<<

THE ROLES OF ERK1 AND ERK2 MAP IN

NEURAL DEVELOPMENT AND DISEASE

by

IVY SAMUELS

Submitted in partial fulfillment of the requirements

For the degree of Doctor of Philosophy

Thesis Advisor: Dr. Gary E. Landreth

Department of Neurosciences

CASE WESTERN RESERVE UNIVERSITY

August 2008

CASE WESTERN RESERVE UNIVERSITY

SCHOOL OF GRADUATE STUDIES

We hereby approve the thesis/dissertation of

Ivy S. Samuels ______candidate for the Ph.D. degree.*

Stephen O’Gorman ______chair of the committee

Gary E. Landreth ______

Robert H. Miller ______

Ruth E. Siegel ______

06/19/2008

* We also certify that written approval has been obtained for any proprietary material contained therein.

For my teachers, past and present, who inspired me.

TABLE OF CONTENTS

Table of Contents………………………………………………………………….1

List of Tables………………………………………………...……………………4

List of Figures……………………………………………………………………..5

Acknowledgements……………………………………………………………..…7

Abstract…………………………………………………………………………....9

Chapter 1: Introduction..……………………………………………………….11

Mitogen-Activated Kinase Signal Transduction……………………..12

Structural Deformities of the Cerebral Cortex and MR: A Historical

Perspective…………………………………………………………………....17

Controlling Cortical Development: The …………………………...19

ERK1/2 and the Cell Cycle…………………………………………………...22

Cortical Development I: The Expansion of Progenitor Cells………………...26

Cortical Development II: Neurogenesis……………………………………...28

Modes of Neuronal Generation…………………………………………..29

Cortical Malformations and Disrupted Division…………………………30

Cortical Development III: Migration………………………………………....32

Interkinetic migration during Neurogenesis……………………………...32

Cortical Malformations and Disrupted Migration………………………..34

Cortical Development IV: Differentiation of Neurons and Synaptogenesis....35

Molecular Mechanisms of Cortical Development…………………………....36

MAPK and Cell Fate Determination…………………………………………39

Disruptions in upstream elements of the MAPK signaling cascade lead to

1

anatomical and behavioral impairments in mice………………………....41

Aberrant ERK1/2 Signaling leads to Developmental Disorders…………45

ERK in Learning and Memory……………………………………………….47

Individual Roles for the ERK1 and ERK2 Isoforms………………………....49

Research Goals……………………………………………………………….54

Literature Cited……………………………………………………………….56

Chapter 2: Conditional Inactivation of ERK2 Identifies Its Key Roles in Cortical

Neurogenesis and Cognitive Function………………….……………………...68

Abstract………………………………………………………………………69

Introduction…………………………………………………………………..69

Materials and Methods.………………………………………………………72

Results………………………………………………………………………..81

ERK2 is required for normal development of the cerebral cortex……….81

ERK2 inactivation alters the cellular composition of the cortex ………..83

ERK2 inactivation alters the dynamics of neurogenesis………………...86

ERK2 CKO neural progenitors generate fewer neurons in vitro…...... 88

Mature ERK2 CKO mice display impaired associate learning………….89

MAPK1/ERK2 deficiency is associated with learning deficits………….90

Discussion…………………………………………………………………...92

Figures………………………………………………………………………98

Literature Cited……………………………………………………………..128

Chapter 3: ERK1 and ERK2 are required for cell cycle regulation of neural progenitor cells………………………………………………………………...132

2

Abstract……………………………………………………………………..132

Introduction…………………………………………………………………133

Materials and Methods………...……………………………………………136

Results……..………………………………………………………………..140

ERK1/2 DKO mice die embryonically, but display normal brain

morphology……………………………………………………………..140

Loss of both ERK1 and ERK2 does not exacerbate the alterations in cellular

composition identified in the ERK2 CKO cortex……………………....141

ERK1/2 DKO NPCs display inhibited cell cycle progression attributable to

reduced in cyclin D2 and increased p27kip1 expression…...…………..142

Discussion…………………………………………………………………..144

Figures………………………………………………………………………150

Literature Cited……………………………………………………………...166

Chapter 4: Discussion…………………..……………………………………...169

Isoform specific expression regulates isoform specific function…………...169

ERK and Neurogenesis/the Cell Cycle……………………………………..174

ERK and Cognitive Function/Mental Retardation…………………………179

ERK and 22q11 Deletion Syndrome……………………………………….182

22q11 and psychopathology……………………………………………183

22q11 physical pathology………………………………………………185

ERK and Affective Function/Mood Disorders……………………………..187

Literature Cited……………………………………………………………..191

Chapter 5: Literature Cited…………………………………………………..198

3

LIST OF TABLES

Table 3-1 Inactivation of ERK2 does not alter the morphology of Layer III cortical neurons………………………………………...………………………………..127

Table 3-2 Inactivation of ERK2 does not alter the morphology of Layer V cortical neurons………………………………………………………………………….127

4

LIST OF FIGURES

Figure 1-1 MAPK Signal Transduction…………………………………………13

Figure 1-2 General cycle cycle regulation……………………………………….20

Figure 1-3 Temporal development of the murine cerebral cortex...……………..21

Figure 1-4 ERK1/2 regulates cell cycle entry and progression………………….23

Figure 1-5 Modes of Cellular Generation in the Developing Cerebral Cortex….30

Figure 1-6 ERK1/2 Activity Regulates Cell Fate Determination………………..40

Figure 1-7 Mutations in elements of the MAPK Signaling Cascade lead to MR and CFC Syndromes………………………………..……………………………46

Figure 2-1 Generation of ERK2 Conditional Knockout mice…………………...99

Figure 2-2 ERK2 CKO mice display reduced cortical thickness………………101

Figure 2-3 Loss of ERK2 expression and activity in ERK2 CKO cortices…….103

Figure 2-4 Inactivation of ERK2 in neural progenitor cells results in generation of fewer cortical neurons…………………………………………………..……105

Figure 2-5 Inactivation of ERK2 in neural progenitor cells results in the presence of more astrocytes within the cerebral cortex…………………………107

Figure 2-6 ERK2 CKO mice display changes in the dynamics of NPC proliferation……………………………………………………………………..109

Figure 2-7 ERK2 CKO NPCs exhibit reduced neuronal generation…………....111

Figure 2-8 ERK2 CKO cortical progenitors generate more astrocytes in the presence of gliogenic stimuli…………………………………………………...113

Figure 2-9 Male ERK2 CKO mice have deficits in associative learning……....115

5

Figure 2-10 ERK2 expression is abolished in NPCs beginning at E13.5.………117

Figure 2-11 ERK2 expression is reduced in the hippocampus, but not the amygdala……………………………………………………………….………..119

Figure 2-12 The ERK2 CKO cortex does not exhibit apoptosis……….……….121

Figure 2-13 The number of Pax6 and Tbr2 immunoreactive cells at E14.5 is not altered by conditional inactivation of ERK2…………………………………....123

Figure 2-14 Patients with deletions of distal 22q11 have reduced ERK2 levels..125

Figure 3-1 Generation of the ERK1/2 Double Knockout Mouse………………151

Figure 3-2 ERK1/2 Double knockout mice die embryonically but display normal brain morphology...……………………………………………………………..153

Figure 3-3 Loss of ERK1/2 expression and activity in DKO cortices………….155

Figure 3-4 Inactivation of ERK1 and ERK2 in neural progenitor cells results in the generation of fewer cortical neurons…………………………………..……157

Figure 3-5 DKO NPCs exhibit reduced neuronal generation…………………..159

Figure 3-6 DKO neural progenitor cells exhibit reduced proliferation throughout neurogenesis………………………………………………………………….…161

Figure 3-7 DKO mice exhibit reductions in number of mitotic Tbr2+ intermediate progenitor cells………………………………………………………………….163

Figure 3-8 Decreased proliferation of DKO NPCs is mediated by reductions in cyclin D2 and increased p27 expression……………...………………………...165

Figure 4-1 Mutations in elements of the MAPK Signaling Cascade lead to MR

and CFC Syndromes…………...……………………………………………….179

6

ACKNOWLEDGEMENTS

To Dr. Gary Landreth. Without his unwavering belief in my ideas, his enthusiasm for science, and his ability to follow the biology, this dissertation would not have been possible. His patience, understanding and caring have nurtured my life and my career.

To my thesis committee, Dr. Robert Miller, Dr. Stephen O’Gorman, Dr. Ruth Siegel and

Dr. Karl Herrup, their ideas and advice were extremely helpful.

To the Landreth Lab. It has been said that this is a special place to work. It is so much more, because it is a home and a family. I can’t thank them enough for their advice, challenges and support in all endeavors. I would especially like to express my gratitude to Colleen Karlo for her initial and continuous contributions to this project, her endless support, and her contagious laughter.

To the Alzheimer Laboratory. I didn’t study AD, but they didn’t care. I would like to thank them for all their advice, help and the fun times.

To my parents, who recognized and nurtured my potential and encouraged me to become the first doctor in our family. And to Vicki and Mike. The love and support of my family inspires me everyday.

7

To Randall, who is not only an amazing scientist, but also my best friend and cheerleader. I am most appreciative of his support, advice, and love. And to Zander,

Malone and Callie, my munchkins.

To my immediate, extended, and “Cleveland” families and friends. They fill my life with happiness and were always interested, supportive and encouraging. I would specifically like to thank Jessica Lerch, Brandy Wilkinson, Barbara Schwartz, the Kossiver’s, and the

York’s for helping to make Cleveland a wonderful place to work and live.

8

The Roles of ERK1 and ERK2 MAP Kinase in Neural Development and Disease

Abstract

by

IVY SAMUELS

The Activated Protein , ERK1 and ERK2 are critical intracellular signaling intermediates in proliferation and differentiation. ERK1/2 are also essential for memory, learning, and synaptic plasticity in the brain. ERK1/2 activity in neural progenitor cells (NPCs) is required for neuronal cell fate determination; yet little is known about their isoform-specific functions during neural development. We have previously shown that inactivation of ERK1 does not initiate severe phenotypic or

behavioral deficits in mice, suggesting a redundant role for ERK2. Inactivation of ERK2 causes mice to die during embryogenesis. We have therefore examined the role of ERK2 and the combined actions of both isoforms in neural development by targeted and conditional inactivation of /ERK1 and /ERK2 in murine NPCs. Loss of

ERK2 alone resulted in a reduction in cortical thickness attributable to impaired proliferation of NPCs and the generation of fewer neurons. Mutant NPCs remained in an undifferentiated state until gliogenic stimuli induced their differentiation, resulting in the generation of more astrocytes. The ERK2 mutant mice displayed profound deficits in associative learning. Importantly, we identified patients with a 1 Mb microdeletion on 22q11.2 encompassing the MAPK1/ERK2 . These children have reduced ERK2 levels and exhibit microcephaly, impaired cognition, and developmental delay. These findings demonstrate an important role for ERK2 in cellular proliferation

9

and differentiation during neural development as well as in cognition and memory formation.

Simultaneous inactivation of both ERK1 and ERK2 (ERK1/2 DKO mice) lead to a

slightly more severe phenotype than loss of ERK2 alone; however, the loss of both isoforms did not significantly alter the composition of the cortex. ERK1/2 DKO mice

display similar changes in cell fate determination as ERK2 conditional knockouts where

fewer neurons are generated and a greater number of NPCs remain undifferentiated.

These changes are due to elevated levels of p27 and reductions in cyclin D2. Our

findings indicate that in vivo, ERK1 can compensate for ERK2 late in neurogenesis, but

that the normal function of ERK1 in neural development is subordinate to ERK2.

Moreover, ERK1/2 are indispensable for proper neural progenitor cell function as well as

learning and cognition.

10 INTRODUCTION

Congenital developmental abnormalities within the brain are the predominant cause of mental retardation and intellectual disabilities in humans. The incidence of these

disorders is between 1 and 3 percent of children and young adults as classified by the

Diagnostic and Statistical Manual of Mental Disorders- 4th Ed. (DSM-IV). The diagnosis

for mental retardation is defined by an overall intelligence quotient lower than 70, motor

dysfunction, and a lack of adaptive behavior. Investigation into the mechanisms leading

to cognitive impairment has yielded a collection of almost 300 causative (Inlow

and Restifo 2004). Importantly, nearly all of the genetic mutations giving rise to varying

degrees of mental retardation are from the inactivation or functional impairment of

involved in cortical development and plasticity (Chelly et al., 2006).

Furthermore, recent evidence suggests that size and gyrification of the cerebral cortex is

associated with intellectual capacity (Shaw et al., 2006; Witelson et al., 2006). These

data suggest that a delineation of the causes of abnormal cortical development and

function is the key to understanding mental retardation.

The mammalian cerebral cortex is produced through strict temporal and spatial control of

cellular proliferation. Approximately 21.5 billion neurons (Williams and Herrup, 1988)

populate the mature human cortex, and about 35 million neurons in the mouse (Vaccarino

et al., 1999). The generation of these cells depends on an intricate choreography of the

proliferation of founder progenitor cells and committed neuronal progenitors within the

germinal centers of the brain. Therefore, control of the cell cycle in each cellular

11 population is extremely important in development of the cerebral cortex, and more broadly, to cognitive function. A mitogen is defined as a substance which induces a cell to commence cellular division, triggering mitosis. The canonical Mitogen-Activated

Protein Kinases were therefore the target of our investigation. This thesis demonstrates

that the prototypical mitogen activated protein kinases, ERK1 and ERK2, act to regulate

proper cortical development through maintenance of the cell cycle in neural progenitor

cells and that ERK2 is the predominant isoform responsible for this function. Moreover,

this work demonstrates that reduced expression of only the ERK2 isoform is sufficient to

induce physical and behavioral characteristics of mental retardation.

Mitogen-Activated Protein Kinase Signal Transduction

The Mitogen-Activated Protein Kinase (MAPK) superfamily is comprised of a conserved

group of which regulate a diverse range of cellular processes including

proliferation, migration, differentiation, stress, and survival. In mammals, the MAPKs

are divided into three major groups based on sequence similarity, activity, and substrate

specificity. They are the Extra-cellular signal Regulated protein Kinases (ERK 1 and 2;

also termed p44 and p42 MAPK, respectively), the Stress activated protein kinases, also

known as the Jun N terminal kinases (JNK1, 2 and 3), and the p38 MAPKs (α, β and γ).

Enzymatic activity of the MAPKs is tightly regulated by phosphorylation status of the proteins, which is mediated by kinases and phosphatases. Activation of the MAP Kinase superfamily occurs through a classical three-tier kinase cascade whereby extracellular cues stimulate cell surface receptors which then propagate intracellular signaling to the cytoplasm and nucleus (Figure 1-1).

12 In the developing organism, growth factors are a primary mitogenic stimulus and act largely through receptor tyrosine kinases which signal to elements that control the cell cycle. Receptor tyrosine kinases are located on the plasma membrane and

Figure 1-1: Signal transduction from the plasma membrane to the nuclueus is mediated by the canonical Mitogen Acticated Protein Kinases, ERK1/2.

become activated upon ligand binding, provoking them to undergo dimerization, autophosphorylation for full transactivation, and subsequent conformational changes.

These changes result in the recruitment of scaffolding proteins which bind the newly

phosphorylated tyrosine residues located on the intracellular domain of the receptor.

Activation of the ERK signaling pathway leads to recruitment of signaling molecules that

possess Src homology 2 (SH2), phosphotyrosine (PTB), and PDZ domains that recognize

13 and associate with phospho-tyrosine residues. These binding domains enable the formation of signaling complexes through the association of various adaptor proteins, namely Shc and Grb2, which bind additional scaffolding proteins (FRS2, Crk, CrkL) near the plasma membrane to ultimately activate guanine nucleotide exchange factors (SOS,

C3G). Subsequently, members of the Ras family of small G-proteins such as Ras and

Rap1 are activated to facilitate intracellular signaling. The stimulation of these small G- proteins initiates a downstream chain of phosphorylation events whereby MAP Kinase kinase kinases (MAP3Ks: c-Raf, B-Raf, A-Raf) phosphorylate and activate the dual specificity MAP Kinase Kinases (MAP2K: MEK1/2) which in turn finally activate the

MAP Kinases. The prototypical signaling elements of the mitogen-activated protein kinase superfamily are ERK1 and ERK2, proline-directed serine/threonine kinases which become activated by phosphorylation on specific threonine and tyrosine residues

(Threonine 183, Tyrosine 185) located in the activation loop of the enzymes (Pearson et al., 2001; Roux and Blenis, 2004; Rubinfeld and Seger, 2005). The amplification of the signaling cascade is demonstrated by the ability of activation of just 5% of cellular Ras molecules to induce full activation of ERK1/2 (Hallberg et al., 1994).

The three tier structure of the MAP Kinase cascade allows for both diversity in signaling and specificity. Depending on the stimulus, the engaged receptor tyrosine kinase (RTK) recruits an appropriate set of scaffolding proteins to initiate signaling through specific

MAPK3Ks and MAP2Ks, generating a ligand-specific downstream effect and biological response (Pouyssegur et al., 2002).

14 Due to the variety of stimuli which induce MAPK activity, cell type and signaling kinetics generate different cellular responses. For example, rat pheochromocytoma cells,

PC12 cells, have the capacity to differentiate into mature adrenal chromaffin cells by dexamethosone (a MAPK independent event) or depending on the level of MAPK stimulation, differentiate into sympathetic-like neurons or continue to proliferate. When stimulated with either EGF or insulin, PC12 cells rapidly proliferate due to a transient increase in ERK1/2 activity. Alternatively, when stimulated with NGF, FGF or IL-6, sustained ERK1/2 activation leads to differentiation of these cells into sympathetic-like neurons as evidenced by neurite extension and electrical excitability (Marshall 1995).

Additionally, a single stimulus can initiate different biological responses based on the required duration of MAPK activation. Again in PC12 cells, Nerve Growth Factor recruits either a Crk-C3G-Rap1 complex to RTKs which signal through B-Raf for prolonged MAPK activation and differentiation, or RTKs can assemble with a Shc-Grb2-

SOS-Ras complex to signal through c-Raf for transient MAPK activation and proliferation (Marshall, 1995; York et al., 1998; Kao et al., 2001). Other cell types such as T lymphocytes and megakaryocytes, are similarly receptive to transient or prolonged

MAPK activity for generation of distinct biological responses (Katz et al., 2007).

Importantly, the response is cell type specific as NGF-driven prolonged MAPK activity induces differentiation in PC12 cells, but leads to proliferation in olfactory sensory neuron precursors (Simpson et al., 2007). A similar proliferative response to sustained

ERK1/2 activity occurs in fibroblasts as well (Katz et al., 2007). These studies suggested that cell type, regulation of duration, and intensity of ERK1/2 activity is important in differentiation or proliferative responses. Moreover, the signaling complex which leads

15 to the activation of ERK1/2 often dictates biological outcomes and serves as a modulator

of MAPK activity in time and space (Pouyssegur et al., 2002; Sweatt et al., 2003).

Given the differential response that can occur based on MAPK activity it is not surprising

that ERK1/2 plays a role in cell fate decisions. The induction of ERK1/2 activity in embryonic stem (ES) cells dictates the alternative fates of self renewal or lineage commitment, during embryogenesis. Autocrine stimulation of ES cells by LIF leads to proliferation and ES cell renewal, whereas FGF4 stimulation results in cell commitment, prompting naïve ES cells to exit self-renewal and differentiate into neural and mesodermal cell fates (Kunath et al., 2007). Interestingly, the complete lack of ERK1/2

activity or removal of ERK2 expression does not affect ES cell proliferation (Meloche et al., 2004).

It is hypothesized that ERK1/2 similarly affects cellular function and cell fate decisions within the developing telencephalon. Rossant and colleagues (2003) developed two

specific which recognize the di-phosphorylated form of ERK1/2. These tools

were used to map the spatial and temporal domains of FGF activity in the developing

mouse embryo through visualization of phosphorylated ERK1/2 (Corson et al., 2003).

Importantly, they found that among the most reproducible and prominent domains of

sustained ERK1/2 activity from E8.5 to E10.5 were the frontonasal passages, the

forebrain, and the midbrain-hindbrain boundary. This places ERK1/2 at the correct time

and location to influence cortical development.

16 Structural Deformities of the Cerebral Cortex and MR:

A Historical Perspective

Mental retardation was first linked to changes in brain structures as early as the 1600s

when Thomas Willis initiated the field of clinical neurology by dissecting human brains

and mapping their cellular connections. He recognized that patients with congenital

mental retardation have abnormalities in their brain size and/or shape (Molnar, 2004).

However, it was not until the mid 1900s that clinically-defined changes in brain structure

became associated with mental retardation. Patients with mental retardation were initially classified on the basis of known causes, such as chromosomal abnormalities, infection, and identified syndromes (i.e. phenylketonuria, hydrocephalus or ), or undetermined causes. Between 15% and 50% of patients identified with mental disabilities fell into one of the classified causes. Remarkably, of the unclassified cases of MR, 90% were attributable to malformations of the brain or gliotic encephalopathies (Crome, 1972). The malformations documented to be associated with mental retardation included microcephaly which is literally “small brain.” It was thought that “the smaller the brain, the less intelligent the individual and the shorter expectation his life” (Crome, 1972). Another malformation associated with the microcephaly was unclassified hydrocephaly, whereby the lateral ventricles are enlarged and the overlying tissue is smaller, producing an overall larger head. Prosencephaly, or the union of the two telencephalic hemispheres due to the lack of formation of two separate ventricles, was also associated with mental retardation. It was further recognized at this time that neuronal migration defects give rise to mental retardation. The cortical deformities

17 specifically associated with migration were microgyria (smaller brain convolutions or

gyri), pachygyria (diminuation of gyri and sulci) and nodular ectopias (extra-cellular spots within the brain). This group of malformations is now commonly grouped as various forms of lissencephaly, or smooth brain disorders.

In the 1980s, Magnetic Resonance Imaging (MRI) became available and revolutionized diagnosis of MR. Scientists and physicians found that even subtle defects in cortical structures are associated with decreased mental capacity and developmental delays. The question then became one of mechanism and function. How do structural changes of the cerebral cortex arise? And, how do they translate into reduced mental capacity? Thus, the generation of the cortex became the focus of attention.

Cortical development is separated into four major stages: (1) expansion, the exponential

proliferation of founder progenitor cells by symmetric division necessary to generate

sufficient cells to populate the cortex; (2) neurogenesis, the generation of the correct

number of neurons and glia to produce and fill the cortex; (3) neuronal migration, the

traveling of each neuron to its proper location within the cortex; and (4) neuronal

differentiation, the maturation of neurons and establishment of connections to their

correct targets. It is now recognized that reductions in numbers of neurons populating the

cortex (from aberrant neurogenesis or neuronal death) gives rise to cortical hypoplasias

which include microcephaly, and cortical ectopias. Migration defects can also result in

cortical ectopias and lissencephaly. And finally, cortical dysplasias (abnormalities in

number or structure of dendrites, i.e. maturation) are due to abnormal neuronal migration

18 or neuronal maturation (Berger-Sweeney and Hohmann, 1997). The observations and

linkages made early in clinical neuroscience and later with more advanced imaging capabilities have established the connection between malformations of the cerebral cortex and congenital mental retardation. Proper cortical development is crucial for normal

cognitive function, and mutations in the genes fundamental for cortical development

result in mental retardation.

Controlling Cortical Development: The cell cycle

The process by which all eukaryotic cells divide is termed the cell cycle. It is comprised

of 4 major phases, namely mitosis (M phase), DNA synthesis (S phase), gap 1 (G1 phase)

and gap 2 (G2 phase). When cells are not in the process of division they are said to be in

the G0 phase. G0 is a quiescent stage whereby cells cease division or have terminally

differentiated. The G1 phase of the cell cycle is critical for a cell as it is the time in

which a cell surveys its environment and decides whether to commit to another round of

division or undergo differentiation, exiting into G0. As such, a restriction checkpoint

occurs at the end of G1. If a cell passes the checkpoint, it will proceed to the S phase and

prepare to complete another round of division. A second checkpoint is located at G2. If

a cell has undergone faulty replication of its DNA, the cell will not complete division and

will undergo apoptosis. Mitosis occurs through a precise and defined sequence of events,

in which structures within the nucleus of a cell separate both the nucleus and cytoplasm

into two compartments providing a means for division by cytokinesis.

19 Progression between the stages of the cell cycle is controlled by regulatory mechanisms which integrates extracellular signals that control proliferation with the intricate orchestration of events within the cell cycle. The induction of a cell cycle is tightly

regulated through the expression of cyclins which act to initiate the cell cycle and sustain

Figure 1-2: Progression through the cell cycle is controlled by inductive cyclin-dependent kinases and antiproliferative cyclin dependent kinase inhibitors.

its progression (Ohnuma and Harris, 2003). In order for a mammalian cell to transition from the cellular growth phase or first gap, G1, to DNA synthesis (S), cyclins D1-3, and cyclin E are required. The G2-M checkpoint is controlled by a separate group comprised of cyclin A, B1 and B2, in an intricate signaling cascade leading to mitosis and resulting

20 in proliferation (Ross, 1996). The function of the cyclins is regulated by the formation of a complex with specific cyclin-dependent kinase (CDK) molecules. This interaction controls the transition between the phases of the cell cycle. Cell cycle kinetics can further be modulated by tumor suppressors, p53 and pRb as well as the Cyclin-dependent

Kinase Inhibitors (CKIs), p15, p16, p21 and p27 (Dehay and Kennedy, 2007). These antiproliferative genes suppress progress through the cell cycle by sequestering Cyclin

D/E from their Cdk binding partners, and prevent phosphorylation of Rb, thus inhibiting

E2F-dependent transcription of additional cyclins required for G1 to S transition.

In the mouse, the neurogenic phase of cortical development includes the neural progenitor cells which have been shown to undergo 11 cell cycles spanning 6 embryonic days (E11-E17, Figure 1-3). Each cell cycle lengthens in duration through regulation of

Figure 1-3: Development of the Murine Cerebral Cortex occurs over a 6 day interval during which successively longer cell cycles generate the majority of cells.

21 G1 phase, so that the first cycle lasts 8.1 hours and the final cycle lasts 18.4 hours

(Caviness et al., 1995; Takahashi et al., 1995). The increase in cell cycle length is attributed to increases in the length of G1 as S phase is continuously about 4 hours long and G2-M is consistently 2.5hrs. Post-mitotic neurons exit the cell cycle and enter G0, where they become quiescent and unable to further divide. The exquisite regulation of both the progenitor cell expansion and neuronal proliferation to populate the cerebral cortex ensures its proper generation and function.

ERK1/2 and the Cell Cycle

As the original nomenclature implies, ERK1 and 2 were identified as mitogen-activated protein kinases which play a substantial role in mitotic regulation. Before the proteins were cloned and identified as protein kinases, it was recognized that they were phosphorylated and activated upon the presentation of mitogenic stimuli and their long- term phosphorylation through the G0 and G1 phases in the cell cycle were mandatory for cell cycle entry (Pages G 1993). The extent of ERK1/2 activity required for S phase entry has been debated. It is generally accepted that sustained activity promotes progression through G1, although ERK1/2 activity early in G1 is dispensable for transition in fibroblasts (Villanueva et al., 2007). Treatment of mouse embryonic fibroblasts with the MEK inhibitor U0126 at varying timpoints post serum-stimulation demonstrated that ERK1/2 activity is critical from mid-G1 to the end of G1 (as marked by cyclin A expression). Treatment of FGF-stimulated NIH3T3 cells with U0126 similarly showed that sustained ERK1/2 activity is not required until 2-3 hours prior to S phase entry (Yamamoto et al., 2006; Villanueva et al., 2007). These studies demonstrate

22 that ERK1/2 activity should be sustained through late G1 to ensure cell cycle progression into S phase. Further analysis of cell cycle progression by in vitro siRNA studies have demonstrated that ERK1 is necessary for completion of the G2 phase of the cell cycle, while ERK2 is reported to be required for G1 progression in HeLa cells (Liu et al., 2004), suggesting different roles for the ERK isoforms in cellular proliferation.

There are many known targets of ERK1/2 activity which are important for cell proliferation (Figure 1-2, 1-4). The primary mechanism by which ERK1/2 regulates cell cycle entry and progression is through induction of the critical G1 restriction checkpoint

Figure 1-4: ERK1/2 regulates cell cycle entry and progression by phosphorylation and transcription of numerous targets.

23 cyclin, Cyclin D1 (Roovers and Assoian, 2000; Wilkinson and Millar, 2000). In response

to serum or growth factor stimulation, ERK1/2 activation leads to phosphorylation of Ets

transcription factors such as Elk-1 and SRF which form the AP-1 transcriptional complex

with c-. This ternary complex binds the Serum Response Element on the cyclin D1

promoter to facilitate transcription of the cyclin. ERK1/2 can also regulate cyclin D1 by

phosphorylation of the transcription factor, c-myc. ERK-dependent expression of cyclin

D1 and its complex formation with cdk4/6 is therefore capable of stimulating cells to exit from G0 and promote entry into the cell cycle. Although cyclin D1 is an essential component of cell cycle progression and the primary target of ERK activity in response to , as cells prepare for replication and division, ERK activity also contributes to the production of multiple constituents of cellular growth. ERK1/2 facilitates the generation of new DNA and rRNA through the phosphorylation of carbamoyl phosphate synthetase, an which catalyzes the synthesis of new pyrimidine nucleotides

(Graves et al., 2000). ERK1/2 aids in transcription of ribosomal RNA genes through its phosphorylation of HMG box DNA binding domains on the transcription factor UBF

(Stefanovsky et al., 2006a; Stefanovsky et al., 2006b). ERK1/2 also directs protein translation by the phosphorylation of serine 209 on the translation initiation factor eIF4E through its immediate downstream target, the protein kinase, MNK1 (Chambard et al.,

2007).

ERK1/2 activity is critical for both cell cycle entry and progression through the cell cycle. Once in G1, cells will not enter the S phase if ERK activity is abrogated. This action is mediated by the suppression of antiproliferative genes (Yamamoto et al., 2006).

24 ERK1/2 can therefore regulate cell cycle entry and progression through phosphorylation

of both pRb and p27. ERK1/2 acts to regulate p27KIP1 levels at multiple levels. ERK1/2

directly phosphorylate p27KIP1, inhibiting its function, and stimulating its degradation

through induction of the ubiquitin E3 , SKP-2 (Aktas et al., 1997; Kawada et al.,

1997; Takuwa and Takuwa, 1997; Villanueva et al., 2007). ERK1/2 have also been

reported to regulate p27KIP1 translation (Miskimins et al., 2001) or mRNA stability

(Sakakibara et al., 2005). ERK1/2 activation is correlated with reduced cellular p27KIP1

levels (Teixeira et al., 2000; Sakakibara et al., 2005) and enhanced cellular proliferation.

Conversely, inhibition of ERK activity results in elevation of p27KIP1 levels and cell cycle arrest (Gysin et al., 2005). Another potential level of ERK regulation on cell cycle entry and progression is through phosphorylation of the antiproliferative gene, tis21. It is of particular importance due to its expression in cells which are undergoing terminal mitosis

(Iacopetti et al., 1999) and downregulation in post-mitotic neurons. Overexpression of tis21 inhibits G1-S progression through transcriptional regulation of cyclin D1 in the presence of Rb, while in the absence of Rb, tis21 downregulates cyclinE-cdk4. Of particular relevance is the fact that ERK1/2 phosphorylates tis21 (Hong et al., 2005), providing a potential mechanism by which ERK affects terminal mitosis of neural progenitor cells.

The ERKs were first identified as microtubule associated protein (MAP), kinases owing to their phosphorylation of MAP itself. ERK1/2 has also been implicated in regulation of

the cell cycle through phosphorylation of and association with cytoskeletal proteins such

as neurofilament and paxillin (Chambard et al., 2007). ERK activity has been shown to

25 be important in regulating mitotic progression at the G1/S boundary (Guadagno and

Ferrell, 1998; Roberts et al., 2002; Horne and Guadagno, 2003), as well as in microtubule

reorganization during M phase (Gotoh et al., 1991). ERK has also been localized to spindle poles during pro-metaphase, the mid-body during cytokinesis and is generally

thought to contribute to normal mitotic progression in somatic cells (Willard and Crouch,

2001).

Cortical Development I: The Expansion of Progenitor

Cells

Cortical development begins upon closure of the neural tube. Initially, the single layer of

multipotent neural progenitor cells which define the proliferative neuroepithelium

(neuroepithelial cells) must undergo an exponential series of symmetrical divisions. This

expansion in the number of neuroepithelial cells in turn increases the number of

committed neural precursors that will comprise the telencephalic ventricular zone. The

neuroepithelial or founder cell expansion is a critical component of cortical development

as insufficient proliferation results inadequate numbers of committed neural progenitors

and the generation of a microcephalic brain (Rakic, 1988). Furthermore, the expansion of

neuroepithelial cells is responsible for the evolutionary enlargement of brains between

species (Caviness et al., 1995; Rakic, 1995).

Neuroepithelial cells are highly polarized in their apical-basal axis and are identifiable

due to their epithelial features including expression of nestin and the presence of tight

26 junctions. The symmetrical division of the neuroepithelial cells is ensured by the

orientation of the cleavage plane in cytokinesis (Figure 1-5, division A). A vertical cleavage plane allows for the equal distribution of apical and basal cellular matter and the production of two equal neuroepithelial daughter cells (Gotz and Huttner, 2005). After expansion of the progenitor pool is complete, the neuroepithelial cells transition into a committed neural progenitor cell, downregulating some of their epithelial features and adopting astroglial properties such as glycogen granules (Figure 1-5, division B). This cell has also been termed a radial glial cell. Committed neural progenitor cells are marked by the expression of numerous proteins and their maturation into restricted- precursors and can be identified by the inclusion or exclusion of a subset of these markers. Most of the markers for radial glial cells are intermediate filament proteins including nestin, (which is maintained from the neuroepithelial cell), vimentin, and the glial fibrillary acidic protein (GFAP). Additional markers of radial glia include expression of the transcription factor Pax6, the astrocyte-specific glutamate transporter,

GLAST, and the brain-lipid binding protein, BLBP (Gotz and Huttner, 2005). A population of secondary or subventricular progenitors is further derived from neuroepithelial cells and the radial glia (Fishell and Kriegstein, 2005). These cells are successively committed intermediate progenitor cells identifiable by downregulation of

Pax6, and expression of the transcription factor Tbr2 (Englund et al., 2005; Gal et al.,

2006); (Figure 1-5, division C). The expansion of the founder neuroepithelial cell population through self-renewal and generation of intermediate progenitor cells is essential for the generation of the millions to billions of neurons which populate the mature cortex (Rakic, 1995; Martinez-Cerdeno et al., 2006). It is of significance that

27 IPCs are the only symmetrically neurogenic progenitor cells. Furthermore, they

contribute neurons to all cortical lamina and are postulated to underlie the amplification

of the cortex throughout development and evolution (Pontious et al., 2008).

Cortical Development II: Neurogenesis

Upon completion of the expansion phase and concomitant with the appearance of radial

glia, successive groups of neural progenitors re-enter the cell cycle and subsequently

initiate neurogenesis. The first glimpses of the cellular dynamics and the level of

complexity involved in cortical neurogenesis came in the late 19th century when neuroanatomists such as Wilhelm His recognized that two populations of cells within the

embryonic cerebral cortex underwent mitosis. One was located at the ventricular surface

and the other away from the ventricular lumen. Each of these cell types divided,

producing the non-mitotic cells which resided in the cortical plate. Rudolf Albert von

Kolliker further identified the progenitor cells which lined the ventricles in the pseduostratified epithelium and ultimately changed their shape. These were the surface dividing cells that His described as germinal cells and are now known as the radial glia.

The non-surface dividing or abventricular mitotic cells, which His also documented, are now thought to be the intermediate progenitor cells that reside in the sub-ventricular zone. Notably, it has been recognized that the excitatory projection neurons which populate the cerebral cortex are derived from and can migrate along His’ germinal cells

(radial glia). Extensive lineage mapping has shown that nearly all divisions producing

the excitatory projection neurons are derived from a radial glia. Studies by Messing and

colleagues revealed the multipotential properties of these cells following generation and

28 analysis of transgenic mice expressing the cre-recombinase by the human GFAP promoter (hGFAP-cre) (Zhuo et al., 2001). Subsequently, the hGFAPcre/+ transgenic mice have been used for a detailed fate mapping of the developing telencephalon (Malatesta et al., 2003; Anthony et al., 2004). Malatesta et al. have shown that over 90% of neurons in the neocortex and hippocampus are derived from hGFAP-cre expressing progenitors, including all long range projection neurons, but few interneurons (Malatesta et al., 2003).

Modes of neuronal generation

After the initial expansion phase of cortical development, whereby radial glia are produced, these cells then begin to produce neurons. Multiple modes of division underlie the generation of cortical projection neurons (see Figure 1-5): (1) Radial glia which are located in the ventricular zone undergo asymmetric divisions producing one neuron and maintaining one progenitor cell (65.8%). (2) Radial glial cells divide to produce one self-renewed radial glia and a cell which becomes an intermediate progenitor cell and goes on to divide away from the ventricular zone (7.3%). (3) Radial glia symmetrically divide into two radial glia,

29 Figure 1-5: Modes of Cellular Generation in the Developing Cerebral Cortex.

one which divides again and one which will translocate to the cortical plate and differentiate (26.9%). (4) Radial glia symmetrically divide to produce two more radial

glia which will both undergo asymmetric divisions (9.6%) (5) Abventricular

intermediate progenitors symmetrically divide to produce two neurons (89.5% of

abventricular divisions). (6) Abventricular intermediate progenitors divide, producing

two cells which will each divide again (10.5% of abeventricular divisions) (Haydar et al.,

2003; Noctor et al., 2004; Gotz and Huttner, 2005).

Cortical malformations and disrupted division

Disruption in either phase of proliferation, expansion of the progenitor pool or

neurogenesis, leads to significant structural changes in the cerebral cortex and are

30 associated with cognitive deficits. For example, Autosomal Recessive Primary

Microcephaly, MCPH, is characterized by reduced growth of the cerebral cortex, albeit with normal lamination, a small head circumference, and mental retardation. Four genetic loci have been mapped and each gene is important in the centrosomal regulation of neural progenitor cellular mitosis (Bond and Woods, 2006). The genes,

Microcephalin, ASPM (the abnormal spindle-like microcephaly associated protein),

Cdk5Rap2, and CenpJ are highly expressed and localized to the neuroepithelium lining the lateral ventricles of the developing cortex. Mutations in each of these genes individually results in a reduction in numbers of neurons stemming from either a reduction in neural progenitor cells themselves or an inability of the progenitor to divide appropriately. Nde1 (a Lis1-interacting protein) mutants are also microcephalic due to a loss of microtubule organization and spindle pole orientation which results in aberrent neuroepithelial cell mitosis (Feng and Walsh, 2004). One of the most recognizable forms of mental retardation, Down’s syndrome, is also characterized by small brain size stemming from reduced numbers of neural progenitors due to an increase in cell cycle length (Haydar et al., 2000). The conditional knockout of the cell cycle regulator Rb in the developing telencephalon also causes a robust (7 fold) increase in progenitor proliferation resulting in a dramatic increase in telencephalon size (Ferguson et al., 2002).

This and previous studies therefore demonstrate that the Rb protein is critical for commited neural progenitors to undergo terminal neurogenic mitotsis and survive (Slack et al., 1998). The CKI, p27KIP1 is of particular interest and importance due to its role in cortical development (Delalle et al., 1999; Mitsuhashi et al., 2001; Goto et al., 2004;

Lukaszewicz et al., 2005; Tarui et al., 2005). Animals in which p27KIP1 has been

31 knocked out exhibit increased cortical thickness resulting from a 27% increase in neurons

in upper cortical lamina (Caviness et al., 2003; Goto et al., 2004). Conversely,

overexpression of p27KIP1 in the developing telencephalon slows progenitor proliferation with fewer neurons populating the upper cortical layers, resulting in reduced cortical thickness (Tarui et al., 2005).

Cortical Development III: Migration

Newly generated neurons and self-renewed progenitor cells each have a separate fate.

This is most obvious in the third stage of neurogenesis, neuronal migration.

Interkinetic migration during neurogenesis:

Within the ventricular zone, radial glial cells begin the cell cycle located immediately

adjacent to the ventricle. During S phase, the cells migrate away from the ventricle and subsequently return to their original position along the ventricle to undergo mitosis

(Takahashi et al., 1993; Noctor et al., 2004). After division, committed neural progenitor

cells or radial glia can maintain their position within the apical ventricular zone and

divide again, or they can migrate away from the ventricle where they can mature into an intermediate or basal progenitor cell. In contrast to radial glia, intermediate progenitor

cells do not undergo interkinetic nuclear movement. Upon their generation from a radial

glial division, they migrate away from the ventricle and undergo all phases of the cell

cycle in an abventricular position. Despite the progenitor from which a neuron is born,

both the ventricular neurons and abventricular-derived neurons migrate toward the pial

surface of the brain upon their birth. The laminar organization of the cerebral cortex is

32 derived from the migration of these newly born neurons in a stereotypical “inside-out”

pattern. The neurons utilize radial glial fibers or self-generated radial fibers as scaffolds

to migrate toward their final locations. The earliest generated neurons split the preplate and subplate and remain closest to the ventricle, forming layer VI. Later born cells migrate outward past the earlier migrants and are positioned at successively more superficial positions (demonstrated in Figure 1-3).

Within each discrete wave of division, newly born neurons carry intrinsic signals which dictate their location in the cortical plate. If neurons are harvested from a mouse embryo at E14 and transplanted into a distinct ventricular zone, they can migrate to their correct location. However, after gliogenesis has begun, the newly born neurons can maintain their transcription factor identity but they do not migrate to the proper position

(McConnell and Kaznowski, 1991; Desai and McConnell, 2000). Takahashi et al. found that a cell cycle counting mechanism is intrinsic to the laminar fate of the neuron

(Takahashi et al., 1999).

