N. Rahman, K. Jacobs, R. Prentice, G. L. Anderson, A. Rajkovic, J. D. Curb, R. G. Ziegler, C. D. Berg, S. S. Buys, C. A. McCarty, H. S. Feigelson, E. E. Calle, M. J. Thun, W. R. Diver, S. Bojesen, B. G. Nordestgaard, H. Flyger, T. Dork, P. Schurmann, P. Hillemanns, J. H. Karstens, N. V. Bogdanova, N. N. Antonenkova, I. V. Zalutsky, M. Bermisheva, S. Fedorova, E. Khusnutdinova, D. Kang, K. Y. Yoo, D. Y. Noh, S. H. Ahn, P. Devilee, C. J. van Asperen, R. A. Tollenaar, C. Seynaeve, M. Garcia-Closas, J. Lissowska, L. Brinton, B. Peplonska, H. Nevanlinna, T. Heikkinen, K. Aittomaki, C. Blomqvist, J. L. Hopper, M. C. Southey, L. Smith, A. B. Spurdle, M. K. Schmidt, A. Broeks, R. R. van Hien, S. Cornelissen, R. L. Milne, G. Ribas, A. Gonzalez-Neira, J. Benitez, R. K. Schmutzler, B. Burwinkel, C. R. Bartram, A. Meindl, H. Brauch, C. Justenhoven, U. Hamann, J. Chang-Claude, R. Hein, S. Wang-Gohrke, A. Lindblom, S. Margolin, A. Mannermaa, V. M. Kosma, V. Kataja, J. E. Olson, X. Wang, Z. Fredericksen, G. G. Giles, G. Severi, L. Baglietto, D. R. English, S. E. Hankinson, D. G. Cox, P. Kraft, L. J. Vatten, K. Hveem, M. Kumle, et al. 2009. Newly discovered breast cancer susceptibility loci on 3p24 and 17q23.2. Nat Genet. 5. Ahn, J., M. Urist, and C. Prives. 2004. The Chk2 protein kinase. DNA Repair 3:1039-1047. 6. Ahn, J. Y., X. Li, H. L. Davis, and C. E. Canman. 2002. Phosphorylation of threonine 68 promotes oligomerization and autophosphorylation of the Chk2 protein kinase via the forkhead-associated domain. J Biol Chem 277:19389-95. 7. al-Khodairy, F., and A. M. Carr. 1992. DNA repair mutants defining G2 checkpoint pathways in Schizosaccharomyces pombe. Embo J 11:1343-50. 8. Alessi, D. R., M. Andjelkovic, B. Caudwell, P. Cron, N. Morrice, P. Cohen, and B. A. Hemmings. 1996. Mechanism of activation of protein kinase B by insulin and IGF-1. Embo J 15:6541-51. 9. Alessi, D. R., A. Cuenda, P. Cohen, D. T. Dudley, and A. R. Saltiel. 1995. PD 098059 is a specific inhibitor of the activation of mitogen-activated protein kinase kinase in vitro and in vivo. J Biol Chem 270:27489-94. 10. Altun, D. H. H. a. Z. F. 2008. C.elegans atlas. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

131 11. Andersen, S. S., A. J. Ashford, R. Tournebize, O. Gavet, A. Sobel, A. A. Hyman, and E. Karsenti. 1997. Mitotic chromatin regulates phosphorylation of Stathmin/Op18. Nature 389:640-3. 12. Archambault, V., and D. M. Glover. 2009. Polo-like kinases: conservation and divergence in their functions and regulation (vol 10, pg 265, 2009). Nature Reviews Molecular Cell Biology 10. 13. Astin, J. W., N. J. O'Neil, and P. E. Kuwabara. 2008. Nucleotide excision repair and the degradation of RNA pol II by the Caenorhabditis elegans XPA and Rsp5 orthologues, RAD-3 and WWP-1. DNA Repair (Amst) 7:267-80. 14. Bahe, S., Y. D. Stierhof, C. J. Wilkinson, F. Leiss, and E. A. Nigg. 2005. Rootletin forms centriole-associated filaments and functions in centrosome cohesion. J Cell Biol 171:27-33. 15. Bahmanyar, S., D. D. Kaplan, J. G. Deluca, T. H. Giddings, Jr., E. T. O'Toole, M. Winey, E. D. Salmon, P. J. Casey, W. J. Nelson, and A. I. Barth. 2008. beta- Catenin is a Nek2 substrate involved in centrosome separation. Genes Dev 22:91- 105. 16. Balan, V., D. T. Leicht, J. Zhu, K. Balan, A. Kaplun, V. Singh-Gupta, J. Qin, H. Ruan, M. J. Comb, and G. Tzivion. 2006. Identification of novel in vivo Raf-1 phosphorylation sites mediating positive feedback Raf-1 regulation by extracellular signal-regulated kinase. Mol Biol Cell 17:1141-53. 17. Baldin, V., K. Pelpel, M. Cazales, C. Cans, and B. Ducommun. 2002. Nuclear localization of CDC25B1 and serine 146 integrity are required for induction of mitosis. J Biol Chem 277:35176-82. 18. Balmanno, K., and S. J. Cook. 1999. Sustained MAP kinase activation is required for the expression of cyclin D1, p21Cip1 and a subset of AP-1 proteins in CCL39 cells. Oncogene 18:3085-97. 19. Bartek, J., and J. Lukas. 2003. Chk1 and Chk2 kinases in checkpoint control and cancer. Cancer Cell 3:421-429. 20. Bartolini, F., and G. G. Gundersen. 2006. Generation of noncentrosomal microtubule arrays. J Cell Sci 119:4155-63. 21. Bayliss, R., T. Sardon, I. Vernos, and E. Conti. 2003. Structural basis of Aurora-A activation by TPX2 at the mitotic spindle. Mol Cell 12:851-62. 22. Belanger, L. F., S. Roy, M. Tremblay, B. Brott, A. M. Steff, W. Mourad, P. Hugo, R. Erikson, and J. Charron. 2003. Mek2 is dispensable for mouse growth and development. Mol Cell Biol 23:4778-87. 23. Belham, C., M. J. Comb, and J. Avruch. 2001. Identification of the NIMA family kinases NEK6/7 as regulators of the p70 . Curr Biol 11:1155-67. 24. Belham, C., J. Roig, J. A. Caldwell, Y. Aoyama, B. E. Kemp, M. Comb, and J. Avruch. 2003. A mitotic cascade of NIMA family kinases. Nercc1/Nek9 activates the Nek6 and Nek7 kinases. J Biol Chem 278:34897-909. 25. Bentley, N. J., D. A. Holtzman, G. Flaggs, K. S. Keegan, A. DeMaggio, J. C. Ford, M. Hoekstra, and A. M. Carr. 1996. The Schizosaccharomyces pombe rad3 checkpoint gene. Embo Journal 15:6641-6651. 26. Berbari, N. F., A. K. O'Connor, C. J. Haycraft, and B. K. Yoder. 2009. The primary cilium as a complex signaling center. Curr Biol 19:R526-35.

132 27. Berdnik, D., and J. A. Knoblich. 2002. Drosophila Aurora-A is required for centrosome maturation and actin-dependent asymmetric protein localization during mitosis. Curr Biol 12:640-7. 28. Besson, A., S. F. Dowdy, and J. M. Roberts. 2008. CDK inhibitors: cell cycle regulators and beyond. Dev Cell 14:159-69. 29. Bettencourt-Dias, M., A. Rodrigues-Martins, L. Carpenter, M. Riparbelli, L. Lehmann, M. K. Gatt, N. Carmo, F. Balloux, G. Callaini, and D. M. Glover. 2005. SAK/PLK4 is required for centriole duplication and flagella development. Current Biology 15:2199-2207. 30. Biggs, W. H., 3rd, K. H. Zavitz, B. Dickson, A. van der Straten, D. Brunner, E. Hafen, and S. L. Zipursky. 1994. The Drosophila rolled locus encodes a MAP kinase required in the sevenless signal transduction pathway. Embo J 13:1628-35. 31. Booher, R. N., C. E. Alfa, J. S. Hyams, and D. H. Beach. 1989. The fission yeast cdc2/cdc13/suc1 protein kinase: regulation of catalytic activity and nuclear localization. Cell 58:485-97. 32. Booher, R. N., P. S. Holman, and A. Fattaey. 1997. Human Myt1 is a cell cycle- regulated kinase that inhibits Cdc2 but not Cdk2 activity. J Biol Chem 272:22300-6. 33. Bos, J. L. 1989. ras oncogenes in human cancer: a review. Cancer Res 49:4682-9. 34. Bost, F., M. Aouadi, L. Caron, P. Even, N. Belmonte, M. Prot, C. Dani, P. Hofman, G. Pages, J. Pouyssegur, Y. Le Marchand-Brustel, and B. Binetruy. 2005. The extracellular signal-regulated kinase isoform ERK1 is specifically required for in vitro and in vivo adipogenesis. Diabetes 54:402-11. 35. Boutros, R., C. Dozier, and B. Ducommun. 2006. The when and wheres of CDC25 phosphatases. Current Opinion in Cell Biology 18:185-191. 36. Brondello, J. M., A. Brunet, J. Pouyssegur, and F. R. McKenzie. 1997. The dual specificity mitogen-activated protein kinase phosphatase-1 and -2 are induced by the p42/p44(MAPK) cascade. Journal of Biological Chemistry 272:1368-1376. 37. Brondello, J. M., J. Pouyssegur, and F. R. McKenzie. 1999. Reduced MAP kinase phosphatase-1 degradation after p42/p44(MAPK)-dependent phosphorylation. Science 286:2514-2517. 38. Brunet, A., A. Bonni, M. J. Zigmond, M. Z. Lin, P. Juo, L. S. Hu, M. J. Anderson, K. C. Arden, J. Blenis, and M. E. Greenberg. 1999. Akt promotes cell survival by phosphorylating and inhibiting a Forkhead transcription factor. Cell 96:857-68. 39. Brunet, A., D. Roux, P. Lenormand, S. Dowd, S. Keyse, and J. Pouyssegur. 1999. Nuclear translocation of p42/p44 mitogen-activated protein kinase is required for growth factor-induced gene expression and cell cycle entry. Embo Journal 18:664- 674. 40. Budirahardja, Y., and P. Gonczy. 2008. PLK-1 asymmetry contributes to asynchronous cell division of C. elegans embryos. Development 135:1303-13. 41. Bulavin, D. V., Y. Higashimoto, I. J. Popoff, W. A. Gaarde, V. Basrur, O. Potapova, E. Appella, and A. J. Fornace, Jr. 2001. Initiation of a G2/M checkpoint after ultraviolet radiation requires p38 kinase. Nature 411:102-7. 42. Busino, L., M. Donzelli, M. Chiesa, D. Guardavaccaro, D. Ganoth, N. V. Dorrello, A. Hershko, M. Pagano, and G. F. Draetta. 2003. Degradation of

133 Cdc25A by beta-TrCP during S phase and in response to DNA damage. Nature 426:87-91. 43. Canagarajah, B. J., A. Khokhlatchev, M. H. Cobb, and E. J. Goldsmith. 1997. Activation mechanism of the MAP kinase ERK2 by dual phosphorylation. Cell 90:859-869. 44. Canman, C. E., D. S. Lim, K. A. Cimprich, Y. Taya, K. Tamai, K. Sakaguchi, E. Appella, M. B. Kastan, and J. D. Siliciano. 1998. Activation of the ATM kinase by ionizing radiation and phosphorylation of p53. Science 281:1677-1679. 45. Casar, B., I. Arozarena, V. Sanz-Moreno, A. Pinto, L. Agudo-Ibanez, R. Marais, R. E. Lewis, M. T. Berciano, and P. Crespo. 2009. Ras subcellular localization defines extracellular signal-regulated kinase 1 and 2 substrate specificity through distinct utilization of scaffold proteins. Mol Cell Biol 29:1338-53. 46. Catling, A. D., H. J. Schaeffer, C. W. Reuter, G. R. Reddy, and M. J. Weber. 1995. A proline-rich sequence unique to MEK1 and MEK2 is required for raf binding and regulates MEK function. Mol Cell Biol 15:5214-25. 47. Chambard, J. C., R. Lefloch, J. Pouyssegur, and P. Lenormand. 2007. ERK implication in cell cycle regulation. Biochim Biophys Acta 1773:1299-310. 48. Chang, L., and M. Karin. 2001. Mammalian MAP kinase signalling cascades. Nature 410:37-40. 49. Chen, R. H., C. Sarnecki, and J. Blenis. 1992. Nuclear localization and regulation of erk- and rsk-encoded protein kinases. Mol Cell Biol 12:915-27. 50. Chen, Y., P. L. Chen, C. F. Chen, X. Jiang, and D. J. Riley. 2008. Never-in- mitosis related kinase 1 functions in DNA damage response and checkpoint control. Cell Cycle 7:3194-201. 51. Chen, Y. H., and Y. Sanchez. 2004. Chk1 in the DNA damage response: conserved roles from yeasts to mammals. DNA Repair 3:1025-1032. 52. Chin-Ming Tan, B., and S. C. Lee. 2003. Nek9, a novel FACT-associated protein, modulates interphase progression. J Biol Chem. 53. Choi, K. Y., B. Satterberg, D. M. Lyons, and E. A. Elion. 1994. Ste5 tethers multiple protein kinases in the MAP kinase cascade required for mating in S. cerevisiae. Cell 78:499-512. 54. Cimprich, K. A., and D. Cortez. 2008. ATR: an essential regulator of genome integrity. Nat Rev Mol Cell Biol 9:616-27. 55. Cimprich, K. A., T. B. Shin, C. T. Keith, and S. L. Schreiber. 1996. cDNA cloning and gene mapping of a candidate human cell cycle checkpoint protein. Proceedings of the National Academy of Sciences of the United States of America 93:2850-2855. 56. Cole, A., S. Frame, and P. Cohen. 2004. Further evidence that the tyrosine phosphorylation of glycogen synthase kinase-3 (GSK3) in mammalian cells is an autophosphorylation event. Biochem J 377:249-55. 57. Cooper, J. A., B. M. Sefton, and T. Hunter. 1984. Diverse mitogenic agents induce the phosphorylation of two related 42,000-dalton proteins on tyrosine in quiescent chick cells. Mol Cell Biol 4:30-7. 58. Coppola, J. A., B. A. Lewis, and M. D. Cole. 1990. Increased retinoblastoma gene expression is associated with late stages of differentiation in many different cell types. Oncogene 5:1731-3.

134 59. Corson, L. B., Y. Yamanaka, K. M. Lai, and J. Rossant. 2003. Spatial and temporal patterns of ERK signaling during mouse embryogenesis. Development 130:4527-37. 60. Crasta, K., P. Huang, G. Morgan, M. Winey, and U. Surana. 2006. Cdk1 regulates centrosome separation by restraining proteolysis of microtubule-associated proteins. Embo J 25:2551-63. 61. Curran, K. L., and R. M. Grainger. 2000. Expression of activated MAP kinase in Xenopus laevis embryos: evaluating the roles of FGF and other signaling pathways in early induction and patterning. Dev Biol 228:41-56. 62. Dalal, S. N., C. M. Schweitzer, J. M. Gan, and J. A. DeCaprio. 1999. Cytoplasmic localization of human cdc25C during interphase requires an intact 14-3-3 binding site. Molecular and Cellular Biology 19:4465-4479. 63. Das, S., J. Cho, I. Lambertz, M. A. Kelliher, A. G. Eliopoulos, K. Y. Du, and P. N. Tsichlis. 2005. Tpl2/Cot signals activate ERK, JNK, and NF-kappa B in a cell- type and stimulus-specific manner. Journal of Biological Chemistry 280:23748- 23757. 64. Davezac, N., V. Baldin, B. Gabrielli, A. Forrest, N. Theis-Febvre, M. Yashida, and B. Ducommun. 2000. Regulation of CDC25B phosphatases subcellular localization. Oncogene 19:2179-2185. 65. Davies, H., G. R. Bignell, C. Cox, P. Stephens, S. Edkins, S. Clegg, J. Teague, H. Woffendin, M. J. Garnett, W. Bottomley, N. Davis, E. Dicks, R. Ewing, Y. Floyd, K. Gray, S. Hall, R. Hawes, J. Hughes, V. Kosmidou, A. Menzies, C. Mould, A. Parker, C. Stevens, S. Watt, S. Hooper, R. Wilson, H. Jayatilake, B. A. Gusterson, C. Cooper, J. Shipley, D. Hargrave, K. Pritchard-Jones, N. Maitland, G. Chenevix-Trench, G. J. Riggins, D. D. Bigner, G. Palmieri, A. Cossu, A. Flanagan, A. Nicholson, J. W. Ho, S. Y. Leung, S. T. Yuen, B. L. Weber, H. F. Seigler, T. L. Darrow, H. Paterson, R. Marais, C. J. Marshall, R. Wooster, M. R. Stratton, and P. A. Futreal. 2002. Mutations of the BRAF gene in human cancer. Nature 417:949-54. 66. Davies, H., C. Hunter, R. Smith, P. Stephens, C. Greenman, G. Bignell, J. Teague, A. Butler, S. Edkins, C. Stevens, A. Parker, S. O'Meara, T. Avis, S. Barthorpe, L. Brackenbury, G. Buck, J. Clements, J. Cole, E. Dicks, K. Edwards, S. Forbes, M. Gorton, K. Gray, K. Halliday, R. Harrison, K. Hills, J. Hinton, D. Jones, V. Kosmidou, R. Laman, R. Lugg, A. Menzies, J. Perry, R. Petty, K. Raine, R. Shepherd, A. Small, H. Solomon, Y. Stephens, C. Tofts, J. Varian, A. Webb, S. West, S. Widaa, A. Yates, F. Brasseur, C. S. Cooper, A. M. Flanagan, A. Green, M. Knowles, S. Y. Leung, L. H. Looijenga, B. Malkowicz, M. A. Pierotti, B. T. Teh, S. T. Yuen, S. R. Lakhani, D. F. Easton, B. L. Weber, P. Goldstraw, A. G. Nicholson, R. Wooster, M. R. Stratton, and P. A. Futreal. 2005. Somatic mutations of the protein kinase gene family in human lung cancer. Cancer Res 65:7591-5. 67. Davies, J. R., A. H. Osmani, C. P. De Souza, C. Bachewich, and S. A. Osmani. 2004. Potential link between the NIMA mitotic kinase and nuclear membrane fission during mitotic exit in Aspergillus nidulans. Eukaryot Cell 3:1433-44. 68. Dawe, H. R., H. Farr, and K. Gull. 2007. Centriole/basal body morphogenesis and migration during ciliogenesis in animal cells. J Cell Sci 120:7-15.

135 69. De Souza, C. P., A. H. Osmani, L. P. Wu, J. L. Spotts, and S. A. Osmani. 2000. Mitotic histone H3 phosphorylation by the NIMA kinase in Aspergillus nidulans. Cell 102:293-302. 70. DeGregori, J., T. Kowalik, and J. R. Nevins. 1995. Cellular targets for activation by the E2F1 transcription factor include DNA synthesis- and G1/S-regulatory genes. Mol Cell Biol 15:4215-24. 71. Dent, P., D. B. Reardon, D. K. Morrison, and T. W. Sturgill. 1995. Regulation of Raf-1 and Raf-1 Mutants by Ras-Dependent and Ras-Independent Mechanisms in- Vitro. Molecular and Cellular Biology 15:4125-4135. 72. Dessev, G., C. Iovcheva-Dessev, J. R. Bischoff, D. Beach, and R. Goldman. 1991. A Complex Containing P34-Cdc-2 and Cyclin B Phosphorylates the Nuclear Lamin and Disassembles Nuclei of Clam Oocytes in-Vitro. Journal of Cell Biology 112:523- 534. 73. Dhanasekaran, D. N., K. Kashef, C. M. Lee, H. Xu, and E. P. Reddy. 2007. Scaffold proteins of MAP-kinase modules. Oncogene 26:3185-202. 74. Dhillon, A. S., and W. Kolch. 2002. Untying the regulation of the Raf-1 kinase. Archives of Biochemistry and Biophysics 404:3-9. 75. Dimitri, C. A., W. Dowdle, J. P. MacKeigan, J. Blenis, and L. O. Murphy. 2005. Spatially separate docking sites on ERK2 regulate distinct signaling events in vivo. Curr Biol 15:1319-24. 76. Ditchfield, C., V. L. Johnson, A. Tighe, R. Ellston, C. Haworth, T. Johnson, A. Mortlock, N. Keen, and S. S. Taylor. 2003. Aurora B couples chromosome alignment with anaphase by targeting BubR1, Mad2, and Cenp-E to kinetochores. J Cell Biol 161:267-80. 77. Dobles, M., V. Liberal, M. L. Scott, R. Benezra, and P. K. Sorger. 2000. Chromosome missegregation and apoptosis in mice lacking the mitotic checkpoint protein Mad2. Cell 101:635-45. 78. Donella-Deana, A., L. Cesaro, S. Sarno, A. M. Brunati, M. Ruzzene, and L. A. Pinna. 2001. Autocatalytic tyrosine-phosphorylation of protein kinase CK2 alpha and alpha' subunits: implication of Tyr182. Biochem J 357:563-7. 79. Donzelli, M., L. Busino, M. Chiesa, D. Ganoth, A. Hershko, and G. F. Draetta. 2004. Hierarchical order of phosphorylation events commits Cdc25A to beta TrCP- dependent degradation. Cell Cycle 3:469-471. 80. Donzelli, M., and G. F. Draetta. 2003. Regulating mammalian checkpoints through Cdc25 inactivation. EMBO Rep 4:671-7. 81. Dou, Q. P., P. J. Markell, and A. B. Pardee. 1992. Thymidine kinase transcription is regulated at G1/S phase by a complex that contains retinoblastoma-like protein and a cdc2 kinase. Proc Natl Acad Sci U S A 89:3256-60. 82. Dou, Q. P., S. Zhao, A. H. Levin, J. Wang, K. Helin, and A. B. Pardee. 1994. G1/S-regulated E2F-containing protein complexes bind to the mouse thymidine kinase gene promoter. J Biol Chem 269:1306-13. 83. Dougherty, M. K., J. Muller, D. A. Ritt, M. Zhou, X. Z. Zhou, T. D. Copeland, T. P. Conrads, T. D. Veenstra, K. P. Lu, and D. K. Morrison. 2005. Regulation of Raf-1 by direct feedback phosphorylation. Mol Cell 17:215-24. 84. Draviam, V. M., S. Orrechia, M. Lowe, R. Pardi, and J. Pines. 2001. The localization of human cyclins B1 and B2 determines CDK1 substrate specificity and

136 neither enzyme requires MEK to disassemble the Golgi apparatus. Journal of Cell Biology 152:945-958. 85. Du, J., X. Cai, J. Yao, X. Ding, Q. Wu, S. Pei, K. Jiang, Y. Zhang, W. Wang, Y. Shi, Y. Lai, J. Shen, M. Teng, H. Huang, Q. Fei, E. S. Reddy, J. Zhu, C. Jin, and X. Yao. 2008. The mitotic checkpoint kinase NEK2A regulates kinetochore microtubule attachment stability. Oncogene 27:4107-14. 86. Dulic, V., W. K. Kaufmann, S. J. Wilson, T. D. Tlsty, E. Lees, J. W. Harper, S. J. Elledge, and S. I. Reed. 1994. P53-Dependent Inhibition of Cyclin-Dependent Kinase-Activities in Human Fibroblasts During Radiation-Induced G1 Arrest. Cell 76:1013-1023. 87. Dumitru, C. D., J. D. Ceci, C. Tsatsanis, D. Kontoyiannis, K. Stamatakis, J. H. Lin, C. Patriotis, N. A. Jenkins, N. G. Copeland, G. Kollias, and P. N. Tsichlis. 2000. TNF-alpha induction by LPS is regulated posttranscriptionally via a Tpl2/ERK- dependent pathway. Cell 103:1071-1083. 88. Dunaway, S., H. Y. Liu, and N. C. Walworth. 2005. Interaction of 14-3-3 protein with Chk1 affects localization and checkpoint function. J Cell Sci 118:39-50. 89. Durocher, D., J. Henckel, A. R. Fersht, and S. P. Jackson. 1999. The FHA domain is a modular phosphopeptide recognition motif. Mol Cell 4:387-94. 90. Dutertre, S., M. Cazales, M. Quaranta, C. Froment, V. Trabut, C. Dozier, G. Mirey, J. P. Bouche, N. Theis-Febvre, E. Schmitt, B. Monsarrat, C. Prigent, and B. Ducommun. 2004. Phosphorylation of CDC25B by Aurora-A at the centrosome contributes to the G2-M transition. Journal of Cell Science 117:2523-2531. 91. Eccles, S. A. 2001. The role of c-erbB-2/HER2/neu in breast cancer progression and metastasis. J Mammary Gland Biol Neoplasia 6:393-406. 92. Ekerot, M., M. P. Stavridis, L. Delavaine, M. P. Mitchell, C. Staples, D. M. Owens, I. D. Keenan, R. J. Dickinson, K. G. Storey, and S. M. Keyse. 2008. Negative-feedback regulation of FGF signalling by DUSP6/MKP-3 is driven by ERK1/2 and mediated by Ets factor binding to a conserved site within the DUSP6/MKP-3 gene promoter. Biochem J 412:287-98. 93. Eley, L., L. M. Yates, and J. A. Goodship. 2005. Cilia and disease. Curr Opin Genet Dev 15:308-14. 94. Evans, T., E. T. Rosenthal, J. Youngblom, D. Distel, and T. Hunt. 1983. Cyclin: a protein specified by maternal mRNA in sea urchin eggs that is destroyed at each cleavage division. Cell 33:389-96. 95. Eyers, P. A., E. Erikson, L. G. Chen, and J. L. Maller. 2003. A novel mechanism for activation of the protein kinase Aurora A. Curr Biol 13:691-7. 96. Fabian, J. R., I. O. Daar, and D. K. Morrison. 1993. Critical Tyrosine Residues Regulate the Enzymatic and Biological-Activity of Raf-1 Kinase. Molecular and Cellular Biology 13:7170-7179. 97. Fant, X., V. Srsen, A. Espigat-Georger, and A. Merdes. 2009. Nuclei of non- muscle cells bind centrosome proteins upon fusion with differentiating myoblasts. PLoS One 4:e8303. 98. Faragher, A. J., and A. M. Fry. 2003. Nek2A kinase stimulates centrosome disjunction and is required for formation of bipolar mitotic spindles. Mol Biol Cell 14:2876-89.

137 99. Faure, M., T. A. Voyno-Yasenetskaya, and H. R. Bourne. 1994. cAMP and beta gamma subunits of heterotrimeric G proteins stimulate the mitogen-activated protein kinase pathway in COS-7 cells. J Biol Chem 269:7851-4. 100. Favata, M. F., K. Y. Horiuchi, E. J. Manos, A. J. Daulerio, D. A. Stradley, W. S. Feeser, D. E. Van Dyk, W. J. Pitts, R. A. Earl, F. Hobbs, R. A. Copeland, R. L. Magolda, P. A. Scherle, and J. M. Trzaskos. 1998. Identification of a novel inhibitor of mitogen-activated protein kinase kinase. J Biol Chem 273:18623-32. 101. Feige, E., and B. Motro. 2002. The related murine kinases, Nek6 and Nek7, display distinct patterns of expression. Mech Dev 110:219-23. 102. Fesquet, D., J. C. Labbe, J. Derancourt, J. P. Capony, S. Galas, F. Girard, T. Lorca, J. Shuttleworth, M. Doree, and J. C. Cavadore. 1993. The MO15 gene encodes the catalytic subunit of a protein kinase that activates cdc2 and other cyclin- dependent kinases (CDKs) through phosphorylation of Thr161 and its homologues. Embo J 12:3111-21. 103. Flemington, E. K., S. H. Speck, and W. G. Kaelin, Jr. 1993. E2F-1-mediated transactivation is inhibited by complex formation with the retinoblastoma susceptibility gene product. Proc Natl Acad Sci U S A 90:6914-8. 104. Fletcher, L., G. J. Cerniglia, E. A. Nigg, T. J. Yend, and R. J. Muschel. 2004. Inhibition of centrosome separation after DNA damage: a role for Nek2. Radiat Res 162:128-35. 105. Franklin, D. S., and Y. Xiong. 1996. Induction of p18INK4c and its predominant association with CDK4 and CDK6 during myogenic differentiation. Mol Biol Cell 7:1587-99. 106. Friedel, A. M., B. L. Pike, and S. M. Gasser. 2009. ATR/Mec1: coordinating fork stability and repair. Curr Opin Cell Biol 21:237-44. 107. Frisch, S. M., and H. Francis. 1994. Disruption of epithelial cell-matrix interactions induces apoptosis. J Cell Biol 124:619-26. 108. Frodin, M., T. L. Antal, B. A. Dummler, C. J. Jensen, M. Deak, S. Gammeltoft, and R. M. Biondi. 2002. A phosphoserine/threonine-binding pocket in AGC kinases and PDK1 mediates activation by hydrophobic motif phosphorylation. Embo J 21:5396-407. 109. Frost, J. A., H. Steen, P. Shapiro, T. Lewis, N. Ahn, P. E. Shaw, and M. H. Cobb. 1997. Cross-cascade activation of ERKs and ternary complex factors by Rho family proteins. Embo Journal 16:6426-6438. 110. Fry, A. M., T. Mayor, P. Meraldi, Y. D. Stierhof, K. Tanaka, and E. A. Nigg. 1998. C-Nap1, a novel centrosomal coiled-coil protein and candidate substrate of the cell cycle-regulated protein kinase Nek2. J Cell Biol 141:1563-74. 111. Fry, A. M., P. Meraldi, and E. A. Nigg. 1998. A centrosomal function for the human Nek2 protein kinase, a member of the NIMA family of cell cycle regulators. Embo J 17:470-81. 112. Fry, A. M., S. J. Schultz, J. Bartek, and E. A. Nigg. 1995. Substrate specificity and cell cycle regulation of the Nek2 protein kinase, a potential human homolog of the mitotic regulator NIMA of Aspergillus nidulans. J Biol Chem 270:12899-905. 113. Fukuda, M., Y. Gotoh, and E. Nishida. 1997. Interaction of MAP kinase with MAP kinase kinase: Its possible role in the control of nucleocytoplasmic transport of MAP kinase. Embo Journal 16:1901-1908.