The generation of neurons and glia which fill the cortical plate therefore arises from a highly regulated process of symmetrical division, asymmetrical division and migration.

The radial glia which do not divide either differentiate into neurons early in neurogenesis or eventually translocate into the cortical plate where they encounter extrinsic cues to direct their differentiation into astrocytes or oligodendrocyte precursors (Qian et al.,

2000; Sauvageot and Stiles, 2002). GABAergic interneurons, which also reside within the cortex, are not descended from GFAP-expressing radial glia. They are derived

33 principally from the ganglionic eminence and migrate tangentially into the cortical plate

(Malatesta et al., 2003; Anthony et al., 2004).

Cortical malformations and disrupted migration

Aberrations in neuronal migration have been shown to be responsible for lissencephaly and developmental disorders which are characterized by cognitive impairment (Rakic,

2000). When neurons are not properly positioned after migration, heterotopias of neurons are formed and numerous clinical syndromes can result. In humans, these disorders are associated with the cortical malformations of microgyria, pachygria, and nodular ectopias, all of which are more generally classified as lissencephaly. Mutations in Reelin, Lis1 and DCX lead to lissencephaly and mental retardation by affecting the process of neuronal migration or the substrates on which neurons migrate (Vallee and

Tsai 2006). Lis1 and doublecortin encode proteins which associate with microtublules and cytoskeletal elements like the microtubule motor protein dynein. These proteins ensure that neuronal migration is possible and that neurons migrate properly to form the laminar organization of the cortex. Reelin is an extracellular matrix glycoprotein which is important in defining the pial surface and generating a stop signal for neurons migrating out of the germinal zones (Haydar, 2005). When neuronal migration is faulty, heterotopias often form, leading to improper synaptic connectivity. This often results in cortical malformations associated with epilepsy (Rakic, 2000; Guerrini and Marini,

2006).

34 Cortical Development IV: Differentiation of Neurons

and Synaptogenesis.

Upon reaching its final destination, a newly-born cortical neuron differentiates into a

specific cell-type by expressing neurotransmitters and corresponding receptors,

elaborating dendrites and dendritic spines, and sending axons to their synaptic targets.

These processes are important and ensure that communication between cells occurs properly and efficiently. Mutations in FMRP, the fragile X mental retardation protein, is responsible for abnormal dendritic spines found in patients with Fragile X MR due to disrupted translation of mRNA at synapses. Disruption of dendritic arborization and spine density is also associated with Rett syndrome and Down syndrome. The mechanisms underlying the malformations seen in these diseases is unknown, however, synapse formation and elaboration of dendrites is critically linked to neuronal activity and therefore both neuronal proliferation and migration (Haydar, 2005). This phase of development will not be covered in depth.

In summary, cortical development can be divided into four very broad stages, founder cell expansion, neurogenesis, neuronal migration and differentiation. When neurogenesis is disrupted, the result is cortical hypoplasia or microcephaly; when migration is disrupted, cortical ectopias or lissencephaly results, and when differentiation is disrupted, cortical dysplasia or a loss of dendritic arborization results. Abrogation of any stage of cortical development through cell cycle machinery or directive proteins leads to alterations in cortical structures associated with cognitive impairment, providing

35 convincing evidence that the two are coupled (Berger-Sweeney and Hohmann, 1997).

Accordingly, the genetic mutations linked to structural defects and MR are involved in

regulation of mitotic structural proteins or the intrinsic and extrinsic cues required for

proper progenitor expansion and division. It is clear that normal division of these cells

through the cell cycle is critical to cortical development.

The Molecular Mechanisms of Cortical Development

Following the neurogenic phase of cortical development, a gliogenic phase arises and peaks at birth to produce first astroglia, followed by oligodendroglia (Qian et al., 2000).

This transition from expansion of progenitor pools to neurogenesis and then gliogenesis

is regulated by extrinsic and instrinsic factors (Cameron et al., 1998; Ross et al., 2003).

Basic Fibroblast Growth Factor (bFGF or FGF2) and Epidermal Growth Factor (EGF)

are responsible for directing neural progenitor cell expansion in concert with the basic

Helix Loop Helix (bHLH) transcription factors (Vaccarino et al., 1999). The principal

bHLH factors that act to regulate early stages of cortical development are members of the

Hes (Nakamura et al., 2000; Ohtsuka et al., 2001) and Id (Lyden et al., 1999) families.

These transcription factors act by promoting the progenitor proliferation through

suppression of neuronal differentiation. They functionally antagonize the actions of the

proneural bHLH factors, Ngn1/2 and Mash1 (Nieto et al., 2001). Neurogenesis is

induced when the Hes/Id factors are dowregulated and the proneural factors dominate.

Inactivation of the proneural bHLH factors Ngn2 and Mash1 results in a smaller cortex

with fewer neurons and many more astrocytes (Nieto et al., 2001). Inactivation of either

the Hes1 or Hes5 gene leads to premature differentiation of neurons due to the absence of

36 this intrinsic brake on neural differentiation. This results in a two-fold increase in the

number of neurons in the brain during early development (Ishibashi et al., 1995;

Nakamura et al., 2000; Ishibashi, 2004). Double knockout mice in which both Hes1 and

Hes5 are inactivated results in an even more severe phenotype, in which radial glia

differentiate prematurely and lose their structural integrity to organize the neural tube.

This phenotype results from the loss in compensation by the other gene. To further

assess compensation by Hes3, triple knockout mice were generated and demonstrate that

these Hes genes are required for maintenance of radial glia as their ablation leads to

complete premature neuronal generation and loss of all progenitors (Hatakeyama et al.,

2004). Conversely, overexpression of Hes1 increases the number of mitotically active progenitors in the developing telencephalon (Ishibashi et al., 1994) stemming from the delayed arrest of proliferation, and causing hypercellularity of the cortex.

The Id family of bHLH factors consists of 4 members and is functionally similar to the

Hes family in their ability to promote progenitor proliferation and suppress the action of

proneural factors (Lyden et al., 1999). Moreover, Id sustains Hes1 expression to allow

for expansion of the progenitor pool and suppress precocious neuronal differentiation

(Bai et al., 2007). Specifically, Id2 overexpression in chick neural tube enhances Hes1

expression and inhibits expression of the proneural genes and therefore neuronal

differentiation. Moreover, Id1/3 double knockout mice or chicks electroporated for Id1-3

RNAi display precocious neurogenesis whereby Hes1 expression is suppressed and

proneural expands (Bai et al., 2007). In addition to regulation of bHLH

transcription factors, Id proteins stimulate cell cycle progression by inhibiting the action

37 of the retinoblastoma protein (Toma et al., 2000). Rb-deficient neural progenitor cells exhibit significant delays in cell cycle exit compared with wildtype cells, suggesting that

Rb is essential for cell cycle control of E2F activity required for cell cycle progression

(Callaghan et al., 1999). Mutation of the Rb protein leads to hypercellularity of the cortex due to increased proliferation of the neural progenitor cells (Ferguson et al., 2002).

The expression of Id1/4 is high in neural precursors and expression is extinguished upon exit from the cell cycle and differentiation (Iavarone and Lasorella, 2004). Interestingly,

Id1-4 are targets of MeCP2, the mutated gene in Rett Syndrome which is characterized by MR. In both a Mecp2 deficient mouse model, and human brain tissue from Rett

Syndrome patients, all four Id proteins were found to be elevated (Peddada et al., 2006).

The transition into neurogenesis is critically dependent upon expression of the proneural genes Mash1 and Ngn1/2. When the activity of these transcription factors is inhibited, a loss of neurons and neural progenitors results, accompanied by the precocious generation of glial-restricted progenitors and astrocytes (Guillemot, 2007). Interestingly, a novel role for p27 has recently identified its ability to promote neuronal differentiation through stabilization of Neurogenin 2 (Ngn2), and migration of post-mitotic projection neurons through inhibition of RhoA. Although both of these activities are independent of p27’s role in cell cycle regulation, they demonstrate a previously unappreciated role for p27 in neuronal differentiation at the level of bHLH transcription factors (Nguyen L 2006).

The transition from neurogenesis to gliogenesis is largely controlled by JAK-STAT signaling. Gliogenesis is suppressed during early neurogenic periods due to epigenetic factors such as DNA methylation, but more importantly, by proneural gene-mediated inhibition. Ngn1/2 inhibits STAT phosphorylation by sequestering CBP/p300 complexes

38 away from STAT activation sites. Once Hes/Id bHLH transcription factors dominate over the proneural transcription factors, STAT1/3 is phosphorylated and activated to promote astrogliogenesis through transcription of gfap (He et al., 2005).

MAPK and Cell Fate Determination

ERK1/2 is critical for development of the neocortex and specifically in cell fate

determination of neural progenitor cells. A series of elegant experiments by Miller and

colleagues have delineated the pathway by which endogenous growth factors such as

bFGF, NT-3 and BDNF activate a MEK-ERK-RSK-C/EBP signal transduction cascade

within neural progenitor cells to induce neuronal differentiation. Specifically, MEK

activation leads to activation of both ERK1/2 and RSK, which ultimately phosphorylates

the C/AAT Enhancer Binding Protein (C/EBP) to induce transcription of the neuron-

specific Tα1 tubulin isoform. The expression of Tα1 tubulin represents the initial phase

of and specifies neuronal fate. MEK regulation of C/EBP activity through ERK1/2 and RSK phsophorylation promotes neurogenesis and inihbits

gliogenesis without affecting progenitor proliferation or survival (Menard et al., 2002;

Barnabe-Heider and Miller, 2003; Paquin et al., 2005; Bartkowska et al., 2007; Calella et

al., 2007). In neural progenitor cells, ablation or mutation of a single member of the

scaffolding complex which links receptor tyrosine kinases to the activation of the MAP

Kinase cascade, such as SHP2 (Gauthier et al., 2007; Ke et al., 2007), FRS2 (Yamamoto

et al., 2005) or ShcA (McFarland et al., 2006) leads to reduced ERK1/2 phosphorylation and reduced neuronal differentiation. These data will be discussed in detail below.

39 Greenberg and colleagues (Bonni et al., 1997) have demonstrated that CNTF promotes

differentiation of cortical precursors into astrocytes through JAK-STAT and transient

ERK1/2 signaling; ERK1/2 inhibition was shown to augment CNTF-mediated GFAP

expression. C/EBP inhibition also leads to enhanced CNTF-mediated GFAP

Figure 1-6: Neurogenic versus gliogenic cell fate determination is regulated by ERK1/2 Activity.

transcription. These data demonstrate that ERK1/2 serves to mediate neurogenic cell fate determination and actively suppresses gliogenesis until the end of neurogenesis when proliferative and neurotrophic growth factors are downregulated. The neurogenic/

gliogenic switch identified by ERK1/2 activity is illustrated above in Figure 1-6.

40 ERK1/2 has also been shown to regulate bHLH transcription factors. Inhibition of

Ras/MAPK activity with MEK inhibitors reduces TGFβ-stimulated Hes1 transcription

(Stockhausen et al., 2005). Of particular importance is that Id 1/3 expression is positively regulated by phosphorylation (Deed et al., 1997; Hara et al., 1997) and specifically by

ERK activity (Bain et al., 2001; Ruzinova and Benezra, 2003). As Hes1 and the Id transcription factors suppress neurogenic bHLH transcription factors during the proliferative phases, these data support the idea that ERK1/2 may also play a role in ensuring sufficient proliferation of founder progenitor cells to enable the generation of a fully populated cortex.

Together, these data define a role for ERK1/2 in cell fate determination within the developing cerebral cortex whereby ERK1/2 activation promotes neurogenesis and suppresses gliogenesis to ensure the proper timing and generation of cells in the neocortex.

Disruptions in upstream elements of the MAPK signaling cascade lead to anatomical and behavioral impairments in mice

The effect of targeted mutations or deletions in the components of the MAPK cascade on cell fate determination leads to significant repercussions in formation and function of the cerebral cortex. Disruption or mutation of elements at each level of the MAPK cascade, from receptors at the plasma membrane to guanine nucleotide exchange factors, results in cortical malformations. Mice lacking the extrinsic ligand, FGF2, display reductions in neuronal density and cortical thickness due to a severe reduction in the number of neural

41 progenitor cell divisions. The loss of progenitors limits the numbers of neurons that are

able to be generated (50% fewer in FGF2 nulls as compared to wildtypes) (Dono et al.,

1998; Ortega et al., 1998; Vaccarino et al., 1999; Qian et al., 2000; Raballo et al., 2000;

Korada et al., 2002). Likewise, mice overexpressing a mutated FGF2 receptor, kinase

domain deficient FGFR1 mice, exhibit similar phenotypes with drastically reduced numbers of neural progenitors (Shin et al., 2004). Interestingly, the converse experiment

whereby a constitutively activated FGFR3 mutation was expressed in cortical progenitor

cells revealed that the activated receptor induced the generation of enlarged brains due to increased progenitor cell proliferation (Inglis-Broadgate et al., 2005; Thomson et al.,

2007). This effect was due to ERK1/2 activity, as treatment with MEK inhibitiors reversed the proliferative effects. These data reveal that ERK1/2 are critical components of FGF2 signaling necessary for the expansion of the neural progenitor cell pool and subsequent neurogenesis.

Recently, Miller and colleagues have demonstrated that inhibition or downregulation of the TrkB/C receptors also abrogates the proliferation and neuronal differentiation of neural progenitor cells (Bartkowska et al., 2007). These data elucidated a role for the neurotrophins BDNF and NT-3 in neurogenesis through their activation of TrkB and

TrkC. Trk activation leads to signaling through the MEK-ERK-RSK-C/EBP cascade.

Downstream of the RTKs, mutation and inactivation of the MAP3Ks, C-Raf/B-Raf/A-

Raf results in neurological perturbations. Activation of each isoform has been implicated in the neuron-like differentiation of PC12 cells, and is coupled to signaling from RTKs through to MEK1/2 and ERK1/2 (Wixler et al., 1996). Therefore, it is of importance to note the effects of inactivation of each. B-Raf knockouts die during embryogenesis and

42 display specific loss in both sensory and motoneurons (Wiese et al., 2001). A-Raf

knockout mice survive birth; however, they display distinct motor deficits and die in

early perinatal life (Pritchard et al., 1996). C-Raf knockouts also die embyronically,

displaying gross growth retardation (Wojnowski et al., 1998; Mikula et al., 2001). B-Raf

and C-Raf can compensate for each other, however. When A-Raf is expressed from the

B-Raf (in a B-Raf null mouse) although lethality is rescued, neural defects remain.

These mice continued to display deficits in neural progenitor cell proliferation, neuronal migration and dendritic maturation, all of which contribute to reductions in cortical thickness and disorganization of cortical lamination. These data demonstrate that B-Raf is specifically responsible for regulation of neural progenitor cell differentiation associated with cortical layering (Camarero et al., 2006). In a separate attempt to circumvent the lethality of B-Raf knockout mice, behavioral assessment of a spatially restricted B-Raf mutant, where inactivation occurs in the forebrain (predominantly hippocampus), spatial learning and hippocampal-dependent memory deficits were identified by the Morris Water Maze and context discrimination tasks, but not the formation or extinction of cued or contextual fear conditioning (Chen et al., 2006). These data further show that B-Raf has specific roles in cortical function. The generation of additional B-raf and Raf-1 (C-Raf) conditonal knockouts, where these genes are lost after recomibination driven by a nestin-cre transgene, lead to thinner cortices because of hypothalamic dysfunction and embryonic death (Zhong et al., 2007).

Further insight into the mechanism behind the cortical phenotypes associated with reductions in Raf function come from the inactivation of the MAP2Ks, MEK1/2. The

43 expression of dominant negative MEK constructs in cortical progenitor cells either

through in utero electroporation or by transfection/infection techniques demonstrates that

MEK-ERK activity is required for neuronal differentiation but not neural progenitor cell

survival or proliferation. Progenitors remain in an undifferentiated state (as defined by

nestin immunoreactivity) until gliogenic stimuli induce their differentiation into

astrocytes (Menard et al., 2002; Paquin et al., 2005). These data demonstrate a direct role

for MEK-ERK activity in promoting neuronal fate. In behavioral assays, mice which

express dominant negative MEK1 under the Tα1 tubulin promoter (expression occurs in

post-mitotic neurons) have poor performance on contextual fear conditioning paradigms,

despite displaying normal performance in motor coordination and learning, and cued fear conditioning (Shalin et al., 2004). These data demonstrate that MEK1 activity is required for hippocampal-dependent learning and memory. Ablation of various components of the signaling cascade which couples extracellular stimuli to MAPK activity supports the importance of MAPK signal transduction in cortical development and function.

Inactivation of the adaptor protein ShcA in the developing forebrain by Nestin-cre

recombination leads to a normally organized cortex of significantly reduced size. This

microcephalic brain arises from apoptosis of neural progenitor cells during early

neurogenesis (E10.5-12.5) and the specific reduction in neurons of the superficial layers

of the cortex. Interestingly, the neural progenitor cells do not display proliferation

defects in the ventricular or subventricular zones, nor do they undergo premature

differentiation (McFarland et al., 2006). SHP2 is the Src-homology 2 domain containing

tyrosine phosphatase which also acts as an adaptor protein for ERK activation. Two

independent reports demonstrate that Shp2 ablation or loss of function elicits impaired

44 corticogenesis, whereby neural progenitor cells lose their capacity to proliferate and

generate neurons, therefore giving rise to an abnormally large number of astrocytes

(Gauthier et al., 2007; Ke et al., 2007). Both studies identify a loss of ERK1/2 activation

as a result of Shp2 inactivation. Mice which have mutation in the Shp2 binding sites on

the FRS2α adaptor protein (FRS2α2F/2F mice) have normal gross morphology, but exhibit

a 10% reduction in brain size compared to wildtype mice, or those with mutations in

Grb2 binding sites on FRS2. These mice also have specific deficits in cortical thickness and lamination of the cortical plate at E18.5, which was attributed to the presence of many fewer cells. The number of intermediate progenitor cells and mitotic cells found in the subventricular zone were both found to be significantly reduced in FRS2α2F/2F mice.

Importantly, the level of ERK1/2 activity was dramatically lower throughout neurogenesis in FRS2α2F/2F mice (Hadari et al., 2001; Yamamoto et al., 2005).

Mice hypomorphic for the guanine nucleotide exchange factor, C3G, die at E14.5 due to

defects in blood vessel maturation; however, even at this age they exhibit increased neural progenitor proliferation and decreased neuronal differentiation (Voss et al., 2003;

Voss et al., 2006). These results were due to a delay or failure of the progenitors to exit the cell cycle.

Aberrant ERK1/2 Signaling leads to Developmental Disorders

45 Significantly, many components of the MAPK superfamily are responsible for genetic

Figure1-7: Mutation of nearly every component of the MAPK signaling pathway leads to a congenital disorder characterized by mental retardation and/or cardio-facio-cutaneous abnormalities.

developmental disorders, each of which is characterized by a form of mental retardation

(See Figure 7). As reviewed by Weeber and Sweatt (2002), Rsk2 mutations leads to

Coffin-Lowry Syndrome (Zeniou 2002, Dufresne 2000) and Rsk4 mutations are associated with X-linked MR. Mutations in the Ras-GAP, NF1Gap, leads to

Neurofibromatisis1 MR, while CREB Binding Protein (CBP) mutations cause

Rubenstein-Taybi Syndrome, and MSK2 has been mapped within the Bardet Biedl locus.

46 Mutations in MNK also contribute to Fragile X Syndrome. Furthermore, it has recently been appreciated that mutations in genes at each level of the MAPK signaling cascade are associated with Neuro-Cardio-Facio-Cutaneous (NCFC) Syndromes (Duesbery N and

Woude 2006, Bentires-Alj 2006). These include neurofibromatosis 1, NF1

(neurofibromin), Noonan Syndrome (SHP2), LEOPARD Syndrome (SHP2 and K-Ras),

Costello Syndrome (H-Ras), and CFC Syndrome (B-Raf, MEK1/2, and K-Ras).

Importantly, each of these syndromes presents with varying degrees of mental retardation.

ERK in Learning and Memory

There is now a significant body of work demonstrating the critical role the ERKs play in synaptic plasticity, learning and memory (reviewed in Adams and Sweatt, 2002). Sweatt and colleagues performed a series of seminal experiments demonstrating the requirement for ERK activation in synaptic plasticity by showing that electrical stimulation of the hippocampus resulted in ERK activation, the development of hippocampal LTP and expression of long term memory (English and Sweatt, 1996, 1997; Atkins et al., 1998).

Subsequent studies have demonstrated that ERKs are required for other forms of plasticity including the development of LTP in the dentate gyrus (Coogan et al., 1999), amygdala (Huang et al., 2000; Schafe et al., 2001), insular cortex (Berman et al., 1998) and visual cortex (Di Cristo et al., 2001). ERK signaling has been shown to be essential for spatial learning (Blum et al., 1999; Hebert and Dash, 2002), fear conditioning, (Atkins et al., 1998; Schafe et al., 2001) and conditioned taste aversion (Berman et al., 1998;

47 Schaeffer et al., 1998; Sharma et al., 2004; Vomastek et al., 2004). The molecular basis

of these effects is the subject of intense interest and a number of mechanisms have been

investigated, including ERK-dependent regulation of AMPA receptor insertion into the postsynaptic membrane (Lu et al., 2001; Shi et al., 2001), physical remodeling (Wu et al.,

2001; Goldin and Segal, 2003; Arendt et al., 2004) and generation (Futter et al., 2005) of dendritic spines, Kv4.2 potassium channel function (Yuan et al., 2002; Morozov et al.,

2003) and the local regulation of protein synthesis (Kelleher et al., 2004). The ERKs also serve to regulate the expression and function of transcription factors such as CREB, c-fos,

Zif268, Arc and Homer that regulate synaptic plasticity (Rosenblum et al., 2002;

Vazdarjanova et al., 2002; Bozon et al., 2003; Igaz et al., 2004). Of particular relevance is the report that mice expressing a dominant negative MEK1, the immediate upstream regulator of the ERKs, in cortical neurons display deficits in contextual memory formation, spatial reference memory, contextual fear conditioning (Shalin et al., 2004), as well as the late, translation-dependent, phase of LTP induction (Kelleher et al., 2004). It is unclear whether structural abnormalities within the cortex, or the molecular mechanism by which ERK regulates synaptic plasticity in mature neurons underlies these conditions.

As it has been recognized that mutation in both upstream and downstream components of

ERK1/2 are involved in cognitive ability, and ERK1/2 is a critical component of the mechanism underlying memory formation and retrieval, it seems likely that ERK1/2 are also involved in higher order cognition. Intriguingly, the ERK2 locus is at chromosome

22q11.2 in the human. This locus is responsible for 22q11Deletion Syndrome which is a

1.5 or 3 Mb microdeletion resulting in a congenital disorder that has been variously

48 termed DiGeorge Syndrome, Velocardiofacial Syndromes, Conotruncal Anomaly Face

Syndrome (and others) all manifesting as a Cardio-Facio-Cutaneous syndrome. In addition to craniofacial patterning and cardiac defects, these patients present with varying levels of cognitive deficits ranging from mild to severe mental retardation, a dramatically increased risk of schizophrenia or major depressive disorder and they exhibit small but significant reduced cortical volumes (Swillen et al., 2000; Maynard et al., 2002;

Campbell et al., 2006). ERK2 maps to a locus just distal to the typical 3Mb deletion breakpoint critical to DiGeorge Syndrome (Saitta et al., 1999) but is similarly flanked by low copy repeats which facilitate the larger microdeletion (Edelmann et al., 1999). A small cohort of individuals with a 1 Mb micodeletion encompassing ERK2 and who have a classic DiGeorge Syndrome phenotype has been identified by 2 independent studies

(Saitta et al., 1999; Shaikh et al., 2007; Ben-Shachar et al., 2008). The identification of these individuals is of relevance to the idea that ERK1/2 signaling is critical for cortical development and the subsequent learning deficits and cognitive disabilities associated with these syndromes.

Individual roles for the ERK1 and ERK2 isoforms?

The MAPKs, ERK1 and ERK2, share 84% sequence identity and are conserved from yeast through man (Boulton et al., 1990). Both isoforms are ubiquitously expressed, have identical substrate specificity and kinetics of activation. The ERKs have been implicated in an extraordinary range of biological phenomena and are among the best studied protein kinases (Pearson et al., 2001; Roux and Blenis, 2004). Our lab (Selcher et al., 2001;

Selcher et al., 2003; Nekrasova et al., 2005) and others (Pages et al., 1999) have

49 generated and characterized ERK1 knockout mice. These animals exhibit no overt

phenotype, are viable, fertile and have normal life spans. Pages et al. reported a modest

reduction in T-cell differentiation (Pages et al., 1999), which we have been unable to confirm in our ERK1 null animals (Nekrasova et al., 2005). We have found that ERK1 null animals exhibit a greater sensitivity to experimental allergic encephalitis (Nekrasova et al., 2005) and defects in innate immune system responses (Dillon et al., 2004). ERK1 is specifically required for adipogenesis (Bost et al., 2005). The ERK1 null animals are behaviorally normal with respect to acquisition and retention of contextual or cued fear conditioning and passive avoidance (Selcher et al., 2001) and did not have any deficits in their ability to induce hippocampal LTP (Selcher et al., 2001; Mazzucchelli et al., 2002).

These animals are hyperactive. A detailed examination of the ERK1 null animals by

Brambilla and colleagues revealed that the mutant animals showed a paradoxical improvement in a striatal-based long term memory task and facilitation of LTP in the nucleus accumbens (Mazzucchelli et al., 2002). ERK2 activation in the brains of the mutant animals was only modestly greater, but was sustained for longer periods. This was interpreted as a compensatory action of ERK2 within specific brain regions and that

ERK1 deletion enhances the action of ERK2 in the brain. ERK2 activity in the absence of ERK1 is clearly illustrated in ERK1 knockout mouse models which display enhanced behavioral plasticity and gene expression in response to cocaine (Ferguson et al., 2006).

The ERK pathway had been previously implicated in the behavioral and molecular responses of cocaine as the MEK selective inhibitor SL327 abolished cocaine-induced hyper-locomotion, reward-seeking behaviors, and the expression of c-fos after acute administration of cocaine to mice (Valjent et al., 2000); however, data from the ERK1

50 knockout mouse line suggests that this response can be facilitated solely by ERK2.

ERK1 null mice were also noted to exhibit behavioral excitement profiles similar to that evoked by psychostimulants. When these mice were treated with lithium, olanzapine, or valproate, drugs commonly used to treat manic phases of ADHD or bipolar disorder, this behavior was reversed (Engel et al., 2008). These data suggest that ERK1 and ERK2 may have specific roles in behavior, or that a threshold of ERK activity may be required for proper signaling which is not fulfilled in the absence of ERK1.

Several groups have generated animals in which the ERK2 gene has been knocked out.

Inactivation of the ERK2 gene results in embryonic lethality as early as E6.5 (Hatano et al., 2003; Saba-El-Leil et al., 2003; Yao et al., 2003); our unpublished observations. Two of these recent reports identify the cause of lethality as a defect in trophoblast development, resulting in failure of placenta formation. In contrast, Yao et al. found that in their ERK2 null line, embryonic lethality occurred by E11.5 resulting from the failure to induce mesoderm formation. Examination of ERK2 null embryonic stem cells and chimeric embryos revealed that ERK2 is not required for stem cell proliferation but has been argued to act principally to promote cellular differentiation (Kuida and Boucher,

2004; Meloche et al., 2004). A recent report has suggested that ERK1 and ERK2 specific actions may not exist and the localization and expression level of each isoform has made it appear that each isoform has specific functions (Lefloch et al., 2008). This hypothesis stems from the identification of skewed expression patterns of ERK1 and ERK2 in specific tissues and cell types.

51 Localization of the ERK isoforms is conferred by scaffolding proteins and specific

docking sites which hold the kinases at particular locations. For example, KSR1 binds

both ERK isoforms to ensure their localization at the plasma membrane where they can

be phosphorylated by Raf isoforms. Furthermore, translocation of ERK to the nucleus

generally leads to gene transcription and differentiation. In this regard, PEA-15 is an adaptor protein which binds to ERK1/2 exclusively in the cytoplasm and prevents

ERK1/2 nuclear translocation. This is of biological relevance due to the ability of PEA15

to prevent proliferation which requires the nuclear translocation of ERK1/2. β-arrestin is

another scaffolding protein which binds ERK1/2 to facilitate signaling by the kinases

through the generation of signaling complexes at the endosome. The localization of

ERK1/2 at yet a different cellular compartment, the Golgi apparatus, is mediated by its

interation with Sef, which captures activated MEK-ERK1/2 complexes at the Golgi and

restricts signaling to cytosolic targets. The MEK partner-1, MP-1, is a small scaffolding

protein which also localizes ERK1/2 to the endosome, and facilitates nuclear signaling.

It is of interest that MP-1 is selective for association with ERK1 thereby restricting

signaling by isoform specificity (Kolch, 2005). These data demonstrate that both tissue

expression of each isoform as well as the compartmentalization of ERK1/2 helps to

regulate the intensity and duration of ERK activity as well as the target specificity.

It has been of substantial interest to determine the specific targets and actions of the

ERKs. ERK1/2 activation in neurons is elicited during the normal processes of synaptic

transmission by membrane depolarization following neurotransmitter receptor activation

(Adams and Sweatt, 2002; Thomas and Huganir, 2004). Importantly, although both

52 ERK1 and ERK2 are highly expressed in hippocampus and cortex, ERK2 is the

predominant isoform that is expressed (in a 13:1 ratio in nude mice (Lefloch et al., 2008)) and activated by these stimuli, with very little activation of ERK1 observed (English and

Sweatt, 1996, 1997). In contrast, both ERK isoforms are activated by neurotrophin stimulation. The recent generation of transgenic mice which are hypomorphic for only

ERK2 has implicated this isoform in a dominant role for memory and behavior. This mouse loses 20% of its ERK2 expression within the hippocampus and displays significant deficits in cued and contextual fear conditioning. These data suggest that a minor loss in ERK2 levels is sufficient to induce measurable behavioral changes.

53 Research Goals:

The evidence reviewed above demonstrates a gap in the understanding of how

ERK2 regulates cortical development. Three facts about ERK1 and ERK2

underscore the importance of this research. First, the ERKs regulate both signaling

components and machinery important to the cell cycle. Moreover, ERK activity is a

critical determinant in neural progenitor cell fate determination. And, mutations in

both upstream and downstream components of the MAPK cascade are causative for

MR. Because ERK1 knockout animals are largely normal and ERK2 activity is

prevalent throughout the developing telencephalon, it is likely that ERK2 plays a significant role in contributing to normal cognitive function through development of the cerebral cortex and postnatally, through downstream signaling. The goal of this dissertation is to discern if loss of only the ERK2 MAP Kinase isoform affects cortical development and cognitive function.

We have investigated the specific actions of ERK1 and ERK2 in newly developed animal

models. Initially, we investigated cortical development of a mouse in which ERK2 is

deleted in neural progenitor cells. We have found that the deletion of ERK2 results in

profound perturbations of cortical development and learning/memory tasks.

Furthermore, we have identified individuals with chromosomal microdeletions including

the MAPK1 gene that exhibit neurocognitive deficits. These observations provide the

first definitive evidence for ERK2-dependent regulation of neural progenitor cell

proliferation and differentiation, and emphasize its critical role in these processes.

Second, we assessed the outcome of loss of ERK2 in neural progenitor cells in an animal

54 which has a complete loss of ERK1. Our findings suggest that ERK2 plays a dominant

role over ERK1 in processes critical to cortical development as these brains are largely similar to those missing only ERK2.

55

Literature Cited

Adams JP, Sweatt JD (2002) Molecular psychology: roles for the ERK MAP kinase cascade in memory. Annu Rev Pharmacol Toxicol 42:135-163. Aktas H, Cai H, Cooper GM (1997) Ras links growth factor signaling to the cell cycle machinery via regulation of cyclin D1 and the Cdk inhibitor p27KIP1. Mol Cell Biol 17:3850-3857. Anthony TE, Klein C, Fishell G, Heintz N (2004) Radial glia serve as neuronal progenitors in all regions of the central nervous system. Neuron 41:881-890. Arendt T, Gartner U, Seeger G, Barmashenko G, Palm K, Mittmann T, Yan L, Hummeke M, Behrbohm J, Bruckner MK, Holzer M, Wahle P, Heumann R (2004) Neuronal activation of Ras regulates synaptic connectivity. Eur J Neurosci 19:2953-2966. Atkins CM, Selcher JC, Petraitis JJ, Trzaskos JM, Sweatt JD (1998) The MAPK cascade is required for mammalian associative learning. Nat Neurosci 1:602-609. Bai G, Sheng N, Xie Z, Bian W, Yokota Y, Benezra R, Kageyama R, Guillemot F, Jing N (2007) Id sustains Hes1 expression to inhibit precocious neurogenesis by releasing negative autoregulation of Hes1. Dev Cell 13:283-297. Bain G, Cravatt CB, Loomans C, Alberola-Ila J, Hedrick SM, Murre C (2001) Regulation of the helix-loop-helix proteins, E2A and Id3, by the Ras-ERK MAPK cascade. Nat Immunol 2:165-171. Barnabe-Heider F, Miller FD (2003) Endogenously produced neurotrophins regulate survival and differentiation of cortical progenitors via distinct signaling pathways. J Neurosci 23:5149-5160. Bartkowska K, Paquin A, Gauthier AS, Kaplan DR, Miller FD (2007) Trk signaling regulates neural precursor cell proliferation and differentiation during cortical development. Development 134:4369-4380. Ben-Shachar S, Ou Z, Shaw CA, Belmont JW, Patel MS, Hummel M, Amato S, Tartaglia N, Berg J, Sutton VR, Lalani SR, Chinault AC, Cheung SW, Lupski JR, Patel A (2008) 22q11.2 distal deletion: a recurrent genomic disorder distinct from DiGeorge syndrome and velocardiofacial syndrome. Am J Hum Genet 82:214- 221. Berger-Sweeney J, Hohmann CF (1997) Behavioral consequences of abnormal cortical development: insights into developmental disabilities. Behav Brain Res 86:121- 142. Berman DE, Hazvi S, Rosenblum K, Seger R, Dudai Y (1998) Specific and differential activation of mitogen-activated protein kinase cascades by unfamiliar taste in the insular cortex of the behaving rat. J Neurosci 18:10037-10044. Blum S, Moore AN, Adams F, Dash PK (1999) A mitogen-activated protein kinase cascade in the CA1/CA2 subfield of the dorsal hippocampus is essential for long- term spatial memory. J Neurosci 19:3535-3544. Bond J, Woods CG (2006) Cytoskeletal genes regulating brain size. Curr Opin Cell Biol 18:95-101. Bonni A, Sun Y, Nadal-Vicens M, Bhatt A, Frank DA, Rozovsky I, Stahl N, Yancopoulos GD, Greenberg ME (1997) Regulation of gliogenesis in the central nervous system by the JAK-STAT signaling pathway. Science 278:477-483.

56 Bost F, Aouadi M, Caron L, Even P, Belmonte N, Prot M, Dani C, Hofman P, Pages G, Pouyssegur J, Le Marchand-Brustel Y, Binetruy B (2005) The extracellular signal-regulated kinase isoform ERK1 is specifically required for in vitro and in vivo adipogenesis. Diabetes 54:402-411. Boulton TG, Yancopoulos GD, Gregory JS, Slaughter C, Moomaw C, Hsu J, Cobb MH (1990) An insulin-stimulated protein kinase similar to yeast kinases involved in cell cycle control. Science 249:64-67. Bozon B, Kelly A, Josselyn SA, Silva AJ, Davis S, Laroche S (2003) MAPK, CREB and zif268 are all required for the consolidation of recognition memory. Philos Trans R Soc Lond B Biol Sci 358:805-814. Calella AM, Nerlov C, Lopez RG, Sciarretta C, von Bohlen Und Halbach O, Bereshchenko O, Minichiello L (2007) Neurotrophin/Trk receptor signaling mediates C/EBPalpha, -beta and NeuroD recruitment to immediate-early gene promoters in neuronal cells and requires C/EBPs to induce immediate-early gene transcription. Neural Develop 2:4. Callaghan DA, Dong L, Callaghan SM, Hou YX, Dagnino L, Slack RS (1999) Neural precursor cells differentiating in the absence of Rb exhibit delayed terminal mitosis and deregulated E2F 1 and 3 activity. Dev Biol 207:257-270. Camarero G, Tyrsin OY, Xiang C, Pfeiffer V, Pleiser S, Wiese S, Gotz R, Rapp UR (2006) Cortical migration defects in mice expressing A-RAF from the B-RAF locus. Mol Cell Biol 26:7103-7115. Cameron HA, Hazel TG, McKay RD (1998) Regulation of neurogenesis by growth factors and neurotransmitters. J Neurobiol 36:287-306. Campbell LE, Daly E, Toal F, Stevens A, Azuma R, Catani M, Ng V, van Amelsvoort T, Chitnis X, Cutter W, Murphy DG, Murphy KC (2006) Brain and behaviour in children with 22q11.2 deletion syndrome: a volumetric and voxel-based morphometry MRI study. Brain 129:1218-1228. Caviness VS, Jr., Takahashi T, Nowakowski RS (1995) Numbers, time and neocortical neuronogenesis: a general developmental and evolutionary model. Trends Neurosci 18:379-383. Caviness VS, Jr., Goto T, Tarui T, Takahashi T, Bhide PG, Nowakowski RS (2003) Cell output, cell cycle duration and neuronal specification: a model of integrated mechanisms of the neocortical proliferative process. Cereb Cortex 13:592-598. Chambard JC, Lefloch R, Pouyssegur J, Lenormand P (2007) ERK implication in cell cycle regulation. Biochim Biophys Acta 1773:1299-1310. Chelly J, Khelfaoui M, Francis F, Cherif B, Bienvenu T (2006) Genetics and pathophysiology of mental retardation. Eur J Hum Genet 14:701-713. Chen AP, Ohno M, Giese KP, Kuhn R, Chen RL, Silva AJ (2006) Forebrain-specific knockout of B- leads to deficits in hippocampal long-term potentiation, learning, and memory. J Neurosci Res 83:28-38. Coogan AN, O'Leary DM, O'Connor JJ (1999) P42/44 MAP kinase inhibitor PD98059 attenuates multiple forms of synaptic plasticity in rat dentate gyrus in vitro. J Neurophysiol 81:103-110. Corson LB, Yamanaka Y, Lai KM, Rossant J (2003) Spatial and temporal patterns of ERK signaling during mouse embryogenesis. Development 130:4527-4537.

57 Crome L (1972) Morphological aspects of mental retardation. Proc R Soc Med 65:587- 588. Deed RW, Hara E, Atherton GT, Peters G, Norton JD (1997) Regulation of Id3 cell cycle function by Cdk-2-dependent phosphorylation. Mol Cell Biol 17:6815-6821. Dehay C, Kennedy H (2007) Cell-cycle control and cortical development. Nat Rev Neurosci 8:438-450. Delalle I, Takahashi T, Nowakowski RS, Tsai LH, Caviness VS, Jr. (1999) Cyclin E-p27 opposition and regulation of the G1 phase of the cell cycle in the murine neocortical PVE: a quantitative analysis of mRNA in situ hybridization. Cereb Cortex 9:824-832. Desai AR, McConnell SK (2000) Progressive restriction in fate potential by neural progenitors during cerebral cortical development. Development 127:2863-2872. Di Cristo G, Berardi N, Cancedda L, Pizzorusso T, Putignano E, Ratto GM, Maffei L (2001) Requirement of ERK activation for visual cortical plasticity. Science 292:2337-2340. Dillon S, Agrawal A, Van Dyke T, Landreth G, McCauley L, Koh A, Maliszewski C, Akira S, Pulendran B (2004) A Toll-like receptor 2 ligand stimulates Th2 responses in vivo, via induction of extracellular signal-regulated kinase mitogen- activated protein kinase and c-Fos in dendritic cells. J Immunol 172:4733-4743. Dono R, Texido G, Dussel R, Ehmke H, Zeller R (1998) Impaired cerebral cortex development and blood pressure regulation in FGF-2-deficient mice. Embo J 17:4213-4225. Edelmann L, Pandita RK, Morrow BE (1999) Low-copy repeats mediate the common 3- Mb deletion in patients with velo-cardio-facial syndrome. Am J Hum Genet 64:1076-1086. Engel SR, Creson TK, Hao Y, Shen Y, Maeng S, Nekrasova T, Landreth GE, Manji HK, Chen G (2008) The extracellular signal-regulated kinase pathway contributes to the control of behavioral excitement. Mol Psychiatry. English JD, Sweatt JD (1996) Activation of p42 mitogen-activated protein kinase in hippocampal long term potentiation. J Biol Chem 271:24329-24332. English JD, Sweatt JD (1997) A requirement for the mitogen-activated protein kinase cascade in hippocampal long term potentiation. J Biol Chem 272:19103-19106. Englund C, Fink A, Lau C, Pham D, Daza RA, Bulfone A, Kowalczyk T, Hevner RF (2005) Pax6, Tbr2, and Tbr1 are expressed sequentially by radial glia, intermediate progenitor cells, and postmitotic neurons in developing neocortex. J Neurosci 25:247-251. Feng Y, Walsh CA (2004) Mitotic spindle regulation by Nde1 controls cerebral cortical size. Neuron 44:279-293. Ferguson KL, Vanderluit JL, Hebert JM, McIntosh WC, Tibbo E, MacLaurin JG, Park DS, Wallace VA, Vooijs M, McConnell SK, Slack RS (2002) Telencephalon- specific Rb knockouts reveal enhanced neurogenesis, survival and abnormal cortical development. Embo J 21:3337-3346. Ferguson SM, Fasano S, Yang P, Brambilla R, Robinson TE (2006) Knockout of ERK1 enhances cocaine-evoked immediate early gene expression and behavioral plasticity. Neuropsychopharmacology 31:2660-2668. Fishell G, Kriegstein A (2005) Cortical development: new concepts. Neuron 46:361-362.