138 114. Furthauer, M., J. Van Celst, C. Thisse, and B. Thisse. 2004. Fgf signalling controls the dorsoventral patterning of the zebrafish embryo. Development 131:2853- 64. 115. Gabay, L., R. Seger, and B. Z. Shilo. 1997. MAP kinase in situ activation atlas during Drosophila embryogenesis. Development 124:3535-41. 116. Galabova-Kovacs, G., A. Kolbus, D. Matzen, K. Meissl, D. Piazzolla, C. Rubiolo, K. Steinitz, and M. Baccarini. 2006. ERK and beyond: insights from B-Raf and Raf-1 conditional knockouts. Cell Cycle 5:1514-8. 117. Garnett, M. J., S. Rana, H. Paterson, D. Barford, and R. Marais. 2005. Wild-type and mutant B-RAF activate C-RAF through distinct mechanisms involving heterodimerization. Mol Cell 20:963-9. 118. Gartner, A., A. J. MacQueen, and A. M. Villeneuve. 2004. Methods for analyzing checkpoint responses in Caenorhabditis elegans. Methods Mol Biol 280:257-74. 119. Gartner, A., S. Milstein, S. Ahmed, J. Hodgkin, and M. O. Hengartner. 2000. A conserved checkpoint pathway mediates DNA damage--induced apoptosis and cell cycle arrest in C. elegans. Mol Cell 5:435-43. 120. Gatei, M., K. Sloper, C. Sorensen, R. Syljuasen, J. Falck, K. Hobson, K. Savage, J. Lukas, B. B. Zhou, J. Bartek, and K. K. Khanna. 2003. Ataxia-telangiectasia- mutated (ATM) and NBS1-dependent phosphorylation of Chk1 on Ser-317 in response to ionizing radiation. J Biol Chem 278:14806-11. 121. Gautier, J., M. J. Solomon, R. N. Booher, J. F. Bazan, and M. W. Kirschner. 1991. cdc25 is a specific tyrosine phosphatase that directly activates p34cdc2. Cell 67:197-211. 122. Ge, B., H. Gram, F. Di Padova, B. Huang, L. New, R. J. Ulevitch, Y. Luo, and J. Han. 2002. MAPKK-independent activation of p38alpha mediated by TAB1- dependent autophosphorylation of p38alpha. Science 295:1291-4. 123. Geng, Y., Q. Yu, E. Sicinska, M. Das, R. T. Bronson, and P. Sicinski. 2001. Deletion of the p27Kip1 gene restores normal development in cyclin D1-deficient mice. Proc Natl Acad Sci U S A 98:194-9. 124. Giacinti, C., and A. Giordano. 2006. RB and cell cycle progression. Oncogene 25:5220-5227. 125. Giet, R., and D. M. Glover. 2001. Drosophila is required for histone H3 phosphorylation and condensin recruitment during chromosome condensation and to organize the central spindle during cytokinesis. J Cell Biol 152:669-82. 126. Giet, R., and D. M. Glover. 2001. Drosophila Aurora B kinase is required for histone H3 phosphorylation and condensin recruitment during chromosome condensation and to organize the central spindle during cytokinesis. Journal of Cell Biology 152:669-681. 127. Giroux, S., M. Tremblay, D. Bernard, J. F. Cardin-Girard, S. Aubry, L. Larouche, S. Rousseau, J. Huot, J. Landry, L. Jeannotte, and J. Charron. 1999. Embryonic death of Mek1-deficient mice reveals a role for this kinase in angiogenesis in the labyrinthine region of the placenta. Curr Biol 9:369-72. 128. Glover, D. M., M. H. Leibowitz, D. A. McLean, and H. Parry. 1995. Mutations in aurora prevent centrosome separation leading to the formation of monopolar spindles. Cell 81:95-105.

139 129. Golan, A., Y. Yudkovsky, and A. Hershko. 2002. The cyclin-ubiquitin ligase activity of cyclosome/APC is jointly activated by protein kinases Cdk1-cyclin B and Plk. Journal of Biological Chemistry 277:15552-15557. 130. Goldberg, J., A. C. Nairn, and J. Kuriyan. 1996. Structural basis for the autoinhibition of calcium/calmodulin-dependent protein kinase I. Cell 84:875-87. 131. Goloudina, A., H. Yamaguchi, D. B. Chervyakova, E. Appella, A. J. Fornace, Jr., and D. V. Bulavin. 2003. Regulation of human Cdc25A stability by Serine 75 phosphorylation is not sufficient to activate a S phase checkpoint. Cell Cycle 2:473-8. 132. Gonczy, P., C. Echeverri, K. Oegema, A. Coulson, S. J. Jones, R. R. Copley, J. Duperon, J. Oegema, M. Brehm, E. Cassin, E. Hannak, M. Kirkham, S. Pichler, K. Flohrs, A. Goessen, S. Leidel, A. M. Alleaume, C. Martin, N. Ozlu, P. Bork, and A. A. Hyman. 2000. Functional genomic analysis of cell division in C. elegans using RNAi of genes on chromosome III. Nature 408:331-6. 133. Gonzalez, F. A., D. L. Raden, and R. J. Davis. 1991. Identification of substrate recognition determinants for human ERK1 and ERK2 protein kinases. J Biol Chem 266:22159-63. 134. Goodship, J., H. Gill, J. Carter, A. Jackson, M. Splitt, and M. Wright. 2000. Autozygosity mapping of a seckel syndrome locus to chromosome 3q22. 1-q24. Am J Hum Genet 67:498-503. 135. Gould, K. L., and P. Nurse. 1989. Tyrosine phosphorylation of the fission yeast cdc2+ protein kinase regulates entry into mitosis. Nature 342:39-45. 136. Greenman, C., P. Stephens, R. Smith, G. L. Dalgliesh, C. Hunter, G. Bignell, H. Davies, J. Teague, A. Butler, C. Stevens, S. Edkins, S. O'Meara, I. Vastrik, E. E. Schmidt, T. Avis, S. Barthorpe, G. Bhamra, G. Buck, B. Choudhury, J. Clements, J. Cole, E. Dicks, S. Forbes, K. Gray, K. Halliday, R. Harrison, K. Hills, J. Hinton, A. Jenkinson, D. Jones, A. Menzies, T. Mironenko, J. Perry, K. Raine, D. Richardson, R. Shepherd, A. Small, C. Tofts, J. Varian, T. Webb, S. West, S. Widaa, A. Yates, D. P. Cahill, D. N. Louis, P. Goldstraw, A. G. Nicholson, F. Brasseur, L. Looijenga, B. L. Weber, Y. E. Chiew, A. DeFazio, M. F. Greaves, A. R. Green, P. Campbell, E. Birney, D. F. Easton, G. Chenevix- Trench, M. H. Tan, S. K. Khoo, B. T. Teh, S. T. Yuen, S. Y. Leung, R. Wooster, P. A. Futreal, and M. R. Stratton. 2007. Patterns of somatic mutation in human cancer genomes. Nature 446:153-8. 137. Guo, Z. J., A. Kumagai, S. X. Wang, and W. G. Dunphy. 2000. Requirement for Atr in phosphorylation of Chk1 and cell cycle regulation in response to DNA replication blocks and UV-damaged DNA in Xenopus egg extracts. Genes & Development 14:2745-2756. 138. Ha Kim, Y., J. Yeol Choi, Y. Jeong, D. J. Wolgemuth, and K. Rhee. 2002. Nek2 localizes to multiple sites in mitotic cells, suggesting its involvement in multiple cellular functions during the cell cycle. Biochem Biophys Res Commun 290:730-6. 139. Habedanck, R., Y. D. Stierhof, C. J. Wilkinson, and E. A. Nigg. 2005. The Plk4 functions in centriole duplication. Nature Cell Biology 7:1140-1146. 140. Haccard, O., B. Sarcevic, A. Lewellyn, R. Hartley, L. Roy, T. Izumi, E. Erikson, and J. L. Maller. 1993. Induction of metaphase arrest in cleaving Xenopus embryos by MAP kinase. Science 262:1262-5.

140 141. Han, J., J. D. Lee, L. Bibbs, and R. J. Ulevitch. 1994. A MAP kinase targeted by endotoxin and hyperosmolarity in mammalian cells. Science 265:808-11. 142. Hannak, E., M. Kirkham, A. A. Hyman, and K. Oegema. 2001. Aurora-A kinase is required for centrosome maturation in Caenorhabditis elegans. J Cell Biol 155:1109-16. 143. Harbour, J. W., and D. C. Dean. 2000. The Rb/E2F pathway: expanding roles and emerging paradigms. Genes & Development 14:2393-2409. 144. Harbour, J. W., R. X. Luo, A. Dei Santi, A. A. Postigo, and D. C. Dean. 1999. Cdk phosphorylation triggers sequential intramolecular interactions that progressively block Rb functions as cells move through G1. Cell 98:859-69. 145. Hartman, P. S. 1984. UV irradiation of wild type and radiation-sensitive mutants of the nematode Caenorhabditis elegans: fertilities, survival, and parental effects. Photochem Photobiol 39:169-75. 146. Hartman, P. S., and R. K. Herman. 1982. Radiation-sensitive mutants of Caenorhabditis elegans. Genetics 102:159-78. 147. Hashimoto, Y., H. Akita, M. Hibino, K. Kohri, and M. Nakanishi. 2002. Identification and characterization of Nek6 protein kinase, a potential human homolog of NIMA histone H3 kinase. Biochem Biophys Res Commun 293:753-8. 148. Hatano, N., Y. Mori, M. Oh-hora, A. Kosugi, T. Fujikawa, N. Nakai, H. Niwa, J. Miyazaki, T. Hamaoka, and M. Ogata. 2003. Essential role for ERK2 mitogen- activated protein kinase in placental development. Genes Cells 8:847-56. 149. Hauf, S., R. W. Cole, S. LaTerra, C. Zimmer, G. Schnapp, R. Walter, A. Heckel, J. van Meel, C. L. Rieder, and J. M. Peters. 2003. The small molecule Hesperadin reveals a role for Aurora B in correcting kinetochore-microtubule attachment and in maintaining the spindle assembly checkpoint. J Cell Biol 161:281-94. 150. Helps, N. R., X. Luo, H. M. Barker, and P. T. Cohen. 2000. NIMA-related kinase 2 (Nek2), a cell-cycle-regulated protein kinase localized to centrosomes, is complexed to protein phosphatase 1. Biochem J 349:509-18. 151. Herold, A., R. Truant, H. Wiegand, and B. R. Cullen. 1998. Determination of the functional domain organization of the importin alpha nuclear import factor. J Cell Biol 143:309-18. 152. Hibi, M., A. Lin, T. Smeal, A. Minden, and M. Karin. 1993. Identification of an oncoprotein- and UV-responsive protein kinase that binds and potentiates the c-Jun activation domain. Genes Dev 7:2135-48. 153. Hilton, L. K., M. C. White, and L. M. Quarmby. 2009. The NIMA-related kinase NEK1 cycles through the nucleus. Biochem Biophys Res Commun 389:52-6. 154. Hirai, H., M. F. Roussel, J. Y. Kato, R. A. Ashmun, and C. J. Sherr. 1995. Novel INK4 proteins, p19 and p18, are specific inhibitors of the cyclin D-dependent kinases CDK4 and CDK6. Mol Cell Biol 15:2672-81. 155. Hirao, A., A. Cheung, G. Duncan, P. M. Girard, A. J. Elia, A. Wakeham, H. Okada, T. Sarkissian, J. A. Wong, T. Sakai, E. de Stanchina, R. G. Bristow, T. Suda, S. W. Lowe, P. A. Jeggo, S. J. Elledge, and T. W. Mak. 2002. Chk2 is a tumor suppressor that regulates apoptosis in both an ataxia telangiectasia mutated (ATM)-dependent and an ATM-independent manner. Molecular and Cellular Biology 22:6521-6532.

141 156. Hirschl, D., P. Bayer, and O. Muller. 1996. Secondary structure of an armadillo single repeat from the APC protein. Febs Letters 383:31-36. 157. Honda, K., H. Mihara, Y. Kato, A. Yamaguchi, H. Tanaka, H. Yasuda, K. Furukawa, and T. Urano. 2000. Degradation of human Aurora2 protein kinase by the anaphase-promoting complex-ubiquitin-proteasome pathway. Oncogene 19:2812- 2819. 158. Honegger, A. M., R. M. Kris, A. Ullrich, and J. Schlessinger. 1989. Evidence that autophosphorylation of solubilized receptors for epidermal growth factor is mediated by intermolecular cross-phosphorylation. Proc Natl Acad Sci U S A 86:925-9. 159. Hoshi, M., E. Nishida, and H. Sakai. 1989. Characterization of a Mitogen- Activated, Ca-2+-Sensitive Microtubule-Associated Protein-2 Kinase. European Journal of Biochemistry 184:477-486. 160. Hoshi, M., K. Ohta, Y. Gotoh, A. Mori, H. Murofushi, H. Sakai, and E. Nishida. 1992. Mitogen-activated-protein-kinase-catalyzed phosphorylation of microtubule- associated proteins, microtubule-associated protein 2 and microtubule-associated protein 4, induces an alteration in their function. Eur J Biochem 203:43-52. 161. Houldsworth, J., and M. F. Lavin. 1980. Effect of ionizing radiation on DNA synthesis in ataxia telangiectasia cells. Nucleic Acids Res 8:3709-20. 162. Huber, A. H., W. J. Nelson, and W. I. Weis. 1997. Three-dimensional structure of the armadillo repeat region of beta-catenin. Cell 90:871-882. 163. Huse, M., and J. Kuriyan. 2002. The conformational plasticity of protein kinases. Cell 109:275-82. 164. Huser, M., J. Luckett, A. Chiloeches, K. Mercer, M. Iwobi, S. Giblett, X. M. Sun, J. Brown, R. Marais, and C. Pritchard. 2001. MEK kinase activity is not necessary for Raf-1 function. Embo J 20:1940-51. 165. Ikawa, S., M. Fukui, Y. Ueyama, N. Tamaoki, T. Yamamoto, and K. Toyoshima. 1988. B-raf, a new member of the raf family, is activated by DNA rearrangement. Mol Cell Biol 8:2651-4. 166. Inoue, D., and N. Sagata. 2005. The Polo-like kinase Plx1 interacts with and inhibits Myt1 after fertilization of Xenopus eggs. Cell Structure and Function 30:27-27. 167. Ishibe, S., D. Joly, X. Zhu, and L. G. Cantley. 2003. Phosphorylation-dependent paxillin-ERK association mediates hepatocyte growth factor-stimulated epithelial morphogenesis. Mol Cell 12:1275-85. 168. Isoda, M., Y. Kanemori, N. Nakajo, S. Uchida, K. Yamashita, H. Ueno, and N. Sagata. 2009. The extracellular signal-regulated kinase-mitogen-activated protein kinase pathway phosphorylates and targets Cdc25A for SCF beta-TrCP-dependent degradation for cell cycle arrest. Mol Biol Cell 20:2186-95. 169. Jack, M. T., R. A. Woo, A. Hirao, A. Cheung, T. W. Mak, and P. W. K. Lee. 2002. Chk2 is dispensable for p53-mediated G(1) arrest but is required for a latent p53-mediated apoptotic response. Proceedings of the National Academy of Sciences of the United States of America 99:9825-9829. 170. Jacobs, D., D. Glossip, H. Xing, A. J. Muslin, and K. Kornfeld. 1999. Multiple docking sites on substrate proteins form a modular system that mediates recognition by ERK MAP kinase. Genes Dev 13:163-75.

142 171. James, M. K., A. Ray, D. Leznova, and S. W. Blain. 2008. Differential modification of p27Kip1 controls its cyclin D-cdk4 inhibitory activity. Mol Cell Biol 28:498-510. 172. Jang, Y. J., S. Ma, Y. Terada, and R. L. Erikson. 2002. Phosphorylation of threonine 210 and the role of serine 137 in the regulation of mammalian polo-like kinase. J Biol Chem 277:44115-20. 173. Jazayeri, A., J. Falck, C. Lukas, J. Bartek, G. C. M. Smith, J. Lukas, and S. P. Jackson. 2006. ATM- and cell cycle-dependent regulation of ATR in response to DNA double-strand breaks. Nature Cell Biology 8:37-U13. 174. Jeffrey, P. D., A. A. Russo, K. Polyak, E. Gibbs, J. Hurwitz, J. Massague, and N. P. Pavletich. 1995. Mechanism of CDK activation revealed by the structure of a cyclinA-CDK2 complex. Nature 376:313-20. 175. Jelinek, T., A. D. Catling, C. W. Reuter, S. A. Moodie, A. Wolfman, and M. J. Weber. 1994. RAS and RAF-1 form a signalling complex with MEK-1 but not MEK-2. Mol Cell Biol 14:8212-8. 176. Jin, J. P., T. Shirogane, L. Xu, G. Nalepa, J. Qin, S. J. Elledge, and J. W. Harper. 2003. SCF beta-TRCP links Chk1 signaling to degradation of the Cdc25A protein phosphatase. Genes & Development 17:3062-3074. 177. Jinno, S., K. Suto, A. Nagata, M. Igarashi, Y. Kanaoka, H. Nojima, and H. Okayama. 1994. Cdc25A is a novel phosphatase functioning early in the cell cycle. Embo J 13:1549-56. 178. Jones, C. A., and P. S. Hartman. 1996. Replication in UV-irradiated Caenorhabditis elegans embryos. Photochem Photobiol 63:187-92. 179. Kahan, C., K. Seuwen, S. Meloche, and J. Pouyssegur. 1992. Coordinate, biphasic activation of p44 mitogen-activated protein kinase and S6 kinase by growth factors in hamster fibroblasts. Evidence for thrombin-induced signals different from phosphoinositide turnover and adenylylcyclase inhibition. J Biol Chem 267:13369- 75. 180. Kamath, R. S., A. G. Fraser, Y. Dong, G. Poulin, R. Durbin, M. Gotta, A. Kanapin, N. Le Bot, S. Moreno, M. Sohrmann, D. P. Welchman, P. Zipperlen, and J. Ahringer. 2003. Systematic functional analysis of the Caenorhabditis elegans genome using RNAi. Nature 421:231-7. 181. Kandli, M., E. Feige, A. Chen, G. Kilfin, and B. Motro. 2000. Isolation and characterization of two evolutionarily conserved murine kinases (Nek6 and nek7) related to the fungal mitotic regulator, NIMA. Genomics 68:187-96. 182. Karin, M. 1998. Mitogen-activated protein kinase cascades as regulators of stress responses. Ann N Y Acad Sci 851:139-46. 183. Karreth, F. A., and D. A. Tuveson. 2009. Modelling oncogenic Ras/Raf signalling in the mouse. Curr Opin Genet Dev 19:4-11. 184. Kastan, M. B., Q. Zhan, W. S. el-Deiry, F. Carrier, T. Jacks, W. V. Walsh, B. S. Plunkett, B. Vogelstein, and A. J. Fornace, Jr. 1992. A mammalian cell cycle checkpoint pathway utilizing p53 and GADD45 is defective in ataxia-telangiectasia. Cell 71:587-97. 185. Kato, J., H. Matsushime, S. W. Hiebert, M. E. Ewen, and C. J. Sherr. 1993. Direct binding of cyclin D to the retinoblastoma gene product (pRb) and pRb phosphorylation by the cyclin D-dependent kinase CDK4. Genes Dev 7:331-42.

143 186. Katsuragi, Y., and N. Sagata. 2004. Regulation of Chk1 kinase by autoinhibition and ATR-mediated phosphorylation. Mol Biol Cell 15:1680-9. 187. Keyse, S. M. 2008. Dual-specificity MAP kinase phosphatases (MKPs) and cancer. Cancer Metastasis Rev 27:253-61. 188. Kim, S., K. Lee, and K. Rhee. 2007. NEK7 is a centrosomal kinase critical for microtubule nucleation. Biochem Biophys Res Commun 360:56-62. 189. Kimmins, S., C. Crosio, N. Kotaja, J. Hirayama, L. Monaco, C. Hoog, M. van Duin, J. A. Gossen, and P. Sassone-Corsi. 2007. Differential functions of the Aurora-B and Aurora-C kinases in mammalian spermatogenesis. Mol Endocrinol 21:726-39. 190. Kimura, K., M. Hirano, R. Kobayashi, and T. Hirano. 1998. Phosphorylation and activation of 13S condensin by Cdc2 in vitro. Science 282:487-490. 191. King, A. J., H. Y. Sun, B. Diaz, D. Barnard, W. Y. Miao, S. Bagrodia, and M. S. Marshall. 1998. The protein kinase Pak3 positively regulates Raf-1 activity through phosphorylation of serine 338. Nature 396:180-183. 192. Kittler, R., A. K. Heninger, K. Franke, B. Habermann, and F. Buchholz. 2005. Production of endoribonuclease-prepared short interfering RNAs for gene silencing in mammalian cells. Nat Methods 2:779-84. 193. Kitzmann, M., and A. Fernandez. 2001. Crosstalk between cell cycle regulators and the myogenic factor MyoD in skeletal myoblasts. Cell Mol Life Sci 58:571-9. 194. Koch, C. A., D. Anderson, M. F. Moran, C. Ellis, and T. Pawson. 1991. SH2 and SH3 domains: elements that control interactions of cytoplasmic signaling proteins. Science 252:668-74. 195. Kolch, W. 2005. Coordinating ERK/MAPK signalling through scaffolds and inhibitors. Nat Rev Mol Cell Biol 6:827-37. 196. Kornfeld, K., D. B. Hom, and H. R. Horvitz. 1995. The ksr-1 gene encodes a novel protein kinase involved in Ras-mediated signaling in C. elegans. Cell 83:903-13. 197. Kortum, R. L., and R. E. Lewis. 2004. The molecular scaffold KSR1 regulates the proliferative and oncogenic potential of cells. Mol Cell Biol 24:4407-16. 198. Kraft, C., F. Herzog, C. Gieffers, K. Mechtler, A. Hagting, J. Pines, and J. M. Peters. 2003. Mitotic regulation of the human anaphase-promoting complex by phosphorylation. Embo Journal 22:6598-6609. 199. Kramer, E. R., N. Scheuringer, A. V. Podtelejnikov, M. Mann, and J. M. Peters. 2000. Mitotic regulation of the APC activator proteins CDC20 and CDH1. Mol Biol Cell 11:1555-69. 200. Kranz, J. E., B. Satterberg, and E. A. Elion. 1994. The MAP kinase Fus3 associates with and phosphorylates the upstream signaling component Ste5. Genes Dev 8:313-27. 201. Krek, W., and E. A. Nigg. 1991. Mutations of p34cdc2 phosphorylation sites induce premature mitotic events in HeLa cells: evidence for a double block to p34cdc2 kinase activation in vertebrates. Embo J 10:3331-41. 202. Krien, M. J., S. J. Bugg, M. Palatsides, G. Asouline, M. Morimyo, and M. J. O'Connell. 1998. A NIMA homologue promotes chromatin condensation in fission yeast. J Cell Sci 111 ( Pt 7):967-76. 203. Krishna, M., and H. Narang. 2008. The complexity of mitogen-activated protein kinases (MAPKs) made simple. Cell Mol Life Sci 65:3525-44.

144 204. Kubbutat, M. H. G., S. N. Jones, and K. H. Vousden. 1997. Regulation of p53 stability by Mdm2. Nature 387:299-303. 205. Kumagai, A., and W. G. Dunphy. 1996. Purification and molecular cloning of Plx1, a Cdc25-regulatory kinase from Xenopus egg extracts. Science 273:1377-1380. 206. Kumar, R. A., S. KaraMohamed, J. Sudi, D. F. Conrad, C. Brune, J. A. Badner, T. C. Gilliam, N. J. Nowak, E. H. Cook, Jr., W. B. Dobyns, and S. L. Christian. 2008. Recurrent 16p11.2 microdeletions in autism. Hum Mol Genet 17:628-38. 207. LaBaer, J., M. D. Garrett, L. F. Stevenson, J. M. Slingerland, C. Sandhu, H. S. Chou, A. Fattaey, and E. Harlow. 1997. New functional activities for the p21 family of CDK inhibitors. Genes & Development 11:847-862. 208. Lackner, M. R., K. Kornfeld, L. M. Miller, H. R. Horvitz, and S. K. Kim. 1994. A MAP kinase homolog, mpk-1, is involved in ras-mediated induction of vulval cell fates in Caenorhabditis elegans. Genes Dev 8:160-73. 209. Lammer, C., S. Wagerer, R. Saffrich, D. Mertens, W. Ansorge, and I. Hoffmann. 1998. The cdc25B phosphatase is essential for the G2/M phase transition in human cells. J Cell Sci 111 ( Pt 16):2445-53. 210. Lan, W. J., X. Zhang, S. L. Kline-Smith, S. E. Rosasco, G. A. Barrett-Wilt, J. Shabanowitz, C. E. Walczak, and P. T. Stukenberg. 2004. Aurora B phosphorylates centromeric MCAK and regulates its localization and microtubule depolymerization activity. Current Biology 14:273-286. 211. Lane, H. A., and E. A. Nigg. 1996. Antibody microinjection reveals an essential role for human polo-like kinase 1 (Plk1) in the functional maturation of mitotic centrosomes. Journal of Cell Biology 135:1701-1713. 212. Lane, H. A., and E. A. Nigg. 1996. Antibody microinjection reveals an essential role for human polo-like kinase 1 (Plk1) in the functional maturation of mitotic centrosomes. J Cell Biol 135:1701-13. 213. Lane, M. E., M. Elend, D. Heidmann, A. Herr, S. Marzodko, A. Herzig, and C. F. Lehner. 2000. A screen for modifiers of cyclin E function in Drosophila melanogaster identifies Cdk2 mutations, revealing the insignificance of putative phosphorylation sites in Cdk2. Genetics 155:233-44. 214. Lassar, A., and A. Munsterberg. 1994. Wiring diagrams: regulatory circuits and the control of skeletal myogenesis. Curr Opin Cell Biol 6:432-42. 215. Lavin, M. F. 2008. Ataxia-telangiectasia: from a rare disorder to a paradigm for cell signalling and cancer. Nat Rev Mol Cell Biol 9:759-69. 216. Lee, C. H., and J. H. Chung. 2001. The hCds1 (Chk2)-FHA domain is essential for a chain of phosphorylation events on hCds1 that is induced by ionizing radiation. J Biol Chem 276:30537-41. 217. Lee, K. Y., I. Yang, J. E. Park, O. R. Baek, K. Y. Chung, and H. S. Koo. 2007. Developmental stage- and DNA damage-specific functions of C. elegans FANCD2. Biochem Biophys Res Commun 352:479-85. 218. Lee, M. Y., H. J. Kim, M. A. Kim, H. J. Jee, A. J. Kim, Y. S. Bae, J. I. Park, J. H. Chung, and J. Yun. 2008. Nek6 is involved in G2/M phase cell cycle arrest through DNA damage-induced phosphorylation. Cell Cycle 7:2705-9. 219. Lee, T., A. N. Hoofnagle, Y. Kabuyama, J. Stroud, X. Min, E. J. Goldsmith, L. Chen, K. A. Resing, and N. G. Ahn. 2004. Docking motif interactions in MAP kinases revealed by hydrogen exchange mass spectrometry. Mol Cell 14:43-55.