58 Futter M, Uematsu K, Bullock SA, Kim Y, Hemmings HC, Jr., Nishi A, Greengard P, Nairn AC (2005) Phosphorylation of spinophilin by ERK and cyclin-dependent PK 5 (Cdk5). Proc Natl Acad Sci U S A 102:3489-3494. Gal JS, Morozov YM, Ayoub AE, Chatterjee M, Rakic P, Haydar TF (2006) Molecular and morphological heterogeneity of neural precursors in the mouse neocortical proliferative zones. J Neurosci 26:1045-1056. Gauthier AS, Furstoss O, Araki T, Chan R, Neel BG, Kaplan DR, Miller FD (2007) Control of CNS cell-fate decisions by SHP-2 and its dysregulation in Noonan syndrome. Neuron 54:245-262. Goldin M, Segal M (2003) and ERK involvement in dendritic spine plasticity in cultured rodent hippocampal neurons. Eur J Neurosci 17:2529-2539. Goto T, Mitsuhashi T, Takahashi T (2004) Altered patterns of neuron production in the p27 knockout mouse. Dev Neurosci 26:208-217. Gotoh Y, Nishida E, Matsuda S, Shiina N, Kosako H, Shiokawa K, Akiyama T, Ohta K, Sakai H (1991) In vitro effects on microtubule dynamics of purified Xenopus M phase-activated MAP kinase. Nature 349:251-254. Gotz M, Huttner WB (2005) The cell biology of neurogenesis. Nat Rev Mol Cell Biol 6:777-788. Graves LM, Guy HI, Kozlowski P, Huang M, Lazarowski E, Pope RM, Collins MA, Dahlstrand EN, Earp HS, 3rd, Evans DR (2000) Regulation of carbamoyl phosphate synthetase by MAP kinase. Nature 403:328-332. Guadagno TM, Ferrell JE, Jr. (1998) Requirement for MAPK activation for normal mitotic progression in Xenopus egg extracts. Science 282:1312-1315. Guerrini R, Marini C (2006) Genetic malformations of cortical development. Exp Brain Res 173:322-333. Guillemot F (2007) Cell fate specification in the mammalian telencephalon. Prog Neurobiol 83:37-52. Gysin S, Lee SH, Dean NM, McMahon M (2005) Pharmacologic inhibition of RAF-- >MEK-->ERK signaling elicits pancreatic cancer cell cycle arrest through induced expression of p27Kip1. Cancer Res 65:4870-4880. Hadari YR, Gotoh N, Kouhara H, Lax I, Schlessinger J (2001) Critical role for the docking-protein FRS2 alpha in FGF receptor-mediated signal transduction pathways. Proc Natl Acad Sci U S A 98:8578-8583. Hallberg B, Rayter SI, Downward J (1994) Interaction of Ras and Raf in intact mammalian cells upon extracellular stimulation. J Biol Chem 269:3913-3916. Hara E, Hall M, Peters G (1997) Cdk2-dependent phosphorylation of Id2 modulates activity of E2A-related transcription factors. Embo J 16:332-342. Hatakeyama J, Bessho Y, Katoh K, Ookawara S, Fujioka M, Guillemot F, Kageyama R (2004) Hes genes regulate size, shape and histogenesis of the nervous system by control of the timing of neural stem cell differentiation. Development 131:5539- 5550. Hatano N, Mori Y, Oh-hora M, Kosugi A, Fujikawa T, Nakai N, Niwa H, Miyazaki J, Hamaoka T, Ogata M (2003) Essential role for ERK2 mitogen-activated protein kinase in placental development. Genes Cells 8:847-856.

59 Haydar TF (2005) Advanced microscopic imaging methods to investigate cortical development and the etiology of mental retardation. Ment Retard Dev Disabil Res Rev 11:303-316. Haydar TF, Ang E, Jr., Rakic P (2003) Mitotic spindle rotation and mode of cell division in the developing telencephalon. Proc Natl Acad Sci U S A 100:2890-2895. Haydar TF, Nowakowski RS, Yarowsky PJ, Krueger BK (2000) Role of founder cell deficit and delayed neuronogenesis in microencephaly of the trisomy 16 mouse. J Neurosci 20:4156-4164. He F, Ge W, Martinowich K, Becker-Catania S, Coskun V, Zhu W, Wu H, Castro D, Guillemot F, Fan G, de Vellis J, Sun YE (2005) A positive autoregulatory loop of Jak-STAT signaling controls the onset of astrogliogenesis. Nat Neurosci 8:616- 625. Hebert AE, Dash PK (2002) Extracellular signal-regulated kinase activity in the entorhinal cortex is necessary for long-term spatial memory. Learn Mem 9:156- 166. Hong JW, Ryu MS, Lim IK (2005) Phosphorylation of serine 147 of tis21/BTG2/pc3 by p-Erk1/2 induces Pin-1 binding in cytoplasm and cell death. J Biol Chem 280:21256-21263. Horne MM, Guadagno TM (2003) A requirement for MAP kinase in the assembly and maintenance of the mitotic spindle. J Cell Biol 161:1021-1028. Huang YY, Martin KC, Kandel ER (2000) Both and mitogen-activated protein kinase are required in the amygdala for the macromolecular synthesis- dependent late phase of long-term potentiation. J Neurosci 20:6317-6325. Iacopetti P, Michelini M, Stuckmann I, Oback B, Aaku-Saraste E, Huttner WB (1999) Expression of the antiproliferative gene TIS21 at the onset of neurogenesis identifies single neuroepithelial cells that switch from proliferative to neuron- generating division. Proc Natl Acad Sci U S A 96:4639-4644. Iavarone A, Lasorella A (2004) Id proteins in neural cancer. Cancer Lett 204:189-196. Igaz LM, Bekinschtein P, Izquierdo I, Medina JH (2004) One-trial aversive learning induces late changes in hippocampal CaMKIIalpha, Homer 1a, Syntaxin 1a and ERK2 protein levels. Brain Res Mol Brain Res 132:1-12. Inglis-Broadgate SL, Thomson RE, Pellicano F, Tartaglia MA, Pontikis CC, Cooper JD, Iwata T (2005) FGFR3 regulates brain size by controlling progenitor cell proliferation and apoptosis during embryonic development. Dev Biol 279:73-85. Ishibashi M (2004) Molecular mechanisms for morphogenesis of the central nervous system in mammals. Anat Sci Int 79:226-234. Ishibashi M, Moriyoshi K, Sasai Y, Shiota K, Nakanishi S, Kageyama R (1994) Persistent expression of helix-loop-helix factor HES-1 prevents mammalian neural differentiation in the central nervous system. Embo J 13:1799-1805. Ishibashi M, Ang SL, Shiota K, Nakanishi S, Kageyama R, Guillemot F (1995) Targeted disruption of mammalian hairy and Enhancer of split homolog-1 (HES-1) leads to up-regulation of neural helix-loop-helix factors, premature neurogenesis, and severe neural tube defects. Genes Dev 9:3136-3148. Kao S, Jaiswal RK, Kolch W, Landreth GE (2001) Identification of the mechanisms regulating the differential activation of the mapk cascade by epidermal growth factor and nerve growth factor in PC12 cells. J Biol Chem 276:18169-18177.

60 Katz M, Amit I, Yarden Y (2007) Regulation of MAPKs by growth factors and receptor tyrosine kinases. Biochim Biophys Acta 1773:1161-1176. Kawada M, Yamagoe S, Murakami Y, Suzuki K, Mizuno S, Uehara Y (1997) Induction of p27Kip1 degradation and anchorage independence by Ras through the MAP kinase signaling pathway. 15:629-637. Ke Y, Zhang EE, Hagihara K, Wu D, Pang Y, Klein R, Curran T, Ranscht B, Feng GS (2007) Deletion of Shp2 in the brain leads to defective proliferation and differentiation in neural stem cells and early postnatal lethality. Mol Cell Biol 27:6706-6717. Kelleher RJ, 3rd, Govindarajan A, Jung HY, Kang H, Tonegawa S (2004) Translational control by MAPK signaling in long-term synaptic plasticity and memory. Cell 116:467-479. Kolch W (2005) Coordinating ERK/MAPK signalling through scaffolds and inhibitors. Nat Rev Mol Cell Biol 6:827-837. Korada S, Zheng W, Basilico C, Schwartz ML, Vaccarino FM (2002) Fibroblast growth factor 2 is necessary for the growth of glutamate projection neurons in the anterior neocortex. J Neurosci 22:863-875. Kuida K, Boucher DM (2004) Functions of MAP kinases: insights from gene-targeting studies. J Biochem 135:653-656. Kunath T, Saba-El-Leil MK, Almousailleakh M, Wray J, Meloche S, Smith A (2007) FGF stimulation of the Erk1/2 signalling cascade triggers transition of pluripotent embryonic stem cells from self-renewal to lineage commitment. Development 134:2895-2902. Lefloch R, Pouyssegur J, Lenormand P (2008) Single and combined silencing of ERK1 and ERK2 reveals their positive contribution to growth signaling depending on their expression levels. Mol Cell Biol 28:511-527. Liu X, Yan S, Zhou T, Terada Y, Erikson RL (2004) The MAP kinase pathway is required for entry into mitosis and cell survival. Oncogene 23:763-776. Lu W, Man H, Ju W, Trimble WS, MacDonald JF, Wang YT (2001) Activation of synaptic NMDA receptors induces membrane insertion of new AMPA receptors and LTP in cultured hippocampal neurons. Neuron 29:243-254. Lukaszewicz A, Savatier P, Cortay V, Giroud P, Huissoud C, Berland M, Kennedy H, Dehay C (2005) G1 phase regulation, area-specific cell cycle control, and cytoarchitectonics in the primate cortex. Neuron 47:353-364. Lyden D, Young AZ, Zagzag D, Yan W, Gerald W, O'Reilly R, Bader BL, Hynes RO, Zhuang Y, Manova K, Benezra R (1999) Id1 and Id3 are required for neurogenesis, angiogenesis and vascularization of tumour xenografts. Nature 401:670-677. Malatesta P, Hack MA, Hartfuss E, Kettenmann H, Klinkert W, Kirchhoff F, Gotz M (2003) Neuronal or glial progeny: regional differences in radial glia fate. Neuron 37:751-764. Marshall CJ (1995) Specificity of receptor tyrosine kinase signaling: transient versus sustained extracellular signal-regulated kinase activation. Cell 80:179-185. Martinez-Cerdeno V, Noctor SC, Kriegstein AR (2006) The role of intermediate progenitor cells in the evolutionary expansion of the cerebral cortex. Cereb Cortex 16 Suppl 1:i152-161.

61 Maynard TM, Haskell GT, Lieberman JA, LaMantia AS (2002) 22q11 DS: genomic mechanisms and gene function in DiGeorge/velocardiofacial syndrome. Int J Dev Neurosci 20:407-419. Mazzucchelli C, Vantaggiato C, Ciamei A, Fasano S, Pakhotin P, Krezel W, Welzl H, Wolfer DP, Pages G, Valverde O, Marowsky A, Porrazzo A, Orban PC, Maldonado R, Ehrengruber MU, Cestari V, Lipp HP, Chapman PF, Pouyssegur J, Brambilla R (2002) Knockout of ERK1 MAP kinase enhances synaptic plasticity in the striatum and facilitates striatal-mediated learning and memory. Neuron 34:807-820. McConnell SK, Kaznowski CE (1991) Cell cycle dependence of laminar determination in developing neocortex. Science 254:282-285. McFarland KN, Wilkes SR, Koss SE, Ravichandran KS, Mandell JW (2006) Neural- specific inactivation of ShcA results in increased embryonic neural progenitor apoptosis and microencephaly. J Neurosci 26:7885-7897. Meloche S, Vella FD, Voisin L, Ang SL, Saba-El-Leil M (2004) Erk2 signaling and early embryo stem cell self-renewal. Cell Cycle 3:241-243. Menard C, Hein P, Paquin A, Savelson A, Yang XM, Lederfein D, Barnabe-Heider F, Mir AA, Sterneck E, Peterson AC, Johnson PF, Vinson C, Miller FD (2002) An essential role for a MEK-C/EBP pathway during growth factor-regulated cortical neurogenesis. Neuron 36:597-610. Mikula M, Schreiber M, Husak Z, Kucerova L, Ruth J, Wieser R, Zatloukal K, Beug H, Wagner EF, Baccarini M (2001) Embryonic lethality and fetal liver apoptosis in mice lacking the c-raf-1 gene. Embo J 20:1952-1962. Miskimins WK, Wang G, Hawkinson M, Miskimins R (2001) Control of cyclin- dependent kinase inhibitor p27 expression by cap-independent translation. Mol Cell Biol 21:4960-4967. Mitsuhashi T, Aoki Y, Eksioglu YZ, Takahashi T, Bhide PG, Reeves SA, Caviness VS, Jr. (2001) Overexpression of p27Kip1 lengthens the G1 phase in a mouse model that targets inducible gene expression to central nervous system progenitor cells. Proc Natl Acad Sci U S A 98:6435-6440. Molnar Z (2004) Thomas Willis (1621-1675), the founder of clinical neuroscience. Nat Rev Neurosci 5:329-335. Morozov A, Muzzio IA, Bourtchouladze R, Van-Strien N, Lapidus K, Yin D, Winder DG, Adams JP, Sweatt JD, Kandel ER (2003) Rap1 couples cAMP signaling to a distinct pool of p42/44MAPK regulating excitability, synaptic plasticity, learning, and memory. Neuron 39:309-325. Nakamura Y, Sakakibara S, Miyata T, Ogawa M, Shimazaki T, Weiss S, Kageyama R, Okano H (2000) The bHLH gene hes1 as a repressor of the neuronal commitment of CNS stem cells. J Neurosci 20:283-293. Nekrasova T, Shive C, Gao Y, Kawamura K, Guardia R, Landreth G, Forsthuber T (2005) ERK1-deficient mice show normal T cell effector function and are highly susceptible to experimental autoimmune encephalomyelitis. J Immunol 175:2374- 2380. Nieto M, Schuurmans C, Britz O, Guillemot F (2001) Neural bHLH genes control the neuronal versus glial fate decision in cortical progenitors. Neuron 29:401-413.

62 Noctor SC, Martinez-Cerdeno V, Ivic L, Kriegstein AR (2004) Cortical neurons arise in symmetric and asymmetric division zones and migrate through specific phases. Nat Neurosci 7:136-144. Ohnuma S, Harris WA (2003) Neurogenesis and the cell cycle. Neuron 40:199-208. Ohtsuka T, Sakamoto M, Guillemot F, Kageyama R (2001) Roles of the basic helix-loop- helix genes Hes1 and Hes5 in expansion of neural stem cells of the developing brain. J Biol Chem 276:30467-30474. Ortega S, Ittmann M, Tsang SH, Ehrlich M, Basilico C (1998) Neuronal defects and delayed wound healing in mice lacking fibroblast growth factor 2. Proc Natl Acad Sci U S A 95:5672-5677. Pages G, Guerin S, Grall D, Bonino F, Smith A, Anjuere F, Auberger P, Pouyssegur J (1999) Defective thymocyte maturation in p44 MAP kinase (Erk 1) knockout mice. Science 286:1374-1377. Paquin A, Barnabe-Heider F, Kageyama R, Miller FD (2005) CCAAT/enhancer-binding protein phosphorylation biases cortical precursors to generate neurons rather than astrocytes in vivo. J Neurosci 25:10747-10758. Pearson G, Robinson F, Beers Gibson T, Xu BE, Karandikar M, Berman K, Cobb MH (2001) Mitogen-activated protein (MAP) kinase pathways: regulation and physiological functions. Endocr Rev 22:153-183. Peddada S, Yasui DH, LaSalle JM (2006) Inhibitors of differentiation (ID1, ID2, ID3 and ID4) genes are neuronal targets of MeCP2 that are elevated in Rett syndrome. Hum Mol Genet 15:2003-2014. Pontious A, Kowalczyk T, Englund C, Hevner RF (2008) Role of intermediate progenitor cells in cerebral cortex development. Dev Neurosci 30:24-32. Pouyssegur J, Volmat V, Lenormand P (2002) Fidelity and spatio-temporal control in MAP kinase (ERKs) signalling. Biochem Pharmacol 64:755-763. Pritchard CA, Bolin L, Slattery R, Murray R, McMahon M (1996) Post-natal lethality and neurological and gastrointestinal defects in mice with targeted disruption of the A-Raf protein kinase gene. Curr Biol 6:614-617. Qian X, Shen Q, Goderie SK, He W, Capela A, Davis AA, Temple S (2000) Timing of CNS cell generation: a programmed sequence of neuron and glial cell production from isolated murine cortical stem cells. Neuron 28:69-80. Raballo R, Rhee J, Lyn-Cook R, Leckman JF, Schwartz ML, Vaccarino FM (2000) Basic fibroblast growth factor (Fgf2) is necessary for cell proliferation and neurogenesis in the developing cerebral cortex. J Neurosci 20:5012-5023. Rakic P (1988) Defects of neuronal migration and the pathogenesis of cortical malformations. Prog Brain Res 73:15-37. Rakic P (1995) A small step for the cell, a giant leap for mankind: a hypothesis of neocortical expansion during evolution. Trends Neurosci 18:383-388. Rakic P (2000) Molecular and cellular mechanisms of neuronal migration: relevance to cortical epilepsies. Adv Neurol 84:1-14. Roberts EC, Shapiro PS, Nahreini TS, Pages G, Pouyssegur J, Ahn NG (2002) Distinct cell cycle timing requirements for extracellular signal-regulated kinase and phosphoinositide 3-kinase signaling pathways in somatic cell mitosis. Mol Cell Biol 22:7226-7241.

63 Roovers K, Assoian RK (2000) Integrating the MAP kinase signal into the G1 phase cell cycle machinery. Bioessays 22:818-826. Rosenblum K, Futter M, Voss K, Erent M, Skehel PA, French P, Obosi L, Jones MW, Bliss TV (2002) The role of extracellular regulated kinases I/II in late-phase long- term potentiation. J Neurosci 22:5432-5441. Ross M (1996) Cell division and the nervous system: regulating the cycle from neural differentiation to death. Trends Neurosci 19:62-68. Ross SE, Greenberg ME, Stiles CD (2003) Basic helix-loop-helix factors in cortical development. Neuron 39:13-25. Roux PP, Blenis J (2004) ERK and p38 MAPK-activated protein kinases: a family of protein kinases with diverse biological functions. Microbiol Mol Biol Rev 68:320-344. Rubinfeld H, Seger R (2005) The ERK cascade: a prototype of MAPK signaling. Mol Biotechnol 31:151-174. Ruzinova MB, Benezra R (2003) Id proteins in development, cell cycle and cancer. Trends Cell Biol 13:410-418. Saba-El-Leil MK, Vella FD, Vernay B, Voisin L, Chen L, Labrecque N, Ang SL, Meloche S (2003) An essential function of the mitogen-activated protein kinase Erk2 in mouse trophoblast development. EMBO Rep 4:964-968. Saitta SC, McGrath JM, Mensch H, Shaikh TH, Zackai EH, Emanuel BS (1999) A 22q11.2 deletion that excludes UFD1L and CDC45L in a patient with conotruncal and craniofacial defects. Am J Hum Genet 65:562-566. Sakakibara K, Kubota K, Worku B, Ryer EJ, Miller JP, Koff A, Kent KC, Liu B (2005) PDGF-BB regulates p27 expression through ERK-dependent RNA turn-over in vascular smooth muscle cells. J Biol Chem 280:25470-25477. Sauvageot CM, Stiles CD (2002) Molecular mechanisms controlling cortical gliogenesis. Curr Opin Neurobiol 12:244-249. Schaeffer HJ, Catling AD, Eblen ST, Collier LS, Krauss A, Weber MJ (1998) MP1: a MEK binding partner that enhances enzymatic activation of the MAP kinase cascade. Science 281:1668-1671. Schafe GE, Nader K, Blair HT, LeDoux JE (2001) Memory consolidation of Pavlovian fear conditioning: a cellular and molecular perspective. Trends Neurosci 24:540- 546. Selcher JC, Nekrasova T, Paylor R, Landreth GE, Sweatt JD (2001) Mice lacking the ERK1 isoform of MAP kinase are unimpaired in emotional learning. Learn Mem 8:11-19. Selcher JC, Weeber EJ, Christian J, Nekrasova T, Landreth GE, Sweatt JD (2003) A role for ERK MAP kinase in physiologic temporal integration in hippocampal area CA1. Learn Mem 10:26-39. Shaikh TH, O'Connor RJ, Pierpont ME, McGrath J, Hacker AM, Nimmakayalu M, Geiger E, Emanuel BS, Saitta SC (2007) Low copy repeats mediate distal chromosome 22q11.2 deletions: sequence analysis predicts breakpoint mechanisms. Genome Res 17:482-491. Shalin SC, Zirrgiebel U, Honsa KJ, Julien JP, Miller FD, Kaplan DR, Sweatt JD (2004) Neuronal MEK is important for normal fear conditioning in mice. J Neurosci Res 75:760-770.

64 Sharma C, Vomastek T, Tarcsafalvi A, Catling AD, Schaeffer HJ, Eblen ST, Weber MJ (2004) MEK partner 1 (MP1): Regulation of oligomerization in MAP kinase signaling. J Cell Biochem. Shaw P, Greenstein D, Lerch J, Clasen L, Lenroot R, Gogtay N, Evans A, Rapoport J, Giedd J (2006) Intellectual ability and cortical development in children and adolescents. Nature 440:676-679. Shi S, Hayashi Y, Esteban JA, Malinow R (2001) Subunit-specific rules governing AMPA receptor trafficking to synapses in hippocampal pyramidal neurons. Cell 105:331-343. Shin DM, Korada S, Raballo R, Shashikant CS, Simeone A, Taylor JR, Vaccarino F (2004) Loss of glutamatergic pyramidal neurons in frontal and temporal cortex resulting from attenuation of FGFR1 signaling is associated with spontaneous hyperactivity in mice. J Neurosci 24:2247-2258. Simpson PJ, Moon C, Kleman AM, Connolly E, Ronnett GV (2007) Progressive and inhibitory cell cycle proteins act simultaneously to regulate neurotrophin- mediated proliferation and maturation of neuronal precursors. Cell Cycle 6:1077- 1089. Slack RS, El-Bizri H, Wong J, Belliveau DJ, Miller FD (1998) A critical temporal requirement for the retinoblastoma during neuronal determination. J Cell Biol 140:1497-1509. Stefanovsky V, Langlois F, Gagnon-Kugler T, Rothblum LI, Moss T (2006a) Growth factor signaling regulates elongation of RNA polymerase I transcription in mammals via UBF phosphorylation and r-chromatin remodeling. Mol Cell 21:629-639. Stefanovsky VY, Langlois F, Bazett-Jones D, Pelletier G, Moss T (2006b) ERK modulates DNA bending and enhancesome structure by phosphorylating HMG1- boxes 1 and 2 of the RNA polymerase I transcription factor UBF. Biochemistry 45:3626-3634. Stockhausen MT, Sjolund J, Axelson H (2005) Regulation of the Notch target gene Hes-1 by TGFalpha induced Ras/MAPK signaling in human neuroblastoma cells. Exp Cell Res 310:218-228. Sweatt JD, Weeber EJ, Lombroso PJ (2003) Genetics of childhood disorders: LI. Learning and memory, Part 4: Human cognitive disorders and the ras/ERK/CREB pathway. J Am Acad Child Adolesc Psychiatry 42:741-744. Swillen A, Vogels A, Devriendt K, Fryns JP (2000) Chromosome 22q11 deletion syndrome: update and review of the clinical features, cognitive-behavioral spectrum, and psychiatric complications. Am J Med Genet 97:128-135. Takahashi T, Nowakowski RS, Caviness VS, Jr. (1993) Cell cycle parameters and patterns of nuclear movement in the neocortical proliferative zone of the fetal mouse. J Neurosci 13:820-833. Takahashi T, Nowakowski RS, Caviness VS, Jr. (1995) The cell cycle of the pseudostratified ventricular epithelium of the embryonic murine cerebral wall. J Neurosci 15:6046-6057. Takahashi T, Goto T, Miyama S, Nowakowski RS, Caviness VS, Jr. (1999) Sequence of neuron origin and neocortical laminar fate: relation to cell cycle of origin in the developing murine cerebral wall. J Neurosci 19:10357-10371.

65 Takuwa N, Takuwa Y (1997) Ras activity late in G1 phase required for p27kip1 downregulation, passage through the restriction point, and entry into S phase in growth factor-stimulated NIH 3T3 fibroblasts. Mol Cell Biol 17:5348-5358. Tarui T, Takahashi T, Nowakowski RS, Hayes NL, Bhide PG, Caviness VS (2005) Overexpression of p27 Kip 1, probability of cell cycle exit, and laminar destination of neocortical neurons. Cereb Cortex 15:1343-1355. Teixeira A, Chaverot N, Strosberg AD, Cazaubon S (2000) Differential regulation of cyclin D1 and D3 expression in the control of astrocyte proliferation induced by endothelin-1. J Neurochem 74:1034-1040. Thomas GM, Huganir RL (2004) MAPK cascade signalling and synaptic plasticity. Nat Rev Neurosci 5:173-183. Thomson RE, Pellicano F, Iwata T (2007) Fibroblast growth factor receptor 3 kinase domain mutation increases cortical progenitor proliferation via mitogen-activated protein kinase activation. J Neurochem 100:1565-1578. Toma JG, El-Bizri H, Barnabe-Heider F, Aloyz R, Miller FD (2000) Evidence that helix- loop-helix proteins collaborate with retinoblastoma tumor suppressor protein to regulate cortical neurogenesis. J Neurosci 20:7648-7656. Vaccarino FM, Schwartz ML, Raballo R, Nilsen J, Rhee J, Zhou M, Doetschman T, Coffin JD, Wyland JJ, Hung YT (1999) Changes in cerebral cortex size are governed by fibroblast growth factor during embryogenesis. Nat Neurosci 2:246- 253. Valjent E, Corvol JC, Pages C, Besson MJ, Maldonado R, Caboche J (2000) Involvement of the extracellular signal-regulated kinase cascade for cocaine-rewarding properties. J Neurosci 20:8701-8709. Vazdarjanova A, McNaughton BL, Barnes CA, Worley PF, Guzowski JF (2002) Experience-dependent coincident expression of the effector immediate-early genes arc and Homer 1a in hippocampal and neocortical neuronal networks. J Neurosci 22:10067-10071. Villanueva J, Yung Y, Walker JL, Assoian RK (2007) ERK activity and G1 phase progression: identifying dispensable versus essential activities and primary versus secondary targets. Mol Biol Cell 18:1457-1463. Vomastek T, Schaeffer HJ, Tarcsafalvi A, Smolkin ME, Bissonette EA, Weber MJ (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 101:6981-6986. Voss AK, Gruss P, Thomas T (2003) The guanine nucleotide exchange factor C3G is necessary for the formation of focal adhesions and vascular maturation. Development 130:355-367. Voss AK, Krebs DL, Thomas T (2006) C3G regulates the size of the cerebral cortex neural precursor population. Embo J 25:3652-3663. Wiese S, Pei G, Karch C, Troppmair J, Holtmann B, Rapp UR, Sendtner M (2001) Specific function of B-Raf in mediating survival of embryonic motoneurons and sensory neurons. Nat Neurosci 4:137-142. Wilkinson MG, Millar JB (2000) Control of the eukaryotic cell cycle by MAP kinase signaling pathways. Faseb J 14:2147-2157.

66 Willard FS, Crouch MF (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 13:653-664. Williams RW, Herrup K (1988) The control of neuron number. Annu Rev Neurosci 11:423-453. Witelson SF, Beresh H, Kigar DL (2006) Intelligence and brain size in 100 postmortem brains: sex, lateralization and age factors. Brain 129:386-398. Wixler V, Smola U, Schuler M, Rapp U (1996) Differential regulation of Raf isozymes by growth versus differentiation inducing factors in PC12 pheochromocytoma cells. FEBS Lett 385:131-137. Wojnowski L, Stancato LF, Zimmer AM, Hahn H, Beck TW, Larner AC, Rapp UR, Zimmer A (1998) Craf-1 protein kinase is essential for mouse development. Mech Dev 76:141-149. Wu GY, Deisseroth K, Tsien RW (2001) Spaced stimuli stabilize MAPK pathway activation and its effects on dendritic morphology. Nat Neurosci 4:151-158. Yamamoto S, Yoshino I, Shimazaki T, Murohashi M, Hevner RF, Lax I, Okano H, Shibuya M, Schlessinger J, Gotoh N (2005) Essential role of Shp2-binding sites on FRS2alpha for corticogenesis and for FGF2-dependent proliferation of neural progenitor cells. Proc Natl Acad Sci U S A 102:15983-15988. Yamamoto T, Ebisuya M, Ashida F, Okamoto K, Yonehara S, Nishida E (2006) Continuous ERK activation downregulates antiproliferative genes throughout G1 phase to allow cell-cycle progression. Curr Biol 16:1171-1182. Yao Y, Li W, Wu J, Germann UA, Su MS, Kuida K, Boucher DM (2003) Extracellular signal-regulated kinase 2 is necessary for mesoderm differentiation. Proc Natl Acad Sci U S A 100:12759-12764. York RD, Yao H, Dillon T, Ellig CL, Eckert SP, McCleskey EW, Stork PJ (1998) Rap1 mediates sustained MAP kinase activation induced by nerve growth factor. Nature 392:622-626. Yuan LL, Adams JP, Swank M, Sweatt JD, Johnston D (2002) Protein kinase modulation of dendritic K+ channels in hippocampus involves a mitogen-activated protein kinase pathway. J Neurosci 22:4860-4868. Zhong J, Li X, McNamee C, Chen AP, Baccarini M, Snider WD (2007) Raf kinase signaling functions in sensory neuron differentiation and axon growth in vivo. Nat Neurosci 10:598-607. Zhuo L, Theis M, Alvarez-Maya I, Brenner M, Willecke K, Messing A (2001) hGFAP- cre transgenic mice for manipulation of glial and neuronal function in vivo. Genesis 31:85-94.

67 DELETION OF ERK2 MAP KINASE IDENTIFIES ITS KEY ROLES IN

CORTICAL NEUROGENESIS AND COGNITIVE FUNCTION

Ivy S. Samuels1, J. Colleen Karlo1, Alicia N. Faruzzi2, Kathryn Pickering3, Karl Herrup4,

J. David Sweatt2 , Sulagna C. Saitta3, and Gary E. Landreth1

1Department of Neurosciences

Case Western Reserve University

Cleveland, OH 44106

2Department of Neurobiology and McKnight Brain Institute

University of Alabama, Birmingham

Birmingham, AL 35294

3Division of Human Genetics

The Children's Hospital of Philadelphia

University of Pennsylvania School of Medicine

Philadelphia, PA 19104

4Department of Cell Biology & Neuroscience

Rutgers University

Piscataway, NJ 08855-8082

68 Abstract

The MAP kinases, ERK1 and ERK2 are critical intracellular signaling intermediates;

however, little is known about their isoform-specific functions in vivo. We have examined the role of ERK2 in neural development by conditional inactivation of the murine mapk1/ERK2 gene in neural progenitor cells of the developing cortex. ERK

MAPK activity in neural progenitor cells is required for neuronal cell fate determination.

Loss of ERK2 resulted in a reduction in cortical thickness attributable to impaired proliferation of neural progenitors during the neurogenic period and the generation of fewer neurons. Mutant neural progenitor cells remained in an undifferentiated state until gliogenic stimuli induced their differentiation, resulting in the generation of more

astrocytes. The mutant mice displayed profound deficits in associative learning.

Importantly, we have identified patients with a 1 Mb microdeletion on chromosome

22q11.2 encompassing the MAPK1/ERK2 gene. These children, who have reduced

ERK2 levels, exhibit microcephaly, impaired cognition, and developmental delay. These

findings demonstrate an important role for ERK2 in cellular proliferation and differentiation during neural development as well as in cognition and memory formation.

Introduction

The Extracellular signal-Regulated Kinases, ERK1/2, are amongst the most prominent

signal transduction molecules through which extracellular stimuli are propagated from

the cell surface to cytoplasmic and nuclear effectors. ERK1 (mapk3) and ERK2 (mapk1)

exhibit 84% sequence identity and are uniquely and coordinately activated through the

sequential phosphorylation of the classical three-tier mitogen-activated protein kinase

69 (MAPK) cascade. Significantly, although both ERK1 and ERK2 are expressed

throughout the brain (Selcher et al., 2001; Mazzucchelli et al., 2002), genetic inactivation

of ERK1 has only subtle phenotypic effects while ERK2 inactivation results in early

embryonic lethality (Aouadi et al., 2006).

The orchestration of cortical development occurs primarily through the actions of growth

factors which signal through receptor tyrosine kinases (RTKs) to control cell cycle

initiation, progression, and cell fate decisions in neural progenitor cells (NPCs). The

mammalian neocortex arises from self-renewing multipotent NPCs that initially generate

neurons and subsequently produce glia. During neurogenesis, NPCs within the

ventricular zone (VZ) divide symmetrically to produce two neurons, or asymmetrically to

generate a self-renewing progenitor and a daughter cell, which becomes either a neuron

or an intermediate progenitor cell (IPC). The nascent IPC continues to undergo

asymmetric renewal or, more frequently, divides symmetrically yielding two neurons

(Noctor et al., 2004; Pontious et al., 2008). Gliogenesis commences late in

embryogenesis, with undifferentiated NPCs responding to stimuli secreted by newly born

neurons (Miller and Gauthier, 2007). Disruption of the dynamics of NPC or IPC

proliferation and differentiation result in alterations in cell number, cellular identity and

organization of the cortex (Dehay and Kennedy, 2007).

Signal transduction through the sequential activation of elements of the ERK MAPK cascade directs neurogenesis while concurrently suppressing gliogenesis (Miller and

Gauthier, 2007). It has been demonstrated that alteration of MAPK signaling abrogates the generation of a fully populated, normal size cortex (Vaccarino et al., 1999; Menard et

70 al., 2002; Barnabe-Heider and Miller, 2003; Dono, 2003; Ohkubo et al., 2004; Zheng et

al., 2004; Paquin et al., 2005; Thomson et al., 2007). Inhibition of the upstream activator

of the MAPKs, MEK1, causes NPCs to remain in the SVZ/VZ in an undifferentiated

state, blocking neurogenesis (Menard et al., 2002; Barnabe-Heider and Miller, 2003;

Paquin et al., 2005). Similarly, mutation or inactivation of scaffolding proteins such as

FRS2 and SHP-2 which link RTKs to MAPK activation have analogous effects on

cortical progenitors (Hadari et al., 1998; Yamamoto et al., 2005; Gauthier et al., 2007; Ke

et al., 2007).

It has recently been appreciated that a group of genetic disorders termed neuro-cardio-

facial-cutaneous syndromes (including Cardio-Facio-Cutaneous, Costello, LEOPARD

and Noonan Syndromes) are caused by mutations in upstream elements of the

ERK/MAPK signaling cascade (Bentires-Alj et al., 2006; Roberts et al., 2006) and are

collectively characterized by distinctive cardiac and craniofacial defects, developmental

delay, and mental retardation. The mutation of downstream elements in the ERK cascade

has similarly been associated with mental retardation syndromes (Weeber and Sweatt,

2002). These observations suggest that perturbations in ERK MAPK signaling underlie a

diverse range of neurodevelopmental syndromes.

In order to determine the role of ERK2 in cortical development and function, we generated a murine model in which mapk1/ERK2 undergoes conditional inactivation at the peak of neurogenesis. The loss of ERK2 resulted in the generation of fewer neurons and many more astrocytes in the cortex. This change in cell fate is due to alterations in

71 the dynamics of NPC proliferation and specifically with a reduction in IPC division.

Behaviorally, the mutant mice display profound impairments in cued and contextual fear

conditioning. The identification of individuals with haploinsufficiency for the

MAPK1/ERK2 gene as a result of distal microdeletions at 22q11.2 (Shaikh et al, 2007),

enabled the first analysis of humans with reduced expression levels of ERK2. These

patients exhibit microcephaly and neurodevelopmental deficits, consistent with the

phenotype observed in the murine models presented here. Taken together, these studies

provide direct evidence for the unique actions of the ERK2 isoform in neural

development and suggest its involvement in cognitive function.

Materials and Methods

Generation of a conditional ERK2 allele. A targeting construct was designed using the

loxP system to excise exon 2 in the mapk1 gene, resulting in a null allele. A neo cassette

flanked by loxP sites was inserted into an ApaI site upstream of exon 2, and a loxP site

was cloned into an Xba I site between exons 2 and 3. The βADT diphtheria toxin

cassette was used for negative selection. The construct was linearized and electroporated

into 129 ES cells, and clones were selected following treatment with G418. Clones

positive for homologous recombination were further transfected with a Cre-expressing

vector, and resultant clones were analyzed for deletion of the Neo cassette. One clone

had deletion of only the Neo cassette and was expanded and injected in C57BL/6

blastocysts. Resultant chimeras were mated to establish germline transmission, and

heterozygote (Mapk1flox/+) animals were intercrossed to establish a Mapk1flox/flox line of mice. hGFAP-cre mice on an FVB/N background were obtained from Dr. Albee

72 Messing (University of Wisconsin-Madison) and interbred with the ERK2 floxed line to

generate Mapk1flox/flox;hGFAPcre/+ animals in which the ERK2 gene has been

conditionally knocked out (ERK2 CKO). All experiments were performed in accordance

with the Case Western Reserve University Institutional Animal Care and Use Committee.

Southern blotting. Genomic DNA derived from cortex was digested with Xba, resolved

on a 0.8% agarose gel and transferred to a nylon membrane. The DNA was crosslinked

and hybridized to a 32P radiolabelled, 500 bp 3’ probe (located distal to the targeted region of the mapk1 locus, between exons 3 and 4) using ExpressHyb (Clonetech,

Mountain View, CA).

Cell Culture.

Cortical Progenitor Cell Culture. Cortical progenitor cells were cultured from embryonic day 14.5 (E14.5) mouse embryos using protocols previously described by

Miller and colleagues (Menard et al., 2002; Barnabe-Heider and Miller, 2003). Briefly,

cortices were dissected from E14.5 mouse embryos in ice-cold HBSS and transferred to

Neurobasal medium containing 500 µM L-glutamine, 2% B27 supplement, 1% penicillin-

streptomycin (Invitrogen, Gaithersburg, MD) and 40 ng/ml FGF2 (BD, Bedford, MA).

The tissue was mechanically triturated with a plastic pipette and plated into eight-well

chamber slides (Nunc, Naperville, IL) precoated with 2% laminin (BD) and 1% poly-D-

lysine (Sigma, St Louis, MO) at a cell density of 50,000 cells per well. Cells were fed

every 2 days. For experiments with CNTF, cells were plated in FGF2 for 12 hours and

73 then CNTF (3 ng/ml, Peprotech, Rocky Hill, NY) was added to the media for an additional 2 or 4 days.

Astrocyte Culture. Astrocytes were derived from the brains of neonatal mice as previously described but with some modifications (Giulian and Baker, 1986). Cerebral cortices were isolated from postnatal day 1-3 mice. Cortices were minced and cells were dissociated in PBS containing 0.25% trypsin and 1 mM EDTA. Digestion was terminated by adding an equal volume of DMEM/F12 medium (Life Technologies, Gaithersburg,

MD) containing 20% FCS, and cells were triturated to obtain a single-cell suspension.

Cells were plated and media was replaced the next day with DMEM/F12 containing 20%

FCS.

Cell Counts. Quantification of immunocytochemical staining was performed on dissociated cells or cortical progenitor cell cultures. Counts were made from at least 3 embryos of each genotype obtained from a minimum of 3 litters. For each condition,

>300 cells from 9 fields of a randomly selected reference space per well were counted and analyzed from 2 wells per embryo. Digital image acquisition was performed on a

Leitz DM R B microscope with Leica DC 500 camera and manufacturer’s software

(Cambridge, UK).

Viability Assay. Cellular viability was assessed by 3-[4,5-dimethylthiazol-2-yl]-2,5- diphenyltetrazolium bromide (MTT; Sigma) assay. In brief, primary CKO and WT astrocytes (5 x 103/well) were seeded in a 96-well plate. At 2 and 4 days, 10 µl of MTT

74 (5 mg/ml PBS) was added to each well, and plates were incubated at 37°C for 2 h.

Medium was then removed and cells were resuspended in 100 µl of DMSO. Cell

viability was assessed by colorimetric change using the SpectraMax 340 PC plate reader

(Molecular Devices, Sunnyvale, CA) at = 550 nm.

Immunocytochemistry. Immunocytochemistry of cultured cells was carried out by

washing the cultures with HEPES-buffered saline (HBS), followed by fixation with 4%

paraformaldehyde for 15 min. The cells were permeabilized with 0.2% NP-40 in HBS,

followed by incubation in blocking with buffer containing 6% normal goat serum (NGS) and 0.5% bovine serum albumin (BSA) for 1-2 h at room temperature. Cells were then

incubated with primary antibodies in HBS containing 3% NGS and 0.25% BSA at 4°C

overnight. After washing with HBS, cells were incubated with secondary antibodies

prepared in HBS containing 3% NGS and 0.25% BSA at room temperature for 1 h. Cells

were then washed with HBS, and counterstained with 1 µg/ml 4',6'-diamidino-2-

phenylindole (DAPI) for 2 min. Primary antibodies used were anti-ERK2 (BD, 1:200),

anti-Nestin (Chemicon, Temecula CA 1:200), and anti-βIII tubulin (1:4000), anti-MAP2

(1:200), anti-S100β (1:200), anti-Vimentin IgM (1:200, each from Sigma).