145 220. Leevers, S. J., H. F. Paterson, and C. J. Marshall. 1994. Requirement for Ras in Raf Activation Is Overcome by Targeting Raf to the Plasma-Membrane. Nature 369:411-414. 221. Lenormand, P., C. Sardet, G. Pages, G. L'Allemain, A. Brunet, and J. Pouyssegur. 1993. Growth factors induce nuclear translocation of MAP kinases (p42mapk and p44mapk) but not of their activator MAP kinase kinase (p45mapkk) in fibroblasts. J Cell Biol 122:1079-88. 222. Letwin, K., L. Mizzen, B. Motro, Y. Ben-David, A. Bernstein, and T. Pawson. 1992. A mammalian dual specificity protein kinase, Nek1, is related to the NIMA cell cycle regulator and highly expressed in meiotic germ cells. Embo J 11:3521-31. 223. Li, X., G. Sakashita, H. Matsuzaki, K. Sugimoto, K. Kimura, F. Hanaoka, H. Taniguchi, K. Furukawa, and T. Urano. 2004. Direct association with inner centromere protein (INCENP) activates the novel chromosomal passenger protein, Aurora-C. J Biol Chem 279:47201-11. 224. Lindberg, R. A., A. M. Quinn, and T. Hunter. 1992. Dual-specificity protein kinases: will any hydroxyl do? Trends Biochem Sci 17:114-9. 225. Lindqvist, A., V. Rodriguez-Bravo, and R. H. Medema. 2009. The decision to enter mitosis: feedback and redundancy in the mitotic entry network. J Cell Biol 185:193-202. 226. Lipp, J. J., T. Hirota, I. Poser, and J. M. Peters. 2007. Aurora B controls the association of condensin I but not condensin II with mitotic chromosomes. J Cell Sci 120:1245-55. 227. Little, J. B., H. Nagasawa, P. C. Keng, Y. J. Yu, and C. Y. Li. 1995. Absence of Radiation-Induced G(1) Arrest in 2 Closely-Related Human Lymphoblast Cell-Lines That Differ in P53 Status. Journal of Biological Chemistry 270:11033-11036. 228. Liu, F., J. J. Stanton, Z. Wu, and H. Piwnica-Worms. 1997. The human Myt1 kinase preferentially phosphorylates Cdc2 on threonine 14 and localizes to the endoplasmic reticulum and Golgi complex. Mol Cell Biol 17:571-83. 229. Liu, M. K., R. W. Brownsey, and N. E. Reiner. 1994. Gamma interferon induces rapid and coordinate activation of mitogen-activated protein kinase (extracellular signal-regulated kinase) and calcium-independent protein kinase C in human monocytes. Infect Immun 62:2722-31. 230. Liu, Q., and J. V. Ruderman. 2006. Aurora A, mitotic entry, and spindle bipolarity. Proc Natl Acad Sci U S A 103:5811-6. 231. Liu, Q. H., S. Guntuku, X. S. Cui, S. Matsuoka, D. Cortez, K. Tamai, G. B. Luo, S. Carattini-Rivera, F. DeMayo, A. Bradley, L. A. Donehower, and S. J. Elledge. 2000. Chk1 is an essential kinase that is regulated by Atr and required for the G(2)/M DNA damage checkpoint. Genes & Development 14:1448-1459. 232. Liu, S., W. Lu, T. Obara, S. Kuida, J. Lehoczky, K. Dewar, I. A. Drummond, and D. R. Beier. 2002. A defect in a novel Nek-family kinase causes cystic kidney disease in the mouse and in zebrafish. Development 129:5839-46. 233. Liu, Y., K. Z. Guyton, M. Gorospe, Q. Xu, J. C. Lee, and N. J. Holbrook. 1996. Differential activation of ERK, JNK/SAPK and P38/CSBP/RK map kinase family members during the cellular response to arsenite. Free Radic Biol Med 21:771-81. 234. Lizcano, J. M., M. Deak, N. Morrice, A. Kieloch, C. J. Hastie, L. Dong, M. Schutkowski, U. Reimer, and D. R. Alessi. 2002. Molecular basis for the substrate

146 specificity of NIMA-related kinase-6 (NEK6). Evidence that NEK6 does not phosphorylate the hydrophobic motif of ribosomal S6 protein kinase and serum- and glucocorticoid-induced protein kinase in vivo. J Biol Chem 277:27839-49. 235. Lobrich, M., and P. A. Jeggo. 2007. The impact of a negligent G2/M checkpoint on genomic instability and cancer induction. Nat Rev Cancer 7:861-9. 236. Lochhead, P. A., G. Sibbet, R. Kinstrie, T. Cleghon, M. Rylatt, D. K. Morrison, and V. Cleghon. 2003. dDYRK2: a novel dual-specificity tyrosine-phosphorylation- regulated kinase in Drosophila. Biochem J 374:381-91. 237. Lochhead, P. A., G. Sibbet, N. Morrice, and V. Cleghon. 2005. Activation-loop autophosphorylation is mediated by a novel transitional intermediate form of DYRKs. Cell 121:925-36. 238. Lou, Y., J. Yao, A. Zereshki, Z. Dou, K. Ahmed, H. Wang, J. Hu, Y. Wang, and X. Yao. 2004. NEK2A interacts with MAD1 and possibly functions as a novel integrator of the spindle checkpoint signaling. J Biol Chem 279:20049-57. 239. Lowe, M., C. Rabouille, N. Nakamura, R. Watson, M. Jackman, E. Jamsa, D. Rahman, D. J. Pappin, and G. Warren. 1998. Cdc2 kinase directly phosphorylates the cis-Golgi matrix protein GM130 and is required for Golgi fragmentation in mitosis. Cell 94:783-93. 240. Lu, K. P., and T. Hunter. 1995. Evidence for a NIMA-like mitotic pathway in vertebrate cells. Cell 81:413-24. 241. Lu, K. P., B. E. Kemp, and A. R. Means. 1994. Identification of substrate specificity determinants for the cell cycle-regulated NIMA protein kinase. J Biol Chem 269:6603-7. 242. Lundberg, A. S., and R. A. Weinberg. 1999. Control of the cell cycle and apoptosis. Eur J Cancer 35:531-9. 243. Macaluso, M., M. Montanari, and A. Giordano. 2006. Rb family proteins as modulators of gene expression and new aspects regarding the interaction with chromatin remodeling . Oncogene 25:5263-7. 244. Macurek, L., A. Lindqvist, D. Lim, M. A. Lampson, R. Klompmaker, R. Freire, C. Clouin, S. S. Taylor, M. B. Yaffe, and R. H. Medema. 2008. Polo-like kinase-1 is activated by aurora A to promote checkpoint recovery. Nature 455:119-U88. 245. Maeda, I., Y. Kohara, M. Yamamoto, and A. Sugimoto. 2001. Large-scale analysis of gene function in Caenorhabditis elegans by high-throughput RNAi. Curr Biol 11:171-6. 246. Mahjoub, M. R., B. Montpetit, L. Zhao, R. J. Finst, B. Goh, A. C. Kim, and L. M. Quarmby. 2002. The FA2 gene of Chlamydomonas encodes a NIMA family kinase with roles in cell cycle progression and microtubule severing during deflagellation. J Cell Sci 115:1759-68. 247. Mailand, N., A. V. Podtelejnikov, A. Groth, M. Mann, J. Bartek, and J. Lukas. 2002. Regulation of G(2)/M events by Cdc25A through phosphorylation-dependent modulation of its stability. Embo J 21:5911-20. 248. Manke, I. A., A. Nguyen, D. Lim, M. Q. Stewart, A. E. Elia, and M. B. Yaffe. 2005. MAPKAP kinase-2 is a cell cycle checkpoint kinase that regulates the G2/M transition and S phase progression in response to UV irradiation. Mol Cell 17:37-48. 249. Manning, G., D. B. Whyte, R. Martinez, T. Hunter, and S. Sudarsanam. 2002. The protein kinase complement of the . Science 298:1912-34.

147 250. Mansour, S. J., J. M. Candia, K. K. Gloor, and N. G. Ahn. 1996. Constitutively active mitogen-activated protein kinase kinase 1 (MAPKK1) and MAPKK2 mediate similar transcriptional and morphological responses. Cell Growth Differ 7:243-50. 251. Marais, R., Y. Light, H. F. Paterson, and C. J. Marshall. 1995. Ras Recruits Raf-1 to the Plasma-Membrane for Activation by Tyrosine Phosphorylation. Embo Journal 14:3136-3145. 252. Marais, R., Y. Light, H. F. Paterson, C. S. Mason, and C. J. Marshall. 1997. Differential regulation of Raf-1, A-Raf, and B-Raf by oncogenic ras and tyrosine kinases. Journal of Biological Chemistry 272:4378-4383. 253. Mason, C. S., C. J. Springer, R. G. Cooper, G. Superti-Furga, C. J. Marshall, and R. Marais. 1999. Serine and tyrosine phosphorylations cooperate in Raf-1, but not B-Raf activation. Embo J 18:2137-48. 254. Matsuoka, S., B. A. Ballif, A. Smogorzewska, E. R. McDonald, 3rd, K. E. Hurov, J. Luo, C. E. Bakalarski, Z. Zhao, N. Solimini, Y. Lerenthal, Y. Shiloh, S. P. Gygi, and S. J. Elledge. 2007. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 316:1160-6. 255. Matsuoka, S., M. Huang, and S. J. Elledge. 1998. Linkage of ATM to cell cycle regulation by the Chk2 protein kinase. Science 282:1893-7. 256. Matsuoka, S., G. Rotman, A. Ogawa, Y. Shiloh, K. Tamai, and S. J. Elledge. 2000. Ataxia telangiectasia-mutated phosphorylates Chk2 in vivo and in vitro. Proceedings of the National Academy of Sciences of the United States of America 97:10389-10394. 257. Matsushita, T., Y. Y. Chan, A. Kawanami, G. Balmes, G. E. Landreth, and S. Murakami. 2009. Extracellular signal-regulated kinase 1 (ERK1) and ERK2 play essential roles in osteoblast differentiation and in supporting osteoclastogenesis. Mol Cell Biol 29:5843-57. 258. Mazzucchelli, C., C. Vantaggiato, A. Ciamei, S. Fasano, P. Pakhotin, W. Krezel, H. Welzl, D. P. Wolfer, G. Pages, O. Valverde, A. Marowsky, A. Porrazzo, P. C. Orban, R. Maldonado, M. U. Ehrengruber, V. Cestari, H. P. Lipp, P. F. Chapman, J. Pouyssegur, and R. Brambilla. 2002. Knockout of ERK1 MAP kinase enhances synaptic plasticity in the striatum and facilitates striatal-mediated learning and memory. Neuron 34:807-20. 259. McKay, M. M., and D. K. Morrison. 2007. Integrating signals from RTKs to ERK/MAPK. Oncogene 26:3113-21. 260. Melixetian, M., D. K. Klein, C. S. Sorensen, and K. Helin. 2009. NEK11 regulates CDC25A degradation and the IR-induced G2/M checkpoint. Nat Cell Biol 11:1247- 53. 261. Meloche, S., and J. Pouyssegur. 2007. The ERK1/2 mitogen-activated protein kinase pathway as a master regulator of the G1- to S-phase transition. Oncogene 26:3227-3239. 262. Meloche, S., K. Seuwen, G. Pages, and J. Pouyssegur. 1992. Biphasic and synergistic activation of p44mapk (ERK1) by growth factors: correlation between late phase activation and mitogenicity. Mol Endocrinol 6:845-54. 263. Meraldi, P., and E. A. Nigg. 2001. Centrosome cohesion is regulated by a balance of kinase and phosphatase activities. J Cell Sci 114:3749-57.

148 264. Merienne, K., S. Jacquot, M. Zeniou, S. Pannetier, P. Sassone-Corsi, and A. Hanauer. 2000. Activation of RSK by UV-light: phosphorylation dynamics and involvement of the MAPK pathway. Oncogene 19:4221-9. 265. Meulmeester, E., Y. Pereg, Y. Shiloh, and A. G. Jochemsen. 2005. ATM-mediated phosphorylations inhibit Mdmx/Mdm2 stabilization by HAUSP in favor of p53 activation. Cell Cycle 4:1166-1170. 266. Mi, J., C. Guo, D. L. Brautigan, and J. M. Larner. 2007. Protein phosphatase- 1alpha regulates centrosome splitting through Nek2. Cancer Res 67:1082-9. 267. Mikula, M., M. Schreiber, Z. Husak, L. Kucerova, J. Ruth, R. Wieser, K. Zatloukal, H. Beug, E. F. Wagner, and M. Baccarini. 2001. Embryonic lethality and fetal liver apoptosis in mice lacking the c-raf-1 gene. Embo J 20:1952-62. 268. Millar, J. B., J. Blevitt, L. Gerace, K. Sadhu, C. Featherstone, and P. Russell. 1991. p55CDC25 is a nuclear protein required for the initiation of mitosis in human cells. Proc Natl Acad Sci U S A 88:10500-4. 269. Minet, E., T. Arnould, G. Michel, I. Roland, D. Mottet, M. Raes, J. Remacle, and C. Michiels. 2000. ERK activation upon hypoxia: involvement in HIF-1 activation. FEBS Lett 468:53-8. 270. Moarefi, I., M. LaFevre-Bernt, F. Sicheri, M. Huse, C. H. Lee, J. Kuriyan, and W. T. Miller. 1997. Activation of the Src-family tyrosine kinase Hck by SH3 domain displacement. Nature 385:650-3. 271. Mogila, V., F. Xia, and W. X. Li. 2006. An intrinsic cell cycle checkpoint pathway mediated by MEK and ERK in Drosophila. Dev Cell 11:575-82. 272. Molinari, M., C. Mercurio, J. Dominguez, F. Goubin, and G. F. Draetta. 2000. Human Cdc25 A inactivation in response to S phase inhibition and its role in preventing premature mitosis. EMBO Rep 1:71-9. 273. Momand, J., G. P. Zambetti, D. C. Olson, D. George, and A. J. Levine. 1992. The Mdm-2 Oncogene Product Forms a Complex with the P53 Protein and Inhibits P53- Mediated Transactivation. Cell 69:1237-1245. 274. Moodie, S. A., B. M. Willumsen, M. J. Weber, and A. Wolfman. 1993. Complexes of Ras.GTP with Raf-1 and mitogen-activated protein kinase kinase. Science 260:1658-61. 275. Moreno, S., J. Hayles, and P. Nurse. 1989. Regulation of p34cdc2 protein kinase during mitosis. Cell 58:361-72. 276. Morgan, D. O. 1995. Principles of CDK regulation. Nature 374:131-4. 277. Morgan, M. R., M. J. Humphries, and M. D. Bass. 2007. Synergistic control of cell adhesion by integrins and syndecans. Nat Rev Mol Cell Biol 8:957-69. 278. Morris, N. R. 1975. Mitotic mutants of Aspergillus nidulans. Genet Res 26:237-54. 279. Mueller, P. R., T. R. Coleman, A. Kumagai, and W. G. Dunphy. 1995. Myt1: a membrane-associated inhibitory kinase that phosphorylates Cdc2 on both threonine- 14 and tyrosine-15. Science 270:86-90. 280. Murphy, L. O., S. Smith, R. H. Chen, D. C. Fingar, and J. Blenis. 2002. Molecular interpretation of ERK signal duration by immediate early gene products. Nature Cell Biology 4:556-564. 281. Murray, A. W., M. J. Solomon, and M. W. Kirschner. 1989. The Role of Cyclin Synthesis and Degradation in the Control of Maturation Promoting Factor Activity. Nature 339:280-286.

149 282. Musacchio, A., and E. D. Salmon. 2007. The spindle-assembly checkpoint in space and time. Nat Rev Mol Cell Biol 8:379-93. 283. Muslin, A. J., J. W. Tanner, P. M. Allen, and A. S. Shaw. 1996. Interaction of 14- 3-3 with signaling proteins is mediated by the recognition of phosphoserine. Cell 84:889-97. 284. Nakajima, H., S. Yonemura, M. Murata, N. Nakamura, H. Piwnica-Worms, and E. Nishida. 2008. Myt1 protein kinase is essential for Golgi and ER assembly during mitotic exit. J Cell Biol 181:89-103. 285. Nigg, E. A. 2001. Mitotic kinases as regulators of cell division and its checkpoints. Nat Rev Mol Cell Biol 2:21-32. 286. Noguchi, K., H. Fukazawa, Y. Murakami, and Y. Uehara. 2002. Nek11, a new member of the NIMA family of kinases, involved in DNA replication and genotoxic stress responses. J Biol Chem 277:39655-65. 287. Nolen, B., S. Taylor, and G. Ghosh. 2004. Regulation of protein kinases; controlling activity through activation segment conformation. Mol Cell 15:661-75. 288. O'Connell, M. J., M. J. Krien, and T. Hunter. 2003. Never say never. The NIMA- related protein kinases in mitotic control. Trends Cell Biol 13:221-8. 289. O'Connell, M. J., C. Norbury, and P. Nurse. 1994. Premature chromatin condensation upon accumulation of NIMA. Embo J 13:4926-37. 290. O'Driscoll, M., V. L. Ruiz-Perez, C. G. Woods, P. A. Jeggo, and J. A. Goodship. 2003. A splicing mutation affecting expression of ataxia-telangiectasia and Rad3- related protein (ATR) results in Seckel syndrome. Nat Genet 33:497-501. 291. O'Neill, E., L. Rushworth, M. Baccarini, and W. Kolch. 2004. Role of the kinase MST2 in suppression of apoptosis by the proto-oncogene product Raf-1. Science 306:2267-70. 292. O'Regan, L., and A. M. Fry. 2009. The Nek6 and Nek7 protein kinases are required for robust mitotic spindle formation and cytokinesis. Mol Cell Biol 29:3975-90. 293. Oakley, B. R., and N. R. Morris. 1983. A mutation in Aspergillus nidulans that blocks the transition from interphase to prophase. J Cell Biol 96:1155-8. 294. Obaya, A. J., I. Kotenko, M. D. Cole, and J. M. Sedivy. 2002. The proto-oncogene c-myc acts through the cyclin-dependent kinase (Cdk) inhibitor p27(Kip1) to facilitate the activation of Cdk4/6 and early G(1) phase progression. J Biol Chem 277:31263-9. 295. Ohtani, K., J. DeGregori, and J. R. Nevins. 1995. Regulation of the cyclin E gene by transcription factor E2F1. Proc Natl Acad Sci U S A 92:12146-50. 296. Osmani, A. H., S. L. McGuire, and S. A. Osmani. 1991. Parallel activation of the NIMA and p34cdc2 cell cycle-regulated protein kinases is required to initiate mitosis in A. nidulans. Cell 67:283-91. 297. Osmani, S. A., R. T. Pu, and N. R. Morris. 1988. Mitotic induction and maintenance by overexpression of a G2-specific gene that encodes a potential protein kinase. Cell 53:237-44. 298. Osmani, S. A., and X. S. Ye. 1996. Cell cycle regulation in Aspergillus by two protein kinases. Biochem J 317 ( Pt 3):633-41. 299. Pages, G., S. Guerin, D. Grall, F. Bonino, A. Smith, F. Anjuere, P. Auberger, and J. Pouyssegur. 1999. Defective thymocyte maturation in p44 MAP kinase (Erk 1) knockout mice. Science 286:1374-7.

150 300. Painter, R. B., and B. R. Young. 1980. Radiosensitivity in ataxia-telangiectasia: a new explanation. Proc Natl Acad Sci U S A 77:7315-7. 301. Palmer, A., A. C. Gavin, and A. R. Nebreda. 1998. A link between MAP kinase and p(34cdc2) cyclin B during oocyte maturation: p90(rsk) phosphorylates and inactivates the p34(cdc2) inhibitory kinase Myt1. Embo Journal 17:5037-5047. 302. Pandit, B., A. Sarkozy, L. A. Pennacchio, C. Carta, K. Oishi, S. Martinelli, E. A. Pogna, W. Schackwitz, A. Ustaszewska, A. Landstrom, J. M. Bos, S. R. Ommen, G. Esposito, F. Lepri, C. Faul, P. Mundel, J. P. Lopez Siguero, R. Tenconi, A. Selicorni, C. Rossi, L. Mazzanti, I. Torrente, B. Marino, M. C. Digilio, G. Zampino, M. J. Ackerman, B. Dallapiccola, M. Tartaglia, and B. D. Gelb. 2007. Gain-of-function RAF1 mutations cause Noonan and LEOPARD syndromes with hypertrophic cardiomyopathy. Nat Genet 39:1007-12. 303. Park, E. R., S. T. Eblen, and A. D. Catling. 2007. MEK1 activation by PAK: a novel mechanism. Cell Signal 19:1488-96. 304. Parker, L. L., and H. Piwnica-Worms. 1992. Inactivation of the p34cdc2-cyclin B complex by the human WEE1 tyrosine kinase. Science 257:1955-7. 305. Patterson, K. I., T. Brummer, P. M. O'Brien, and R. J. Daly. 2009. Dual- specificity phosphatases: critical regulators with diverse cellular targets. Biochem J 418:475-89. 306. Pei, X. H., and Y. Xiong. 2005. Biochemical and cellular mechanisms of mammalian CDK inhibitors: a few unresolved issues. Oncogene 24:2787-95. 307. Peng, C. Y., P. R. Graves, R. S. Thoma, Z. Q. Wu, A. S. Shaw, and H. PiwnicaWorms. 1997. Mitotic and G(2) checkpoint control: Regulation of 14-3-3 protein binding by phosphorylation of Cdc25C on serine-216. Science 277:1501- 1505. 308. Perez-Roger, I., S. H. Kim, B. Griffiths, A. Sewing, and H. Land. 1999. Cyclins D1 and D2 mediate myc-induced proliferation via sequestration of p27(Kip1) and p21(Cip1). Embo J 18:5310-20. 309. Peters, J. M. 2006. The anaphase promoting complex/cyclosome: a machine designed to destroy. Nature Reviews Molecular Cell Biology 7:644-656. 310. Petronczki, M., M. Glotzer, N. Kraut, and J. M. Peters. 2007. Polo-like kinase 1 triggers the initiation of cytokinesis in human cells by promoting recruitment of the RhoGEF Ect2 to the central spindle. Developmental Cell 12:713-725. 311. Pike, A. C., P. Rellos, F. H. Niesen, A. Turnbull, A. W. Oliver, S. A. Parker, B. E. Turk, L. H. Pearl, and S. Knapp. 2008. Activation segment dimerization: a mechanism for kinase autophosphorylation of non-consensus sites. Embo J 27:704- 14. 312. Planas-Silva, M. D., and R. A. Weinberg. 1997. The restriction point and control of cell proliferation. Curr Opin Cell Biol 9:768-72. 313. Polci, R., A. Peng, P. L. Chen, D. J. Riley, and Y. Chen. 2004. NIMA-related protein kinase 1 is involved early in the ionizing radiation-induced DNA damage response. Cancer Res 64:8800-3. 314. Polesskaya, A., and M. A. Rudnicki. 2002. A MyoD-dependent differentiation checkpoint: ensuring genome integrity. Dev Cell 3:757-8.

151 315. Poon, R. Y., W. Jiang, H. Toyoshima, and T. Hunter. 1996. Cyclin-dependent kinases are inactivated by a combination of p21 and Thr-14/Tyr-15 phosphorylation after UV-induced DNA damage. J Biol Chem 271:13283-91. 316. Posada, J., N. Yew, N. G. Ahn, G. F. Vande Woude, and J. A. Cooper. 1993. Mos stimulates MAP kinase in Xenopus oocytes and activates a MAP kinase kinase in vitro. Mol Cell Biol 13:2546-53. 317. Price, M. A., F. H. Cruzalegui, and R. Treisman. 1996. The p38 and ERK MAP kinase pathways cooperate to activate Ternary Complex Factors and c-fos transcription in response to UV light. Embo J 15:6552-63. 318. Pritchard, C. A., L. Bolin, R. Slattery, R. Murray, and M. McMahon. 1996. Post- natal lethality and neurological and gastrointestinal defects in mice with targeted disruption of the A-Raf protein kinase gene. Curr Biol 6:614-7. 319. Pu, R. T., and S. A. Osmani. 1995. Mitotic destruction of the cell cycle regulated NIMA protein kinase of Aspergillus nidulans is required for mitotic exit. Embo J 14:995-1003. 320. Pu, R. T., G. Xu, L. Wu, J. Vierula, K. O'Donnell, X. S. Ye, and S. A. Osmani. 1995. Isolation of a functional homolog of the cell cycle-specific NIMA protein kinase of Aspergillus nidulans and functional analysis of conserved residues. J Biol Chem 270:18110-6. 321. Pugacheva, E. N., S. A. Jablonski, T. R. Hartman, E. P. Henske, and E. A. Golemis. 2007. HEF1-Dependent aurora a activation induces disassembly of the primary cilium. Cell 129:1351-1363. 322. Qian, Y. W., E. Erikson, C. Li, and J. L. Maller. 1998. Activated polo-like kinase Plx1 is required at multiple points during mitosis in Xenopus laevis. Molecular and Cellular Biology 18:4262-4271. 323. Quarmby, L. M., and M. R. Mahjoub. 2005. Caught Nek-ing: cilia and centrioles. J Cell Sci 118:5161-9. 324. Quarmby, L. M., and J. D. Parker. 2005. Cilia and the cell cycle? J Cell Biol 169:707-10. 325. Radler-Pohl, A., C. Sachsenmaier, S. Gebel, H. P. Auer, J. T. Bruder, U. Rapp, P. Angel, H. J. Rahmsdorf, and P. Herrlich. 1993. UV-induced activation of AP-1 involves obligatory extranuclear steps including Raf-1 kinase. Embo J 12:1005-12. 326. Rajakulendran, T., M. Sahmi, M. Lefrancois, F. Sicheri, and M. Therrien. 2009. A dimerization-dependent mechanism drives RAF catalytic activation. Nature 461:542-5. 327. Raman, M., S. Earnest, K. Zhang, Y. Zhao, and M. H. Cobb. 2007. TAO kinases mediate activation of p38 in response to DNA damage. Embo J 26:2005-14. 328. Rapp, U. R., M. D. Goldsborough, G. E. Mark, T. I. Bonner, J. Groffen, F. H. Reynolds, Jr., and J. R. Stephenson. 1983. Structure and biological activity of v-raf, a unique oncogene transduced by a retrovirus. Proc Natl Acad Sci U S A 80:4218-22. 329. Razidlo, G. L., H. J. Johnson, S. M. Stoeger, K. H. Cowan, T. Bessho, and R. E. Lewis. 2009. KSR1 is required for cell cycle re-initiation following DNA damage. J Biol Chem. 330. Reinhardt, H. C., A. S. Aslanian, J. A. Lees, and M. B. Yaffe. 2007. p53-deficient cells rely on ATM- and ATR-mediated checkpoint signaling through the p38MAPK/MK2 pathway for survival after DNA damage. Cancer Cell 11:175-89.

152 331. Reinhardt, H. C., and M. B. Yaffe. 2009. Kinases that control the cell cycle in response to DNA damage: Chk1, Chk2, and MK2. Current Opinion in Cell Biology 21:245-255. 332. Rellos, P., F. J. Ivins, J. E. Baxter, A. Pike, T. J. Nott, D. M. Parkinson, S. Das, S. Howell, O. Fedorov, Q. Y. Shen, A. M. Fry, S. Knapp, and S. J. Smerdon. 2007. Structure and regulation of the human Nek2 centrosomal kinase. J Biol Chem 282:6833-42. 333. Reszka, A. A., R. Seger, C. D. Diltz, E. G. Krebs, and E. H. Fischer. 1995. Association of mitogen-activated protein kinase with the microtubule cytoskeleton. Proc Natl Acad Sci U S A 92:8881-5. 334. Richards, M. W., L. O'Regan, C. Mas-Droux, J. M. Blot, J. Cheung, S. Hoelder, A. M. Fry, and R. Bayliss. 2009. An autoinhibitory tyrosine motif in the cell-cycle- regulated Nek7 kinase is released through binding of Nek9. Mol Cell 36:560-70. 335. Rieder, C. L., R. W. Cole, A. Khodjakov, and G. Sluder. 1995. The Checkpoint Delaying Anaphase in Response to Chromosome Monoorientation Is Mediated by an Inhibitory Signal Produced by Unattached Kinetochores. Journal of Cell Biology 130:941-948. 336. Rieder, C. L., A. Schultz, R. Cole, and G. Sluder. 1994. Anaphase Onset in Vertebrate Somatic-Cells Is Controlled by a Checkpoint That Monitors Sister Kinetochore Attachment to the Spindle. Journal of Cell Biology 127:1301-1310. 337. Ritt, D. A., D. M. Monson, S. I. Specht, and D. K. Morrison. Impact of feedback phosphorylation and Raf heterodimerization on normal and mutant B-Raf signaling. Mol Cell Biol 30:806-19. 338. Roberts, E. C., P. S. Shapiro, T. S. Nahreini, G. Pages, J. Pouyssegur, and N. G. Ahn. 2002. Distinct cell cycle timing requirements for extracellular signal-regulated kinase and phosphoinositide 3-kinase signaling pathways in somatic cell mitosis. Mol Cell Biol 22:7226-41. 339. Robinson, M. J., S. A. Stippec, E. Goldsmith, M. A. White, and M. H. Cobb. 1998. A constitutively active and nuclear form of the MAP kinase ERK2 is sufficient for neurite outgrowth and cell transformation. Current Biology 8:1141-1150. 340. Roig, J., A. Mikhailov, C. Belham, and J. Avruch. 2002. Nercc1, a mammalian NIMA-family kinase, binds the Ran GTPase and regulates mitotic progression. Genes Dev 16:1640-58. 341. Rossomando, A. J., D. M. Payne, M. J. Weber, and T. W. Sturgill. 1989. Evidence That Pp42, a Major Tyrosine Kinase Target Protein, Is a Mitogen-Activated Serine Threonine Protein-Kinase. Proceedings of the National Academy of Sciences of the United States of America 86:6940-6943. 342. Rossomando, A. J., D. M. Payne, M. J. Weber, and T. W. Sturgill. 1989. Evidence that pp42, a major tyrosine kinase target protein, is a mitogen-activated serine/threonine protein kinase. Proc Natl Acad Sci U S A 86:6940-3. 343. Roux, P. P., and J. Blenis. 2004. ERK and p38 MAPK-activated protein kinases: a family of protein kinases with diverse biological functions. Microbiol Mol Biol Rev 68:320-44. 344. Roy, F., G. Laberge, M. Douziech, D. Ferland-McCollough, and M. Therrien. 2002. KSR is a scaffold required for activation of the ERK/MAPK module. Genes Dev 16:427-38.