Immunohistochemistry. For immunohistochemistry, 10 µm cryosections were rinsed in

PBS, blocked, and permeabilized with 0.1% Triton X-100, 10% goat serum in PBS.

Sections were incubated with primary antibodies at 4°C overnight, washed with PBS, and incubated with secondary antibodies at room temperature for 1 h. The sections were then

counterstained with DAPI for 2 min. The primary antibodies used were anti-Tbr1 (a gift

of Dr. Robert Hevner, 1:2000), anti-Otx1 (Developmental Hybridoma Bank, 1:10), anti-

75 Brn1 (Santa Cruz,1:50), anti-GFAP (DAKO, 1:500), anti-NeuN (Chemicon, 1:300), anti-

phosphoHistone H3 (Millipore, Billerica, MA,1:200), anti-Zebrin (1:50), anti-Reelin

(Santa Cruz, 1:50), anti-GABA (Sigma, 1:500), anti-Caspase 3 (Cell Signaling 1:100).

The secondary antibodies used were Alexa fluorophore-conjugated IgG or IgM. TUNEL staining was performed as described by the manufacturer’s protocol (DeadEnd

Fluorometric TUNEL staining kit, Promega). Golgi staining was carried out using the

Rapid GolgiStain kit from FD Neurotechnologies (Ellicot City, MD).

For confocal analysis of sections, images were obtained using a Zeiss LSM 510 confocal laser microscope (Zeiss, Gottingen, Germany) using Argon and HeNe lasers

(excitation lines, 488 and 594 respectively) and Plan-Neofluar 10x, N.A. 0.3, Plan-

Neofluar 20x, N.A. 0.5, and C-Apochromat 40x, N.A. 1.2W corr objectives.

Photomicrographs were processed using Adobe Photoshop and cell number quantified

with Image ProPlus software.

Cell Counts. Analysis was performed on at least 8 tissue sections obtained from

four mice in each group. Counts were made within the frontal approximately 2.0mm

proximal to Bregma) and parietal lobes (approximately 1.25mm distal to Bregma), owing to the robust recombination in these areas. The primary motor cortex (“medial”) and primary somatosensory cortex (“lateral”) were identified in the same section of each lobe

using standard stereotaxic coordinates (Paxinos and Franklin, 2001). The reference space

was fixed by the reticule of the microscope objective, measuring 150 µm in width. Cell

layers were identified based on morphological distinctions and distance from the pial

76 surface. Quantification of cell numbers was performed using the Image ProPlus software

by blinded observer.

Western blot Analysis. Cells were washed with ice-cold HBSS and lysed by sonication in

lysis buffer (20mM Tris pH 7.5, 150mM NaCl, 1% NP-40, 10% glycerol, 1mM EDTA,

1.5mM MgCl2, 20mM NaF, 20mM β-glycerophosphate) supplemented with protease

inhibitors (1µg/ml leupeptin, 1µg/ml aprotinin, 1mM PMSF, 1mM Na3VO4). Lysates

were cleared by centrifugation and protein concentration was determined using the

bicinchoninic acid assay (Pierce, Rockford, IL) using BSA as a standard. Equal amounts

of protein were boiled in sample buffer, separated by SDS-PAGE gels, and transferred to

Immobilon-P PVDF membranes (Millipore). Membranes were blocked in 3% BSA (or

5% skim milk powder for anti-G3PDH) in TBS and 0.1% Tween 20 (TBS-T) for 2 h at

room temperature and incubated overnight at 4°C with primary antibodies. The primary

antibodies used were rabbit anti-pERK (Cell Signaling, Danvers, MA 1:1000), mouse

anti-ERK2 (BD, 1:3000) directed to the C-terminus of ERK2 or mouse anti-ERK2

that detects the N-terminus (Acris Antibodies, Herford Germany, 0.5 µg/ml),

mouse anti-ERK1 (Zymed 1:1000), anti-CrkL (Santa Cruz, 1:500), anti-G3PDH

(Trevigen, Gaithersburg, MD, 1:5000), anti-β-tubulin (Santa Cruz, 1:5000) and anti-c-src

(Santa Cruz, 1:500). After washing with TBS-T, membranes were incubated with HRP-

conjugated goat anti-mouse or anti-rabbit (1:5,000, GE Healthcare) secondary antibodies

in TBS-T with 5%milk for 2 h at room temperature. Blots were washed with TBS-T and

detection was performed using either Millipore or Pierce chemiluminescence reagents

using BioMax MR X-ray film (Eastman Kodak, Rochester, NY). Densitometric analysis

77 was performed using Adobe Photoshop histogram function, and statistical analysis was

done with GraphPad Prism software. Patient samples and clinical data were collected

after informed consent under an IRB approved protocol at the Children’s Hospital of

Philadelphia.

BrdU Labeling and Analysis. Pregnant mice were injected intraperitoneally with BrdU at

100 mg/kg of body weight. The animals were sacrificed 30 min after the injection.

BrdU-positive cells were detected by immunostaining as described by the manufacturers’

instructions with heat-mediated antigen retrieval in citrate buffer followed by 2N HCl-

treatment. Sections were incubated with a rat anti-BrdU antibody (Abcam, 1:1000)

overnight at 4oC. Quantification of BrdU+ cells during embryogenesis was performed by

dissociation of individual cortices and fixation in 4% paraformaldehyde immediately after sacrifice. Cells were plated at 50,000 cells per well and immunostained and counted as described above. For BrdU labeling of astrocytes, cells were plated at 50,000 cells per well in 8 well chamber slides overnight. Astrocytes were then treated with 1µg/ml BrdU for 24 hours and processed as described above.

Statistical Analysis. Statistical analysis was performed using either two-way ANOVA

with the Bonferroni post hoc test or the Student's t test, as indicated in the figures. In all graphs, the error bars indicate SD.

Behavioral Analysis. Animals used for behavioral analysis were bred at the University of

Alabama at Birmingham. They were housed with same sex siblings and with ad libitum

78 access to food and water, and they were maintained on a 12 hr light:dark cycle. Male

wild-type and ERK2 CKO littermates were compared in all behavioral experiments. All experiments were performed in accordance with the University of Alabama at

Birmingham Institutional Animal Care and Use Committee. All experiments were

performed with the experimenter blind to genotype.

The following assessments were performed similarly to those previously described in

(Selcher et al., 2001). Adult mice (14-40 weeks of age were tested for baseline

locomotor behavior and anxiety-like behavior using an open field test (Med-Associates,

Georgia, VT) which consisted of a 17 x 17 inch square polycarbonate box with three 16-

beam I/R arrays (measuring X, Y, and Z axis movement). Mice were placed into the

center of the open field to start the test, and activity was recorded for 15 min under

fluorescent lighting. Total distance moved was used as a measure of locomotor activity

and thigmotaxis (quantified by calculating the percent of total ambulatory distance that

was traveled in the center zone only of the open field) was used as a measure of baseline

anxiety-like behavior. Nociception and visual ability were performed as previously described (Crawley, 1999). Nociception was measured with a shock threshold test. One- second duration shocks were administered in increasing increments (0.08 mA, 0.10 mA,

0.15 mA, 0.25 mA, 0.35 mA, 0.45 mA, and 0.50 mA), each separated by 15-sec, and mice were observed for flinching, jumping, running, and vocalization at each increment.

The test was terminated when mice performed all four behaviors or when the highest shock increment was reached, whichever came first. Visual ability was measured using a

79 visual cliff test. All mice were observed during fear conditioning experiments for their

startle response to the onset of the tone used as a conditioned stimulus.

Fear Conditioning. Mice were tested for cued and contextual fear conditioning as previously described (Selcher et al., 2001) with some modifications. For training, animals were placed in a conditioning chamber with a metal grid floor (Video Fear

Conditioning, Med-Associates) for eight minutes. Training consisted of a 3-min exploration period followed by 3 tone-shock pairings consisting of a 30 sec-long 90dB tone with a 1-sec 0.5 mA shock occurring during the last second of the tone. Each tone- shock pairing was followed by a 60-sec rest except after the third pairing, which was followed by a 90-sec rest period leading up to the conclusion of the training session.

Twenty-four hours following training, each mouse was placed back into the training chamber for a 5-min contextual learning test, during which time freezing behavior was assessed in the absence of a tone or foot shock. One hour following the contextual test, cued learning was assessed by placing the mouse back into the chamber after the chamber’s context had been masked by altering its size, shape, color, and odor. After a

3-min exploration period of the novel context, the 90 dB conditioned stimulus tone was presented for 3 mins, during which time freezing behavior was assessed. For the repeated-training paradigm, animals were treated as noted, except that the training procedure was performed three times in the same day. Contextual and cued learning was then assessed 24 hours after the second training trial, using the procedure described above.

80 Results

ERK2 is required for normal development of the murine cerebral cortex

To directly assess the role of ERK2 in neural development, we generated a conditional mapk1/ERK2 null mutation. The mapk1flox allele was produced by introducing loxP sites into positions flanking exon 2 of the mapk1/ERK2 gene (Fig. 2-1A). Exon 2 is within the

ERK2 open reading frame and contains the highly conserved subdomain II, which is essential for catalytic activity of the kinase (Robinson et al., 1996). Transgenic mice in which cre recombinase expression is driven by the human glial fibrillary acidic protein

(hGFAP) promoter (Zhuo et al., 2001) were interbred with the mapk1flox/flox mice, resulting in deletion of ERK2 within cortical NPCs. Upon activation of the hGFAP promoter in these animals, cre expression led to the deletion of a 2kb genomic region that includes exon 2 of the mapk1/ERK2 gene. Cre-mediated recombination resulted in production of a 9.5kb fragment on the Southern blots (Fig. 2-1B). Western blot analysis of adult cortical brain homogenates confirmed that ERK2 protein levels in the cortex were reduced. No truncated fragments of ERK2 were detected by immunoblot analysis using an antibody directed to the N-terminus of the protein, confirming that the targeting strategy resulted in a null allele (Figure 2-1D). Expression of ERK1 was analyzed as a control and was not altered (Fig. 2-1C).

Mapk1flox/flox;hGFAPcre/+ mice are viable and fertile mice with normal life spans.

Macroscopic examination of these ERK2 Conditional Knockout (CKO) brains at P2 revealed no gross abnormalities (Fig. 2-2A) or differences in brain mass (CKO= 0.157g, n=6; WT= 0.167g, n=7; unpaired t-test, p=0.1466). Despite the normal overall

81 morphology and size of the ERK2 CKO cortex, we observed a significant reduction in cortical thickness throughout the dorsal telencephalon of CKO mice compared to wild- type littermates. A detailed regional analysis revealed a reduction in thickness of primary motor (0.8968), somatosensory (0.8673), auditory (0.8872) and visual (0.9042) cortices

(Fig. 2-2B). No deficits in the organization of specific cortical lamina were recognized; however, cellular density appeared greater throughout the ERK2 CKO cortex (Fig 2-2C).

ERK1 null mice were analyzed and were not different from wild-type controls (Selcher et al., 2001), establishing that specific inactivation of ERK2 causes the cortical thinning

(Fig. 2-2D).

The hGFAPcre/+ transgene is expressed in NPCs beginning at embryonic day 13.5 (Zhuo et al., 2001; Malatesta et al., 2003). This population of NPCs gives rise to >90% of

projection neurons and astrocytes populating the neocortex and hippocampus (Malatesta

et al., 2003), which we verified by interbreeding of Rosa26 reporter (R26R) mice

(Soriano, 1999) with mice carrying the hGFAPcre/+ transgene and subsequent LacZ

staining in the adult brain of cre+ progeny (Fig. 2-10). ERK2 is deleted throughout the dorsal telencephalon with only a minority of cells retaining expression. The ERK2 expressing cells include Tbr1+ cells populating layer VIb, and Reelin+ Cajal Retzius cells, both of which are generated prior to induction of cre expression. The majority of

ERK2-expressing cells in the mature cortex of the CKO mice are GABAergic interneurons, which arise and migrate from the ganglionic eminence that does not express the hGFAPcre/+ transgene (Supplemental Fig. 2-10). ERK2 expression was absent in the

hippocampus of the mutant mice; however, expression in the amygdala was similar to

82 that of wild-type animals, as the hGFAPcre/+ transgene is not expressed in this structure

(Supplemental Fig. 2-11).

We quantified cortical ERK2 protein levels at E14.5 and found a 50% reduction in ERK2

levels and a corresponding decrease in the levels of activated, phosphoERK2 in the

mutant mice. We did not detect any change in ERK1 expression (Fig. 2-3A,B) or

activation status at this time (Fig. 2-3C,D). Evaluation of these animals at E16.5 revealed

that the expression levels of ERK2 were reduced to 25% of wild-type levels, reflecting

the ongoing recombination of the mapk1/ERK2 gene and the half-life of the protein. The

level of active ERK2 was found to be 75% lower in the ERK2 CKO cells than those from

the wild-type animals (Fig. 2-3A,C) at E16.5. Although no compensatory increase in

ERK1 expression was found, phosphorylated ERK1 levels were dramatically (4-fold)

higher in the ERK2 CKO as compared to the wild-type cortex at this time (Fig. 2-3B,D).

These data provide evidence that while ERK1 expression was not elevated to compensate

for the loss of ERK2, there was a significant increase in ERK1 activity at E16.5.

ERK2 inactivation alters the cellular composition of the cortex

To determine the basis for the cortical thinning and assess the cellular density within the cortex (Fig. 2-2B,C), we quantified both total number of DAPI+ cells per cortical region

as well as the number of neurons (NeuN+ cells) in ERK2 CKO and wild-type littermates at postnatal day 10. The area from the VZ/SVZ boundary to the pial surface of the cortex

was analyzed in 150 µm sectors from sections of the frontal and parietal lobes. We found

no difference in DAPI+ cells per cortical region between genotypes, a finding consistent

83 with the observation of increased cellular density and reduced cortical thickness.

Strikingly, we found 10% fewer NeuN+ neurons per region in the ERK2 CKO as compared to wild-type littermates (Fig. 2-4A). However, there was a 40% increase in non-neuronal (DAPI; NeuN-), presumptive glial cells. These data indicate that ERK2 inactivation resulted in the production of fewer neurons and more non-neuronal cells, which was associated with an overall reduction in cortical thickness.

To determine if neuronal identity was affected by the loss of ERK2, a detailed examination of the number of neurons within each cortical lamina was performed.

Cortices from P2 wild-type and ERK2 CKO littermates were stained with antibodies to transcription factors whose expression is limited to specific cortical layers. We found a

35% reduction in number of Tbr1+ neurons in layer VI and the cortical preplate (Fig. 2-

4B) of ERK2 CKO animals. Similarly, we detected approximately half as many Otx1+ neurons populating layer V in the ERK2 CKO as compared to wild-type animals (Fig. 2-

4C). Evaluation of layers II-IV using Brn1 revealed an analogous, genotype-dependent

30% decrease in neurons populating this layer (Fig. 2-4D). No significant differences were found between medial and lateral regions of the cortex. Thus, inactivation of ERK2 was associated with an overall reduction in the numbers of neurons populating the cortex.

To confirm the identity of the non-neuronal (DAPI+; NeuN-) cells, we performed immunohistochemistry with astrocyte-specific markers. We found a dramatic increase in

Zebrin II+ (Fig. 2-5A) and GFAP+ astrocytes (Fig. 2-5B) throughout the cortex and subventricular zone. In order to assess the proliferative capacity of ERK2 CKO

84 astrocytes as compared to wild-type astrocytes, we cultured these cells from each genotype and evaluated their proliferation by measuring BrdU incorporation, Ki67 immunoreactivity and change in cell number (Fig. 2-5C,D,E respectively). We also assessed astrocyte metabolism and viability using an MTT assay (Fig. 2-5F). No significant differences were found in any of these parameters. We interpret these data to suggest that loss of ERK2 does not significantly affect astrocyte viability or their intrinsic ability to proliferate. We did not observe increasing numbers of astrocytes in older animals.

Thus, these data demonstrate that the loss of ERK2 from NPCs results in the reduction of neurons populating those cortical layers that are generated subsequent to its inactivation.

We further document that the increase in cellular density is caused by reductions in cortical thickness as the total number of DAPI+ cells is equivalent between genotypes.

We also demonstrate that many more astrocytes are present in the ERK2 CKO cortex.

We next investigated whether the reduction in neuronal cell number resulted from enhanced apoptosis. Analysis of E17 embryos by immunostaining for cleaved caspase 3 and by TUNEL assay revealed no differences between genotypes (Fig. 2-12). These data suggest that loss of ERK2 does not provoke neuronal death. Additionally, we assessed whether incomplete neuronal differentiation might contribute to the reduction in cortical thickness. Brains from four month-old CKO and wild-type littermates were Golgi- stained, and individual projection neurons within layers 3 and 5 of the primary motor cortex were traced and reconstructed with Neurolucida. We did not observe any

85 differences in dendritic length, branching, volume, or surface area (Supplemental Table 1

and 2). We also did not identify any structural changes in axonal projections.

ERK2 inactivation alters the dynamics of neurogenesis

The ERKs play critical roles in regulating cellular proliferation (Chambard et al., 2006)

and neuronal differentiation of cortical NPCs (Menard et al., 2002; Barnabe-Heider and

Miller, 2003; Paquin et al., 2005). To determine when the loss of ERK2 affects cell fate

determination, we investigated the proliferation of NPCs by evaluation of BrdU

incorporation. Pregnant dams were injected with BrdU, embryos were harvested 0.5 hr

post injection, and coronal sections of embryonic brains were immunostained with anti-

BrdU and anti-phosphohistone H3 antibodies. During S phase, BrdU is incorporated into

NPCs that are actively replicating DNA. Phosphorylated-Histone H3 (pH3) marks

mitotic cells during G2 and M phase of the cell cycle. At E12.5 (Fig. 2-6A), prior to the

induction of the hGFAPcre/+ transgene, the incorporation of BrdU and the number of ventricular cells undergoing mitosis was indistinguishable between genotypes. However,

when pregnant dams were injected on E14.5, the peak of the neurogenic period and one

day following initiation of hGFAP-cre expression, we observed significantly less BrdU

incorporation and pH3 immunoreactivity in the ERK2 CKO cortex (Fig. 2-6B). To

quantify the difference, individual embryonic cortices were dissected 30 minutes post-

injection, dissociated, plated, and immediately processed for immunostaining. We found

a 44% reduction in the number of BrdU+ cells in the E14.5 ERK2 CKO cortex

(CKO=6.7%; WT=11.87%, n=3 for each; unpaired t-test, p<0.0001). Importantly, we

observed a dramatic reduction in the number of abventricular pH3+ cells following ERK2

86 inactivation. These cells were identified as intermediate progenitor cells by co-labeling with Tbr2, a marker of this cellular population (Hevner et al., 2006) (Fig. 2-6D). There was no significant genotype-related difference in the total number of mitotic pH3+ cells

(CKO=101; WT=104.3, for each n=3, unpaired t-test, p=0.6005) as counted in tissue sections. Similarly, we did not observe any significant difference in the number of Tbr2+

IPCs or in Pax6+ progenitors (Fig. 2-13). However, the number of dividing IPCs (pH3+;

Tbr2+) was significantly reduced (by 37%) in the ERK2 CKO cortex (CKO=13%; WT=

20.67% per section, n=3 animals per genotype, unpaired t-test, p=0.0192).

In contrast to the results at E14.5, analysis of NPC proliferation at E16.5 revealed

dramatically more BrdU incorporation (Fig. 2-6C). Quantification of dissociated cortices

revealed 2.5 fold more BrdU+ cells in the CKO at E16.5 when compared to wild-type

littermates (CKO=9.41%; WT= 4.18%, n=4 animals per genotype, unpaired t-test,

p=0.005). A moderate increase in the numbers of pH3+, mitotic cells lining the ventricle

(Fig. 2-6C, E) was also observed in the ERK2 CKO; however, no increase in

abventricular pH3+ IPCs was identified (Fig 2-6E). Wild-type animals have passed

through the neurogenic period and exhibit few proliferating or dividing cells at this time.

Taken together, these findings suggest that ERK2 inactivation delays NPC proliferation

until E16.5, and suppresses the division of IPCs at the peak of the neurogenic period,

resulting in fewer neurons populating the ERK2 CKO cortex. The data also suggest a

reinitiation of progenitor proliferation during the period of normal gliogenesis in the

mutant cortex, resulting in more astrocytes.

87 ERK2 CKO neural progenitors generate fewer neurons in vitro

In order to evaluate whether ERK2 inactivation affected intrinsic cell fate determination

by NPCs, we performed in vitro analysis of cortical progenitor cell cultures (Menard et al., 2002; Paquin et al., 2005). Cortical progenitor cells from ERK2 CKO and wild-type

littermate embryos were cultured at E14.5 and grown under proliferating conditions in

Neurobasal media supplemented with FGF2 for 2 days. Importantly, we observed a

significant increase (1.3 fold) in the number of uncommitted, Nestin+ NPCs (Fig. 2-7A1-

2, E) in the ERK2 CKO cultures compared to those of wild-type littermates.

Furthermore, there were fewer immature βIII-tubulin+ (0.8 fold) and mature MAP2+ (0.7

fold) neurons in the ERK2 CKO cultures compared to those from wild-type animals (Fig.

2-7B, C, E). To assess if ERK2 CKO progenitors became committed to glial fates in

vitro, we counted S100β+ cells. ERK2 CKO cultures have 20% fewer S100β+ cells than

wild-type cultures (Fig. 2-7D-E), indicating that NPCs lacking ERK2 persist in an

undifferentiated state. Our finding is consistent with a previous report demonstrating that

NPCs transfected with a dominant negative MEK1-construct remain in an

undifferentiated state until gliogenic signals induce astrocyte differentiation (Paquin et al., 2005).

It has been suggested that ERK1/2 activation actively suppresses gliogenesis elicited by

CNTF/LIF (Bonni et al., 1997; Paquin et al., 2005). To assess if the loss of ERK2 affects cellular fate in a gliogenic environment, we added CNTF to cultures 12 hrs after plating.

A significant (1.4 fold) increase in vimentin+ astrocytes in the ERK2 CKO cultures was observed after treatment for either 2 or 4 days with CNTF (Fig. 2-8A1-4, D). Similar

88 results were found at both 2 and 4 days in vitro when cultures were assessed with the

mature astrocyte marker, GFAP (1.7 fold at 2DIV, 1.6 fold at 4DIV) and the immature

astrocyte marker, S100β (1.6 fold at 2DIV, 1.3 fold at 4DIV), (Fig 2-8B-F). These data

indicate that ERK2 normally acts to suppress gliogenesis and that this isoform alone may

be sufficient for directing cell fate decisions. In summary, our in vitro analysis confirms

and validates the conclusion that ERK2 participates in neurogenic cell divisions.

Furthermore, loss of ERK2 does not change the differentiative potential of the NPCs, as

they remain in an undifferentiated nestin+ state until gliogenic stimuli induce astrocytic

differentiation.

Mature ERK2 CKO mice display impaired associative learning

Due to the changes in cellular composition observed in the ERK2 CKO cortex, and the

critical roles the ERKs play in learning and memory (Selcher et al., 2001; Mazzucchelli

et al., 2002), we tested whether ERK2 CKO mice showed impairments in long-term

memory using two associative learning tests. Cued and contextual fear conditioning are

sensitive to disruption of the MAPK signaling pathway as MEK or MAPK inhibitors

have been shown to attenuate both forms of memory (Walz et al., 1999; Walz et al.,

2000; Sweatt, 2001; Shalin et al., 2004). ERK2 CKO mice exhibited significant deficits

in long term memory for both contextual and cued conditioning compared to littermate

wild-types (Fig. 2-9A,B). Further testing was conducted to determine whether

overtraining could rescue the deficits in contextual and cued conditioning; however, despite repeated training, ERK2 CKO mice continued to display significantly less freezing behavior during testing as compared to littermate wild-types (t15 = 3.65 and 3.35,

89 respectively, p<0.05; Fig. 2-9B). These data indicate a striking deficit in associative fear

memory in ERK2-deficient animals. Differences in freezing behavior could not be

explained by deficits in baseline sensory responses and baseline activities, as wild-type

and CKO mice showed no differences in locomotor behavior (CKO = 3300 cm, WT =

2450 cm, t15 = -1.536) or general anxiety-like behavior measured by thigmotaxis

(CKO=73.386%, WT=77.116%, t15 = 0.027) in the open field, in threshold to respond to

a footshock (flinch: CKO= 0.083 mA, WT= 0.087 mA, t15 = 0.98; jump: CKO= 0.369 mA, WT= 0.389 mA, t15 = 0.42; run: CKO= 0.141 mA, WT= 0.183 mA, t15 = 1.27;

vocalize: CKO= 0.275 mA, WT= 0.2560 mA, t15 = -0.45), or in visual ability (CKO=

71.429%, WT= 66.6670%, t15 = -0.19), which was measured with a visual cliff test; n=8

for CKO, n=10 for WT and p>0.05 for all baseline experiments. We conclude that the

selective loss of the ERK2 gene results in a pronounced associative memory deficit, in

the absence of generalized alterations in sensory responses.

MAPK1/ERK2 deficiency in humans is associated with learning deficits

The human MAPK1 gene encoding ERK2 is positioned on the long (q) arm of

. Chromosomal microdeletions frequently occur at 22q11.2 due to the

presence of highly homologous intrachromosomal low copy repeats (LCRs) that mediate

aberrant recombination events (Emanuel et al., 2001). The most common recurrent

deletion syndrome in humans involves the proximal LCRs (A-D) and leads to 22q11

Deletion/DiGeorge Syndrome (Fig. 2-14). Recently, a number of unrelated patients have

been identified with a 1 Mb deletion mediated by the distal LCRs, D-E (Rauch et al.,

1999; Saitta et al., 1999; Shaikh et al., 2007; Ben-Shachar et al., 2008). These deletions

90 include MAPK1/ERK2, and while they do not overlap the proximal 3Mb deletion causing

DiGeorge Syndrome, they independently exhibit a similar spectrum of craniofacial abnormalities, cardiac defects and neurodevelopmental defects. Several other genes including HIC2, YPEL1, PPIL2, UBE2L3, SDF2L1, and PPM1F occupy the D-E interval. A cluster of genes encoding the immunoglobulin lambda light chain are also located in the distal part of the deletion as well; however, none of these genes have been associated with human disorders (SCS, unpublished data). Thus far, the characteristics of

8 individuals that have DiGeorge-like symptoms but which carry atypical distal deletions have been identified (Rauch et al., 1999; Saitta et al., 1999; Shaikh et al., 2007; Ben-

Shachar et al., 2008). We have analyzed two patients with identical distal 1 Mb deletions. Patient 1 is a male with microcephaly, (head circumference consistently <3% for age and sex-matched norms measured over 4 years) and neurocognitive deficits

(Saitta et al., 1999) that include delays in fine motor function, expressive speech development, and in complex language comprehension. Patient 2 is a recently diagnosed female with microcephaly (head circumference <3% for age and sex-matched norms), a history of delays in expressive speech development and decreased academic performance overall. Analysis of cellular levels of ERK2 in lymphoblasts obtained from these patients with the distal deletion exhibit decreased levels of ERK2 expression relative to either normal individuals (n=4), or those with a typical 3 Mb DiGeorge deletion (n=1).

Densitometric analysis of western blots revealed that ERK2 expression in the distal microdeletion patients was 0.60876 of normal levels (p=0.0026). In contrast, levels of

CRKL, encoded by a gene located within the 3Mb classical deletion interval of DiGeorge syndrome were normal in patients with the distal 22q11 deletions. These data

91 demonstrate that patients with chromosomal microdeletions of distal chromosome 22q have a phenotype that includes microcephaly, neurodevelopmental and neurocognitive deficits, but is not due to genes in the DiGeorge interval. These children instead, have a small microdeletion that includes the MAPK1/ERK2 gene, show reduced ERK2 levels and display cognitive deficits. These findings are in line with our data from the murine model generated in this study where cortical hypoplasia and cognitive deficits in associative learning are found with genetic deletion of mapk1/ERK2.

Discussion

We have identified a specific role for ERK2 in the development of the cerebral cortex and shown that its deletion in the developing telencephalon is associated with the loss of higher order functionality in the mature brain. These findings provide concrete evidence for isoform-specific actions of the ERKs and follow from our previous studies demonstrating that mice in which the mapk3/ERK1 gene was knocked out did not have an overt phenotype and exhibited only modest behavioral changes (Pages et al., 1999;

Selcher et al., 2001; Mazzucchelli et al., 2002).

The developing cortex is exquisitely sensitive to perturbations of the tightly regulated patterns of cellular proliferation and differentiation. The conditional inactivation of

ERK2 within neural progenitors at the beginning of the neurogenic period resulted in fewer neurons populating the cortex. This effect resulted from the dramatic reduction in both the number of progenitors proliferating within the ventricular zone and the number of mitotic abventricular intermediate progenitor cells at E14.5. Importantly, the loss of

92 ERK2 appears only to delay initiation or stall the progression of neural progenitors through the cell cycle during the neurogenic period, as these cells neither die, nor migrate prematurely. It is of importance in this regard that isolated NPCs maintain their ability to undergo properly timed mitotic divisions, generating successive and appropriately specified lineages of neurons (Shen et al., 2006). Our findings demonstrate that loss of

ERK2 does not disrupt the lineage specification of neurons populating any of the cortical lamina. These findings are consistent with in vitro studies that have shown that ERK2 regulates initiation and progression through the G1 phase of the cell cycle (Liu et al.,

2004; Chambard et al., 2006; Meloche and Pouyssegur, 2007). The suppression of NPC proliferation as a consequence of ERK2 inactivation is transient, as we observed a robust increase in S phase-cells two days later (E16.5). The effects observed in the ERK2 CKO may reflect specific actions of this isoform or alternatively might result from a net reduction in overall ERK activity. Of particular relevance, Brambilla and colleagues have argued that ERK1 acts to functionally antagonize the actions of ERK2 in stimulating cellular proliferation, thus slowing progression of the cell cycle (Vantaggiato et al., 2006). However, Lenormand and colleagues have shown that each isoform is indiscriminately activated, and contend that the overall abundance of ERK1 and ERK2 dictates biological outcomes including cellular proliferation. Interestingly, the same report demonstrates that ERK2 is 13 times more abundant than ERK1 in the superficial cortex of nude mice (Lefloch et al., 2008). Our findings are consistent with the hypothesis that net ERK1/2 activity governs cellular behaviors. The striking increase in the levels of activated forms of ERK1 in the cortex of our ERK2 CKO mice at E16.5, following the near complete loss of ERK2 protein, suggests that the increase in ERK1

93 activity might reflect a compensatory response to the loss of ERK2, and this is correlated

with a resumption of NPC proliferation. An analogous compensatory effect on ERK2

activation was reported in a number of tissues in the ERK1 knockout mice, including the

brain (Pages et al., 1999; Selcher et al., 2001; Mazzucchelli et al., 2002). These findings

support the possibility that the biological actions of the ERKs are dictated by their spatial

and temporal expression, rather than the specific activity of one isoform or the other.

However, there is compelling evidence that isoform-specific actions of the ERKs may

arise from their selective association with molecular scaffolds that direct their interaction

with protein substrates and with phosphatases (Pouyssegur et al., 2002). These issues are

under active investigation.

The change in cellular generation by NPCs upon mapk1/ERK2 inactivation confirms and expands the findings from experimental manipulation of other components of the MAPK signaling cascade. Inhibition of the upstream activator of the ERKs, MEK1, causes

NPCs to remain in the SVZ/VZ in an undifferentiated state, blocking neurogenesis

(Menard et al., 2002; Barnabe-Heider and Miller, 2003; Paquin et al., 2005). Similarly, mutation or inactivation of scaffolding proteins which link receptor tyrosine kinases to

ERK activation have analogous effects on cortical progenitors. Mutations in the ERK cascade scaffolding protein FRS2 result in animals that exhibit nearly 30% reductions in cortical thickness and fewer cortical neurons (Hadari et al., 1998; Yamamoto et al.,

2005). This study confirmed that loss of cortical neurons was due to reduced numbers of proliferating IPCs and was a consequence of dramatically reduced levels of ERK activation. Two independent studies have recently found that deletion or mutation of

94 SHP-2, an upstream element of the ERK cascade, in cortical progenitors leads to the

inhibition of neuronal differentiation through ERK-dependent mechanisms (Gauthier et al., 2007; Ke et al., 2007). It has recently been proposed that IPCs are the only progenitors to undergo symmetric neurogenic divisions, contributing neurons to all cortical lamina throughout neurogenesis (Pontious et al., 2008). Our findings which document a decrease in neurons occupying each cortical lamina and a specific reduction in mitotic IPCs indicate that ERK2 may play a critical role in regulation of IPC division.

These data warrant further investigation into the precise roles that ERK2 plays in IPC cell cycle progression and terminal differentiation.

The conditional knockout of ERK2 also resulted in more astrocytes populating the neonatal and mature cortex. Bonni and colleagues originally demonstrated that ERK activation blocked the differentiation of NPCs into astrocytes (Bonni et al., 1997) and this conclusion has been validated both in vitro (Menard et al., 2002) and in animal models

(Paquin et al., 2005). Our findings are therefore consistent with the postulated action of the ERK cascade in suppressing glial fate and promoting neurogenesis in the developing cortex (Menard et al., 2002; Barnabe-Heider and Miller, 2003; Paquin et al., 2005;

Gauthier et al., 2007).

The ERKs are known to play critical roles in learning and memory and the absence of a robust behavioral phenotype in ERK1 null mice has suggested that ERK2 is the principal isoform responsible for learning behaviors which are dependent on the neocortex and associated structures (Selcher et al., 2001; Mazzucchelli et al., 2002). Fear conditioning

95 requires ERK-dependent interactions within and between the hippocampus, cortex and

amygdala. We demonstrate that the loss of ERK2 within the telencephalon resulted in

significant impairment on cued and contextual fear conditioning tests. Our findings are

similar to effects of pharmacological inhibition of the ERKs (Sweatt, 2004) or expression

of dominant negative MEK on fear conditioning (Shalin et al., 2004). Thus, the present

findings provide direct evidence that ERK2 is the principal isoform required for learning

in this behavioral task; however, it is unclear if the deficit is due to developmentally-

related structural changes of the cortex or the loss of ERK2 in mature neurons. It is

important to note that glutamatergic or electrical stimulation of neurons selectively

stimulates ERK2 activation (English and Sweatt, 1996; Mazzucchelli et al., 2002). Our

results compliment a recent report by Satoh and colleagues describing the behavior of an

ERK2 hypomorphic mouse that exhibits deficits in long-term memory in contextual and

cued fear conditioning paradigms (Satoh et al., 2007).

We have identified individuals with deletion of a single MAPK1/ERK2 allele and exhibit

decreased ERK2 protein levels. These patients’ clinical presentation include congenital

anomalies and neurocognitive deficits reflected by developmental delays, microcephaly,

and learning disabilities (Shaikh et al., 2007). Our findings in mutant mice suggest that

these abnormalities may arise from perturbation of neural developmental as well as

impaired ERK2 signaling required for memory and cognition. We postulate that the

cognitive deficits observed in the patients and in the mice result from a perturbation in

signal transduction through the ERK signaling cascade. We further suggest that these findings may have broader significance given the recent recognition that genetic

96 mutations within either upstream elements (Bentires-Alj et al., 2006) or downstream

targets (Weeber and Sweatt, 2002) of the ERK cascade are associated with developmental delay or mental retardation.

97 Figure 2-1. Generation of ERK2 Conditional Knockout mice. A, Schematic representation of gene targeting strategy. The shaded boxes on the mapk1/ERK2 locus and the targeting construct designate homologous regions identified for recombination.

Numbered boxes represent mapk1 Exons 2 through 4. The targeting construct contains a floxed Neo cassette upstream of Exon 2 and a loxP site between Exon 2 and 3. A βADT cassette was inserted downstream of Exon 3 for negative selection. Relevant restriction sites and the 3’ probe used for Southern blot analysis are indicated. Recombinant ES cells were transfected with a Cre plasmid to remove the Neo cassette, giving rise to the floxed mapk1 allele. Cre recombinase expression from the hGFAP-cre transgene produces a null allele. B, Southern blot of genomic brain DNA. Following Xba digestion of DNA, the 3’ probe identifies the wild-type allele as a 4.7kb fragment and the mutant allele as 11.5kb before and 9.5kb after recombination. C, Western blot analysis of adult cortical brain homogenates. Tissue was homogenized, separated by SDS-PAGE and immunoblotted with anti-ERK1, anti-ERK2 and β-tubulin antibodies. D, Western blot analysis of embryonic cortical homogenates. Tissue was lysed, separated by SDS-

PAGE and immunoblotted with an antibody directed to an epitope at the N-terminus of

ERK2. A lack of bands below 42 kD in both flox/flox and flox/flox cre+ samples demonstrates that no truncated protein fragments were produced by recombination. An antibody for c-src was utilized as a loading control.

98 Figure 2-1

99 Figure 2-2. ERK2 CKO mice display reduced cortical thickness. A, Dorsal view of P2

CKO (bottom) and wild-type (top) littermates. B, Quantitative analysis of cortical thickness. Corresponding coronal sections from CKO and wild-type brains were used to measure thickness of the frontal, parietal, temporal and occipital lobes corresponding to the motor, somatosensory, auditory and visual cortices. Measurements were taken from the ventricular zone to the pial surface (n=5), Two-way ANOVA p<0.0001 with

Bonferroni post-tests,***p<0.001, **p<0.01. C, Representative coronal sections of CKO and wild-type brains from P2, stained with DAPI. Note the difference in cellular density apparent between genotypes. D, Coronal sections of adult ERK1 -/- and wild type brains stained with the fluorescent Nissl stain, Neurotrace.

100

Figure 2-2

101 Figure 2-3. Loss of ERK2 expression and activity in ERK2 CKO cortices. A-D, E14.5

and E16.5 cortices were microdissected and prepared for Western blot analysis immediately upon dissection. The levels of ERK2, ERK1 and phospho-ERK1/2 were detected by immunoblotting and representative blots are shown. Densitometric analysis of the bands, normalized to β-tubulin, was performed for quantification and expressed as fold difference relative to wild-type. For A, quantification of ERK2 levels at E14.5

(CKO n=15, WT n=21, p<0.0001) and E16.5, (CKO n=7, WT n=8, p<0.0001). B,

Quantification of ERK1 expression levels at E14.5, (CKO n=13, WT n=18, p=0.4019) and E16.5, (CKO n=7, WT n=8, p=0.7772). C, Relative phospho-ERK2 activation at

E14.5, (CKO n=12, WT n=18, p=0.0093) and E16.5, (CKO n=7, WT n=8, p<0.0001), and D, relative phospho-ERK1 activation at E14.5, (CKO n=11, WT n=17, p=0.3594) and E16.5, (CKO n=4, WT n=4, p<0.0001).

102 Figure 2-3

103 Figure 2-4. Inactivation of ERK2 in neural progenitor cells results in the generation of

fewer cortical neurons. A, Analysis of cell density and cell fate in CKO and wild-type

littermates at P10. Corresponding coronal sections of P10 cortices were immunostained

with NeuN (green) and counterstained with DAPI (blue). Total cell number per cortical

region from the VZ to the pial surface was calculated by counting DAPI+ cells in both

medial (primary motor cortex, CKO=1386.167, WT=1398.167 avg cells per region, for

each n=3) and lateral (primary somatosensory cortex, CKO=1534.833, WT= 1571.167 avg cells per region, for each n=3) sections. Two-way ANOVA p=0.6179 and Bonferroni post-test, p>0.05. The fold difference in NeuN+ cells was calculated by Two-way

ANOVA p<0.0001, Bonferroni post-test p<0.0001. Fold difference in DAPI;NeuN- cells was calculated, Two-way ANOVA p=.0004, Bonferroni post-test p<0.01. B-D,

Corresponding coronal sections of P2 CKO (left) and WT (right) littermates were stained with anti-Tbr1 (B, green) anti-Otx1 (C, red) and anti-Brn1 (D, red) antibodies and counterstained with DAPI. Cortical lamina were identified based on morphology and distance from the pial surface. Immunoreactive cells were counted from 2 independent reference spaces of the cortex from at least 8 sections per genotype. For each, Two-way

ANOVA p<0.0001 and Bonferroni post-tests p<0.0001 were performed (n=4).

104 Figure 2-4

105 Figure 2-5. Inactivation of ERK2 in neural progenitor cells results in the presence of

more astrocytes within the cerebral cortex. A-B, Representative images of corresponding

coronal sections from neonatal (A) and mature (B) ERK2 CKO and wild-type littermates

immunostained with anti-Zebrin II (red) anti-GFAP (green) antibodies and counterstained

with DAPI. A higher magnification image is shown on the right. C, Primary astrocyte cultures from CKO and wild-type P2 pups were prepared and maintained for 5 DIV.

Cultures were incubated with BrdU overnight and fixed. Quantification of percent BrdU immunoreactivity (CKO n=4, WT n=3, p=0.9084). D, Quantification of Ki67 immunoreactive cells (CKO n=3, WT n=3, p=0.4166). E, Characterization of astrocyte proliferative index (CKO n=2, WT n=2, p=0.9827). F, MTT viability assay (CKO n=4,

WT n=5, p=0.4556).

106 Figure 2-5

107 Figure 2-6. ERK2 CKO mice display changes in the dynamics of NPC proliferation. A-

C, Pregnant mice were given intraperitoneal injections of BrdU at E12.5, E14.5 and

E16.5, respectively. Embryos were fixed 30 minutes after BrdU injection and

corresponding coronal sections of ERK2 CKO and wild-type littermates were double

labeled with anti-phospho-histone H3 (pH3+, red) and anti-BrdU (green) antibodies

followed by DAPI counterstaining (blue). Confocal images were taken at 10x and 20x

(A-B, scale bar=50 µm) and 10x and 40x (C, scale bar= 20 µm). D-E, Corresponding coronal sections of E14.5 (D) and E16.5 (E) embryos were immunostained with Tbr2 and phosphorylated Histone H3 followed by counterstaining with DAPI. Scale bar= 20 µm.