153 345. Rubin, G. M., M. D. Yandell, J. R. Wortman, G. L. Gabor Miklos, C. R. Nelson, I. K. Hariharan, M. E. Fortini, P. W. Li, R. Apweiler, W. Fleischmann, J. M. Cherry, S. Henikoff, M. P. Skupski, S. Misra, M. Ashburner, E. Birney, M. S. Boguski, T. Brody, P. Brokstein, S. E. Celniker, S. A. Chervitz, D. Coates, A. Cravchik, A. Gabrielian, R. F. Galle, W. M. Gelbart, R. A. George, L. S. Goldstein, F. Gong, P. Guan, N. L. Harris, B. A. Hay, R. A. Hoskins, J. Li, Z. Li, R. O. Hynes, S. J. Jones, P. M. Kuehl, B. Lemaitre, J. T. Littleton, D. K. Morrison, C. Mungall, P. H. O'Farrell, O. K. Pickeral, C. Shue, L. B. Vosshall, J. Zhang, Q. Zhao, X. H. Zheng, and S. Lewis. 2000. Comparative genomics of the eukaryotes. Science 287:2204-15. 346. Rudner, A. D., and A. W. Murray. 2000. Phosphorylation by Cdc28 activates the Cdc20-dependent activity of the anaphase-promoting complex. Journal of Cell Biology 149:1377-1390. 347. Rudolph, J. 2007. Cdc25 phosphatases: structure, specificity, and mechanism. Biochemistry 46:3595-604. 348. Rushworth, L. K., A. D. Hindley, E. O'Neill, and W. Kolch. 2006. Regulation and role of Raf-1/B-Raf heterodimerization. Mol Cell Biol 26:2262-72. 349. Russo, A. A., P. D. Jeffrey, and N. P. Pavletich. 1996. Structural basis of cyclin- dependent kinase activation by phosphorylation. Nat Struct Biol 3:696-700. 350. Sagata, N. 1997. What does Mos do in oocytes and somatic cells? Bioessays 19:13- 21. 351. Samuels, I. S., S. C. Saitta, and G. E. Landreth. 2009. MAP'ing CNS development and cognition: an ERKsome process. Neuron 61:160-7. 352. Samuels, M. L., and M. McMahon. 1994. Inhibition of platelet-derived growth factor- and epidermal growth factor-mediated mitogenesis and signaling in 3T3 cells expressing delta Raf-1:ER, an estradiol-regulated form of Raf-1. Mol Cell Biol 14:7855-66. 353. Sancar, A., L. A. Lindsey-Boltz, K. Unsal-Kacmaz, and S. Linn. 2004. Molecular mechanisms of mammalian DNA repair and the DNA damage checkpoints. Annu Rev Biochem 73:39-85. 354. Sartorelli, V., K. A. Webster, and L. Kedes. 1990. Muscle-specific expression of the cardiac alpha-actin gene requires MyoD1, CArG-box binding factor, and Sp1. Genes Dev 4:1811-22. 355. Savitsky, K., A. Bar-Shira, S. Gilad, G. Rotman, Y. Ziv, L. Vanagaite, D. A. Tagle, S. Smith, T. Uziel, S. Sfez, M. Ashkenazi, I. Pecker, M. Frydman, R. Harnik, S. R. Patanjali, A. Simmons, G. A. Clines, A. Sartiel, R. A. Gatti, L. Chessa, O. Sanal, M. F. Lavin, N. G. Jaspers, A. M. Taylor, C. F. Arlett, T. Miki, S. M. Weissman, M. Lovett, F. S. Collins, and Y. Shiloh. 1995. A single ataxia telangiectasia gene with a product similar to PI-3 kinase. Science 268:1749-53. 356. Schulman, B. A., D. L. Lindstrom, and E. Harlow. 1998. Substrate recruitment to cyclin-dependent kinase 2 by a multipurpose docking site on cyclin A. Proc Natl Acad Sci U S A 95:10453-8. 357. Schulze-Gahmen, U., and S. H. Kim. 2002. Structural basis for CDK6 activation by a virus-encoded cyclin. Nat Struct Biol 9:177-81. 358. Schumacher, J. M., A. Golden, and P. J. Donovan. 1998. AIR-2: An Aurora/Ip11- related protein kinase associated with chromosomes and midbody microtubules is

154 required for polar body extrusion and cytokinesis in Caenorhabditis elegans embryos. Journal of Cell Biology 143:1635-1646. 359. Sears, R., F. Nuckolls, E. Haura, Y. Taya, K. Tamai, and J. R. Nevins. 2000. Multiple Ras-dependent phosphorylation pathways regulate Myc protein stability. Genes & Development 14:2501-2514. 360. Seimiya, H., H. Sawada, Y. Muramatsu, M. Shimizu, K. Ohko, K. Yamane, and T. Tsuruo. 2000. Involvement of 14-3-3 proteins in nuclear localization of telomerase. Embo J 19:2652-61. 361. Seki, A., J. A. Coppinger, C. Y. Jang, J. R. Yates, and G. W. Fang. 2008. Bora and the kinase Aurora A cooperatively activate the kinase Plk1 and control mitotic entry. Science 320:1655-1658. 362. Sengupta, S., J. A. den Boon, I. H. Chen, M. A. Newton, D. B. Dahl, M. Chen, Y. J. Cheng, W. H. Westra, C. J. Chen, A. Hildesheim, B. Sugden, and P. Ahlquist. 2006. Genome-wide expression profiling reveals EBV-associated inhibition of MHC class I expression in nasopharyngeal carcinoma. Cancer Res 66:7999-8006. 363. Serrano, M., A. W. Lin, M. E. McCurrach, D. Beach, and S. W. Lowe. 1997. Oncogenic ras provokes premature cell senescence associated with accumulation of p53 and p16INK4a. Cell 88:593-602. 364. Sessa, F., M. Mapelli, C. Ciferri, C. Tarricone, L. B. Areces, T. R. Schneider, P. T. Stukenberg, and A. Musacchio. 2005. Mechanism of Aurora B activation by INCENP and inhibition by hesperadin. Mol Cell 18:379-91. 365. Shirayama, M., W. Zachariae, R. Ciosk, and K. Nasmyth. 1998. The Polo-like kinase Cdc5p and the WD-repeat protein Cdc20p/fizzy are regulators and substrates of the anaphase promoting complex in Saccharomyces cerevisiae. Embo J 17:1336- 49. 366. Sicheri, F., I. Moarefi, and J. Kuriyan. 1997. Crystal structure of the Src family tyrosine kinase Hck. Nature 385:602-9. 367. Siegel, P. M., E. D. Ryan, R. D. Cardiff, and W. J. Muller. 1999. Elevated expression of activated forms of Neu/ErbB-2 and ErbB-3 are involved in the induction of mammary tumors in transgenic mice: implications for human breast cancer. Embo J 18:2149-64. 368. Sigrist, S., H. Jacobs, R. Stratmann, and C. F. Lehner. 1995. Exit from mitosis is regulated by Drosophila fizzy and the sequential destruction of cyclins A, B and B3. Embo J 14:4827-38. 369. Slack-Davis, J. K., S. T. Eblen, M. Zecevic, S. A. Boerner, A. Tarcsafalvi, H. B. Diaz, M. S. Marshall, M. J. Weber, J. T. Parsons, and A. D. Catling. 2003. PAK1 phosphorylation of MEK1 regulates fibronectin-stimulated MAPK activation. J Cell Biol 162:281-91. 370. Smith, L. A., N. O. Bukanov, H. Husson, R. J. Russo, T. C. Barry, A. L. Taylor, D. R. Beier, and O. Ibraghimov-Beskrovnaya. 2006. Development of polycystic kidney disease in juvenile cystic kidney mice: insights into pathogenesis, ciliary abnormalities, and common features with human disease. J Am Soc Nephrol 17:2821-31. 371. Smolka, M. B., C. P. Albuquerque, S. H. Chen, and H. Zhou. 2007. Proteome- wide identification of in vivo targets of DNA damage checkpoint kinases. Proc Natl Acad Sci U S A 104:10364-9.

155 372. Solomon, M. J., T. Lee, and M. W. Kirschner. 1992. Role of phosphorylation in p34cdc2 activation: identification of an activating kinase. Mol Biol Cell 3:13-27. 373. Songyang, Z. 2001. Analysis of protein kinase specificity by peptide libraries and prediction of in vivo substrates. Methods Enzymol 332:171-83. 374. Sonnichsen, B., L. B. Koski, A. Walsh, P. Marschall, B. Neumann, M. Brehm, A. M. Alleaume, J. Artelt, P. Bettencourt, E. Cassin, M. Hewitson, C. Holz, M. Khan, S. Lazik, C. Martin, B. Nitzsche, M. Ruer, J. Stamford, M. Winzi, R. Heinkel, M. Roder, J. Finell, H. Hantsch, S. J. Jones, M. Jones, F. Piano, K. C. Gunsalus, K. Oegema, P. Gonczy, A. Coulson, A. A. Hyman, and C. J. Echeverri. 2005. Full-genome RNAi profiling of early embryogenesis in Caenorhabditis elegans. Nature 434:462-9. 375. Sorokin, S. 1962. Centrioles and the formation of rudimentary cilia by fibroblasts and smooth muscle cells. J Cell Biol 15:363-77. 376. Sorrentino, V., R. Pepperkok, R. L. Davis, W. Ansorge, and L. Philipson. 1990. Cell proliferation inhibited by MyoD1 independently of myogenic differentiation. Nature 345:813-5. 377. Stanton, V. P., Jr., D. W. Nichols, A. P. Laudano, and G. M. Cooper. 1989. Definition of the human raf amino-terminal regulatory region by deletion mutagenesis. Mol Cell Biol 9:639-47. 378. Stein, G. H., L. F. Drullinger, A. Soulard, and V. Dulic. 1999. Differential roles for cyclin-dependent kinase inhibitors p21 and p16 in the mechanisms of senescence and differentiation in human fibroblasts. Mol Cell Biol 19:2109-17. 379. Stoker, M., C. O'Neill, S. Berryman, and V. Waxman. 1968. Anchorage and growth regulation in normal and virus-transformed cells. Int J Cancer 3:683-93. 380. Stommel, J. M., and G. M. Wahl. 2004. Accelerated MDM2 auto-degradation induced by DNA-damage kinases is required for p53 activation. Embo Journal 23:1547-1556. 381. Sumara, I., J. F. Gimenez-Abian, D. Gerlich, T. Hirota, C. Kraft, C. de la Torre, J. Ellenberg, and J. M. Peters. 2004. Roles of polo-like kinase 1 in the assembly of functional mitotic spindles. Current Biology 14:1712-1722. 382. Sumara, I., E. Vorlaufer, P. T. Stukenberg, O. Kelm, N. Redemann, E. A. Nigg, and J. M. Peters. 2002. The dissociation of cohesin from chromosomes in prophase is regulated by Polo-like kinase. Mol Cell 9:515-25. 383. Sundaram, M., and M. Han. 1995. The C. elegans ksr-1 gene encodes a novel Raf- related kinase involved in Ras-mediated signal transduction. Cell 83:889-901. 384. Sunkel, C. E., and D. M. Glover. 1988. Polo, a Mitotic Mutant of Drosophila Displaying Abnormal Spindle Poles. Journal of Cell Science 89:25-38. 385. Suzuki-Takahashi, I., M. Kitagawa, M. Saijo, H. Higashi, H. Ogino, H. Matsumoto, Y. Taya, S. Nishimura, and A. Okuyama. 1995. The interactions of E2F with pRB and with p107 are regulated via the phosphorylation of pRB and p107 by a cyclin-dependent kinase. Oncogene 10:1691-8. 386. Takai, H., K. Naka, Y. Okada, M. Watanabe, N. Harada, S. Saito, C. W. Anderson, E. Appella, M. Nakanishi, H. Suzuki, K. Nagashima, H. Sawa, K. Ikeda, and N. Motoyama. 2002. Chk2-deficient mice exhibit radioresistance and defective p53-mediated transcription. Embo Journal 21:5195-5205.

156 387. Takai, H., K. Tominaga, N. Motoyama, Y. A. Minamishima, H. Nagahama, T. Tsukiyama, K. Ikeda, K. Nakayama, N. Nakanishi, and K. Nakayama. 2000. Aberrant cell cycle checkpoint function and early embryonic death in Chk1(-/-) mice. Genes & Development 14:1439-1447. 388. Takeda, S., D. L. North, M. M. Lakich, S. D. Russell, and R. G. Whalen. 1992. A possible regulatory role for conserved promoter motifs in an adult-specific muscle myosin gene from mouse. J Biol Chem 267:16957-67. 389. Tanaka, K., and E. A. Nigg. 1999. Cloning and characterization of the murine Nek3 protein kinase, a novel member of the NIMA family of putative cell cycle regulators. J Biol Chem 274:13491-7. 390. Tang, D., D. Wu, A. Hirao, J. M. Lahti, L. Liu, B. Mazza, V. J. Kidd, T. W. Mak, and A. J. Ingram. 2002. ERK activation mediates cell cycle arrest and apoptosis after DNA damage independently of p53. J Biol Chem 277:12710-7. 391. Tanoue, T., M. Adachi, T. Moriguchi, and E. Nishida. 2000. A conserved docking motif in MAP kinases common to substrates, activators and regulators. Nat Cell Biol 2:110-6. 392. Tanoue, T., R. Maeda, M. Adachi, and E. Nishida. 2001. Identification of a docking groove on ERK and p38 MAP kinases that regulates the specificity of docking interactions. Embo J 20:466-79. 393. Tarricone, C., R. Dhavan, J. Peng, L. B. Areces, L. H. Tsai, and A. Musacchio. 2001. Structure and regulation of the CDK5-p25(nck5a) complex. Mol Cell 8:657-69. 394. Tassin, A. M., B. Maro, and M. Bornens. 1985. Fate of microtubule-organizing centers during myogenesis in vitro. J Cell Biol 100:35-46. 395. Teis, D., W. Wunderlich, and L. A. Huber. 2002. Localization of the MP1-MAPK scaffold complex to endosomes is mediated by p14 and required for signal transduction. Dev Cell 3:803-14. 396. Therrien, M., H. C. Chang, N. M. Solomon, F. D. Karim, D. A. Wassarman, and G. M. Rubin. 1995. KSR, a novel protein kinase required for RAS signal transduction. Cell 83:879-88. 397. Therrien, M., N. R. Michaud, G. M. Rubin, and D. K. Morrison. 1996. KSR modulates signal propagation within the MAPK cascade. Genes Dev 10:2684-95. 398. Tibbetts, R. S., K. M. Brumbaugh, J. M. Williams, J. N. Sarkaria, W. A. Cliby, S. Y. Shieh, Y. Taya, C. Prives, and R. T. Abraham. 1999. A role for ATR in the DNA damage-induced phosphorylation of p53. Genes & Development 13:152-157. 399. Tidyman, W. E., and K. A. Rauen. 2009. The RASopathies: developmental syndromes of Ras/MAPK pathway dysregulation. Curr Opin Genet Dev 19:230-6. 400. Timmons, L., D. L. Court, and A. Fire. 2001. Ingestion of bacterially expressed dsRNAs can produce specific and potent genetic interference in Caenorhabditis elegans. Gene 263:103-12. 401. Tone, A. A., H. Begley, M. Sharma, J. Murphy, B. Rosen, T. J. Brown, and P. A. Shaw. 2008. Gene expression profiles of luteal phase fallopian tube epithelium from BRCA mutation carriers resemble high-grade serous carcinoma. Clin Cancer Res 14:4067-78. 402. Torii, S., M. Kusakabe, T. Yamamoto, M. Maekawa, and E. Nishida. 2004. Sef is a spatial regulator for Ras/MAP kinase signaling. Dev Cell 7:33-44.

157 403. Toyoshima-Morimoto, F., E. Taniguchi, and E. Nishida. 2002. Plk1 promotes nuclear translocation of human Cdc25C during prophase. EMBO Rep 3:341-8. 404. Traverse, S., N. Gomez, H. Paterson, C. Marshall, and P. Cohen. 1992. Sustained Activation of the Mitogen-Activated Protein (Map) Kinase Cascade May Be Required for Differentiation of Pc12 Cells - Comparison of the Effects of Nerve Growth-Factor and Epidermal Growth-Factor. Biochemical Journal 288:351-355. 405. Traverse, S., K. Seedorf, H. Paterson, C. J. Marshall, P. Cohen, and A. Ullrich. 1994. Egf Triggers Neuronal Differentiation of Pc12 Cells That Overexpress the Egf Receptor. Current Biology 4:694-701. 406. Tucker, R. W., A. B. Pardee, and K. Fujiwara. 1979. Centriole ciliation is related to quiescence and DNA synthesis in 3T3 cells. Cell 17:527-35. 407. Tzivion, G., Z. J. Luo, and J. Avruch. 1998. A dimeric 14-3-3 protein is an essential for Raf kinase activity. Nature 394:88-92. 408. Upadhya, P., E. H. Birkenmeier, C. S. Birkenmeier, and J. E. Barker. 2000. Mutations in a NIMA-related kinase gene, Nek1, cause pleiotropic effects including a progressive polycystic kidney disease in mice. Proc Natl Acad Sci U S A 97:217-21. 409. Uto, K., D. Inoue, K. Shimuta, N. Nakajo, and N. Sagata. 2004. Chk1, but not Chk2, inhibits Cdc25 phosphatases by a novel common mechanism. Embo Journal 23:3386-3396. 410. Vader, G., and S. M. A. Lens. 2008. The Aurora kinase family in cell division and cancer. Biochimica Et Biophysica Acta-Reviews on Cancer 1786:60-72. 411. van de Vijver, M. J., Y. D. He, L. J. van't Veer, H. Dai, A. A. Hart, D. W. Voskuil, G. J. Schreiber, J. L. Peterse, C. Roberts, M. J. Marton, M. Parrish, D. Atsma, A. Witteveen, A. Glas, L. Delahaye, T. van der Velde, H. Bartelink, S. Rodenhuis, E. T. Rutgers, S. H. Friend, and R. Bernards. 2002. A gene- expression signature as a predictor of survival in breast cancer. N Engl J Med 347:1999-2009. 412. Van Der Hoeven, P. C., J. C. Van Der Wal, P. Ruurs, M. C. Van Dijk, and J. Van Blitterswijk. 2000. 14-3-3 isotypes facilitate coupling of protein kinase C-zeta to Raf-1: negative regulation by 14-3-3 phosphorylation. Biochem J 345 Pt 2:297- 306. 413. van Hemert, M. J., H. Y. Steensma, and G. P. van Heusden. 2001. 14-3-3 proteins: key regulators of cell division, signalling and apoptosis. Bioessays 23:936- 46. 414. Verlhac, M. H., C. Lefebvre, J. Z. Kubiak, M. Umbhauer, P. Rassinier, W. Colledge, and B. Maro. 2000. Mos activates MAP kinase in mouse oocytes through two opposite pathways. Embo J 19:6065-74. 415. Vincenz, C., and V. M. Dixit. 1996. 14-3-3 proteins associate with A20 in an isoform-specific manner and function both as chaperone and adapter molecules. J Biol Chem 271:20029-34. 416. Vomastek, T., H. J. Schaeffer, A. Tarcsafalvi, M. E. Smolkin, E. A. Bissonette, and M. J. Weber. 2004. Modular construction of a signaling scaffold: MORG1 interacts with components of the ERK cascade and links ERK signaling to specific agonists. Proc Natl Acad Sci U S A 101:6981-6. 417. Vousden, K. H., and D. P. Lane. 2007. p53 in health and disease. Nature Reviews Molecular Cell Biology 8:275-283.

158 418. Wagner, E. F., and A. R. Nebreda. 2009. Signal integration by JNK and p38 MAPK pathways in cancer development. Nature Reviews Cancer 9:537-549. 419. Walsh, S., S. S. Margolis, and S. Kornbluth. 2003. Phosphorylation of the cyclin B1 cytoplasmic retention sequence by mitogen-activated protein kinase and PIx. Molecular Cancer Research 1:280-289. 420. Wan, P. T., M. J. Garnett, S. M. Roe, S. Lee, D. Niculescu-Duvaz, V. M. Good, C. M. Jones, C. J. Marshall, C. J. Springer, D. Barford, and R. Marais. 2004. Mechanism of activation of the RAF-ERK signaling pathway by oncogenic mutations of B-RAF. Cell 116:855-67. 421. Wang, J., and B. Nadal-Ginard. 1995. Regulation of cyclins and p34CDC2 expression during terminal differentiation of C2C12 myocytes. Biochem Biophys Res Commun 206:82-8. 422. Wang, R. N., G. G. He, M. Nelman-Gonzalez, C. L. Ashorn, G. E. Gallick, P. T. Stukenberg, M. W. Kirschner, and J. Kuang. 2007. Regulation of Cdc25C by ERK-MAP kinases during the G(2)/M transition. Cell 128:1119-1132. 423. Wang, X., C. H. McGowan, M. Zhao, L. He, J. S. Downey, C. Fearns, Y. Wang, S. Huang, and J. Han. 2000. Involvement of the MKK6-p38gamma cascade in gamma-radiation-induced cell cycle arrest. Mol Cell Biol 20:4543-52. 424. Ward, I. M., K. Minn, and J. Chen. 2004. UV-induced ataxia-telangiectasia- mutated and Rad3-related (ATR) activation requires replication stress. J Biol Chem 279:9677-80. 425. Watanabe, N., H. Arai, Y. Nishihara, M. Taniguchi, N. Watanabe, T. Hunter, and H. Osada. 2004. M-phase kinases induce phospho-dependent ubiquitination of somatic Wee1 by SCF beta-TrCP. Proceedings of the National Academy of Sciences of the United States of America 101:4419-4424. 426. Weber, C. K., J. R. Slupsky, H. A. Kalmes, and U. R. Rapp. 2001. Active Ras induces heterodimerization of cRaf and BRaf. Cancer Res 61:3595-8. 427. Weber, J. D., D. M. Raben, P. J. Phillips, and J. J. Baldassare. 1997. Sustained activation of extracellular-signal-regulated kinase 1 (ERK1) is required for the continued expression of cyclin D1 in G(1) phase. Biochemical Journal 326:61-68. 428. Weiss, L. A., Y. Shen, J. M. Korn, D. E. Arking, D. T. Miller, R. Fossdal, E. Saemundsen, H. Stefansson, M. A. Ferreira, T. Green, O. S. Platt, D. M. Ruderfer, C. A. Walsh, D. Altshuler, A. Chakravarti, R. E. Tanzi, K. Stefansson, S. L. Santangelo, J. F. Gusella, P. Sklar, B. L. Wu, and M. J. Daly. 2008. Association between microdeletion and microduplication at 16p11.2 and autism. N Engl J Med 358:667-75. 429. Wellbrock, C., M. Karasarides, and R. Marais. 2004. The RAF proteins take centre stage. Nat Rev Mol Cell Biol 5:875-85. 430. White, M. C., and L. M. Quarmby. 2008. The NIMA-family kinase, Nek1 affects the stability of centrosomes and ciliogenesis. BMC Cell Biol 9:29. 431. Whitehurst, A., M. H. Cobb, and M. A. White. 2004. Stimulus-coupled spatial restriction of extracellular signal-regulated kinase 1/2 activity contributes to the specificity of signal-response pathways. Mol Cell Biol 24:10145-50. 432. Willard, F. S., and M. F. Crouch. 2001. MEK, ERK, and p90RSK are present on mitotic tubulin in Swiss 3T3 cells: a role for the MAP kinase pathway in regulating mitotic exit. Cell Signal 13:653-64.

159 433. Willard, M. D., F. S. Willard, X. Li, S. D. Cappell, W. D. Snider, and D. P. Siderovski. 2007. Selective role for RGS12 as a Ras/Raf/MEK scaffold in nerve growth factor-mediated differentiation. Embo J 26:2029-40. 434. Wojnowski, L., A. M. Zimmer, T. W. Beck, H. Hahn, R. Bernal, U. R. Rapp, and A. Zimmer. 1997. Endothelial apoptosis in Braf-deficient mice. Nat Genet 16:293-7. 435. Wu, D., B. Chen, K. Parihar, L. He, C. Fan, J. Zhang, L. Liu, A. Gillis, A. Bruce, A. Kapoor, and D. Tang. 2006. ERK activity facilitates activation of the S-phase DNA damage checkpoint by modulating ATR function. Oncogene 25:1153-64. 436. Wu, L., S. A. Osmani, and P. M. Mirabito. 1998. A role for NIMA in the nuclear localization of cyclin B in Aspergillus nidulans. J Cell Biol 141:1575-87. 437. Wu, X., and J. Chen. 2003. Autophosphorylation of checkpoint kinase 2 at serine 516 is required for radiation-induced apoptosis. J Biol Chem 278:36163-8. 438. Xing, H., K. Kornfeld, and A. J. Muslin. 1997. The protein kinase KSR interacts with 14-3-3 protein and Raf. Curr Biol 7:294-300. 439. Xiong, Y., G. J. Hannon, H. Zhang, D. Casso, R. Kobayashi, and D. Beach. 1993. p21 is a universal inhibitor of cyclin kinases. Nature 366:701-4. 440. Xu, W., and D. Kimelman. 2007. Mechanistic insights from structural studies of beta-catenin and its binding partners. J Cell Sci 120:3337-44. 441. Xu, X., L. M. Tsvetkov, and D. F. Stern. 2002. Chk2 activation and phosphorylation-dependent oligomerization. Mol Cell Biol 22:4419-32. 442. Yaffe, M. B., and A. E. Elia. 2001. Phosphoserine/threonine-binding domains. Curr Opin Cell Biol 13:131-8. 443. Yamaguchi, O., T. Watanabe, K. Nishida, K. Kashiwase, Y. Higuchi, T. Takeda, S. Hikoso, S. Hirotani, M. Asahi, M. Taniike, A. Nakai, I. Tsujimoto, Y. Matsumura, J. Miyazaki, K. R. Chien, A. Matsuzawa, C. Sadamitsu, H. Ichijo, M. Baccarini, M. Hori, and K. Otsu. 2004. Cardiac-specific disruption of the c-raf- 1 gene induces cardiac dysfunction and apoptosis. J Clin Invest 114:937-43. 444. Yamamoto, T., M. Ebisuya, F. Ashida, K. Okamoto, S. Yonehara, and E. Nishida. 2006. Continuous ERK activation downregulates anti proliferative genes throughout G1 phase to allow cell-cycle progression. Current Biology 16:1171-1182. 445. Yan, Y., C. P. Black, and K. H. Cowan. 2007. Irradiation-induced G2/M checkpoint response requires ERK1/2 activation. Oncogene 26:4689-98. 446. Yan, Y., R. S. Spieker, M. Kim, S. M. Stoeger, and K. H. Cowan. 2005. BRCA1- mediated G2/M cell cycle arrest requires ERK1/2 kinase activation. Oncogene 24:3285-96. 447. Yang, J., P. Cron, V. Thompson, V. M. Good, D. Hess, B. A. Hemmings, and D. Barford. 2002. Molecular mechanism for the regulation of protein kinase B/Akt by hydrophobic motif phosphorylation. Mol Cell 9:1227-40. 448. Yang, S. H., A. J. Whitmarsh, R. J. Davis, and A. D. Sharrocks. 1998. Differential targeting of MAP kinases to the ETS-domain transcription factor Elk-1. Embo J 17:1740-9. 449. Yang, S. H., P. R. Yates, A. J. Whitmarsh, R. J. Davis, and A. D. Sharrocks. 1998. The Elk-1 ETS-domain transcription factor contains a mitogen-activated protein kinase targeting motif. Mol Cell Biol 18:710-20.

160 450. Yao, Y., W. Li, J. Wu, U. A. Germann, M. S. Su, K. Kuida, and D. M. Boucher. 2003. Extracellular signal-regulated kinase 2 is necessary for mesoderm differentiation. Proc Natl Acad Sci U S A 100:12759-64. 451. Ye, X. S., G. Xu, R. T. Pu, R. R. Fincher, S. L. McGuire, A. H. Osmani, and S. A. Osmani. 1995. The NIMA protein kinase is hyperphosphorylated and activated downstream of p34cdc2/cyclin B: coordination of two mitosis promoting kinases. Embo J 14:986-94. 452. Yin, M. J., L. Shao, D. Voehringer, T. Smeal, and B. Jallal. 2003. The serine/threonine kinase Nek6 is required for cell cycle progression through mitosis. J Biol Chem 278:52454-60. 453. Yissachar, N., H. Salem, T. Tennenbaum, and B. Motro. 2006. Nek7 kinase is enriched at the centrosome, and is required for proper spindle assembly and mitotic progression. FEBS Lett 580:6489-95. 454. Yoon, S., and R. Seger. 2006. The extracellular signal-regulated kinase: multiple substrates regulate diverse cellular functions. Growth Factors 24:21-44. 455. Yu, W., W. J. Fantl, G. Harrowe, and L. T. Williams. 1998. Regulation of the MAP kinase pathway by mammalian Ksr through direct interaction with MEK and ERK. Curr Biol 8:56-64. 456. Zanke, B. W., K. Boudreau, E. Rubie, E. Winnett, L. A. Tibbles, L. Zon, J. Kyriakis, F. F. Liu, and J. R. Woodgett. 1996. The stress-activated protein kinase pathway mediates cell death following injury induced by cis-platinum, UV irradiation or heat. Curr Biol 6:606-13. 457. Zhang, B. H., and K. L. Guan. 2000. Activation of B-Raf kinase requires phosphorylation of the conserved residues Thr598 and Ser601. Embo J 19:5429-39. 458. Zhang, H. S., A. A. Postigo, and D. C. Dean. 1999. Active transcriptional repression by the Rb-E2F complex mediates G1 arrest triggered by p16INK4a, TGFbeta, and contact inhibition. Cell 97:53-61. 459. Zheng, C. F., and K. L. Guan. 1994. Activation of MEK family kinases requires phosphorylation of two conserved Ser/Thr residues. Embo J 13:1123-31. 460. Zheng, C. F., and K. L. Guan. 1993. Cloning and Characterization of 2 Distinct Human Extracellular Signal-Regulated Kinase Activator Kinases, Mek1 and Mek2. Journal of Biological Chemistry 268:11435-11439. 461. Zhu, J., D. Woods, M. McMahon, and J. M. Bishop. 1998. Senescence of human fibroblasts induced by oncogenic Raf. Genes Dev 12:2997-3007. 462. Zimmermann, S., and K. Moelling. 1999. Phosphorylation and regulation of Raf by Akt (protein kinase B). Science 286:1741-1744. 463. Zur, A., and M. Brandeis. 2001. Securin degradation is mediated by fzy and fzr, and is required for complete chromatid separation but not for cytokinesis. Embo J 20:792- 801.

161