108 Figure 2-6

109 Figure 2-7. ERK2 CKO cortical progenitor cells exhibit reduced neuronal generation.

A, Cortical progenitor cell cultures from ERK2 CKO and wild-type E14.5 embryos were grown in Neurobasal with FGF2 for 2 days in vitro. Cells were then fixed and immunostained. Progenitors were identified by Nestin immunoreactivity (A, n=3).

Cultures were assayed for immature neurons with βIII Tubulin (B, n=3), mature neurons with MAP2 (C, n=4), and astrocyte precursors with S-100β (D, n=4). E, Analysis of cellular identity in culture with FGF2. Fold difference is the comparison of average of positive cells/total cell number relative to wild-type. Student’s t-test, p<0.001 for Nestin, p=0.014 for βIII-tubulin, p=0.0003 for MAP2 and p=0.0018 for S100β.

110 Figure 2-7

111 Figure 2-8. ERK2 CKO cortical progenitors generate more astrocytes in the presence of gliogenic stimuli. Cortical progenitor cell cultures from ERK2 CKO and wild-type E14.5 embryos were grown in Neurobasal media. Twelve hours after plating, cells were treated with CNTF to induce astrocyte differentiation. Cells were fixed and immunostained with anti-vimentin (A) anti-GFAP (B) and anti-S100β (C) antibodies after 2 days in vitro (1-2) and 4 days in vitro (3-4). D-F, Analysis of cellular identity with CNTF treatment. Fold difference in astrocyte generation was measured by the average of marker immunopositive cells/total cell number relative to wild-type At 2DIV and 4 DIV, (n=4 for each), Two-way ANOVA, p<0.0001 with Bonferroni post-tests, p<0.0001 for each.

112 Figure 2-8

113 Figure 2-9. Male ERK2 CKO mice have deficits in associative learning. Male ERK2

CKO mice displayed deficits in cued and contextual fear conditioning. A, 24-hr post- training, CKOs showed significantly less freezing behavior when tested for contextual and cued recall, t16 = 2.27 and 2.95, respectively, p<0.05. B, Repeated training, or

“overtraining,” did not overcome the deficit seen in cued and contextual fear conditioning. CKOs showed significantly less freezing behavior in both recall tests when tested the next day, t16 = 3.65 and 3.35, respectively, p<0.05.

114 Figure 2-9

115 Figure 2-10. ERK2 expression is abolished in neural progenitor cells beginning at

E13.5. A-B, Expression of hGFAP-cre activity. Sagittal sections of adult hGFAP- cre;R26R mice were stained for X-gal and counterstained with neutral red. High magnification inset in B demonstrates the extent of recombination throughout the cortex.

C, ERK2 expression in ERK2 CKO and wild-type mice at P2. Coronal sections of ERK2

CKO (left) and wild type littermates (right) were immunostained with an ERK2 specific antibody and visualized using DAB. D-I, Coronal sections of P2 ERK2 CKO and wild- type mice immunostained with an ERK2 specific antibody (red) and colabeled with antibodies directed to (D-E) Tbr1, the preplate/layer VIb marker (green), (F-G)

GABAergic interneuron marker, GABA (green) and (H-I) Reelin (green) which labels

Cajal-Retzius cells. The merged images and inset enlargements are shown in the far left panels of each.

116 Figure 2-10

117 Figure 2-11. ERK2 expression is reduced in the hippocampus, but not the amygdala in

ERK2 CKO mice. A-B, ERK2 expression was assessed by immunostaining with an anti-

ERK2 specific antibody. Corresponding coronal sections of adult ERK2 CKO (A) and

littermate wild-types (B) were evaluated for ERK2 expression in the hippocampus and

amygdala. Note that ERK2 is most highly expressed in the dentate gyrus (DG) and CA3

regions of the hippocampus and was lost in these areas in the mutant mice. No change in

ERK2 expression was identified within the basolateral amygdala (B/LA).

118 Figure 2-11

119 Figure 2-12. The ERK2 CKO cortex does not exhibit apoptosis. Analysis of apoptosis

was performed by immunostaining with anti-Caspase 3 antibody (A, DAB) and TUNEL

staining (B, green). Corresponding coronal sections from E17 ERK2 CKO (left) and wildtype (right) brains were stained with anti-Caspase 3 antibodies (note black arrow) and by TUNEL staining (white arrow).

120 Figure 2-12

121 Figure 2-13. The number of Pax6 and Tbr2 immunoreactive cells at E14.5 is not altered by conditional inactivation of ERK2. Cellular analysis was performed by immunohistochemistry of E14.5 cryosections with anti-Pax6 (red) and anti-Tbr2 (green) antibodies and counting of immunoreactive cells for each marker and for those cells expressing both markers. Two-way ANOVA, p>0.05 for each.

122 Figure 2-13

123 Figure 2-14. Patients with deletions of distal 22q11 have reduced ERK2 levels. A,

Schematic representation of Chromosome 22q11 with the centromere (CEN) to the left and telomeric (TEL) end to the right. Low copy repeats, LCRs are denoted by black boxes. The gene encoding CRKL is located between LCR-C and LCR-D and is included in the common 3Mb DiGeorge Syndrome deletion spanning from LCR-A to LCR-D.

The MAPK1 gene is located between LCR-D and LCR-E and is included in a 1 Mb distal

22q11 deletion occurring between these two LCRs. The D-E deletion interval is expanded to illustrate the other genes located within this 1Mb region. B, Lymphoblasts from normal subjects, distal 22q11 deletion patients, and DiGeorge Syndrome patients were lysed and prepared for Western blot analysis. The levels of ERK1, ERK2, and

CRKL were detected by immunoblotting and representative blots are shown. G3PDH levels were assayed for loading controls. C, Quantification of ERK2 levels from normal,

D-E deletion, and DiGeorge Syndrome patient lymphoblasts (normal, n=4, D-E, n=2,

DiGeorge, n=1, p=0.0026).

124 Figure 2-14

125 Table 2-1 and 2-2. Inactivation of ERK2 does not alter the morphology of cortical neurons.

126

Table 1- Parameters of basal dendrites of layer III pyramidal neurons

Order Quantity Length (µm) Surface area (µm2) Volume (µm3) ______CKO WT CKO WT CKO WT CKO WT ______1 47.67+ 1.76 60.67+ 5.69 122.72+ 9.80 156.52+ 2.62 910.44+ 101.61 1141.69+ 25.23 568.08+ 92.88 762.00+ 29.85 2 47.33+ 4.67 59.67+ 3.48 148.30+ 6.51 153.82+ 4.44 1085.68+ 75.67 901.99+ 48.78 660.29+ 71.69 515.32+ 32.39 3 38.00+ 1.15 25.67+ 0.33 * 158.15+ 3.65 135.16+ 6.58 1021.18+ 42.84 781.59+ 30.18 * 528.76+ 18.40 402.61+ 22.79 * ______

Mean values + SEM. * P< 0.05.

Table 2- Parameters of basal dendrites of Layer V pyramidal neurons

Order Quantity Length (µm) Surface area (µm2) Volume (µm3) ______CKO WT CKO WT CKO WT CKO WT ______1 42.67+ 5.90 61.67+ 1.76 132.53+ 24.96 156.30+ 26.47 1238.69+ 272.60 1151.01+ 279.62 994.20+ 228.48 728.64+ 199.65 2 41.67+ 5.49 42.33+ 3.93 142.68+ 32.83 169.34+ 31.41 1246.00+ 339.24 1323.47+ 310.27 922.62+ 260.06 882.67+ 219.33 3 14.00+ 1.15 18.33+ 1.20 149.35+ 22.53 181.39+ 37.52 1516.03+ 96.15 1409.52+ 328.76 978.03+ 161.41 905.58+ 219.87 ______

Mean values + SEM. * P< 0.05.

127 Literature Cited

Aouadi M, Binetruy B, Caron L, Le Marchand-Brustel Y, Bost F (2006) Role of MAPKs in development and differentiation: lessons from knockout mice. Biochimie 88:1091-1098. Barnabe-Heider F, Miller FD (2003) Endogenously produced neurotrophins regulate survival and differentiation of cortical progenitors via distinct signaling pathways. J Neurosci 23:5149-5160. Ben-Shachar S, Ou Z, Shaw CA, Belmont JW, Patel MS, Hummel M, Amato S, Tartaglia N, Berg J, Sutton VR, Lalani SR, Chinault AC, Cheung SW, Lupski JR, Patel A (2008) 22q11.2 distal deletion: a recurrent genomic disorder distinct from DiGeorge syndrome and velocardiofacial syndrome. Am J Hum Genet 82:214- 221. Bentires-Alj M, Kontaridis MI, Neel BG (2006) Stops along the RAS pathway in human genetic disease. Nat Med 12:283-285. Bonni A, Sun Y, Nadal-Vicens M, Bhatt A, Frank DA, Rozovsky I, Stahl N, Yancopoulos GD, Greenberg ME (1997) Regulation of gliogenesis in the central nervous system by the JAK-STAT signaling pathway. Science 278:477-483. Chambard JC, Lefloch R, Pouyssegur J, Lenormand P (2006) ERK implication in cell cycle regulation. Biochim Biophys Acta. Crawley JN (1999) Behavioral phenotyping of transgenic and knockout mice: experimental design and evaluation of general health, sensory functions, motor abilities, and specific behavioral tests. Brain Res 835:18-26. Dehay C, Kennedy H (2007) Cell-cycle control and cortical development. Nat Rev Neurosci 8:438-450. Dono R (2003) Fibroblast growth factors as regulators of central nervous system development and function. Am J Physiol Regul Integr Comp Physiol 284:R867- 881. Emanuel BS, McDonald-McGinn D, Saitta SC, Zackai EH (2001) The 22q11.2 deletion syndrome. Adv Pediatr 48:39-73. English JD, Sweatt JD (1996) Activation of p42 mitogen-activated protein kinase in hippocampal long term potentiation. J Biol Chem 271:24329-24332. Gauthier AS, Furstoss O, Araki T, Chan R, Neel BG, Kaplan DR, Miller FD (2007) Control of CNS cell-fate decisions by SHP-2 and its dysregulation in Noonan syndrome. Neuron 54:245-262. Giulian D, Baker TJ (1986) Characterization of ameboid microglia isolated from developing mammalian brain. J Neurosci 6:2163-2178. Hadari YR, Kouhara H, Lax I, Schlessinger J (1998) Binding of Shp2 tyrosine phosphatase to FRS2 is essential for fibroblast growth factor-induced PC12 cell differentiation. Mol Cell Biol 18:3966-3973. Hevner RF, Hodge RD, Daza RA, Englund C (2006) Transcription factors in glutamatergic neurogenesis: conserved programs in neocortex, cerebellum, and adult hippocampus. Neurosci Res 55:223-233. Ke Y, Zhang EE, Hagihara K, Wu D, Pang Y, Klein R, Curran T, Ranscht B, Feng GS (2007) Deletion of Shp2 in the brain leads to defective proliferation and

128 differentiation in neural stem cells and early postnatal lethality. Mol Cell Biol 27:6706-6717. Lefloch R, Pouyssegur J, Lenormand P (2008) Single and combined silencing of ERK1 and ERK2 reveals their positive contribution to growth signaling depending on their expression levels. Mol Cell Biol 28:511-527. Liu X, Yan S, Zhou T, Terada Y, Erikson RL (2004) The MAP kinase pathway is required for entry into mitosis and cell survival. Oncogene 23:763-776. Malatesta P, Hack MA, Hartfuss E, Kettenmann H, Klinkert W, Kirchhoff F, Gotz M (2003) Neuronal or glial progeny: regional differences in radial glia fate. Neuron 37:751-764. Mazzucchelli C, Vantaggiato C, Ciamei A, Fasano S, Pakhotin P, Krezel W, Welzl H, Wolfer DP, Pages G, Valverde O, Marowsky A, Porrazzo A, Orban PC, Maldonado R, Ehrengruber MU, Cestari V, Lipp HP, Chapman PF, Pouyssegur J, Brambilla R (2002) Knockout of ERK1 MAP kinase enhances synaptic plasticity in the striatum and facilitates striatal-mediated learning and memory. Neuron 34:807-820. Meloche S, Pouyssegur J (2007) The ERK1/2 mitogen-activated protein kinase pathway as a master regulator of the G1- to S-phase transition. Oncogene 26:3227-3239. Menard C, Hein P, Paquin A, Savelson A, Yang XM, Lederfein D, Barnabe-Heider F, Mir AA, Sterneck E, Peterson AC, Johnson PF, Vinson C, Miller FD (2002) An essential role for a MEK-C/EBP pathway during growth factor-regulated cortical neurogenesis. Neuron 36:597-610. Miller FD, Gauthier AS (2007) Timing is everything: making neurons versus glia in the developing cortex. Neuron 54:357-369. Noctor SC, Martinez-Cerdeno V, Ivic L, Kriegstein AR (2004) Cortical neurons arise in symmetric and asymmetric division zones and migrate through specific phases. Nat Neurosci 7:136-144. Ohkubo Y, Uchida AO, Shin D, Partanen J, Vaccarino FM (2004) Fibroblast growth factor receptor 1 is required for the proliferation of hippocampal progenitor cells and for hippocampal growth in mouse. J Neurosci 24:6057-6069. Pages G, Guerin S, Grall D, Bonino F, Smith A, Anjuere F, Auberger P, Pouyssegur J (1999) Defective thymocyte maturation in p44 MAP kinase (Erk 1) knockout mice. Science 286:1374-1377. Paquin A, Barnabe-Heider F, Kageyama R, Miller FD (2005) CCAAT/enhancer-binding protein phosphorylation biases cortical precursors to generate neurons rather than astrocytes in vivo. J Neurosci 25:10747-10758. Pontious A, Kowalczyk T, Englund C, Hevner RF (2008) Role of intermediate progenitor cells in cerebral cortex development. Dev Neurosci 30:24-32. Pouyssegur J, Volmat V, Lenormand P (2002) Fidelity and spatio-temporal control in MAP kinase (ERKs) signalling. Biochem Pharmacol 64:755-763. Rauch A, Pfeiffer RA, Leipold G, Singer H, Tigges M, Hofbeck M (1999) A novel 22q11.2 microdeletion in DiGeorge syndrome. Am J Hum Genet 64:659-666. Roberts A, Allanson J, Jadico SK, Kavamura MI, Noonan J, Opitz JM, Young T, Neri G (2006) The cardiofaciocutaneous syndrome. J Med Genet 43:833-842.

129 Robinson MJ, Harkins PC, Zhang J, Baer R, Haycock JW, Cobb MH, Goldsmith EJ (1996) Mutation of position 52 in ERK2 creates a nonproductive binding mode for adenosine 5'-triphosphate. Biochemistry 35:5641-5646. Saitta SC, McGrath JM, Mensch H, Shaikh TH, Zackai EH, Emanuel BS (1999) A 22q11.2 deletion that excludes UFD1L and CDC45L in a patient with conotruncal and craniofacial defects. Am J Hum Genet 65:562-566. Satoh Y, Endo S, Ikeda T, Yamada K, Ito M, Kuroki M, Hiramoto T, Imamura O, Kobayashi Y, Watanabe Y, Itohara S, Takishima K (2007) Extracellular signal- regulated kinase 2 (ERK2) knockdown mice show deficits in long-term memory; ERK2 has a specific function in learning and memory. J Neurosci 27:10765- 10776. Selcher JC, Nekrasova T, Paylor R, Landreth GE, Sweatt JD (2001) Mice lacking the ERK1 isoform of MAP kinase are unimpaired in emotional learning. Learn Mem 8:11-19. Shaikh TH, O'Connor RJ, Pierpont ME, McGrath J, Hacker AM, Nimmakayalu M, Geiger E, Emanuel BS, Saitta SC (2007) Low copy repeats mediate distal chromosome 22q11.2 deletions: sequence analysis predicts breakpoint mechanisms. Genome Res 17:482-491. Shalin SC, Zirrgiebel U, Honsa KJ, Julien JP, Miller FD, Kaplan DR, Sweatt JD (2004) Neuronal MEK is important for normal fear conditioning in mice. J Neurosci Res 75:760-770. Shen Q, Wang Y, Dimos JT, Fasano CA, Phoenix TN, Lemischka IR, Ivanova NB, Stifani S, Morrisey EE, Temple S (2006) The timing of cortical neurogenesis is encoded within lineages of individual progenitor cells. Nat Neurosci 9:743-751. Soriano P (1999) Generalized lacZ expression with the ROSA26 Cre reporter strain. Nat Genet 21:70-71. Sweatt JD (2001) The neuronal MAP kinase cascade: a biochemical signal integration system subserving synaptic plasticity and memory. J Neurochem 76:1-10. Sweatt JD (2004) Mitogen-activated protein kinases in synaptic plasticity and memory. Curr Opin Neurobiol 14:311-317. Thomson RE, Pellicano F, Iwata T (2007) Fibroblast growth factor receptor 3 kinase domain mutation increases cortical progenitor proliferation via mitogen-activated protein kinase activation. J Neurochem 100:1565-1578. Vaccarino FM, Schwartz ML, Raballo R, Nilsen J, Rhee J, Zhou M, Doetschman T, Coffin JD, Wyland JJ, Hung YT (1999) Changes in cerebral cortex size are governed by fibroblast growth factor during embryogenesis. Nat Neurosci 2:246- 253. Vantaggiato C, Formentini I, Bondanza A, Bonini C, Naldini L, Brambilla R (2006) ERK1 and ERK2 mitogen-activated protein kinases affect Ras-dependent cell signaling differentially. J Biol 5:14. Walz R, Roesler R, Quevedo J, Sant'Anna MK, Madruga M, Rodrigues C, Gottfried C, Medina JH, Izquierdo I (2000) Time-dependent impairment of inhibitory avoidance retention in rats by posttraining infusion of a mitogen-activated protein kinase kinase inhibitor into cortical and limbic structures. Neurobiol Learn Mem 73:11-20.

130 Walz R, Roesler R, Barros DM, de Souza MM, Rodrigues C, Sant'Anna MK, Quevedo J, Choi HK, Neto WP, DeDavid e Silva TL, Medina JH, Izquierdo I (1999) Effects of post-training infusions of a mitogen-activated protein kinase kinase inhibitor into the hippocampus or entorhinal cortex on short- and long-term retention of inhibitory avoidance. Behav Pharmacol 10:723-730. Weeber EJ, Sweatt JD (2002) Molecular neurobiology of human cognition. Neuron 33:845-848. Yamamoto S, Yoshino I, Shimazaki T, Murohashi M, Hevner RF, Lax I, Okano H, Shibuya M, Schlessinger J, Gotoh N (2005) Essential role of Shp2-binding sites on FRS2alpha for corticogenesis and for FGF2-dependent proliferation of neural progenitor cells. Proc Natl Acad Sci U S A 102:15983-15988. Zheng W, Nowakowski RS, Vaccarino FM (2004) Fibroblast growth factor 2 is required for maintaining the neural stem cell pool in the mouse brain subventricular zone. Dev Neurosci 26:181-196. Zhuo L, Theis M, Alvarez-Maya I, Brenner M, Willecke K, Messing A (2001) hGFAP- cre transgenic mice for manipulation of glial and neuronal function in vivo. Genesis 31:85-94.

131 ERK1 AND ERK2 ARE REQUIRED FOR CELL CYCLE REGULATION OF

NEURAL PROGENITOR CELLS

Ivy S. Samuels1, Joanna Pucilowska1, J. Colleen Karlo1, Natalie A. Cherosky1, Michiko

Watanabe2 and Gary E. Landreth1

1Department of Neurosciences

Case Western Reserve University

Cleveland, OH 44106

2Department of Pediatrics

Case Western Reserve University

Cleveland, OH 44106

Abstract

The canonical MAP Kinases, ERK1 and ERK2, are essential for fundamental cellular

processes such as proliferation and differentiation, as well as actions specific to the

nervous system such as long term potentiation and synaptic plasticity. We have

previously shown that the inactivation of ERK1 does not result in severe phenotypic or

behavioral deficits in mice, suggesting a redundant role for ERK2. Conditional

inactivation of the ERK2 isoform in murine neural progenitor cells causes significant

alterations in cell fate determination within the telencephalon and profound deficits in

associative learning tasks. The latter finding predicts that ERK2 plays a pivotal role in

cortical development and cognition which is not likely to be fully compensated for by

132 ERK1. To investigate the combined contribution of ERK1 and ERK2 to nervous system

development and function, we generated mice in which the genes encoding both ERK1

and ERK2 were inactivated in the telencephalon. ERK1/2 double knockout (DKO) mice

display a more severe phenotype than single knockout of either isoform and die during embryogenesis. ERK1/2 DKO mice exhibit changes in telencephalic cell fate determination similar to those of the ERK2 conditional knockout. We find that fewer

MAP2 neurons are generated and a greater number of neural progenitor cells remain undifferentiated during cortical development. Furthermore, we find a delay in S-phase entry as measured by acute BrdU incorporation during proliferative phases of

neurogenesis. These changes are associated with elevated levels of p27 and reductions in

cyclin D2 levels. Our results demonstrate that despite their role in cell cycle control, the

ERKs are not required for generation of the murine cerebral cortex.

Introduction

The Extracellular signal Regulated protein Kinases, ERK1 (mapk3) and ERK2 (mapk1)

are highly homologous protein kinases that play important roles in cellular proliferation

and differentiation. As the classical Mitogen Activated Protein Kinases (MAPKs), ERK1

and ERK2 are the central components of a canonical three-tier protein kinase cascade in

which extracellular signals emanating from a variety of cell surface receptors at the

plasma membrane stimulate cellular proliferation, differentiation, and survival

(Chambard et al., 2006).

133 The murine cerebral cortex is generated over a 6 day interval in which neural progenitor

cells undergo 11 cell cycles to produce 95% of neurons populating the adult neocortex.

Therefore, intricate regulation of cell cycle progression and termination are essential for

proper cortical development (Nowakowski et al., 2002). ERK1/2 activity has been

identified in the forebrain as early as E8.5 (Corson et al., 2003) and their actions within

neural progenitor cells have been shown to be essential for the proper generation of

pyramidal projection neurons as well as astrocytes that populate the cerebral cortex

(Menard et al., 2002; Paquin et al., 2005; Gauthier et al., 2007). Miller and colleagues have further demonstrated that ERK1/2 is not required for survival of NPCs, and proliferation of cultured NPCs in vitro is not strictly dependent on these kinases either

(Barnabe-Heider and Miller, 2003). Instead, neurotrophin signaling through MEK-

ERK1/2 is required for neuronal differentiation.

A differential role for ERK1 and ERK2 is implied from the individual knockout of each isoform. ERK2 activity in the extra-embryonic tissues at E6.5 is critical for survival as

ERK2 null mice die at this time due to the requirement for ERK2 in mesoderm and trophoblast formation as well as placental development (Hatano et al., 2003; Saba-El-Leil

et al., 2003; Yao et al., 2003). ERK1 knockouts do not exhibit phenotypic defects (Pages

et al., 1999; Selcher et al., 2001; Mazzucchelli et al., 2002). Conditional inactivation of

the ERK2 isoform in the mouse forebrain led to reductions in cortical thickness

associated with the generation of fewer neurons and more astrocytes. The dynamics of

cell cycle progression and exit were altered such that fewer proliferating neural

progenitor cells were identified at the peak of murine neurogenesis, E14.5. These results

134 indicated that ERK2 plays important roles in the control of neurogenic progenitor cell

divisions. However, we observed a dramatic increase in activated, phosphorylated ERK1

at E16.5 in the ERK2 mutant, suggesting that this isoform may have compensated for

ERK2 inactivation and alleviated a more severe developmental defect (Chapter 2).

Insight into the individual roles of each isoform in cell cycle regulation comes primarily

from in vitro studies. Within synchronized HeLa cells, downregulation of ERK1 induced

cell cycle arrest at G2, while loss of ERK2 led to cell cycle arrest and apoptosis in G1

(Liu et al., 2004). Furthermore, in NIH3T3 cells, ERK1 was postulated to suppress

ERK2 activity as the loss of ERK1 induced faster cell proliferation and facilitated cell

growth while down-regulation of ERK2 severely reduced cell proliferation (Vantaggiato et al., 2006). ERK1 null hepatocytes act essentially as wildtype cells; however, in double

null hepatocytes, cell proliferation is inhibited (Fremin et al., 2007).

These findings suggest that ERK2 plays key roles in the early stages of cell proliferation

which are not carried out by ERK1; however, Lenormand and colleagues have argued

that the expression of ERK2 and ERK1 dictates the specificity of isoform activation, as

opposed to individual activation and function of each (Lefloch et al., 2008). Their

analysis of isoform expression suggests that functions previously thought to be ERK isoform-specific may result from the lack of expression of one isoform or the other, not from differential activation.

135 In order to determine if combined inactivation of both ERK isoforms more severely

disrupts development of the cortex than loss of ERK2 alone, we generated a murine

model in which mapk1/ERK2 undergoes conditional inactivation at the peak of neurogenesis in a mapk3/ERK1 null background. We then analyzed cellular composition of the developing cortex of these animals in vivo and cell fate determination by neural progenitors in vitro. We further assessed neural progenitor cell proliferation throughout

neurogenesis and ERK1/2 dependent regulators of the cell cycle.

Materials and Methods

Generation of Mapk3-/-;Mapk1flox/flox;hGFAPcre/+ Mice. Mapk1flox/flox mice (Chapter 2)

were maintained on a mixed 129/B6 background by interbreeding. Congenic B6 mapk3-

/- mice (Nekrasova et al., 2005) were mated with mapk1flox/flox mice to generate doubly

hetereozgous mapk3+/-; mapk1flox/+ mice. These animals were intercrossed to generate

mapk3-/-;mapk1flox/flox mice. This line was maintained and DKO mice were generated by

first mating FVB/N hGFAPcre/+ mice with the mapk3-/-;mapk1flox/flox line. The progeny of

these matings (mapk3+/-; mapk1flox/+ ;hGFAPcre/+ ) were backcrossed to the mapk3-/-;

mapk1flox/flox line to garner litters carrying wildtype (mapk3+/-; mapk1flox/flox or flox/+),

ERK1 null (mapk3-/-;mapk1flox/fox or flox/+) and DKO (mapk3-/-; mapk1flox/flox; hGFAPcre/+) embryos. Primers for mapk3/ERK1 (5’-3’)

ExVI: CCAGGAGGACCTTAATTGCATCATT

3’ERK1: TTAGGGGCCCTCTGGCGCGGGTGGCTG.

Primers for mapk1/ERK2 (5’-3’) E2F-U: AGCCAACAATCCCAAACCTG, E2KOL:

CAACACCACCCCACTTACTA

136

Cortical Progenitor Cell Culture. Cortical progenitor cells were cultured from embryonic day 14.5 (E14.5) mouse embryos using protocols previously described by

Miller and colleagues (Menard et al., 2002; Barnabe-Heider and Miller, 2003). Briefly,

cortices were dissected from E14.5 mouse embryos in ice-cold HBSS and transferred to

Neurobasal medium containing 500 µM L-glutamine, 2% B27 supplement, 1% penicillin-

streptomycin (Invitrogen, Gaithersburg, MD) and 40 ng/ml FGF2 (BD, Bedford, MA).

The tissue was mechanically triturated with a plastic pipette and plated into eight-well

chamber slides (Nunc, Naperville, IL) precoated with 2% laminin (BD) and 1% poly-D-

lysine (Sigma, St Louis, MO) at a cell density of 50,000 cells per well. Cells were fed

every 2 days. For experiments with CNTF, cells were plated in FGF2 for 12 hours and

then CNTF (3 ng/ml, Peprotech, Rocky Hill, NY) was added to the media for an

additional 2 or 4 days.

Quantification of immunocytochemical staining was performed on dissociated cells or

cortical progenitor cell cultures. For each condition, >300 cells per well of at least two

wells were counted and analyzed. Digital image acquisition was performed on a Leitz

DM R B microscope with Leica DC 500 camera and manufacturer’s software

(Cambridge, UK). Photomicrographs were processed using Adobe Photoshop and cell

number quantified with Image ProPlus software.

Immunocytochemistry. Immunocytochemistry of cultured cells was carried out by

washing the cultures with HEPES-buffered saline (HBS), followed by fixation with 4%

137 paraformaldehyde for 15 min. The cells were permeabilized with 0.2% NP-40 in HBS,

followed by incubation in blocking with buffer containing 6% normal goat serum (NGS) and 0.5% bovine serum albumin (BSA) for 1-2 h at room temperature. Cells were then

incubated with primary antibodies in HBS containing 3% NGS and 0.25% BSA at 4°C

overnight. After washing with HBS, cells were incubated with secondary antibodies

prepared in HBS containing 3% NGS and 0.25% BSA at room temperature for 1 h.

Samples were then washed with HBS, and counterstained with 1 µg/ml 4',6'-diamidino-2- phenylindole (DAPI) for 2 min. Primary antibodies used were anti-Nestin (Chemicon,

Temecula CA 1:200), and anti-βIII tubulin (1:4000), anti-MAP2 (1:200), anti-S100β

(1:200). For immunohistochemistry, sections were rinsed in PBS, blocked, and

permeabilized with 0.1% Triton X-100, 10% goat serum in PBS. Sections were

incubated with primary antibodies at 4°C overnight, washed with PBS, and incubated with secondary antibodies at room temperature for 1 h. They were then counterstained

with DAPI for 2 min. The primary antibodies used were anti-Nestin, anti-MAP2, anti-

bIII-tubulin (all as described above), anti-Tbr1 (a gift of Dr. Robert Hevner, 1:2000),

anti-Tbr2 (Chemicon, 1:1000), anti-phosphoHistone H3 (Millipore, Billerica, MA,1:200)

and anti-p27 (BD Transduction Labs, 1:250). The secondary antibodies used were Alexa

fluorophore-conjugated IgG or IgM.

Western blot Analysis. Cells were washed with ice-cold HBSS and lysed by sonication in

lysis buffer (20mM Tris pH 7.5, 150mM NaCl, 1% NP-40, 10% glycerol, 1mM EDTA,

1.5mM MgCl2, 20mM NaF, 20mM β-glycerophosphate) supplemented with protease

inhibitors (1µg/ml leupeptin, 1µg/ml aprotinin, 1mM PMSF, 1mM Na3VO4). Lysates

138 were cleared by centrifugation and protein concentration was determined using the

bicinchoninic acid assay (Pierce, Rockford, IL) using BSA as a standard. Equal amounts

of protein were boiled in sample buffer, separated by SDS-PAGE gels, and transferred to

Immobilon-P PVDF membranes (Millipore). Membranes were blocked in 3% BSA (or

5% skim milk powder for anti-G3PDH) in TBS and 0.1% Tween 20 (TBS-T) for 2 h at

room temperature and incubated overnight at 4°C with primary antibodies. The primary

antibodies used were rabbit anti- pERK (Cell Signaling, Danvers, MA 1:1000), mouse

anti-ERK2 (BD, 1:3000), anti-β-tubulin (Santa Cruz, 1:5000), rabbit anti-G3PDH

(Trevigen, 1:5000), mouse anti-p27 (BD Transduction Labs, 1: 2500) and anti-Cyclin D2

(Thermo Scientific, 1: 1000). After washing with TBS-T, membranes were incubated

with HRP-conjugated goat anti-mouse or anti-rabbit (1:5,000; GE Healthcare) secondary

antibodies in TBS-T with 5%milk for 2 h at room temperature. Blots were washed with

TBS-T and detection was performed using either Millipore or Pierce chemiluminescence

reagents using BioMax MR X-ray film (Eastman Kodak, Rochester, NY). Densitometric

analysis was performed using Adobe Photoshop histogram function, and statistical

analysis was done with GraphPad Prism software.

BrdU Labeling and Analysis. Pregnant mice were injected intraperitoneally with BrdU at

100 mg/kg of body weight. The animals were sacrificed 30 min after the injection.

BrdU-positive cells were detected by immunostaining as described by the manufacturers’ instructions with heat-mediated antigen retrieval in citrate buffer followed by 2N HCl- treatment. Sections were incubated with a rat anti-BrdU antibody (Abcam 1:1000) overnight at 4oC.

139

Results

ERK1/2 DKO mice die embryonically, but display normal brain morphology.

To directly assess the combined roles of mapk3/ERK1 and mapk1/ERK2 in neural

development, we generated a mouse experiencing a conditional mapk1 null mutation with

a mapk3 null background. The mapk1flox/flox mouse line (previously described in Chapter

2) was interbred with a mapk3 null mouse line (Nekrasova et al., 2005) resulting in a

mouse with loxP sites surrounding exon 2 of mapk1 and a neo cassette replacing the span

between exons 1 and 7 of mapk3 (Fig. 3-1A). Transgenic mice in which cre recombinase expression is driven by the human glial fibrillary acidic protein (hGFAPcre/+) promoter

(Zhuo et al., 2001) were interbred with the mapk3-/-;mapk1flox/flox mice, resulting in the deletion of ERK2 within cortical NPCs in an ERK1 null environment at E13.5. PCR analysis confirms the recombination of exon 2 in mapk1/ERK2 represented by the presence of an amplified product, and the loss of mapk3/ERK1 in appropriate genotypes

(Fig 3-1B).

The mapk3-/-;mapk1flox/flox; hGFAPcre/+ (ERK1/2 Double Knockout, DKO) mice die

embryonically between E16.5 and birth. However, the DKO embryos display normal

morphology and size with no obvious phenotypic defects as compared to ERK1 knockout embryos or wildtype embryos at the same age (Fig 3-2A-C). It has been shown the the hGFAPcre/+ transgene is expressed in cardiac neural crest cells (Guris et al., 2001) and we have found that the hearts of the DKO mice have thin ventricular walls, and semi- penetrant ventricular septal defects which may cause the lethality (Fig. 3-2D-F). The

140 overall brain morphology and cortical structure at E14.5 is similar in the DKO, ERK1

knockout and wildtype (Fig. 3-2G-I).

Immunostaining of cortical cryosections from E16.5 DKO and ERK1 null littermates

demonstrated the loss of ERK2 in the developing neocortex (Fig. 3-3A). Western blot analysis of E14.5 cortical brain homogenates confirmed the activity of the hGFAPcre/+

transgene and revealed that ERK2 protein levels in the cortex were reduced to 25% of

mapk3/ERK1 -/- levels (Fig. 3-3B). Activity of ERK2 as measured by its

phosphorylation status was correspondingly reduced by 40% in the DKO at E14.5 (Fig.

3-3C). Cortical homogenates from ERK2 CKO mice revealed an increased level of

active, phosphorylated ERK1 at E16.5, suggesting a compensatory role for ERK1 with

the loss ERK2 (Chapter 2). As ERK1 is never expressed in the ERK1/2 DKO mice, this

phenomenon does not occur (Fig. 3-3C).

Loss of both ERK1 and ERK2 does not exacerbate the alterations in cellular

composition identified in the ERK2 CKO cortex.

Histological analysis of DKO cortices by cresyl violet staining showed no obvious

defects in cortical lamination or morphology (Fig 3-2G-I). At E16.5, it is difficult to

assess the 10-15% reduction in cortical thickness identified within neonatal ERK2 CKO

mice (Chapter 2). However, analysis of the cellular composition of the DKO cortex

revealed that fewer neurons are produced compared to littermate mapk3/ERK1 nulls by

this time point, as displayed by reductions in MAP2 and Tbr1 immunostaining (Fig. 3-4A

and Fig. 3-4B, respectively). This finding is comparable to the 10% reduction in NeuN+

141 neurons identified in ERK2 CKO mice at postnatal day 10. Furthermore, DKO embryos

maintain significantly more Nestin progenitors as compared to mapk3/ERK1 null littermates (Fig. 3-4C), a finding first recognized in the ERK2 CKO. No significant difference in the number of immature neurons immunostained with βIII-tubulin was

revealed between DKO and mapk3/ERK1 null embryos at E16.5 however (Fig. 3-4D).

Cortical progenitor cells cultured at E14.5 and maintained in vitro for 2 days in the

presence of FGF2 revealed analagous differences to the CKO in cellular generation. The

1.418 fold increase in nestin+ progenitors from mapk3/ERK1 -/- cultures as compared to

DKO cultures (p=0.0022, Fig. 3-5A) was correlated with fewer MAP2+ neurons (0.6935

fold, p=0.0001, Fig. 3-5C) and fewer S100β+ glial progenitor cells (0.7904 fold,

p=0.0006, Fig 3-5D). No difference in the percentage of βIII-tubulin+ neurons was

identified between genotypes (0.9501 fold different, p=0.0548, Fig 3-5B). Together, our

in vivo and in vitro findings suggest that ERK1 does not compensate for ERK2 during

cortical development, as the DKO mice do not display a more severe cortical phenotype

than ERK2 CKO mice.

DKO neural progenitor cells display inhibited cell cycle progression attributable to

reductions in cyclin D2 and increased p27Kip1 levels.

Reductions in intermediate progenitor cell divisions and an inhibition of neural

progenitor cells to proceed through G1 to S phase within the ERK2 CKO mice resulted in

the production of fewer neurons and more astrocytes as compared to wildtype littermates

(Chapter 2). To investigate if these processes were similarly affected in ERK1/2 DKO

142 embryos, pregnant dams were injected with BrdU at E14.5 and E16.5 and harvested 0.5

hr post injection. Embryos were either immediately fixed by immersion in 4%

paraformaldehyde or individual embryonic cortices were dissected, dissociated, plated

and immediately processed for quantitative immunostaining. We found that at E14.5,

BrdU incorporation was reduced by 0.6815 fold in ERK1/2 DKO embryos as compared

to mapk3/ERK1 -/- embryos (p= 0.0008, Fig. 3-6C). Coronal sections from embryos

reflect this difference (Fig. 3-6A,B). Interestingly, we found that at E16.5, BrdU

incorporation was also reduced in ERK1/2 DKO embryos as compared to mapk3/ERK1 null embryos, although to a lesser extent, 0.8032 fold (Fig. 3-6D-F). This finding contrasts the observed increase in BrdU incorporation in the ERK2 CKO at E16.5. Our current data confirms the hypothesis that the increase in BrdU incorporation observed in the CKO was due to elevated phospho-ERK1 levels at this time. Our data further suggests that ERK1 and ERK2 play similar roles in proliferation of neural progenitor cells as targeted deletion of both isoforms exacerbates the phenotype found in the ERK2

CKO.

We next investigated the effect of the deletion of both isoforms on progenitor cell division by analyzing the number of mitotic neural progenitor cells and intermediate progenitor cells in ERK1/2 DKO embryos at E14.5. As seen in the ERK2 CKO, fewer

Tbr2+ intermediate progenitor cells were positive for phosphorylated Histone H3 (pH3+) in the DKO as compared to both ERK1 null and wildtype embryos (Fig. 3-7). Our data therefore confirms that the reduction in neuronal production identified in our ERK1/2

DKO embryos is due to slowed cell cycle dynamics and fewer mitotic divisions.

143

ERK1/2 directly promotes cell cycle progression from G1 to S phase by inducing

transcription of the D-type cyclins through the phosphorylation of AP1 and Ets

transcription factors, Elk-1 and c-myc. Cyclin D1 forms a complex with cyclin dependent kinase 4/6 (cdk4/6) to propel the cell past the G1 restriction checkpoint.

Concurrently, ERK1/2 promotes the degradation of the cyclin dependent kinase inhibitor, p27Kip1, through phosphorylation of the ubiquition ligase, SKP2. p27 degradation directly promotes progression to S phase late in G1 as it initially inhibits the rate-limiting enzymatic complex for progression, Cdk2-Cyclin E. To investigate the effect of ERK1/2 loss on transition between G1 and S phase, we measured both CyclinD1/2 and p27 levels at E14.5. We found that ERK1/2 DKO cortices had higher expression levels of p27 than mapk3/ERK1 -/- littermates (Fig. 3-8A). Immunostaining of coronal sections reflects this difference (Fig. 3-8B). We also found that Cyclin D2 levels were lower in ERK1/2

DKOs as compared to mapk3/ERK1 null animals (Fig. 3-8C). Expression of Cyclin D1 was not consistently different between DKO and ERK1 null embryos (Fig. 3-8D).

Discussion

We have investigated the individual roles of ERK1 and ERK2 during neural development through the conditional inactivation of mapk1/ERK2 in a mapk3/ERK1 null mouse model. We find that until embryonic lethality, the cortices of these mice display the same phenotype as ERK2 CKO mice, whereby fewer neurons and more progenitors populate the cortical plate. We further demonstrate similar changes in the dynamics of cell cycle progression, as less BrdU incorporation occurs at E14.5. However, we now establish that

144 the increase in BrdU incorporation identified at E16.5 in the ERK2 CKO is associated with compensatory activation and signaling by ERK1, as the DKO mouse continues to display less BrdU incorporation compared to ERK1 null littermate controls at this stage of neurogenesis.

We have found that the expression levels of two downstream targets of ERK1/2, the cell cycle progressive cyclin, D2 is reduced, and the cyclin dependent kinase inhibitor, p27Kip1 is elevated in the DKO, at E14.5. The disruption of these cell cycle regulators was correlated with a reduction in intermediate progenitor cell division. These findings provide direct evidence for the unique actions of the ERK2 isoform in neural development and demonstrate that the ERK1 isoform does not play a significant role in neurogenesis. Because the DKO mice die before gliogenesis, we were unable to assess the capacity for astrocyte generation and proliferation in vivo. However, we speculate that gliogenesis would occur undisrupted in the ERK1/2 DKO mouse as in the CKO and hypothesize that more astrocytes would populate the cortex.

It is widely accepted that sustained ERK1/2 activity is required for cell cycle progression through G1 phase and for passage past the G1 restriction checkpoint into S phase. The primary mechanism by which ERK1/2 exerts these effects is through production of the

D-type cyclins, a rate-determining molecule in G1 progression (Villanueva et al., 2007).

Cyclin D complexes with cdk4 and leads to both the sequestration and subsequent degradation of the cell cycle inhibitor p27Kip1 as well as the catalytic actions of Cyclin

D-Cdk4. This enzyme complex furthers G1 progression through phosphorylation of the

145 pocket protein pRb, allowing for the release of E2Fs and initiation of E2F dependent gene transcription. The E2F dependent gene Cyclin A then complexes with cdk2 to propel the cell into S phase. Inactivation of the D-type cyclins in mouse models does not cause significant phenotypic changes in the brain due to compensation by Cyclin E (Chen et al., 2005); however, these mice do exhibit neurological abnormalities, overall small size, hypoplastic retinas and pregnancy-insensitive mammary glands which are due to lack of proliferation (Fantl et al., 1995; Sicinski et al., 1995). Our findings demonstrate reductions in Cyclin D2 levels. It has recently been shown that Cyclin D1 and Cyclin D2 display differential expression profiles in forebrain progenitor cell subsets. Specifically,

Cyclin D1 is highly expressed throughout the ventricular zone in early to mid- neurogenesis and decreases as development progresses, while Cyclin D2 expression rises from mid-neurogenesis to late neurogenesis and is more highly expressed in intermediate progenitor cells (Glickstein et al., 2007). It is noteworthy that we find both reductions in

Cyclin D2 and the number of mitotic intermediate progenitor cells. These data suggest that the ERKs may play specific roles in proliferation of intermediate progenitor cells.

Macrophage-colony stimulating Factor-1 (CSF-1) induces cyclin D1 and D2 through combinatorial activation of src, myc, and ERK1/2. Complete inhibition of MEK/ERK

downstream of CSF-1 stimulation causes a 50% decrease in D2 expression, similar to the

level of reduction identified in our mouse model. Cyclin D2 expression is regulated by

the Myc/Max/Mad network (Bouchard et al., 2001) and it has been shown that activation

of myc’s transactivating activity selectively induces D2 but not D1 (Dey et al., 2000).

These findings provide a mechanism by which ERK1/2 inactivation could lead to the

146 selective reduction in Cyclin D2 but not D1. It is of significance that mice harboring a

null mutation in CSF-1 have alterations the cortical circuitry that balances excitation and

inhibition. (Michaelson et al., 1996)

Our data revealed elevations in p27kip1 in the DKO cortex. Animals in which p27Kip1

has been knocked out exhibit increased cortical thickness resulting from a 27% increase

in neurons in upper cortical lamina (Goto et al., 2004). Conversely, overexpression of

p27Kip1 in the developing telencephalon slows progenitor proliferation with fewer

neurons populating the upper cortical layers resulting in reduced cortical thickness (Tarui

et al., 2005). The conditional knockout of the cell cycle regulator pRb in the developing

telencephalon also causes a robust (7 fold) increase in progenitor proliferation resulting in

a dramatic increase in telencephalon size (Ferguson et al., 2002). Intriguingly, removal

of p27 in a cyclin D1 null animal restored cell cycle control and abolished the phenotypic

defects (Geng et al., 2001). Expression levels of p27 are also regulated by c-myc. Myc

stimulates the expression of Cyclin D2 and Cdk4 which in turn sequester p27 in a

p27/CyclinD2/Cdk4 complex, targeting it for degradation (Bouchard et al., 1999; Coller et al., 2000; Dey et al., 2000; Bouchard et al., 2001). These data support our identification of reduced Cyclin D2 levels and increased p27 levels in the DKO cortex.

Furthermore, the alteration in expression of these proteins is sufficient to induce the

change in cellular generation identified in the ERK1/2 DKO. It is of interest to determine

if a reduction in p27 levels would reverse these deficits in neuronal generation.

147 Our findings also reveal that the canonical ERK1/2 signaling pathway is not solely

responsible for generation and differentiation of cortical neurons. In the absence of

ERK1 and near complete loss of ERK2 at the peak of neurogenesis, neurons are still

made. It appears that the ERKs regulate cell fate through the critical maintenance of cell

cycle control. It has been shown that the non-classical ERK5 isoform is important for

neuronal generation and survival but does not regulate cortical progenitor cell

proliferation apoptosis (Liu et al., 2006; Nishimoto and Nishida, 2006). Furthermore,

ERK5 is able to regulate Cyclin D1 expression (Mulloy et al., 2003), providing an

additional potential explanation for the lack of reductions in this cell cycle regulator when ERK1/2 are inactivated. Additionally, Cyclin E can compensate for any reductions in Cyclin D activity, allowing for the generation of a generally normal cerebral cortex.

The ERK1/2 DKO mice are recoverable to E16.5 as ERK2 inactivation does not occur until E13.5 and the ventricular septal defects are not lethal until later in development. A lack of information regarding expression of the GFAP-cre transgene in the heart along with ambiguous expression of GFAP itself in this organ makes it difficult to ascertain exactly where recombination occurs to lead to the VSDs observed. Furthermore, the incomplete penetrance of the phenotype does not clarify the deficit. Importantly, mice in

which the ERKs have been inactivated in the developing cardiac neural crest develop

aortic arch defects and associating ventral septal defects (Newbern et al, submitted).

These mice have been generated as models of 22q11 Deletion Syndrome as they

phenocopy the disease phenotypes. We propose that the ERK1/2 DKOs generated here

can be useful in identifying and understanding the cortical phenotypes associated with

148 this disease. Because over 25% of 22q11 Deletion syndrome patients develop schizophrenia and affective mood disorders, this mouse model is also important for delineating the underlying mechanisms of these disorders.

149 Figure 3-1. Generation of ERK1/2 Double Knockout mice. A, Schematic representation of gene targeting strategy. The numbered boxes represent mapk3 exons 1 through 9 and mapk1 exons 2 through 6. A gene targeting strategy was designed to create a null allele for mapk3. Following homologous recombination, the mapk3 allele contains a floxed

Neo cassette replacing genomic sequence spanning from exon 1 to exon 7. For mapk1, we created a floxed allele with loxP sites flanking exon 2. Following hGFAP cre- mediated recombination, loss of exon 2 results in a null allele. B, PCR of genomic DNA isolated from embryonic cerebral cortices. An ERK2 KO reaction was designed to generate a product in the absence of Exon 2. For ERK1, the upper band represents the endogenous ERK1 allele. The lower band is a non-specific PCR product.

150 Figure 3-1

151 Figure 3-2. ERK1/2 DKO embryos die embryonically but display normal brain morphology. Lateral view of a wildtype (A), ERK1 null (B) and ERK1/2 DKO (C) littermate embryos at E14.5. Horizontal sections of a wildtype B6 mouse heart (D),

ERK1 null heart (E) and the heart from an ERK1/2 DKO (F) embryo at E14.5 stained with hematoxylin and eosin. Coronal sections of cortices from E14.5 wildtype (G),

ERK1 null (H) and ERK1/2 DKO (I) littermates stained with cresyl violet.

152 Figure 3-2

153 Figure 3-3. Loss of ERK2 expression and activity in ERK1/2 DKO cortices. A,

Immunostaining of coronal cortical sections from E16.5 ERK1/2 DKO and ERK1 null littermates using an anti-ERK2 antibody demonstrated the loss of ERK2 in the developing neocortex. B, E14.5 cortices were microdissected and prepared for Western blot analysis immediately upon dissection. The levels of ERK2 and phospho-ERK1/2 were detected by immunoblotting and representative blots are shown. Densitometric analysis of the bands, normalized to β-tubulin, was performed for quantification and expressed as fold difference relative to wild-type. For B, Quantification of ERK2 levels at E14.5 (DKO n=9, ERK1 null n=6, p<0.0001). For C, Relative phospho-ERK1/2 activation at E14.5, (DKO n=5, ERK1 null n=3, p=0.0444).

154 Figure 3-3

155 Figure 3-4. Inactivation of ERK1 and ERK2 in neural progenitor cells results in the generation of fewer cortical neurons. A-D, Analysis of cellular composition at E16.5.

Corresponding coronal sections of ERK1/2 DKO (left) and ERK1 null (right) littermates were immunostained with anti-MAP2 (A, green) anti-Tbr1 (B, red) anti-Nestin (C, green)

and anti-βIII tubulin (D, red) antibodies.

156 Figure 3-4

157 Figure 3-5. ERK1/2 DKO cortical progenitor cells exhibit reduced neuronal generation.

A, Cortical progenitor cell cultures from ERK1/2 DKO and ERK1 null E14.5 embryos were grown in Neurobasal with FGF2 for 2 days in vitro. Analysis of cellular identity in culture was performed by immunostaining with various cellular markers. Fold difference is the comparison of the average number of positive cells/total cell number relative to

ERK1 null. Progenitors were identified by Nestin immunoreactivity (A, n=5 for each, p=0.0022). Cultures were assayed for immature neurons with βIII Tubulin (B, n=5 for each, p=0.0548), mature neurons with MAP2 (C, for each n=4, p=0.0001), and astrocyte precursors with S-100β (D, n=5 for each, p=0.0006).

158 Figure 3-5

159 Figure 3-6. ERK1/2 DKO mice display reductions in proliferation. Pregnant mice were given intraperitoneal injections of BrdU at E14.5 (A-C) and E16.5 (D-F), respectively.

Embryos were fixed 30 minutes after BrdU injection and corresponding coronal sections of ERK1/2 DKO and ERK1 null littermates were double labeled with anti-phospho- histone H3 (pH3+, red) and anti-BrdU (green) antibodies. Cortices were similarly dissected, dissociated, plated and immediately fixed for immunostaining and quantification of BrdU immunoreactivity. For E14.5, DKO n= 5, ERK1 null n=5, p=0.0008 (C), For E16.5, DKO n=3, ERK1 null n=3, p= 0.0028 (F).

160 Figure 3-6

161 Figure 3-7. Fewer abventricular intermediate progenitor cells from ERK1/2 DKO embryos are mitotic at the peak of neurogenesis. Corresponding coronal sections of

E14.5 ERK1/2 DKO (A,D), ERK1 null (B,E) and wildtype (C,F) embryos were immunostained with Tbr2 (green) and phosphorylated Histone H3 (red).

162 Figure 3-7

163 Figure 3-8. DKO neural progenitor cells exhibit reductions in cyclin D2 and increased p27Kip1 levels. A-D, Analysis of cell cycle components. E14.5 cortices were microdissected and prepared for Western blot analysis immediately upon dissection or corresponding coronal sections from littermates were cryosectioned and immunostained.

The levels of p27Kip1 (A, B), Cyclin D2 (C) and Cyclin D1 (D) were detected by immunoblotting and representative blots are shown. Densitometric analysis of the bands, normalized to β-tubulin or G3PDH, was performed for quantification and expressed as fold difference relative to wild-type. For A, Quantification of p27 levels at E14.5 (DKO n=9, ERK1 null n=6, p=0.0829). B, Corresponding coronal sections were immunostained using an anti-p27 antibody (green). For C, Relative Cyclin D2 at E14.5, (DKO n=3,

ERK1 null n=2, p=0.0690) and D, representative western blot analysis of Cyclin D1 expression.

164 Figure 3-8

165 Literature Cited

Barnabe-Heider F, Miller FD (2003) Endogenously produced neurotrophins regulate survival and differentiation of cortical progenitors via distinct signaling pathways. J Neurosci 23:5149-5160. Bouchard C, Dittrich O, Kiermaier A, Dohmann K, Menkel A, Eilers M, Luscher B (2001) Regulation of cyclin D2 gene expression by the Myc/Max/Mad network: Myc-dependent TRRAP recruitment and histone acetylation at the cyclin D2 promoter. Genes Dev 15:2042-2047. Bouchard C, Thieke K, Maier A, Saffrich R, Hanley-Hyde J, Ansorge W, Reed S, Sicinski P, Bartek J, Eilers M (1999) Direct induction of cyclin D2 by Myc contributes to cell cycle progression and sequestration of p27. Embo J 18:5321- 5333. Chambard JC, Lefloch R, Pouyssegur J, Lenormand P (2006) ERK implication in cell cycle regulation. Biochim Biophys Acta. Chen Z, Duan RS, Zhu Y, Folkesson R, Albanese C, Winblad B, Zhu J (2005) Increased cyclin E expression may obviate the role of cyclin D1 during brain development in cyclin D1 knockout mice. J Neurochem 92:1281-1284. Coller HA, Grandori C, Tamayo P, Colbert T, Lander ES, Eisenman RN, Golub TR (2000) Expression analysis with oligonucleotide microarrays reveals that MYC regulates genes involved in growth, cell cycle, signaling, and adhesion. Proc Natl Acad Sci U S A 97:3260-3265. Corson LB, Yamanaka Y, Lai KM, Rossant J (2003) Spatial and temporal patterns of ERK signaling during mouse embryogenesis. Development 130:4527-4537. Dey A, She H, Kim L, Boruch A, Guris DL, Carlberg K, Sebti SM, Woodley DT, Imamoto A, Li W (2000) Colony-stimulating factor-1 receptor utilizes multiple signaling pathways to induce cyclin D2 expression. Mol Biol Cell 11:3835-3848. Fantl V, Stamp G, Andrews A, Rosewell I, Dickson C (1995) Mice lacking cyclin D1 are small and show defects in eye and mammary gland development. Genes Dev 9:2364-2372. Ferguson KL, Vanderluit JL, Hebert JM, McIntosh WC, Tibbo E, MacLaurin JG, Park DS, Wallace VA, Vooijs M, McConnell SK, Slack RS (2002) Telencephalon- specific Rb knockouts reveal enhanced neurogenesis, survival and abnormal cortical development. Embo J 21:3337-3346. Fremin C, Ezan F, Boisselier P, Bessard A, Pages G, Pouyssegur J, Baffet G (2007) ERK2 but not ERK1 plays a key role in hepatocyte replication: an RNAi- mediated ERK2 knockdown approach in wild-type and ERK1 null hepatocytes. Hepatology 45:1035-1045. Gauthier AS, Furstoss O, Araki T, Chan R, Neel BG, Kaplan DR, Miller FD (2007) Control of CNS cell-fate decisions by SHP-2 and its dysregulation in Noonan syndrome. Neuron 54:245-262. Geng Y, Yu Q, Sicinska E, Das M, Bronson RT, Sicinski P (2001) Deletion of the p27Kip1 gene restores normal development in cyclin D1-deficient mice. Proc Natl Acad Sci U S A 98:194-199. Glickstein SB, Alexander S, Ross ME (2007) Differences in cyclin D2 and D1 protein expression distinguish forebrain progenitor subsets. Cereb Cortex 17:632-642.

166 Goto T, Mitsuhashi T, Takahashi T (2004) Altered patterns of neuron production in the p27 knockout mouse. Dev Neurosci 26:208-217. Guris DL, Fantes J, Tara D, Druker BJ, Imamoto A (2001) Mice lacking the homologue of the human 22q11.2 gene CRKL phenocopy neurocristopathies of DiGeorge syndrome. Nat Genet 27:293-298. Hatano N, Mori Y, Oh-hora M, Kosugi A, Fujikawa T, Nakai N, Niwa H, Miyazaki J, Hamaoka T, Ogata M (2003) Essential role for ERK2 mitogen-activated protein kinase in placental development. Genes Cells 8:847-856. Lefloch R, Pouyssegur J, Lenormand P (2008) Single and combined silencing of ERK1 and ERK2 reveals their positive contribution to growth signaling depending on their expression levels. Mol Cell Biol 28:511-527. Liu L, Cundiff P, Abel G, Wang Y, Faigle R, Sakagami H, Xu M, Xia Z (2006) Extracellular signal-regulated kinase (ERK) 5 is necessary and sufficient to specify cortical neuronal fate. Proc Natl Acad Sci U S A 103:9697-9702. Liu X, Yan S, Zhou T, Terada Y, Erikson RL (2004) The MAP kinase pathway is required for entry into mitosis and cell survival. Oncogene 23:763-776. Mazzucchelli C, Vantaggiato C, Ciamei A, Fasano S, Pakhotin P, Krezel W, Welzl H, Wolfer DP, Pages G, Valverde O, Marowsky A, Porrazzo A, Orban PC, Maldonado R, Ehrengruber MU, Cestari V, Lipp HP, Chapman PF, Pouyssegur J, Brambilla R (2002) Knockout of ERK1 MAP kinase enhances synaptic plasticity in the striatum and facilitates striatal-mediated learning and memory. Neuron 34:807-820. Menard C, Hein P, Paquin A, Savelson A, Yang XM, Lederfein D, Barnabe-Heider F, Mir AA, Sterneck E, Peterson AC, Johnson PF, Vinson C, Miller FD (2002) An essential role for a MEK-C/EBP pathway during growth factor-regulated cortical neurogenesis. Neuron 36:597-610. Michaelson MD, Bieri PL, Mehler MF, Xu H, Arezzo JC, Pollard JW, Kessler JA (1996) CSF-1 deficiency in mice results in abnormal brain development. Development 122:2661-2672. Mulloy R, Salinas S, Philips A, Hipskind RA (2003) Activation of cyclin D1 expression by the ERK5 cascade. Oncogene 22:5387-5398. Nekrasova T, Shive C, Gao Y, Kawamura K, Guardia R, Landreth G, Forsthuber TG (2005) ERK1-deficient mice show normal T cell effector function and are highly susceptible to experimental autoimmune encephalomyelitis. J Immunol 175:2374- 2380. Nishimoto S, Nishida E (2006) MAPK signalling: ERK5 versus ERK1/2. EMBO Rep 7:782-786. Nowakowski RS, Caviness VS, Jr., Takahashi T, Hayes NL (2002) Population dynamics during cell proliferation and neuronogenesis in the developing murine neocortex. Results Probl Cell Differ 39:1-25. Pages G, Guerin S, Grall D, Bonino F, Smith A, Anjuere F, Auberger P, Pouyssegur J (1999) Defective thymocyte maturation in p44 MAP kinase (Erk 1) knockout mice. Science 286:1374-1377. Paquin A, Barnabe-Heider F, Kageyama R, Miller FD (2005) CCAAT/enhancer-binding protein phosphorylation biases cortical precursors to generate neurons rather than astrocytes in vivo. J Neurosci 25:10747-10758.

167 Saba-El-Leil MK, Vella FD, Vernay B, Voisin L, Chen L, Labrecque N, Ang SL, Meloche S (2003) An essential function of the mitogen-activated protein kinase Erk2 in mouse trophoblast development. EMBO Rep 4:964-968. Selcher JC, Nekrasova T, Paylor R, Landreth GE, Sweatt JD (2001) Mice lacking the ERK1 isoform of MAP kinase are unimpaired in emotional learning. Learn Mem 8:11-19. Sicinski P, Donaher JL, Parker SB, Li T, Fazeli A, Gardner H, Haslam SZ, Bronson RT, Elledge SJ, Weinberg RA (1995) Cyclin D1 provides a link between development and oncogenesis in the retina and breast. Cell 82:621-630. Tarui T, Takahashi T, Nowakowski RS, Hayes NL, Bhide PG, Caviness VS (2005) Overexpression of p27 Kip 1, probability of cell cycle exit, and laminar destination of neocortical neurons. Cereb Cortex 15:1343-1355. Vantaggiato C, Formentini I, Bondanza A, Bonini C, Naldini L, Brambilla R (2006) ERK1 and ERK2 mitogen-activated protein kinases affect Ras-dependent cell signaling differentially. J Biol 5:14. Villanueva J, Yung Y, Walker JL, Assoian RK (2007) ERK activity and G1 phase progression: identifying dispensable versus essential activities and primary versus secondary targets. Mol Biol Cell 18:1457-1463. Yao Y, Li W, Wu J, Germann UA, Su MS, Kuida K, Boucher DM (2003) Extracellular signal-regulated kinase 2 is necessary for mesoderm differentiation. Proc Natl Acad Sci U S A 100:12759-12764. Zhuo L, Theis M, Alvarez-Maya I, Brenner M, Willecke K, Messing A (2001) hGFAP- cre transgenic mice for manipulation of glial and neuronal function in vivo. Genesis 31:85-94.

168 DISCUSSION

ERK1 and ERK2 are ubiquitously expressed kinases which are activated by a diverse

variety of stimuli (Cobb et al., 1991). The importance of their expression and activation

is demonstrated by the existence of two isoforms which are 90% homologous when the

N-terminus is excluded from comparison (Boulton and Cobb, 1991). Both isoforms are

expressed in all species from yeast through human, with the exception of only the frog

and chick (which express only ERK2; (Lefloch et al., 2008). These facts suggest that either isoform specific functions exist, or that the two isoforms provide functional redundancy which ensures signaling fidelity.

Both ERK1 and ERK2 are expressed at high levels in the mature rodent cortex (Boulton et al., 1990; Boulton and Cobb, 1991; Ortiz et al., 1995). Furthermore, ERK signaling is required for both cortical development (Miller and Gauthier, 2007) and for long term potentiation and learning and memory in adult animals (Sweatt, 2001). Therefore, we utilized the murine cerebral cortex as a system in which to investigate the potential of isoform-specific actions of ERK1 and ERK2.

Isoform specific expression regulates isoform specific

function

We, and others, have previously demonstrated that ERK1 knockout mice do not display

significant phenotypic changes in the cortex (Pages et al., 1999; Selcher et al., 2001;

169 Mazzucchelli et al., 2002). However, it was established that ERK2 activity is elevated in

the absence of ERK1, demonstrating a potential compensatory role for ERK2. In ERK2 conditional knockout mice we found a reduction in cortical thickness which is associated with changes neural progenitor cell proliferation and differentiation such that fewer neurons and more astrocytes populate the cortex (Chapter 2). While no difference in

ERK1 activity at E14.5 was observed, we found a significant increase in ERK1 activity at

E16.5, three days after recombination was initiated. This elevated level of activity suggests that ERK1 may compensate for the absence of ERK2. However, the identification of specific perturbations in cortical cell fate which cause a defect in cortical thickness demonstrates the possibility of an ERK2-specific role in neural development.

To address this possibility, we generated double knockout mice, where both isoforms are inactivated in the developing cortex. The phenotype of the cortex in this mouse is slightly more perturbed than the single ERK2 conditional knockout at the time of embryonic death occurring in utero at about E17 (Chapter 3). The lethality of this mouse can be attributed to heart defects and is unlikely to be related to ERK1/2 inactivation within the nervous system. Because the loss of both ERK isoforms did not significantly exacerbate the cortical defects identified in the single ERK2 conditional knockout, it appears that ERK1 may not compensate for ERK2. This suggests that either ERK1 does not play a substantial role in cortical development, or that ERK1 is not expressed at high levels in the cells contributing to elaboration of the cortical mantle. We have visualized expression of ERK1 from dissociated cortices at E14.5 and E16.5 by western blotting techniques, but have been unable to observe ERK1 expression by immunohistochemistry on cryosections from animals of the same age. Therefore, it is difficult to assess the

170 location of ERK1 action at this time. The lack of a more severe phenotype in the DKO as

compared to the ERK2 CKO can be explained if ERK1 is expressed in post-mitotic

neurons but not neural progenitor cells. Lennormand et al., have recently assessed expression of ERK1 and ERK2 in adult nude mice. They demonstrate that ERK2 is 13 times more abundant than ERK1 in the prefrontal cortex (Lefloch et al., 2008).

Furthermore, studies of ERK1/2 protein expression in rats demonstrates that the

ERK1:ERK2 ratio is 0.16 and 0.28 in the frontal and parietal cortices, respectively (Ortiz et al., 1995). Analysis of ERK1 and ERK2 mRNA in the CNS by in situ hybridization revealed that there were no regions where ERK1 mRNA was present which did not also exhibit ERK2 mRNA localization. Furthermore, ERK2 mRNA was more widely distributed and displayed a higher level of expression than ERK1 throughout all cortical layers (Thomas and Hunt, 1993). These data argue that the modestly exacerbated phenotype that we documented in the double knockout as compared to the ERK2 CKO

(Chapter 3) may result from the low level of expression and activity of ERK1 at this time and place, not from an ERK2 specific function. However, within these experiments,

ERK1 and ERK2 mRNA was not measured at the critical developmental periods or locations.

A few examples of putative ERK1 and ERK2 specific functions have been described.

The most compelling evidence is the potent activation of ERK2 but not ERK1 in response to high frequency tetanic stimulation, which induces LTP in the hippocampus

(English and Sweatt, 1996, 1997). ERK1 deficient mice display deficits in thymocyte maturation (Pages et al., 1999), adipocyte differentiation (Bost et al., 2005a; Bost et al.,

171 2005b), and striatum-dependent synaptic plasticity (Mazzucchelli et al., 2002).

Additionally, ERK1 and ERK2 have been shown to play differential roles in certain phases of the cell cycle, depending on the cell line. In hepatocytes, ERK2 ablation inhibits proliferation while ERK1 silencing has no effect (Fremin et al., 2007). In HeLa cells, ERK1 inactivation arrests the cell cycle at the G2/M transition, while ERK2 inactivation results in cell arrest within G1 phase (Liu et al., 2004). Contradictory results arise from experiments performed with NIH 3T3 fibroblasts: silencing of ERK1 initially demonstrated that this isoform was involved in suppression of cell proliferation and

ERK2 was involved in the promotion of cell division; however, later studies demonstrated that differential expression of each isoform contributed to these results and that both ERK1 and ERK2 positively control cell proliferation (Vantaggiato et al., 2006;

Lefloch et al., 2008). In embryonic stem cells, ERK1/2 activity is required for self- renewal but only loss of ERK2 affects trophoblast stem cells (Meloche et al., 2004).

These data correspond with a specific role for ERK2 in trophoblast and mesoderm development of the embryo, as ERK2 knockout mice die during development and ERK1 cannot compensate (Hatano et al., 2003; Saba-El-Leil et al., 2003; Yao et al., 2003).

Both isoforms play a role in the positive selection for T-cell differentiation (Nekrasova et al., 2005; Agrawal et al., 2006), whereby inactivation of ERK1 and ERK2 results in more severe defects than loss of ERK1 alone. These results support the hypothesis that at least a significant proportion of presumed ERK1 and ERK2 specific functions are related to expression levels in a given environment (Shaul and Seger, 2007). It is important to note that the relative expression of each isoform has not been assessed in detail in many of the

172 tissues and cell types noted above, mainly due to the high between the isoforms and antibody cross-reactivity.

It is of interest however, that the immediate activators of ERK1/2, MEK1 and MEK2 have both redundant and distinct functions. Like the ERKs, knockout of each isoform leads to vastly different results. MEK1 knockout mice die due to placental defects while

MEK2 knockouts are overtly normal in growth and development (Giroux et al., 1999;

Belanger et al., 2003). Furthermore, knockdown of each isoform leads to different outcomes in cell cycle control. MEK1 seems to promote proliferation while MEK2 inhibits this process through upregulation of the CKI, p21 (Ussar and Voss, 2004). A recent report has further demonstrated that MEK1 shuttles ERK2 to the nucleus and

MEK2 differentially sorts ERK2 to the cytoplasm (Skarpen et al., 2008). Although these data point out the differential roles for each isoform, it is important to note that MEK1 and MEK2 do not share the same extent of homology as ERK1 and ERK2, especially in their regulatory domains. These data also support the idea of differential activation of each ERK isoform in dictating specific functions.

While a definitive answer about isoform-specific functions cannot be gleaned from previous studies or the work in this dissertation, we propose that the regulation of isoform-specific expression and possibly activation is responsible for the apparent isoform specific functions for ERK1 and ERK2 identified to date. A detailed analysis of

ERK1 and ERK2 mRNA and protein expression throughout development and on a tissue and cell type basis is required for a complete understanding of ERK1/2 activity. An

173 additional line of experiments which would shed some light on the redundancy of the

isoforms would be the generation of knockin mouse lines in which each isoform was

expressed under the other’s promoter and regulatory elements. The phenotype of these

mice would reveal isoform-specific actions and redundant functions.

ERK and Neurogenesis/the cell cycle

Surprisingly, neither the loss of ERK1 or ERK2 alone, nor both ERK1 and ERK2 in the

developing cortex, lead to gross structural defects in the brain (Chapter 2 and 3

respectively). Furthermore, while we could not assess behavior in the double knockout

due to embryonic death, the ERK2 CKO displays behavioral impairments in fear

conditioning but not sensory tests, motor function, reproduction, or longevity. Due to the

relatively minor defects we did identify, two hypotheses can be entertained: (1) ERK1

and ERK2 do not play a significant role in neurogenesis or (2) Other signaling molecules

compensate for the loss of these isoforms.

To address the first possibility, the phenotype of the single and double knockouts must be

considered. The phenotype of the ERK2 CKO and ERK1/2 DKO cortex is similar to that

of other mouse models in which cell cycle components have been mutated or lost. The

knockout of the proto-oncogene n-myc in neural progenitor cells leads to microcephaly

due to severely disrupted proliferation. Like ERK2, n-myc is highly expressed in the

developing CNS. When inactivated in response to recombination with a Nestin-cre

trangene, reductions in neuroepithelial cell proliferation result, but not apoptosis

174 (Knoepfler et al., 2002). This mouse also exhibited high levels of the CKI, p27. A mutant mouse model in which p27 is over-expressed in neural progenitor cells is also phenotypically similar to the ERK2 CKO and ERK1/2 DKO mice. Specifically, significantly thinner cortices were identified which corresponded to fewer neurons populating the cortical lamina (Tarui et al., 2005). ERK2 CKO and ERK1/2 DKO (even more dramatically) mice display proliferation deficits, without apoptosis, and increased levels of p27. These results describe a mitogenic role for the ERKs in neurogenesis and predict that other cell cycle mutants will display similar findings.

Studies of mice in which all 3 D cyclins, or both cyclins E, or cdk4/6 and cdk2 have been knocked out demonstrate that these seemingly important proteins are dispensable for the majority of fetal development (Sherr and Roberts, 2004; Kozar and Sicinski, 2005).

Cyclin D1-/-D2-/-D3-/- mice are phenotypically similar to our ERK1/2 DKO mice as they die around embryonic day 16, but are of normal size and exhibit cardiac malformations characterized by ventricular septal defects and reductions in wall thickness (Kozar et al.,

2004). Furthermore, cdk4/6 double knockout mice display similar defects, owing to the ability of cyclin E-cdk2 and cyclin A-cdk2 to compensate for the loss of cyclin D or cdk4/6 (Malumbres et al., 2004). The role of cyclinD1-cdk4 activity is to phosphorylate and inactivate the retinoblastoma tumor suppressor gene, Rb. This subsequently allows

E2F transcription factors to induce S phase entry. It has been demonstrated in fibroblasts that loss of Rb allows for cell cycle progression in the absence of ERK1/2 activity

(D'Abaco et al., 2002). Individual silencing of each cyclin D isoform leads to specific phenotypic perturbations, where only cyclin D1 knockouts display neurological defects.

175 This finding was characterized as a cyclin D1-specific effect because mice expressing only cyclin D2, displayed the same phenotype (Ciemerych et al., 2002). Importantly,

deletion of p27 restores normal development (Geng et al., 2001) in cyclin D1 knockouts.

Furthermore, knockin of cyclin D2 into the cyclin D1 locus rescues most of the

abnormalities of the cyclin D1 knockout mouse, demonstrating that the tissue-specific

expression of each isoform mainly dictates their functions (Carthon et al., 2005).

Differential expression of the G1 phase cyclins, D1 and D2, in specific neural progenitor

subsets may contribute to the phenotype observed in both cyclin D1 knockouts and in our

mice. Different patterns of cyclin D1 and D2 expression in neurogenic proliferative zones was first identified in 1996 (Ross et al., 1996). A more recent detailed analysis in

the developing mouse demonstrated that cyclin D2 has a much more restricted pattern of

expression than cyclin D1 throughout neural development (Glickstein et al., 2007).

During early cortical neurogenesis, D2 is mainly expressed in cells of the subventricular zone. Subsequently, at E17.5, its expression becomes more widespread. These cells which make up the subventricular zone at E14.5 are mostly intermediate progenitor cells.

Importantly, we found a severe reduction in mitotic intermediate progenitor cells in the

ERK2 CKO mouse. We hypothesize that loss of ERK2 selectively perturbs cyclin D2 function and/or expression affecting mitotic progression.

The phenotype of the ERK2 CKO and ERK1/2 DKO is similar to other cell cycle component mutants. These data support a role for ERK1/2 in cell cycle control of neural progenitor cells and possibly more importantly, intermediate progenitor cells. To assess a

176 role for ERK1 and ERK2 in this latter cell type, ERK2 can be inactivated through lox/cre

mediated recombination under control of the transcription factor, Tbr2. Additionally,

Tbr2 expressing cells can be isolated and ERK1/2 expression and activity can be individually assessed. Selective inactivation of each isoform in the isolated cells can be achieved through transduction of shRNA or RNAi for each isoform and subsequent

analysis of cell cycle parameters and components.

Because fidelity in proliferation is critical for maintenance of cellular populations and

tissue development, fail-safe mechanisms ensure proper control of mitosis. Hence, the

compensation observed at multiple levels in cdk and cyclin knockouts. Therefore, it is likely that ERK1 and ERK2 are redundant isoforms who are expressed to ensure proper

proliferation of neural progenitor cells. This predicts that ERK1/2 are important for

neurogenesis and that additional levels of regulation may exist to ensure proper

proliferation in their combined absence or mutation.

In consideration of the compensation by other signaling molecules, several alternatively

spliced isoforms of ERK1 are generated in development. In rodents, ERK1b is a 46 kD

ERK isoform which is generated by a 26 amino acid insertion in ERK1. ERK1b has been shown to be activated and behave very similarly to ERK1 and ERK2 (Yung et al., 2000) and therefore may be able to compensate for the loss of each. ERK1c is another alternatively spliced variant which is found in primates but has been demonstrated to have unique functions such as golgi fragmentation during mitosis (Aebersold et al., 2004;

Shaul and Seger, 2006). Despite this latter finding, it remains possible that in the absence

177 of the prototypical ERK isoforms, the alternatively spliced forms may compensate in the normal functioning of ERK1 and ERK2.

Another ERK isoform which may be more relevant to issues of redundancy and

compensation is ERK5. Although ERK5 is similar in function to ERK1/2, it is not classified as a classic ERK MAP kinase due to differences in activation, structure and

identity as the Big MAP kinase (Bmk1). While ERK5 is activated by WNK instead of

Raf, its activity is similarly regulated by growth factor signaling and is important in cell

proliferation and differentiation (Nishimoto and Nishida, 2006). ERK5 shares many

downstream targets with ERK1/2, including phosphorylation of RSK, c-myc and

transcription of fos. Therefore, in the absence of ERK2 and/or ERK1 and 2, signaling

crosstalk to ERK5 can add a level of fidelity to cell cycle signaling. It has also been

recognized that ERK5 is also able to drive Cyclin D1 activity (Mulloy et al., 2003). In

this regard, it is important to note that studies have shown that ERK5 is both necessary

and sufficient for neuronal generation in the cortex (Liu et al., 2006). Knockout studies

have shown that ERK5 directs the survival of neural progenitor cells. Therefore it is

possible that ERK5 compensates for the loss of ERK2 and/or ERK1 in our knockout

models. Further analysis of neural progenitor cell survival and proliferation in a mouse

in which ERK5 is conditionally inactivated along with ERK1 and ERK2 would confirm

this proposal.

178 ERK and Cognitive Function/Mental Retardation

The molecular mechanisms underlying normal cognitive function and their dysfunction in mental retardation are unclear; however, it is well established that structure and plasticity of the cerebral cortex are associated with mental retardation. The ERKs have a long standing role in learning and memory. A substantial body of evidence in organisms ranging from Aplysia to humans implies that the Ras-MAPK pathway plays a role in cognitive function (Mazzucchelli and Brambilla, 2000). Furthermore, disruption of numerous elements both upstream and downstream in MAPK signal transduction leads to decreased intellectual ability and mental retardation in individuals (Thomas and Huganir,

2004).

Figure 4-1: Mutations in upstream and downstream members of the MAP Kinase signaling pathway lead to mental retardation and NCFC Syndromes.

179 Mutation of RSK2 leads to Coffin Lowry Syndrome, and is associated with reductions in total brain volume, with hippocampal and cerebellar volumes most

drastically affected, and significant mental retardation (Kesler et al., 2007; Poirier et al.,

2007). Other forms of X-linked mental retardation are also caused by downstream targets of the ERK signaling cascade as mutation of RSK4 causes X-linked Mental Retardation

(Yntema et al., 1999) and MNK1/2 mutations and aberrant ERK signaling has been associated with Fragile X Syndrome (Kim et al., 2008). Mutations in CREB Binding

Protein (CBP) are causative for Rubenstein Taybi Syndrome. Individuals with this disorder have severe mental retardation and characteristic dysmorphology including craniofacial abnormalities (Hallam and Bourtchouladze, 2006; Gervasini et al., 2007;

Roelfsema and Peters, 2007).

Our data extends these findings by demonstrating that the loss of ERK2 alone produces significant deficits in mouse models of learning. It has also recently been demonstrated that ERK2 hypomorphs display deficits in long term memory (Satoh et al., 2007).

These data provide a clear link between perturbations in ERK signaling and cognitive function. However, due to the limitations of our analysis, mechanistic questions remain.

Are the learning and memory deficits seen in ERK2 CKO and hypomorphic animals (as well as the other models of MAPK-related MR such as Coffin Lowry Syndrome and

Rubenstein Taybi Syndrome) developmental in nature (Do they result from improper connectivity or the reduction in neuron number identified in the CKO?), or do they result from impaired synaptic transmission and plasticity in post-mitotic neurons? Cognitive

180 function is thought to be based on both structural connectivity (cortico-thalamic cortico-

cortico, cortico-amygdalar) as well as neuronal activity via synaptic transmission.

Because ERK1 and 2 are involved in both, it is unclear to what the cognitive deficits are attributed.

Evidence from Down’s Syndrome (DS) patients suggests that this form of mental retardation is due to abnormalities in both structural connectivity and neuronal activity.

In DS patients, cortical size is reduced, and neuronal density is increased. Furthermore, small cerebellum, loss of cholinergic neurons of the basal forebrain, and alterations in dendritic spines have been reported (Branchi et al., 2003). These alterations are similarly found in the numerous trisomy 16 mouse models of DS in which reductions in brain weight, small cerebellum and reductions in specific neuronal populations are found

(Branchi et al., 2003). A subset of the models also revealed functional abnormalities in noradrenergic projections to the cortex and hippocampus. Ts65Dn mice, which have a large triplication of , exhibit learning and behavioral defects but normal morphology of the brain. These mice also have abnormal long-term potentiation (LTP) and depression (LTD) in the dentate gyrus and CA1 regions of the hippocampus. Mice with smaller triplications encompassing the Down’s Syndrome critical region, Ts1Cje mice, also exhibit behavioral and cognitive impairment as well as LTP and LTD. These reductions are generally attributed to decreases in dendritic spine density and spine neck length with an accompanying increase in spine head size. Connectivity was likewise abnormal with decreased inhibitory inputs to the dendritic shafts and spines with more inputs to the necks (Belichenko et al., 2007). Within the cortex, it was documented that

181 subplate neurons and neurons populating the cortical plate were generated simultaneously. This disruption in the temporal dynamics of cortical development was correlated with a significant decrease in thalamo-cortical innervation of the subplate in trisomy 16 mice (Cheng et al., 2004).

In our mouse model, we hypothesize that changes in both structural connectivity and synaptic plasticity play a role here as well. Intra-cortical and thalamocortical connectivity has not been assessed, nor has the excitability of ERK2 null neurons been measured. It is of great interest to assess the extent of thalamo-cortical and cortico- cortico connectivity through lipophilic dye labeling. To circumvent the possibility of abnormal connectivity as a cause for cognitive defects, analysis of mice in which ERK1/2 expression and activity is reduced shortly after development (by use of a CamKII-cre mouse line, currently being generated in collaboration with J. David Sweatt) will be illuminating. Additionally, analysis and comparison of LTP and LTD in CKO and DKOs should be performed. The activation of transcription factors (such as CREB, Elk-1 and c- myc) by luciferase reporter assay and the expression of immediate early genes (c-fos, zif268, BDNF), after glutamatergic stimulation can also be assessed to determine the contribution of the ERKs to these downstream targets.

ERK and 22q11 Deletion Syndrome

We have proposed for the first time a role specifically for ERK1/2 in the etiology of

22q11 Deletion Syndrome (DS). To date, numerous elements of the MAPK signaling

182 cascade have been implicated in the etiology of other Neuro-Cardio-Facial-Cutaneous syndromes. Noonan Syndrome occurs from mutations in K-Ras, SHP-2, C-Raf and

SOS1, while LEOPARD Syndrome results from SHP-2, K-Ras and C-Raf mutations.

Costello Syndrome results from H-Ras mutations, and MEK1/2, B-Raf and H-Ras mutations give rise to Cardio-facio-cutaneous Syndrome. This conglomerate of genotypically-related (MAPK) disorders shares phenotypic pathology which is also displayed by 22q11DS patients. These defects include conotruncal, cardiac and craniofacial anomalies, hypo- or aplastic thymus and thyroid, short stature, microcephaly and developmental delay.

22q11 and psychopathology

One of the most consistent features of 22q11 DS is impaired cognitive function. Easily recognizable in children, the cognitive defects displayed include developmental delays and learning disabilities (verbal – language, speech, reading; and nonverbal - motor skills, math, visuo-spatial organization), mild to moderate mental retardation, behavioral disorders (emotional instability, social withdrawal, anxiety disorder and depression) and psychiatric disorders (schizophrenia and bipolar disorder, schizoaffective disorder, schizotypal personality, Attention Deficit and Hyperactivity Disorder, and Obsessive

Compulsive Disorder). 22q11 DS patients usually have an IQ in the range of 70-85, which is one full standard deviation below the general population.

Children and adolescent 22q11 DS patients display social withdrawal, poor social skills, emotional instability, anxiety, ADHD, and autistic behaviors. As mature adults, their

183 behaviors develop into schizophrenia and bipolar spectrum disorders. The incidence of

22q11 DS patients developing one of these schizoaffective disorders as an adult is a staggering 25%-35%. One in 4000 live births results in a child with 22q11 DS. If 25% of these children develop a schizoaffective disorder, then 1 in every 16,000 people will have a 22q11-related psychiatric disorder. Therefore it is important to understand the basis for these disorders in the context of 22q11 Deletion Syndrome.

Intruigingly, the most effective treatment of the bipolar spectrum disorders in 22q11 DS patients is with valproic acid (valproate) (Vogels et al., 2002). This drug has been shown to activate ERK1/2 and induce neurogenesis in mice (Hao et al., 2004).

Psychiatric core traits of 22q11 DS can be assessed in mice. Paired pulse inhibition (PPI)

is a measure of sensorimotor gating, an established schizophrenia endophenotype. 22q11

DS patients exhibit reduced PPI (Sobin et al., 2005). Mouse models of 22q11 DS have also been tested for PPI and exhibit reductions which have been associated with the haploinsufficiency in Tbx1 (Paylor et al., 2006; Paylor and Lindsay, 2006). Analysis of psychiatric behavior profiles by paired pulse inhibition, contextual fear conditioning and working memory testing in ERK1/2 knockouts as well as the B-Raf/C-Raf and MEK1/2 knockouts would point toward a role for these proteins in cognitive function and disease states.

As numerous mouse models of 22q11 DS have already been generated, it is interesting to

assess ERK activity in both Df1/+ mice (a 21 gene hemizygous deletion of the DiGeorge

184 Critical region, including Tbx1 but not CrkL) and in a mouse model which possesses a deletion encompassing both CrkL and Tbx1. A rather intriguing set of experiments would be to feed pregnant females of the Df1/+ or CrkL/Tbx1 deletion mouse models clinically relevant doses of valproate and subsequently assess the pups for physical defects and psychological deficits.

22q11 physical pathology

One of the major physical impairments associated with 22q11 DS are structural defects of

the heart. They manifest as outflow tract defects such as overriding aortic arch, double

outlet right ventricle, Tetralogy of Fallot among others. These have been found to arise

from the deletion of Tbx1, the major causative gene for the disease pathology.

Importantly, Tbx1 is critical for Fgf8/10 expression. These growth factors are required

for the normal development of the pharyngeal arches and cardiac neural crest which give

rise to craniofacial structures, thymus, thyroid and parathyroid, in addition to the heart.

Recently, we (Newbern et al., submitted) have described the phenotype of B-Raf/C-Raf,

MEK1/2 and ERK1/2 single and double conditional knockout mice in which these genes

were targeted for deletion within the neural crest. They each present with a similar

spectrum of heart, craniofacial, and thymus/parathyroid defects. This represents a body

of work which confirms that the ERK signaling pathway is critical for proper neural crest

development, and its disruption recapitulates the physical characteristics of 22q11

Deletion Syndrome.

185 Due to the cognitive, psychiatric and behavioral impairments associated with 22q11 DS patients, structural abnormalities of the CNS have also been investigated. Within the brain, 22q11 DS is associated with overall reductions in brain volume (Deboer et al.,

2007). Neuroanatomical studies have reported multiple perturbations including changes in both gray and white matter in the temporal lobe (including the amygdala and hippocampus) lissencephaly, cerebral atrophy, cerebellar hypoplasia, ventricular enlargement, partial olfactory tract absence, and “simple gyral pattern” (Bird and

Scambler, 2000). Reductions in region specific cortical volumes, reductions in hippocampal volumes, and polymicrogyria are the most common abnormalities identified

(Bird and Scambler, 2000; Swillen et al., 2000; Campbell et al., 2006; Debbane et al.,

2006; Robin et al., 2006; Schaer et al., 2006; Deboer et al., 2007). Specific reductions in gyrification indices identified by MRI were found to be localized to the frontal and parietal lobes. While the prevalence and penetrance of each malformation is variable, it is generally accepted that brain structure and function is perturbed within these patients.

The mouse models described within this dissertation display reductions in thickness of the frontal and parietal lobes. Furthermore, the mouse models described in Newbern et al. demonstrate small or absent cerebella and olfactory bulbs. These data are the most compelling link to the cortical dysgensis observed in patients as cortical and hippocampal volume was not assessed and loss of gyrification or small gyri cannot be assessed in the mouse.

It has been proposed that reductions in cortical volume and polymicrogyria result from a lack of thalamic innervation. As noted above, Fgf8 expression is induced by Tbx1. It is

186 important to note that Fgf8 signaling controls both intracortical and cortico-thalamic

connectivity in the murine brain (Krubitzer and Kahn, 2003). It has generally been accepted that the cognitive disabilities displayed by afflicted individuals are associated with deficits in visuo-spatial processing and attention, both of which are mediated by cortical networks in the parietal and temporal lobes. It is of significant interest therefore to determine if these connections are intact in the ERK1/2 knockouts (as previously noted) as well as the B-Raf/C-Raf and MEK1/2 knockouts. Measurements of cortical and hippocampal volume would also be illuminating.

ERK and Affective Function/ Mood Disorders

Multiple findings lead one to consider ERK signaling as a component of affective

disorders. First, as mentioned above, the incidence of individuals with 22q11 DS and

schizoaffective disorder is extremely high. Second, there is a significant incidence of reduced ERK1/2 activity and expression (protein and mRNA) in the prefrontal cortex

(Brodmans area 8-10) and hippocampus of depressed suicide subjects. An ERK phosphatase, MKP2, was also elevated in these areas of suicide subjects (Dwivedi et al.,

2001; Dwivedi et al., 2006). And third, neurotrophic factor signaling through ERK1/2

has been shown to regulate the pathophysiology of mood disorders.

Given these facts, it is important to consider the role of ERK1/2 in both the development

and treatment of these illnesses. Chen and colleagues have demonstrated that ERK1 null

mice exhibit signs of mania (behaviors associated with antidepressant and

187 psychostimulant induced action) as they exhibit reduced immobility time in tail

suspension tests and increased swim time in forced swim tests (Engel et al., 2008). These

mice also demonstrate enhanced exploratory behavior in an open field as well as

increased home cage activity on a running wheel. When treated with amphetamines,

these actions were elevated even higher. Chronic treatment with clinically relevant does

of lithium or valproate as well as acute treatment with olanzapine attenuated the

amphetamine-induced hyperactivity, suggesting that the ERK1 null mice exhibit

behaviors that are characteristic of manic phases of bipolar disorder (Einat et al., 2003a;

Einat et al., 2003b; Engel et al., 2008). Independent studies have demonstrated that valproate treatment induces ERK activity and neurite outgrowth in vitro and in vivo

(Yuan et al., 2001; Hao et al., 2004; Rosenberg, 2007).

Others have demonstrated that inhibition of MEK with SL327 increases locomotion in a novel environment and reduces immobility in forced swim tests, further demonstrating that alteration in ERK signaling induces these phenotypes. Additionally, these data demonstrate that developmental perturbations may not be involved in these behaviors

(Einat et al., 2003a). Peripheral injection of MEK inhibitors produce similar helplessness phenotypes and eliminate the responsiveness to antidepressant treatments (Duman et al.,

2007).

Importantly, dysregulation of the FGF signaling cascade has been shown to be involved in Major Depressive Disorder. Specifically, a cohort of human subjects with major depressive disorder had significantly reduced expression levels of FGF1, FGF2, FGFR2

188 and FGFR3 in the dorsolateral prefrontal cortex and/or anterior cingulate cortex as

compared to control subjects (Evans et al., 2004). As the FGFs signal through their

cognate receptors to ERK1/2 these data support the notion that a disruption in ERK signaling underlies mood disorders. Depressive disorders can also be modeled in rats to further study mechanisms and drug therapies. In rats, chronic forced swim tests are used to induce depressive behaviors (less time in open arms, more time in closed arms; decreased locomotor activity, increased grooming, reduced weight). In animals which peformed the chronic forced swim task, phosphorylated ERK2 levels were reduced in both hippocampus and prefrontal cortex (Qi et al., 2006). Amitriptyline (antidepressant

treatment) restores ERK2 phosphorylation in the amygdala and hippocampal regions,

ERK1 phosphorylation in the striatum, and importantly, rescues depressive behaviors

after corticosterone-induced depression (Gourley et al., 2008).

Recently, mice which exhibit a polymorphism in DISC1, disrupted in schizophrenia,

demonstrate reduced ERK phosphorylation and reductions in gray matter volume

associated with Major Depressive Disorder (Hashimoto et al., 2006). These findings

suggest that disruptions in ERK1/2 signaling or improper a loss of ERK1/2 activity could

underlie depressive disorders as well as be causative in the manic and depressive

episodes in bipolar disorder. Taken together, these findings demonstrate that not only do

we need to consider the developmental effects of improper ERK activity on mental

retardation, but we should consider ERK2 and ERK1 hypomorphic effects on other

mental illnesses, especially bipolar disorder and major depressive disorder.

189 An interesting and unexplored line of investigation is the role of the ERK MAP kinases

in autism. Interestingly, two independent autism consortiums have found that

microdeletions and duplications on chromosome 16p11.2 are linked to hereditary forms

of autism (Lucarelli et al., 2003; Kumar et al., 2008; Weiss et al., 2008). Of significance

is the fact that mapk3/ERK1 is located within these chromosomal abnormalities.

Autistism spectrum disorders are generally characterized by language delay, repetitive

actions and deficits in socio-communicative behavior, psychological manifestations

identified in 22q11 DS children as well. These children also have phenotypic

craniofacial defects common to neuro-cardio-facio-cutaneous syndrome patients (Ballif et

al., 2007; Ghebranious et al., 2007; Weiss et al., 2008). Moreover, recent reports suggest

that improper and under-connectivity in thalamocortical and intracortical circuits may be

associated with the disorder (Mizuno et al., 2006; Hughes, 2007; Just et al., 2007). It

may be noteworthy that thinning of the corpus collosum is present in both autism and

22q11DS. Given the possibility of under or improper connectivity due to the loss of

ERK2, it is also worth investigating connectivity and synaptic plasticity during development of ERK1 null mice. This is an intriguing line of investigation as autism and

schizophrenia share common social-behavioral characteristics, and they are thought to

potentially share common etiology (Pinkham et al., 2008). The underlying signaling

pathway leading to both disorders may be the same, namely ERK MAP kinase signal

transduction.

190 Literature Cited

Aebersold DM, Shaul YD, Yung Y, Yarom N, Yao Z, Hanoch T, Seger R (2004) Extracellular signal-regulated kinase 1c (ERK1c), a novel 42-kilodalton ERK, demonstrates unique modes of regulation, localization, and function. Mol Cell Biol 24:10000-10015. Agrawal A, Dillon S, Denning TL, Pulendran B (2006) ERK1-/- mice exhibit Th1 cell polarization and increased susceptibility to experimental autoimmune encephalomyelitis. J Immunol 176:5788-5796. Ballif BC, Hornor SA, Jenkins E, Madan-Khetarpal S, Surti U, Jackson KE, Asamoah A, Brock PL, Gowans GC, Conway RL, Graham JM, Jr., Medne L, Zackai EH, Shaikh TH, Geoghegan J, Selzer RR, Eis PS, Bejjani BA, Shaffer LG (2007) Discovery of a previously unrecognized microdeletion syndrome of 16p11.2- p12.2. Nat Genet 39:1071-1073. Belanger LF, Roy S, Tremblay M, Brott B, Steff AM, Mourad W, Hugo P, Erikson R, Charron J (2003) Mek2 is dispensable for mouse growth and development. Mol Cell Biol 23:4778-4787. Belichenko PV, Kleschevnikov AM, Salehi A, Epstein CJ, Mobley WC (2007) Synaptic and cognitive abnormalities in mouse models of Down syndrome: exploring genotype-phenotype relationships. J Comp Neurol 504:329-345. Bird LM, Scambler P (2000) Cortical dysgenesis in 2 patients with chromosome 22q11 deletion. Clin Genet 58:64-68. Bost F, Aouadi M, Caron L, Binetruy B (2005a) The role of MAPKs in adipocyte differentiation and obesity. Biochimie 87:51-56. Bost F, Aouadi M, Caron L, Even P, Belmonte N, Prot M, Dani C, Hofman P, Pages G, Pouyssegur J, Le Marchand-Brustel Y, Binetruy B (2005b) The extracellular signal-regulated kinase isoform ERK1 is specifically required for in vitro and in vivo adipogenesis. Diabetes 54:402-411. Boulton TG, Cobb MH (1991) Identification of multiple extracellular signal-regulated kinases (ERKs) with antipeptide antibodies. Cell Regul 2:357-371. Boulton TG, Yancopoulos GD, Gregory JS, Slaughter C, Moomaw C, Hsu J, Cobb MH (1990) An insulin-stimulated protein kinase similar to yeast kinases involved in cell cycle control. Science 249:64-67. Branchi I, Bichler Z, Berger-Sweeney J, Ricceri L (2003) Animal models of mental retardation: from gene to cognitive function. Neurosci Biobehav Rev 27:141-153. Campbell LE, Daly E, Toal F, Stevens A, Azuma R, Catani M, Ng V, van Amelsvoort T, Chitnis X, Cutter W, Murphy DG, Murphy KC (2006) Brain and behaviour in children with 22q11.2 deletion syndrome: a volumetric and voxel-based morphometry MRI study. Brain 129:1218-1228. Carthon BC, Neumann CA, Das M, Pawlyk B, Li T, Geng Y, Sicinski P (2005) Genetic replacement of cyclin D1 function in mouse development by cyclin D2. Mol Cell Biol 25:1081-1088. Cheng A, Haydar TF, Yarowsky PJ, Krueger BK (2004) Concurrent generation of subplate and cortical plate neurons in developing trisomy 16 mouse cortex. Dev Neurosci 26:255-265.

191 Ciemerych MA, Kenney AM, Sicinska E, Kalaszczynska I, Bronson RT, Rowitch DH, Gardner H, Sicinski P (2002) Development of mice expressing a single D-type cyclin. Genes Dev 16:3277-3289. Cobb MH, Boulton TG, Robbins DJ (1991) Extracellular signal-regulated kinases: ERKs in progress. Cell Regul 2:965-978. D'Abaco GM, Hooper S, Paterson H, Marshall CJ (2002) Loss of Rb overrides the requirement for ERK activity for cell proliferation. J Cell Sci 115:4607-4616. Debbane M, Schaer M, Farhoumand R, Glaser B, Eliez S (2006) Hippocampal volume reduction in 22q11.2 deletion syndrome. Neuropsychologia 44:2360-2365. Deboer T, Wu Z, Lee A, Simon TJ (2007) Hippocampal volume reduction in children with chromosome 22q11.2 deletion syndrome is associated with cognitive impairment. Behav Brain Funct 3:54. Duman CH, Schlesinger L, Kodama M, Russell DS, Duman RS (2007) A role for MAP kinase signaling in behavioral models of depression and antidepressant treatment. Biol Psychiatry 61:661-670. Dwivedi Y, Rizavi HS, Conley RR, Pandey GN (2006) ERK MAP kinase signaling in post-mortem brain of suicide subjects: differential regulation of upstream Raf kinases Raf-1 and B-Raf. Mol Psychiatry 11:86-98. Dwivedi Y, Rizavi HS, Roberts RC, Conley RC, Tamminga CA, Pandey GN (2001) Reduced activation and expression of ERK1/2 MAP kinase in the post-mortem brain of depressed suicide subjects. J Neurochem 77:916-928. Einat H, Manji HK, Gould TD, Du J, Chen G (2003a) Possible involvement of the ERK signaling cascade in bipolar disorder: behavioral leads from the study of mutant mice. Drug News Perspect 16:453-463. Einat H, Yuan P, Gould TD, Li J, Du J, Zhang L, Manji HK, Chen G (2003b) The role of the extracellular signal-regulated kinase signaling pathway in mood modulation. J Neurosci 23:7311-7316. Engel SR, Creson TK, Hao Y, Shen Y, Maeng S, Nekrasova T, Landreth GE, Manji HK, Chen G (2008) The extracellular signal-regulated kinase pathway contributes to the control of behavioral excitement. Mol Psychiatry. English JD, Sweatt JD (1996) Activation of p42 mitogen-activated protein kinase in hippocampal long term potentiation. J Biol Chem 271:24329-24332. English JD, Sweatt JD (1997) A requirement for the mitogen-activated protein kinase cascade in hippocampal long term potentiation. J Biol Chem 272:19103-19106. Evans SJ, Choudary PV, Neal CR, Li JZ, Vawter MP, Tomita H, Lopez JF, Thompson RC, Meng F, Stead JD, Walsh DM, Myers RM, Bunney WE, Watson SJ, Jones EG, Akil H (2004) Dysregulation of the fibroblast growth factor system in major depression. Proc Natl Acad Sci U S A 101:15506-15511. Fremin C, Ezan F, Boisselier P, Bessard A, Pages G, Pouyssegur J, Baffet G (2007) ERK2 but not ERK1 plays a key role in hepatocyte replication: an RNAi- mediated ERK2 knockdown approach in wild-type and ERK1 null hepatocytes. Hepatology 45:1035-1045. Geng Y, Yu Q, Sicinska E, Das M, Bronson RT, Sicinski P (2001) Deletion of the p27Kip1 gene restores normal development in cyclin D1-deficient mice. Proc Natl Acad Sci U S A 98:194-199.

192 Gervasini C, Castronovo P, Bentivegna A, Mottadelli F, Faravelli F, Giovannucci-Uzielli ML, Pessagno A, Lucci-Cordisco E, Pinto AM, Salviati L, Selicorni A, Tenconi R, Neri G, Larizza L (2007) High frequency of mosaic CREBBP deletions in Rubinstein-Taybi syndrome patients and mapping of somatic and germ-line breakpoints. Genomics 90:567-573. Ghebranious N, Giampietro PF, Wesbrook FP, Rezkalla SH (2007) A novel microdeletion at 16p11.2 harbors candidate genes for aortic valve development, seizure disorder, and mild mental retardation. Am J Med Genet A 143A:1462- 1471. Giroux S, Tremblay M, Bernard D, Cardin-Girard JF, Aubry S, Larouche L, Rousseau S, Huot J, Landry J, Jeannotte L, Charron J (1999) Embryonic death of Mek1- deficient mice reveals a role for this kinase in angiogenesis in the labyrinthine region of the placenta. Curr Biol 9:369-372. Glickstein SB, Alexander S, Ross ME (2007) Differences in cyclin D2 and D1 protein expression distinguish forebrain progenitor subsets. Cereb Cortex 17:632-642. Gourley SL, Wu FJ, Kiraly DD, Ploski JE, Kedves AT, Duman RS, Taylor JR (2008) Regionally specific regulation of ERK MAP kinase in a model of antidepressant- sensitive chronic depression. Biol Psychiatry 63:353-359. Hallam TM, Bourtchouladze R (2006) Rubinstein-Taybi syndrome: molecular findings and therapeutic approaches to improve cognitive dysfunction. Cell Mol Life Sci 63:1725-1735. Hao Y, Creson T, Zhang L, Li P, Du F, Yuan P, Gould TD, Manji HK, Chen G (2004) Mood stabilizer valproate promotes ERK pathway-dependent cortical neuronal growth and neurogenesis. J Neurosci 24:6590-6599. Hashimoto R, Numakawa T, Ohnishi T, Kumamaru E, Yagasaki Y, Ishimoto T, Mori T, Nemoto K, Adachi N, Izumi A, Chiba S, Noguchi H, Suzuki T, Iwata N, Ozaki N, Taguchi T, Kamiya A, Kosuga A, Tatsumi M, Kamijima K, Weinberger DR, Sawa A, Kunugi H (2006) Impact of the DISC1 Ser704Cys polymorphism on risk for major depression, brain morphology and ERK signaling. Hum Mol Genet 15:3024-3033. Hatano N, Mori Y, Oh-hora M, Kosugi A, Fujikawa T, Nakai N, Niwa H, Miyazaki J, Hamaoka T, Ogata M (2003) Essential role for ERK2 mitogen-activated protein kinase in placental development. Genes Cells 8:847-856. Hughes JR (2007) Autism: the first firm finding = underconnectivity? Epilepsy Behav 11:20-24. Just MA, Cherkassky VL, Keller TA, Kana RK, Minshew NJ (2007) Functional and anatomical cortical underconnectivity in autism: evidence from an FMRI study of an executive function task and corpus callosum morphometry. Cereb Cortex 17:951-961. Kesler SR, Simensen RJ, Voeller K, Abidi F, Stevenson RE, Schwartz CE, Reiss AL (2007) Altered neurodevelopment associated with mutations of RSK2: a morphometric MRI study of Coffin-Lowry syndrome. Neurogenetics 8:143-147. Kim SH, Markham JA, Weiler IJ, Greenough WT (2008) Aberrant early-phase ERK inactivation impedes neuronal function in fragile X syndrome. Proc Natl Acad Sci U S A 105:4429-4434.

193 Knoepfler PS, Cheng PF, Eisenman RN (2002) N-myc is essential during neurogenesis for the rapid expansion of progenitor cell populations and the inhibition of neuronal differentiation. Genes Dev 16:2699-2712. Kozar K, Sicinski P (2005) Cell cycle progression without cyclin D-CDK4 and cyclin D- CDK6 complexes. Cell Cycle 4:388-391. Kozar K, Ciemerych MA, Rebel VI, Shigematsu H, Zagozdzon A, Sicinska E, Geng Y, Yu Q, Bhattacharya S, Bronson RT, Akashi K, Sicinski P (2004) Mouse development and cell proliferation in the absence of D-cyclins. Cell 118:477-491. Krubitzer L, Kahn DM (2003) Nature versus nurture revisited: an old idea with a new twist. Prog Neurobiol 70:33-52. Kumar RA, KaraMohamed S, Sudi J, Conrad DF, Brune C, Badner JA, Gilliam TC, Nowak NJ, Cook EH, Jr., Dobyns WB, Christian SL (2008) Recurrent 16p11.2 microdeletions in autism. Hum Mol Genet 17:628-638. Lefloch R, Pouyssegur J, Lenormand P (2008) Single and combined silencing of ERK1 and ERK2 reveals their positive contribution to growth signaling depending on their expression levels. Mol Cell Biol 28:511-527. Liu L, Cundiff P, Abel G, Wang Y, Faigle R, Sakagami H, Xu M, Xia Z (2006) Extracellular signal-regulated kinase (ERK) 5 is necessary and sufficient to specify cortical neuronal fate. Proc Natl Acad Sci U S A 103:9697-9702. Liu X, Yan S, Zhou T, Terada Y, Erikson RL (2004) The MAP kinase pathway is required for entry into mitosis and cell survival. Oncogene 23:763-776. Lucarelli P, Palminiello S, Saccucci P, Bottini N, De Luca D, Elia M, Fiumara A, Curatolo P (2003) Association study of autistic disorder and chromosome 16p. Am J Med Genet A 119:242-246. Malumbres M, Sotillo R, Santamaria D, Galan J, Cerezo A, Ortega S, Dubus P, Barbacid M (2004) Mammalian cells cycle without the D-type cyclin-dependent kinases Cdk4 and Cdk6. Cell 118:493-504. Mazzucchelli C, Brambilla R (2000) Ras-related and MAPK signalling in neuronal plasticity and memory formation. Cell Mol Life Sci 57:604-611. Mazzucchelli C, Vantaggiato C, Ciamei A, Fasano S, Pakhotin P, Krezel W, Welzl H, Wolfer DP, Pages G, Valverde O, Marowsky A, Porrazzo A, Orban PC, Maldonado R, Ehrengruber MU, Cestari V, Lipp HP, Chapman PF, Pouyssegur J, Brambilla R (2002) Knockout of ERK1 MAP kinase enhances synaptic plasticity in the striatum and facilitates striatal-mediated learning and memory. Neuron 34:807-820. Meloche S, Vella FD, Voisin L, Ang SL, Saba-El-Leil M (2004) Erk2 signaling and early embryo stem cell self-renewal. Cell Cycle 3:241-243. Miller FD, Gauthier AS (2007) Timing is everything: making neurons versus glia in the developing cortex. Neuron 54:357-369. Mizuno A, Villalobos ME, Davies MM, Dahl BC, Muller RA (2006) Partially enhanced thalamocortical functional connectivity in autism. Brain Res 1104:160-174. Mulloy R, Salinas S, Philips A, Hipskind RA (2003) Activation of cyclin D1 expression by the ERK5 cascade. Oncogene 22:5387-5398. Nekrasova T, Shive C, Gao Y, Kawamura K, Guardia R, Landreth G, Forsthuber TG (2005) ERK1-deficient mice show normal T cell effector function and are highly

194 susceptible to experimental autoimmune encephalomyelitis. J Immunol 175:2374- 2380. Nishimoto S, Nishida E (2006) MAPK signalling: ERK5 versus ERK1/2. EMBO Rep 7:782-786. Ortiz J, Harris HW, Guitart X, Terwilliger RZ, Haycock JW, Nestler EJ (1995) Extracellular signal-regulated protein kinases (ERKs) and ERK kinase (MEK) in brain: regional distribution and regulation by chronic morphine. J Neurosci 15:1285-1297. Pages G, Guerin S, Grall D, Bonino F, Smith A, Anjuere F, Auberger P, Pouyssegur J (1999) Defective thymocyte maturation in p44 MAP kinase (Erk 1) knockout mice. Science 286:1374-1377. Paylor R, Lindsay E (2006) Mouse models of 22q11 deletion syndrome. Biol Psychiatry 59:1172-1179. Paylor R, Glaser B, Mupo A, Ataliotis P, Spencer C, Sobotka A, Sparks C, Choi CH, Oghalai J, Curran S, Murphy KC, Monks S, Williams N, O'Donovan MC, Owen MJ, Scambler PJ, Lindsay E (2006) Tbx1 haploinsufficiency is linked to behavioral disorders in mice and humans: implications for 22q11 deletion syndrome. Proc Natl Acad Sci U S A 103:7729-7734. Pinkham AE, Hopfinger JB, Pelphrey KA, Piven J, Penn DL (2008) Neural bases for impaired social cognition in schizophrenia and autism spectrum disorders. Schizophr Res 99:164-175. Poirier R, Jacquot S, Vaillend C, Soutthiphong AA, Libbey M, Davis S, Laroche S, Hanauer A, Welzl H, Lipp HP, Wolfer DP (2007) Deletion of the Coffin-Lowry syndrome gene Rsk2 in mice is associated with impaired spatial learning and reduced control of exploratory behavior. Behav Genet 37:31-50. Qi X, Lin W, Li J, Pan Y, Wang W (2006) The depressive-like behaviors are correlated with decreased phosphorylation of mitogen-activated protein kinases in rat brain following chronic forced swim stress. Behav Brain Res 175:233-240. Robin NH, Taylor CJ, McDonald-McGinn DM, Zackai EH, Bingham P, Collins KJ, Earl D, Gill D, Granata T, Guerrini R, Katz N, Kimonis V, Lin JP, Lynch DR, Mohammed SN, Massey RF, McDonald M, Rogers RC, Splitt M, Stevens CA, Tischkowitz MD, Stoodley N, Leventer RJ, Pilz DT, Dobyns WB (2006) Polymicrogyria and deletion 22q11.2 syndrome: window to the etiology of a common cortical malformation. Am J Med Genet A 140:2416-2425. Roelfsema JH, Peters DJ (2007) Rubinstein-Taybi syndrome: clinical and molecular overview. Expert Rev Mol Med 9:1-16. Rosenberg G (2007) The mechanisms of action of valproate in neuropsychiatric disorders: can we see the forest for the trees? Cell Mol Life Sci 64:2090-2103. Ross ME, Carter ML, Lee JH (1996) MN20, a D2 cyclin, is transiently expressed in selected neural populations during embryogenesis. J Neurosci 16:210-219. Saba-El-Leil MK, Vella FD, Vernay B, Voisin L, Chen L, Labrecque N, Ang SL, Meloche S (2003) An essential function of the mitogen-activated protein kinase Erk2 in mouse trophoblast development. EMBO Rep 4:964-968. Satoh Y, Endo S, Ikeda T, Yamada K, Ito M, Kuroki M, Hiramoto T, Imamura O, Kobayashi Y, Watanabe Y, Itohara S, Takishima K (2007) Extracellular signal- regulated kinase 2 (ERK2) knockdown mice show deficits in long-term memory;

195 ERK2 has a specific function in learning and memory. J Neurosci 27:10765- 10776. Schaer M, Schmitt JE, Glaser B, Lazeyras F, Delavelle J, Eliez S (2006) Abnormal patterns of cortical gyrification in velo-cardio-facial syndrome (deletion 22q11.2): an MRI study. Psychiatry Res 146:1-11. Selcher JC, Nekrasova T, Paylor R, Landreth GE, Sweatt JD (2001) Mice lacking the ERK1 isoform of MAP kinase are unimpaired in emotional learning. Learn Mem 8:11-19. Shaul YD, Seger R (2006) ERK1c regulates Golgi fragmentation during mitosis. J Cell Biol 172:885-897. Shaul YD, Seger R (2007) The MEK/ERK cascade: from signaling specificity to diverse functions. Biochim Biophys Acta 1773:1213-1226. Sherr CJ, Roberts JM (2004) Living with or without cyclins and cyclin-dependent kinases. Genes Dev 18:2699-2711. Skarpen E, Flinder LI, Rosseland CM, Orstavik S, Wierod L, Oksvold MP, Skalhegg BS, Huitfeldt HS (2008) MEK1 and MEK2 regulate distinct functions by sorting ERK2 to different intracellular compartments. Faseb J 22:466-476. Sobin C, Kiley-Brabeck K, Karayiorgou M (2005) Lower prepulse inhibition in children with the 22q11 deletion syndrome. Am J Psychiatry 162:1090-1099. Sweatt JD (2001) The neuronal MAP kinase cascade: a biochemical signal integration system subserving synaptic plasticity and memory. J Neurochem 76:1-10. Swillen A, Vogels A, Devriendt K, Fryns JP (2000) Chromosome 22q11 deletion syndrome: update and review of the clinical features, cognitive-behavioral spectrum, and psychiatric complications. Am J Med Genet 97:128-135. Tarui T, Takahashi T, Nowakowski RS, Hayes NL, Bhide PG, Caviness VS (2005) Overexpression of p27 Kip 1, probability of cell cycle exit, and laminar destination of neocortical neurons. Cereb Cortex 15:1343-1355. Thomas GM, Huganir RL (2004) MAPK cascade signalling and synaptic plasticity. Nat Rev Neurosci 5:173-183. Thomas KL, Hunt SP (1993) The regional distribution of extracellularly regulated kinase- 1 and -2 messenger RNA in the adult rat central nervous system. Neuroscience 56:741-757. Ussar S, Voss T (2004) MEK1 and MEK2, different regulators of the G1/S transition. J Biol Chem 279:43861-43869. Vantaggiato C, Formentini I, Bondanza A, Bonini C, Naldini L, Brambilla R (2006) ERK1 and ERK2 mitogen-activated protein kinases affect Ras-dependent cell signaling differentially. J Biol Chem 5:14. Vogels A, Verhoeven WM, Tuinier S, DeVriendt K, Swillen A, Curfs LM, Frijns JP (2002) The psychopathological phenotype of velo-cardio-facial syndrome. Ann Genet 45:89-95. Weiss LA, Shen Y, Korn JM, Arking DE, Miller DT, Fossdal R, Saemundsen E, Stefansson H, Ferreira MA, Green T, Platt OS, Ruderfer DM, Walsh CA, Altshuler D, Chakravarti A, Tanzi RE, Stefansson K, Santangelo SL, Gusella JF, Sklar P, Wu BL, Daly MJ (2008) Association between microdeletion and microduplication at 16p11.2 and autism. N Engl J Med 358:667-675.

196 Yao Y, Li W, Wu J, Germann UA, Su MS, Kuida K, Boucher DM (2003) Extracellular signal-regulated kinase 2 is necessary for mesoderm differentiation. Proc Natl Acad Sci U S A 100:12759-12764. Yntema HG, van den Helm B, Kissing J, van Duijnhoven G, Poppelaars F, Chelly J, Moraine C, Fryns JP, Hamel BC, Heilbronner H, Pander HJ, Brunner HG, Ropers HH, Cremers FP, van Bokhoven H (1999) A novel ribosomal S6-kinase (RSK4; RPS6KA6) is commonly deleted in patients with complex X-linked mental retardation. Genomics 62:332-343. Yuan PX, Huang LD, Jiang YM, Gutkind JS, Manji HK, Chen G (2001) The mood stabilizer valproic acid activates mitogen-activated protein kinases and promotes neurite growth. J Biol Chem 276:31674-31683. Yung Y, Yao Z, Hanoch T, Seger R (2000) ERK1b, a 46-kDa ERK isoform that is differentially regulated by MEK. J Biol Chem 275:15799-15808.

197 LITERATURE CITED

Adams JP, Sweatt JD (2002) Molecular psychology: roles for the ERK MAP kinase cascade in memory. Annu Rev Pharmacol Toxicol 42:135-163. Aebersold DM, Shaul YD, Yung Y, Yarom N, Yao Z, Hanoch T, Seger R (2004) Extracellular signal-regulated kinase 1c (ERK1c), a novel 42-kilodalton ERK, demonstrates unique modes of regulation, localization, and function. Mol Cell Biol 24:10000-10015. Agrawal A, Dillon S, Denning TL, Pulendran B (2006) ERK1-/- mice exhibit Th1 cell polarization and increased susceptibility to experimental autoimmune encephalomyelitis. J Immunol 176:5788-5796. Aktas H, Cai H, Cooper GM (1997) Ras links growth factor signaling to the cell cycle machinery via regulation of cyclin D1 and the Cdk inhibitor p27KIP1. Mol Cell Biol 17:3850-3857. Anthony TE, Klein C, Fishell G, Heintz N (2004) Radial glia serve as neuronal progenitors in all regions of the central nervous system. Neuron 41:881-890. Aouadi M, Binetruy B, Caron L, Le Marchand-Brustel Y, Bost F (2006) Role of MAPKs in development and differentiation: lessons from knockout mice. Biochimie 88:1091-1098. Arendt T, Gartner U, Seeger G, Barmashenko G, Palm K, Mittmann T, Yan L, Hummeke M, Behrbohm J, Bruckner MK, Holzer M, Wahle P, Heumann R (2004) Neuronal activation of Ras regulates synaptic connectivity. Eur J Neurosci 19:2953-2966. Atkins CM, Selcher JC, Petraitis JJ, Trzaskos JM, Sweatt JD (1998) The MAPK cascade is required for mammalian associative learning. Nat Neurosci 1:602-609. Bai G, Sheng N, Xie Z, Bian W, Yokota Y, Benezra R, Kageyama R, Guillemot F, Jing N (2007) Id sustains Hes1 expression to inhibit precocious neurogenesis by releasing negative autoregulation of Hes1. Dev Cell 13:283-297. Bain G, Cravatt CB, Loomans C, Alberola-Ila J, Hedrick SM, Murre C (2001) Regulation of the helix-loop-helix proteins, E2A and Id3, by the Ras-ERK MAPK cascade. Nat Immunol 2:165-171. Ballif BC, Hornor SA, Jenkins E, Madan-Khetarpal S, Surti U, Jackson KE, Asamoah A, Brock PL, Gowans GC, Conway RL, Graham JM, Jr., Medne L, Zackai EH, Shaikh TH, Geoghegan J, Selzer RR, Eis PS, Bejjani BA, Shaffer LG (2007) Discovery of a previously unrecognized microdeletion syndrome of 16p11.2- p12.2. Nat Genet 39:1071-1073. Barnabe-Heider F, Miller FD (2003) Endogenously produced neurotrophins regulate survival and differentiation of cortical progenitors via distinct signaling pathways. J Neurosci 23:5149-5160. Bartkowska K, Paquin A, Gauthier AS, Kaplan DR, Miller FD (2007) Trk signaling regulates neural precursor cell proliferation and differentiation during cortical development. Development 134:4369-4380. Belanger LF, Roy S, Tremblay M, Brott B, Steff AM, Mourad W, Hugo P, Erikson R, Charron J (2003) Mek2 is dispensable for mouse growth and development. Mol Cell Biol 23:4778-4787.

198 Belichenko PV, Kleschevnikov AM, Salehi A, Epstein CJ, Mobley WC (2007) Synaptic and cognitive abnormalities in mouse models of Down syndrome: exploring genotype-phenotype relationships. J Comp Neurol 504:329-345. Berman DE, Hazvi S, Rosenblum K, Seger R, Dudai Y (1998) Specific and differential activation of mitogen-activated protein kinase cascades by unfamiliar taste in the insular cortex of the behaving rat. J Neurosci 18:10037-10044. Ben-Shachar S, Ou Z, Shaw CA, Belmont JW, Patel MS, Hummel M, Amato S, Tartaglia N, Berg J, Sutton VR, Lalani SR, Chinault AC, Cheung SW, Lupski JR, Patel A (2008) 22q11.2 distal deletion: a recurrent genomic disorder distinct from DiGeorge syndrome and velocardiofacial syndrome. Am J Hum Genet 82:214- 221. Bentires-Alj M, Kontaridis MI, Neel BG (2006) Stops along the RAS pathway in human genetic disease. Nat Med 12:283-285. Berger-Sweeney J, Hohmann CF (1997) Behavioral consequences of abnormal cortical development: insights into developmental disabilities. Behav Brain Res 86:121- 142. Bird LM, Scambler P (2000) Cortical dysgenesis in 2 patients with chromosome 22q11 deletion. Clin Genet 58:64-68. Blum S, Moore AN, Adams F, Dash PK (1999) A mitogen-activated protein kinase cascade in the CA1/CA2 subfield of the dorsal hippocampus is essential for long- term spatial memory. J Neurosci 19:3535-3544. Bond J, Woods CG (2006) Cytoskeletal genes regulating brain size. Curr Opin Cell Biol 18:95-101. Bonni A, Sun Y, Nadal-Vicens M, Bhatt A, Frank DA, Rozovsky I, Stahl N, Yancopoulos GD, Greenberg ME (1997) Regulation of gliogenesis in the central nervous system by the JAK-STAT signaling pathway. Science 278:477-483. Bost F, Aouadi M, Caron L, Binetruy B (2005a) The role of MAPKs in adipocyte differentiation and obesity. Biochimie 87:51-56. Bost F, Aouadi M, Caron L, Even P, Belmonte N, Prot M, Dani C, Hofman P, Pages G, Pouyssegur J, Le Marchand-Brustel Y, Binetruy B (2005b) The extracellular signal-regulated kinase isoform ERK1 is specifically required for in vitro and in vivo adipogenesis. Diabetes 54:402-411. Bouchard C, Dittrich O, Kiermaier A, Dohmann K, Menkel A, Eilers M, Luscher B (2001) Regulation of cyclin D2 gene expression by the Myc/Max/Mad network: Myc-dependent TRRAP recruitment and histone acetylation at the cyclin D2 promoter. Genes Dev 15:2042-2047. Bouchard C, Thieke K, Maier A, Saffrich R, Hanley-Hyde J, Ansorge W, Reed S, Sicinski P, Bartek J, Eilers M (1999) Direct induction of cyclin D2 by Myc contributes to cell cycle progression and sequestration of p27. Embo J 18:5321- 5333. Boulton TG, Cobb MH (1991) Identification of multiple extracellular signal-regulated kinases (ERKs) with antipeptide antibodies. Cell Regul 2:357-371. Boulton TG, Yancopoulos GD, Gregory JS, Slaughter C, Moomaw C, Hsu J, Cobb MH (1990) An insulin-stimulated protein kinase similar to yeast kinases involved in cell cycle control. Science 249:64-67.

199 Bozon B, Kelly A, Josselyn SA, Silva AJ, Davis S, Laroche S (2003) MAPK, CREB and zif268 are all required for the consolidation of recognition memory. Philos Trans R Soc Lond B Biol Sci 358:805-814. Branchi I, Bichler Z, Berger-Sweeney J, Ricceri L (2003) Animal models of mental retardation: from gene to cognitive function. Neurosci Biobehav Rev 27:141-153. Calella AM, Nerlov C, Lopez RG, Sciarretta C, von Bohlen Und Halbach O, Bereshchenko O, Minichiello L (2007) Neurotrophin/Trk receptor signaling mediates C/EBPalpha, -beta and NeuroD recruitment to immediate-early gene promoters in neuronal cells and requires C/EBPs to induce immediate-early gene transcription. Neural Develop 2:4. Callaghan DA, Dong L, Callaghan SM, Hou YX, Dagnino L, Slack RS (1999) Neural precursor cells differentiating in the absence of Rb exhibit delayed terminal mitosis and deregulated E2F 1 and 3 activity. Dev Biol 207:257-270. Camarero G, Tyrsin OY, Xiang C, Pfeiffer V, Pleiser S, Wiese S, Gotz R, Rapp UR (2006) Cortical migration defects in mice expressing A-RAF from the B-RAF locus. Mol Cell Biol 26:7103-7115. Cameron HA, Hazel TG, McKay RD (1998) Regulation of neurogenesis by growth factors and neurotransmitters. J Neurobiol 36:287-306. Campbell LE, Daly E, Toal F, Stevens A, Azuma R, Catani M, Ng V, van Amelsvoort T, Chitnis X, Cutter W, Murphy DG, Murphy KC (2006) Brain and behaviour in children with 22q11.2 deletion syndrome: a volumetric and voxel-based morphometry MRI study. Brain 129:1218-1228. Caviness VS, Jr., Takahashi T, Nowakowski RS (1995) Numbers, time and neocortical neuronogenesis: a general developmental and evolutionary model. Trends Neurosci 18:379-383. Caviness VS, Jr., Goto T, Tarui T, Takahashi T, Bhide PG, Nowakowski RS (2003) Cell output, cell cycle duration and neuronal specification: a model of integrated mechanisms of the neocortical proliferative process. Cereb Cortex 13:592-598. Chambard JC, Lefloch R, Pouyssegur J, Lenormand P (2007) ERK implication in cell cycle regulation. Biochim Biophys Acta 1773:1299-1310. Chelly J, Khelfaoui M, Francis F, Cherif B, Bienvenu T (2006) Genetics and pathophysiology of mental retardation. Eur J Hum Genet 14:701-713. Chen AP, Ohno M, Giese KP, Kuhn R, Chen RL, Silva AJ (2006) Forebrain-specific knockout of B-raf kinase leads to deficits in hippocampal long-term potentiation, learning, and memory. J Neurosci Res 83:28-38. Chen Z, Duan RS, Zhu Y, Folkesson R, Albanese C, Winblad B, Zhu J (2005) Increased cyclin E expression may obviate the role of cyclin D1 during brain development in cyclin D1 knockout mice. J Neurochem 92:1281-1284. Cheng A, Haydar TF, Yarowsky PJ, Krueger BK (2004) Concurrent generation of subplate and cortical plate neurons in developing trisomy 16 mouse cortex. Dev Neurosci 26:255-265. Ciemerych MA, Kenney AM, Sicinska E, Kalaszczynska I, Bronson RT, Rowitch DH, Gardner H, Sicinski P (2002) Development of mice expressing a single D-type cyclin. Genes Dev 16:3277-3289. Cobb MH, Boulton TG, Robbins DJ (1991) Extracellular signal-regulated kinases: ERKs in progress. Cell Regul 2:965-978.

200 Coller HA, Grandori C, Tamayo P, Colbert T, Lander ES, Eisenman RN, Golub TR (2000) Expression analysis with oligonucleotide microarrays reveals that MYC regulates genes involved in growth, cell cycle, signaling, and adhesion. Proc Natl Acad Sci U S A 97:3260-3265. Coogan AN, O'Leary DM, O'Connor JJ (1999) P42/44 MAP kinase inhibitor PD98059 attenuates multiple forms of synaptic plasticity in rat dentate gyrus in vitro. J Neurophysiol 81:103-110. Corson LB, Yamanaka Y, Lai KM, Rossant J (2003) Spatial and temporal patterns of ERK signaling during mouse embryogenesis. Development 130:4527-4537. Crawley JN (1999) Behavioral phenotyping of transgenic and knockout mice: experimental design and evaluation of general health, sensory functions, motor abilities, and specific behavioral tests. Brain Res 835:18-26. Crome L (1972) Morphological aspects of mental retardation. Proc R Soc Med 65:587- 588. D'Abaco GM, Hooper S, Paterson H, Marshall CJ (2002) Loss of Rb overrides the requirement for ERK activity for cell proliferation. J Cell Sci 115:4607-4616. Debbane M, Schaer M, Farhoumand R, Glaser B, Eliez S (2006) Hippocampal volume reduction in 22q11.2 deletion syndrome. Neuropsychologia 44:2360-2365. Deboer T, Wu Z, Lee A, Simon TJ (2007) Hippocampal volume reduction in children with chromosome 22q11.2 deletion syndrome is associated with cognitive impairment. Behav Brain Funct 3:54. Deed RW, Hara E, Atherton GT, Peters G, Norton JD (1997) Regulation of Id3 cell cycle function by Cdk-2-dependent phosphorylation. Mol Cell Biol 17:6815-6821. Dehay C, Kennedy H (2007) Cell-cycle control and cortical development. Nat Rev Neurosci 8:438-450. Delalle I, Takahashi T, Nowakowski RS, Tsai LH, Caviness VS, Jr. (1999) Cyclin E-p27 opposition and regulation of the G1 phase of the cell cycle in the murine neocortical PVE: a quantitative analysis of mRNA in situ hybridization. Cereb Cortex 9:824-832. Desai AR, McConnell SK (2000) Progressive restriction in fate potential by neural progenitors during cerebral cortical development. Development 127:2863-2872. Dey A, She H, Kim L, Boruch A, Guris DL, Carlberg K, Sebti SM, Woodley DT, Imamoto A, Li W (2000) Colony-stimulating factor-1 receptor utilizes multiple signaling pathways to induce cyclin D2 expression. Mol Biol Cell 11:3835-3848. Di Cristo G, Berardi N, Cancedda L, Pizzorusso T, Putignano E, Ratto GM, Maffei L (2001) Requirement of ERK activation for visual cortical plasticity. Science 292:2337-2340. Dillon S, Agrawal A, Van Dyke T, Landreth G, McCauley L, Koh A, Maliszewski C, Akira S, Pulendran B (2004) A Toll-like receptor 2 ligand stimulates Th2 responses in vivo, via induction of extracellular signal-regulated kinase mitogen- activated protein kinase and c-Fos in dendritic cells. J Immunol 172:4733-4743. Dono R (2003) Fibroblast growth factors as regulators of central nervous system development and function. Am J Physiol Regul Integr Comp Physiol 284:R867- 881.

201 Duman CH, Schlesinger L, Kodama M, Russell DS, Duman RS (2007) A role for MAP kinase signaling in behavioral models of depression and antidepressant treatment. Biol Psychiatry 61:661-670. Dwivedi Y, Rizavi HS, Conley RR, Pandey GN (2006) ERK MAP kinase signaling in post-mortem brain of suicide subjects: differential regulation of upstream Raf kinases Raf-1 and B-Raf. Mol Psychiatry 11:86-98. Dwivedi Y, Rizavi HS, Roberts RC, Conley RC, Tamminga CA, Pandey GN (2001) Reduced activation and expression of ERK1/2 MAP kinase in the post-mortem brain of depressed suicide subjects. J Neurochem 77:916-928. Einat H, Manji HK, Gould TD, Du J, Chen G (2003a) Possible involvement of the ERK signaling cascade in bipolar disorder: behavioral leads from the study of mutant mice. Drug News Perspect 16:453-463. Einat H, Yuan P, Gould TD, Li J, Du J, Zhang L, Manji HK, Chen G (2003b) The role of the extracellular signal-regulated kinase signaling pathway in mood modulation. J Neurosci 23:7311-7316. Emanuel BS, McDonald-McGinn D, Saitta SC, Zackai EH (2001) The 22q11.2 deletion syndrome. Adv Pediatr 48:39-73. Edelmann L, Pandita RK, Morrow BE (1999) Low-copy repeats mediate the common 3- Mb deletion in patients with velo-cardio-facial syndrome. Am J Hum Genet 64:1076-1086. Engel SR, Creson TK, Hao Y, Shen Y, Maeng S, Nekrasova T, Landreth GE, Manji HK, Chen G (2008) The extracellular signal-regulated kinase pathway contributes to the control of behavioral excitement. Mol Psychiatry. English JD, Sweatt JD (1996) Activation of p42 mitogen-activated protein kinase in hippocampal long term potentiation. J Biol Chem 271:24329-24332. English JD, Sweatt JD (1997) A requirement for the mitogen-activated protein kinase cascade in hippocampal long term potentiation. J Biol Chem 272:19103-19106. Englund C, Fink A, Lau C, Pham D, Daza RA, Bulfone A, Kowalczyk T, Hevner RF (2005) Pax6, Tbr2, and Tbr1 are expressed sequentially by radial glia, intermediate progenitor cells, and postmitotic neurons in developing neocortex. J Neurosci 25:247-251. Evans SJ, Choudary PV, Neal CR, Li JZ, Vawter MP, Tomita H, Lopez JF, Thompson RC, Meng F, Stead JD, Walsh DM, Myers RM, Bunney WE, Watson SJ, Jones EG, Akil H (2004) Dysregulation of the fibroblast growth factor system in major depression. Proc Natl Acad Sci U S A 101:15506-15511. Fantl V, Stamp G, Andrews A, Rosewell I, Dickson C (1995) Mice lacking cyclin D1 are small and show defects in eye and mammary gland development. Genes Dev 9:2364-2372. Feng Y, Walsh CA (2004) Mitotic spindle regulation by Nde1 controls cerebral cortical size. Neuron 44:279-293. Ferguson KL, Vanderluit JL, Hebert JM, McIntosh WC, Tibbo E, MacLaurin JG, Park DS, Wallace VA, Vooijs M, McConnell SK, Slack RS (2002) Telencephalon- specific Rb knockouts reveal enhanced neurogenesis, survival and abnormal cortical development. Embo J 21:3337-3346. Fremin C, Ezan F, Boisselier P, Bessard A, Pages G, Pouyssegur J, Baffet G (2007) ERK2 but not ERK1 plays a key role in hepatocyte replication: an RNAi-

202 mediated ERK2 knockdown approach in wild-type and ERK1 null hepatocytes. Hepatology 45:1035-1045. Ferguson SM, Fasano S, Yang P, Brambilla R, Robinson TE (2006) Knockout of ERK1 enhances cocaine-evoked immediate early gene expression and behavioral plasticity. Neuropsychopharmacology 31:2660-2668. Fishell G, Kriegstein A (2005) Cortical development: new concepts. Neuron 46:361-362. Futter M, Uematsu K, Bullock SA, Kim Y, Hemmings HC, Jr., Nishi A, Greengard P, Nairn AC (2005) Phosphorylation of spinophilin by ERK and cyclin-dependent PK 5 (Cdk5). Proc Natl Acad Sci U S A 102:3489-3494. Gal JS, Morozov YM, Ayoub AE, Chatterjee M, Rakic P, Haydar TF (2006) Molecular and morphological heterogeneity of neural precursors in the mouse neocortical proliferative zones. J Neurosci 26:1045-1056. Gauthier AS, Furstoss O, Araki T, Chan R, Neel BG, Kaplan DR, Miller FD (2007) Control of CNS cell-fate decisions by SHP-2 and its dysregulation in Noonan syndrome. Neuron 54:245-262. Geng Y, Yu Q, Sicinska E, Das M, Bronson RT, Sicinski P (2001) Deletion of the p27Kip1 gene restores normal development in cyclin D1-deficient mice. Proc Natl Acad Sci U S A 98:194-199. Gervasini C, Castronovo P, Bentivegna A, Mottadelli F, Faravelli F, Giovannucci-Uzielli ML, Pessagno A, Lucci-Cordisco E, Pinto AM, Salviati L, Selicorni A, Tenconi R, Neri G, Larizza L (2007) High frequency of mosaic CREBBP deletions in Rubinstein-Taybi syndrome patients and mapping of somatic and germ-line breakpoints. Genomics 90:567-573. Ghebranious N, Giampietro PF, Wesbrook FP, Rezkalla SH (2007) A novel microdeletion at 16p11.2 harbors candidate genes for aortic valve development, seizure disorder, and mild mental retardation. Am J Med Genet A 143A:1462- 1471. Giroux S, Tremblay M, Bernard D, Cardin-Girard JF, Aubry S, Larouche L, Rousseau S, Huot J, Landry J, Jeannotte L, Charron J (1999) Embryonic death of Mek1- deficient mice reveals a role for this kinase in angiogenesis in the labyrinthine region of the placenta. Curr Biol 9:369-372. Giulian D, Baker TJ (1986) Characterization of ameboid microglia isolated from developing mammalian brain. J Neurosci 6:2163-2178. Glickstein SB, Alexander S, Ross ME (2007) Differences in cyclin D2 and D1 protein expression distinguish forebrain progenitor subsets. Cereb Cortex 17:632-642. Goldin M, Segal M (2003) Protein kinase C and ERK involvement in dendritic spine plasticity in cultured rodent hippocampal neurons. Eur J Neurosci 17:2529-2539. Goto T, Mitsuhashi T, Takahashi T (2004) Altered patterns of neuron production in the p27 knockout mouse. Dev Neurosci 26:208-217. Gotoh Y, Nishida E, Matsuda S, Shiina N, Kosako H, Shiokawa K, Akiyama T, Ohta K, Sakai H (1991) In vitro effects on microtubule dynamics of purified Xenopus M phase-activated MAP kinase. Nature 349:251-254. Gotz M, Huttner WB (2005) The cell biology of neurogenesis. Nat Rev Mol Cell Biol 6:777-788.

203 Gourley SL, Wu FJ, Kiraly DD, Ploski JE, Kedves AT, Duman RS, Taylor JR (2008) Regionally specific regulation of ERK MAP kinase in a model of antidepressant- sensitive chronic depression. Biol Psychiatry 63:353-359. Graves LM, Guy HI, Kozlowski P, Huang M, Lazarowski E, Pope RM, Collins MA, Dahlstrand EN, Earp HS, 3rd, Evans DR (2000) Regulation of carbamoyl phosphate synthetase by MAP kinase. Nature 403:328-332. Guadagno TM, Ferrell JE, Jr. (1998) Requirement for MAPK activation for normal mitotic progression in Xenopus egg extracts. Science 282:1312-1315. Guerrini R, Marini C (2006) Genetic malformations of cortical development. Exp Brain Res 173:322-333. Guillemot F (2007) Cell fate specification in the mammalian telencephalon. Prog Neurobiol 83:37-52. Guris DL, Fantes J, Tara D, Druker BJ, Imamoto A (2001) Mice lacking the homologue of the human 22q11.2 gene CRKL phenocopy neurocristopathies of DiGeorge syndrome. Nat Genet 27:293-298. Gysin S, Lee SH, Dean NM, McMahon M (2005) Pharmacologic inhibition of RAF-- >MEK-->ERK signaling elicits pancreatic cancer cell cycle arrest through induced expression of p27Kip1. Cancer Res 65:4870-4880. Hadari YR, Kouhara H, Lax I, Schlessinger J (1998) Binding of Shp2 tyrosine phosphatase to FRS2 is essential for fibroblast growth factor-induced PC12 cell differentiation. Mol Cell Biol 18:3966-3973. Hadari YR, Gotoh N, Kouhara H, Lax I, Schlessinger J (2001) Critical role for the docking-protein FRS2 alpha in FGF receptor-mediated signal transduction pathways. Proc Natl Acad Sci U S A 98:8578-8583. Hallam TM, Bourtchouladze R (2006) Rubinstein-Taybi syndrome: molecular findings and therapeutic approaches to improve cognitive dysfunction. Cell Mol Life Sci 63:1725-1735. Hallberg B, Rayter SI, Downward J (1994) Interaction of Ras and Raf in intact mammalian cells upon extracellular stimulation. J Biol Chem 269:3913-3916. Hara E, Hall M, Peters G (1997) Cdk2-dependent phosphorylation of Id2 modulates activity of E2A-related transcription factors. Embo J 16:332-342. Hatakeyama J, Bessho Y, Katoh K, Ookawara S, Fujioka M, Guillemot F, Kageyama R (2004) Hes genes regulate size, shape and histogenesis of the nervous system by control of the timing of neural stem cell differentiation. Development 131:5539- 5550. Hao Y, Creson T, Zhang L, Li P, Du F, Yuan P, Gould TD, Manji HK, Chen G (2004) Mood stabilizer valproate promotes ERK pathway-dependent cortical neuronal growth and neurogenesis. J Neurosci 24:6590-6599. Hashimoto R, Numakawa T, Ohnishi T, Kumamaru E, Yagasaki Y, Ishimoto T, Mori T, Nemoto K, Adachi N, Izumi A, Chiba S, Noguchi H, Suzuki T, Iwata N, Ozaki N, Taguchi T, Kamiya A, Kosuga A, Tatsumi M, Kamijima K, Weinberger DR, Sawa A, Kunugi H (2006) Impact of the DISC1 Ser704Cys polymorphism on risk for major depression, brain morphology and ERK signaling. Hum Mol Genet 15:3024-3033.

204 Hatano N, Mori Y, Oh-hora M, Kosugi A, Fujikawa T, Nakai N, Niwa H, Miyazaki J, Hamaoka T, Ogata M (2003) Essential role for ERK2 mitogen-activated protein kinase in placental development. Genes Cells 8:847-856. Haydar TF (2005) Advanced microscopic imaging methods to investigate cortical development and the etiology of mental retardation. Ment Retard Dev Disabil Res Rev 11:303-316. Haydar TF, Ang E, Jr., Rakic P (2003) Mitotic spindle rotation and mode of cell division in the developing telencephalon. Proc Natl Acad Sci U S A 100:2890-2895. Haydar TF, Nowakowski RS, Yarowsky PJ, Krueger BK (2000) Role of founder cell deficit and delayed neuronogenesis in microencephaly of the trisomy 16 mouse. J Neurosci 20:4156-4164. He F, Ge W, Martinowich K, Becker-Catania S, Coskun V, Zhu W, Wu H, Castro D, Guillemot F, Fan G, de Vellis J, Sun YE (2005) A positive autoregulatory loop of Jak-STAT signaling controls the onset of astrogliogenesis. Nat Neurosci 8:616- 625. Hebert AE, Dash PK (2002) Extracellular signal-regulated kinase activity in the entorhinal cortex is necessary for long-term spatial memory. Learn Mem 9:156- 166. Hevner RF, Hodge RD, Daza RA, Englund C (2006) Transcription factors in glutamatergic neurogenesis: conserved programs in neocortex, cerebellum, and adult hippocampus. Neurosci Res 55:223-233. Hong JW, Ryu MS, Lim IK (2005) Phosphorylation of serine 147 of tis21/BTG2/pc3 by p-Erk1/2 induces Pin-1 binding in cytoplasm and cell death. J Biol Chem 280:21256-21263. Horne MM, Guadagno TM (2003) A requirement for MAP kinase in the assembly and maintenance of the mitotic spindle. J Cell Biol 161:1021-1028. Huang YY, Martin KC, Kandel ER (2000) Both protein kinase A and mitogen-activated protein kinase are required in the amygdala for the macromolecular synthesis- dependent late phase of long-term potentiation. J Neurosci 20:6317-6325. Hughes JR (2007) Autism: the first firm finding = underconnectivity? Epilepsy Behav 11:20-24. Iacopetti P, Michelini M, Stuckmann I, Oback B, Aaku-Saraste E, Huttner WB (1999) Expression of the antiproliferative gene TIS21 at the onset of neurogenesis identifies single neuroepithelial cells that switch from proliferative to neuron- generating division. Proc Natl Acad Sci U S A 96:4639-4644. Iavarone A, Lasorella A (2004) Id proteins in neural cancer. Cancer Lett 204:189-196. Igaz LM, Bekinschtein P, Izquierdo I, Medina JH (2004) One-trial aversive learning induces late changes in hippocampal CaMKIIalpha, Homer 1a, Syntaxin 1a and ERK2 protein levels. Brain Res Mol Brain Res 132:1-12. Inglis-Broadgate SL, Thomson RE, Pellicano F, Tartaglia MA, Pontikis CC, Cooper JD, Iwata T (2005) FGFR3 regulates brain size by controlling progenitor cell proliferation and apoptosis during embryonic development. Dev Biol 279:73-85. Ishibashi M (2004) Molecular mechanisms for morphogenesis of the central nervous system in mammals. Anat Sci Int 79:226-234.

205 Ishibashi M, Moriyoshi K, Sasai Y, Shiota K, Nakanishi S, Kageyama R (1994) Persistent expression of helix-loop-helix factor HES-1 prevents mammalian neural differentiation in the central nervous system. Embo J 13:1799-1805. Ishibashi M, Ang SL, Shiota K, Nakanishi S, Kageyama R, Guillemot F (1995) Targeted disruption of mammalian hairy and Enhancer of split homolog-1 (HES-1) leads to up-regulation of neural helix-loop-helix factors, premature neurogenesis, and severe neural tube defects. Genes Dev 9:3136-3148. Just MA, Cherkassky VL, Keller TA, Kana RK, Minshew NJ (2007) Functional and anatomical cortical underconnectivity in autism: evidence from an FMRI study of an executive function task and corpus callosum morphometry. Cereb Cortex 17:951-961. Kao S, Jaiswal RK, Kolch W, Landreth GE (2001) Identification of the mechanisms regulating the differential activation of the mapk cascade by epidermal growth factor and nerve growth factor in PC12 cells. J Biol Chem 276:18169-18177. Katz M, Amit I, Yarden Y (2007) Regulation of MAPKs by growth factors and receptor tyrosine kinases. Biochim Biophys Acta 1773:1161-1176. Kawada M, Yamagoe S, Murakami Y, Suzuki K, Mizuno S, Uehara Y (1997) Induction of p27Kip1 degradation and anchorage independence by Ras through the MAP kinase signaling pathway. Oncogene 15:629-637. Ke Y, Zhang EE, Hagihara K, Wu D, Pang Y, Klein R, Curran T, Ranscht B, Feng GS (2007) Deletion of Shp2 in the brain leads to defective proliferation and differentiation in neural stem cells and early postnatal lethality. Mol Cell Biol 27:6706-6717. Kelleher RJ, 3rd, Govindarajan A, Jung HY, Kang H, Tonegawa S (2004) Translational control by MAPK signaling in long-term synaptic plasticity and memory. Cell 116:467-479. Kesler SR, Simensen RJ, Voeller K, Abidi F, Stevenson RE, Schwartz CE, Reiss AL (2007) Altered neurodevelopment associated with mutations of RSK2: a morphometric MRI study of Coffin-Lowry syndrome. Neurogenetics 8:143-147. Kim SH, Markham JA, Weiler IJ, Greenough WT (2008) Aberrant early-phase ERK inactivation impedes neuronal function in fragile X syndrome. Proc Natl Acad Sci U S A 105:4429-4434. Knoepfler PS, Cheng PF, Eisenman RN (2002) N-myc is essential during neurogenesis for the rapid expansion of progenitor cell populations and the inhibition of neuronal differentiation. Genes Dev 16:2699-2712. Kolch W (2005) Coordinating ERK/MAPK signalling through scaffolds and inhibitors. Nat Rev Mol Cell Biol 6:827-837. Korada S, Zheng W, Basilico C, Schwartz ML, Vaccarino FM (2002) Fibroblast growth factor 2 is necessary for the growth of glutamate projection neurons in the anterior neocortex. J Neurosci 22:863-875. Kozar K, Sicinski P (2005) Cell cycle progression without cyclin D-CDK4 and cyclin D- CDK6 complexes. Cell Cycle 4:388-391. Kozar K, Ciemerych MA, Rebel VI, Shigematsu H, Zagozdzon A, Sicinska E, Geng Y, Yu Q, Bhattacharya S, Bronson RT, Akashi K, Sicinski P (2004) Mouse development and cell proliferation in the absence of D-cyclins. Cell 118:477-491.

206 Krubitzer L, Kahn DM (2003) Nature versus nurture revisited: an old idea with a new twist. Prog Neurobiol 70:33-52. Kuida K, Boucher DM (2004) Functions of MAP kinases: insights from gene-targeting studies. J Biochem 135:653-656. Kumar RA, KaraMohamed S, Sudi J, Conrad DF, Brune C, Badner JA, Gilliam TC, Nowak NJ, Cook EH, Jr., Dobyns WB, Christian SL (2008) Recurrent 16p11.2 microdeletions in autism. Hum Mol Genet 17:628-638. Kunath T, Saba-El-Leil MK, Almousailleakh M, Wray J, Meloche S, Smith A (2007) FGF stimulation of the Erk1/2 signalling cascade triggers transition of pluripotent embryonic stem cells from self-renewal to lineage commitment. Development 134:2895-2902. Lefloch R, Pouyssegur J, Lenormand P (2008) Single and combined silencing of ERK1 and ERK2 reveals their positive contribution to growth signaling depending on their expression levels. Mol Cell Biol 28:511-527. Liu L, Cundiff P, Abel G, Wang Y, Faigle R, Sakagami H, Xu M, Xia Z (2006) Extracellular signal-regulated kinase (ERK) 5 is necessary and sufficient to specify cortical neuronal fate. Proc Natl Acad Sci U S A 103:9697-9702. Liu X, Yan S, Zhou T, Terada Y, Erikson RL (2004) The MAP kinase pathway is required for entry into mitosis and cell survival. Oncogene 23:763-776. Lu W, Man H, Ju W, Trimble WS, MacDonald JF, Wang YT (2001) Activation of synaptic NMDA receptors induces membrane insertion of new AMPA receptors and LTP in cultured hippocampal neurons. Neuron 29:243-254. Lucarelli P, Palminiello S, Saccucci P, Bottini N, De Luca D, Elia M, Fiumara A, Curatolo P (2003) Association study of autistic disorder and chromosome 16p. Am J Med Genet A 119:242-246. Lukaszewicz A, Savatier P, Cortay V, Giroud P, Huissoud C, Berland M, Kennedy H, Dehay C (2005) G1 phase regulation, area-specific cell cycle control, and cytoarchitectonics in the primate cortex. Neuron 47:353-364. Lyden D, Young AZ, Zagzag D, Yan W, Gerald W, O'Reilly R, Bader BL, Hynes RO, Zhuang Y, Manova K, Benezra R (1999) Id1 and Id3 are required for neurogenesis, angiogenesis and vascularization of tumour xenografts. Nature 401:670-677. Malatesta P, Hack MA, Hartfuss E, Kettenmann H, Klinkert W, Kirchhoff F, Gotz M (2003) Neuronal or glial progeny: regional differences in radial glia fate. Neuron 37:751-764. Malumbres M, Sotillo R, Santamaria D, Galan J, Cerezo A, Ortega S, Dubus P, Barbacid M (2004) Mammalian cells cycle without the D-type cyclin-dependent kinases Cdk4 and Cdk6. Cell 118:493-504. Marshall CJ (1995) Specificity of receptor tyrosine kinase signaling: transient versus sustained extracellular signal-regulated kinase activation. Cell 80:179-185. Martinez-Cerdeno V, Noctor SC, Kriegstein AR (2006) The role of intermediate progenitor cells in the evolutionary expansion of the cerebral cortex. Cereb Cortex 16 Suppl 1:i152-161. Maynard TM, Haskell GT, Lieberman JA, LaMantia AS (2002) 22q11 DS: genomic mechanisms and gene function in DiGeorge/velocardiofacial syndrome. Int J Dev Neurosci 20:407-419.

207 Mazzucchelli C, Brambilla R (2000) Ras-related and MAPK signalling in neuronal plasticity and memory formation. Cell Mol Life Sci 57:604-611. Mazzucchelli C, Vantaggiato C, Ciamei A, Fasano S, Pakhotin P, Krezel W, Welzl H, Wolfer DP, Pages G, Valverde O, Marowsky A, Porrazzo A, Orban PC, Maldonado R, Ehrengruber MU, Cestari V, Lipp HP, Chapman PF, Pouyssegur J, Brambilla R (2002) Knockout of ERK1 MAP kinase enhances synaptic plasticity in the striatum and facilitates striatal-mediated learning and memory. Neuron 34:807-820. McConnell SK, Kaznowski CE (1991) Cell cycle dependence of laminar determination in developing neocortex. Science 254:282-285. McFarland KN, Wilkes SR, Koss SE, Ravichandran KS, Mandell JW (2006) Neural- specific inactivation of ShcA results in increased embryonic neural progenitor apoptosis and microencephaly. J Neurosci 26:7885-7897. Meloche S, Vella FD, Voisin L, Ang SL, Saba-El-Leil M (2004) Erk2 signaling and early embryo stem cell self-renewal. Cell Cycle 3:241-243. Meloche S, Pouyssegur J (2007) The ERK1/2 mitogen-activated protein kinase pathway as a master regulator of the G1- to S-phase transition. Oncogene 26:3227-3239. Menard C, Hein P, Paquin A, Savelson A, Yang XM, Lederfein D, Barnabe-Heider F, Mir AA, Sterneck E, Peterson AC, Johnson PF, Vinson C, Miller FD (2002) An essential role for a MEK-C/EBP pathway during growth factor-regulated cortical neurogenesis. Neuron 36:597-610. Michaelson MD, Bieri PL, Mehler MF, Xu H, Arezzo JC, Pollard JW, Kessler JA (1996) CSF-1 deficiency in mice results in abnormal brain development. Development 122:2661-2672. Mikula M, Schreiber M, Husak Z, Kucerova L, Ruth J, Wieser R, Zatloukal K, Beug H, Wagner EF, Baccarini M (2001) Embryonic lethality and fetal liver apoptosis in mice lacking the c-raf-1 gene. Embo J 20:1952-1962. Miller FD, Gauthier AS (2007) Timing is everything: making neurons versus glia in the developing cortex. Neuron 54:357-369. Miskimins WK, Wang G, Hawkinson M, Miskimins R (2001) Control of cyclin- dependent kinase inhibitor p27 expression by cap-independent translation. Mol Cell Biol 21:4960-4967. Mitsuhashi T, Aoki Y, Eksioglu YZ, Takahashi T, Bhide PG, Reeves SA, Caviness VS, Jr. (2001) Overexpression of p27Kip1 lengthens the G1 phase in a mouse model that targets inducible gene expression to central nervous system progenitor cells. Proc Natl Acad Sci U S A 98:6435-6440. Mizuno A, Villalobos ME, Davies MM, Dahl BC, Muller RA (2006) Partially enhanced thalamocortical functional connectivity in autism. Brain Res 1104:160-174. Molnar Z (2004) Thomas Willis (1621-1675), the founder of clinical neuroscience. Nat Rev Neurosci 5:329-335. Morozov A, Muzzio IA, Bourtchouladze R, Van-Strien N, Lapidus K, Yin D, Winder DG, Adams JP, Sweatt JD, Kandel ER (2003) Rap1 couples cAMP signaling to a distinct pool of p42/44MAPK regulating excitability, synaptic plasticity, learning, and memory. Neuron 39:309-325. Mulloy R, Salinas S, Philips A, Hipskind RA (2003) Activation of cyclin D1 expression by the ERK5 cascade. Oncogene 22:5387-5398.

208 Nakamura Y, Sakakibara S, Miyata T, Ogawa M, Shimazaki T, Weiss S, Kageyama R, Okano H (2000) The bHLH gene hes1 as a repressor of the neuronal commitment of CNS stem cells. J Neurosci 20:283-293. Nekrasova T, Shive C, Gao Y, Kawamura K, Guardia R, Landreth G, Forsthuber T (2005) ERK1-deficient mice show normal T cell effector function and are highly susceptible to experimental autoimmune encephalomyelitis. J Immunol 175:2374- 2380. Nieto M, Schuurmans C, Britz O, Guillemot F (2001) Neural bHLH genes control the neuronal versus glial fate decision in cortical progenitors. Neuron 29:401-413. Nishimoto S, Nishida E (2006) MAPK signalling: ERK5 versus ERK1/2. EMBO Rep 7:782-786. Noctor SC, Martinez-Cerdeno V, Ivic L, Kriegstein AR (2004) Cortical neurons arise in symmetric and asymmetric division zones and migrate through specific phases. Nat Neurosci 7:136-144. Nowakowski RS, Caviness VS, Jr., Takahashi T, Hayes NL (2002) Population dynamics during cell proliferation and neuronogenesis in the developing murine neocortex. Results Probl Cell Differ 39:1-25. Ohkubo Y, Uchida AO, Shin D, Partanen J, Vaccarino FM (2004) Fibroblast growth factor receptor 1 is required for the proliferation of hippocampal progenitor cells and for hippocampal growth in mouse. J Neurosci 24:6057-6069. Ohnuma S, Harris WA (2003) Neurogenesis and the cell cycle. Neuron 40:199-208. Ohtsuka T, Sakamoto M, Guillemot F, Kageyama R (2001) Roles of the basic helix-loop- helix genes Hes1 and Hes5 in expansion of neural stem cells of the developing brain. J Biol Chem 276:30467-30474. Ortega S, Ittmann M, Tsang SH, Ehrlich M, Basilico C (1998) Neuronal defects and delayed wound healing in mice lacking fibroblast growth factor 2. Proc Natl Acad Sci U S A 95:5672-5677. Ortiz J, Harris HW, Guitart X, Terwilliger RZ, Haycock JW, Nestler EJ (1995) Extracellular signal-regulated protein kinases (ERKs) and ERK kinase (MEK) in brain: regional distribution and regulation by chronic morphine. J Neurosci 15:1285-1297. Pages G, Guerin S, Grall D, Bonino F, Smith A, Anjuere F, Auberger P, Pouyssegur J (1999) Defective thymocyte maturation in p44 MAP kinase (Erk 1) knockout mice. Science 286:1374-1377. Paquin A, Barnabe-Heider F, Kageyama R, Miller FD (2005) CCAAT/enhancer-binding protein phosphorylation biases cortical precursors to generate neurons rather than astrocytes in vivo. J Neurosci 25:10747-10758. Paxinos G, and Franklin, KBJ (2001) The Mouse Brain in Stereotaxic Coordinates. 2nd Edition. San Diego: Academic Press. Paylor R, Lindsay E (2006) Mouse models of 22q11 deletion syndrome. Biol Psychiatry 59:1172-1179. Paylor R, Glaser B, Mupo A, Ataliotis P, Spencer C, Sobotka A, Sparks C, Choi CH, Oghalai J, Curran S, Murphy KC, Monks S, Williams N, O'Donovan MC, Owen MJ, Scambler PJ, Lindsay E (2006) Tbx1 haploinsufficiency is linked to behavioral disorders in mice and humans: implications for 22q11 deletion syndrome. Proc Natl Acad Sci U S A 103:7729-7734.

209 Pearson G, Robinson F, Beers Gibson T, Xu BE, Karandikar M, Berman K, Cobb MH (2001) Mitogen-activated protein (MAP) kinase pathways: regulation and physiological functions. Endocr Rev 22:153-183. Peddada S, Yasui DH, LaSalle JM (2006) Inhibitors of differentiation (ID1, ID2, ID3 and ID4) genes are neuronal targets of MeCP2 that are elevated in Rett syndrome. Hum Mol Genet 15:2003-2014. Pinkham AE, Hopfinger JB, Pelphrey KA, Piven J, Penn DL (2008) Neural bases for impaired social cognition in schizophrenia and autism spectrum disorders. Schizophr Res 99:164-175. Poirier R, Jacquot S, Vaillend C, Soutthiphong AA, Libbey M, Davis S, Laroche S, Hanauer A, Welzl H, Lipp HP, Wolfer DP (2007) Deletion of the Coffin-Lowry syndrome gene Rsk2 in mice is associated with impaired spatial learning and reduced control of exploratory behavior. Behav Genet 37:31-50. Pontious A, Kowalczyk T, Englund C, Hevner RF (2008) Role of intermediate progenitor cells in cerebral cortex development. Dev Neurosci 30:24-32. Pouyssegur J, Volmat V, Lenormand P (2002) Fidelity and spatio-temporal control in MAP kinase (ERKs) signalling. Biochem Pharmacol 64:755-763. Pritchard CA, Bolin L, Slattery R, Murray R, McMahon M (1996) Post-natal lethality and neurological and gastrointestinal defects in mice with targeted disruption of the A-Raf protein kinase gene. Curr Biol 6:614-617. Qi X, Lin W, Li J, Pan Y, Wang W (2006) The depressive-like behaviors are correlated with decreased phosphorylation of mitogen-activated protein kinases in rat brain following chronic forced swim stress. Behav Brain Res 175:233-240. Qian X, Shen Q, Goderie SK, He W, Capela A, Davis AA, Temple S (2000) Timing of CNS cell generation: a programmed sequence of neuron and glial cell production from isolated murine cortical stem cells. Neuron 28:69-80. Raballo R, Rhee J, Lyn-Cook R, Leckman JF, Schwartz ML, Vaccarino FM (2000) Basic fibroblast growth factor (Fgf2) is necessary for cell proliferation and neurogenesis in the developing cerebral cortex. J Neurosci 20:5012-5023. Rakic P (1988) Defects of neuronal migration and the pathogenesis of cortical malformations. Prog Brain Res 73:15-37. Rakic P (1995) A small step for the cell, a giant leap for mankind: a hypothesis of neocortical expansion during evolution. Trends Neurosci 18:383-388. Rakic P (2000) Molecular and cellular mechanisms of neuronal migration: relevance to cortical epilepsies. Adv Neurol 84:1-14. Rauch A, Pfeiffer RA, Leipold G, Singer H, Tigges M, Hofbeck M (1999) A novel 22q11.2 microdeletion in DiGeorge syndrome. Am J Hum Genet 64:659-666. Roberts A, Allanson J, Jadico SK, Kavamura MI, Noonan J, Opitz JM, Young T, Neri G (2006) The cardiofaciocutaneous syndrome. J Med Genet 43:833-842. Roberts EC, Shapiro PS, Nahreini TS, Pages G, Pouyssegur J, Ahn NG (2002) Distinct cell cycle timing requirements for extracellular signal-regulated kinase and phosphoinositide 3-kinase signaling pathways in somatic cell mitosis. Mol Cell Biol 22:7226-7241. Robin NH, Taylor CJ, McDonald-McGinn DM, Zackai EH, Bingham P, Collins KJ, Earl D, Gill D, Granata T, Guerrini R, Katz N, Kimonis V, Lin JP, Lynch DR, Mohammed SN, Massey RF, McDonald M, Rogers RC, Splitt M, Stevens CA,

210 Tischkowitz MD, Stoodley N, Leventer RJ, Pilz DT, Dobyns WB (2006) Polymicrogyria and deletion 22q11.2 syndrome: window to the etiology of a common cortical malformation. Am J Med Genet A 140:2416-2425. Robinson MJ, Harkins PC, Zhang J, Baer R, Haycock JW, Cobb MH, Goldsmith EJ (1996) Mutation of position 52 in ERK2 creates a nonproductive binding mode for adenosine 5'-triphosphate. Biochemistry 35:5641-5646. Roelfsema JH, Peters DJ (2007) Rubinstein-Taybi syndrome: clinical and molecular overview. Expert Rev Mol Med 9:1-16. Roovers K, Assoian RK (2000) Integrating the MAP kinase signal into the G1 phase cell cycle machinery. Bioessays 22:818-826. Rosenberg G (2007) The mechanisms of action of valproate in neuropsychiatric disorders: can we see the forest for the trees? Cell Mol Life Sci 64:2090-2103. Rosenblum K, Futter M, Voss K, Erent M, Skehel PA, French P, Obosi L, Jones MW, Bliss TV (2002) The role of extracellular regulated kinases I/II in late-phase long- term potentiation. J Neurosci 22:5432-5441. Ross M (1996) Cell division and the nervous system: regulating the cycle from neural differentiation to death. Trends Neurosci 19:62-68. Ross ME, Carter ML, Lee JH (1996) MN20, a D2 cyclin, is transiently expressed in selected neural populations during embryogenesis. J Neurosci 16:210-219. Ross SE, Greenberg ME, Stiles CD (2003) Basic helix-loop-helix factors in cortical development. Neuron 39:13-25. Roux PP, Blenis J (2004) ERK and p38 MAPK-activated protein kinases: a family of protein kinases with diverse biological functions. Microbiol Mol Biol Rev 68:320-344. Rubinfeld H, Seger R (2005) The ERK cascade: a prototype of MAPK signaling. Mol Biotechnol 31:151-174. Ruzinova MB, Benezra R (2003) Id proteins in development, cell cycle and cancer. Trends Cell Biol 13:410-418. Saba-El-Leil MK, Vella FD, Vernay B, Voisin L, Chen L, Labrecque N, Ang SL, Meloche S (2003) An essential function of the mitogen-activated protein kinase Erk2 in mouse trophoblast development. EMBO Rep 4:964-968. Saitta SC, McGrath JM, Mensch H, Shaikh TH, Zackai EH, Emanuel BS (1999) A 22q11.2 deletion that excludes UFD1L and CDC45L in a patient with conotruncal and craniofacial defects. Am J Hum Genet 65:562-566. Sakakibara K, Kubota K, Worku B, Ryer EJ, Miller JP, Koff A, Kent KC, Liu B (2005) PDGF-BB regulates p27 expression through ERK-dependent RNA turn-over in vascular smooth muscle cells. J Biol Chem 280:25470-25477. Satoh Y, Endo S, Ikeda T, Yamada K, Ito M, Kuroki M, Hiramoto T, Imamura O, Kobayashi Y, Watanabe Y, Itohara S, Takishima K (2007) Extracellular signal- regulated kinase 2 (ERK2) knockdown mice show deficits in long-term memory; ERK2 has a specific function in learning and memory. J Neurosci 27:10765- 10776. Sauvageot CM, Stiles CD (2002) Molecular mechanisms controlling cortical gliogenesis. Curr Opin Neurobiol 12:244-249.

211 Schaeffer HJ, Catling AD, Eblen ST, Collier LS, Krauss A, Weber MJ (1998) MP1: a MEK binding partner that enhances enzymatic activation of the MAP kinase cascade. Science 281:1668-1671. Schaer M, Schmitt JE, Glaser B, Lazeyras F, Delavelle J, Eliez S (2006) Abnormal patterns of cortical gyrification in velo-cardio-facial syndrome (deletion 22q11.2): an MRI study. Psychiatry Res 146:1-11. Schafe GE, Nader K, Blair HT, LeDoux JE (2001) Memory consolidation of Pavlovian fear conditioning: a cellular and molecular perspective. Trends Neurosci 24:540- 546. Selcher JC, Nekrasova T, Paylor R, Landreth GE, Sweatt JD (2001) Mice lacking the ERK1 isoform of MAP kinase are unimpaired in emotional learning. Learn Mem 8:11-19. Selcher JC, Weeber EJ, Christian J, Nekrasova T, Landreth GE, Sweatt JD (2003) A role for ERK MAP kinase in physiologic temporal integration in hippocampal area CA1. Learn Mem 10:26-39. Shaikh TH, O'Connor RJ, Pierpont ME, McGrath J, Hacker AM, Nimmakayalu M, Geiger E, Emanuel BS, Saitta SC (2007) Low copy repeats mediate distal chromosome 22q11.2 deletions: sequence analysis predicts breakpoint mechanisms. Genome Res 17:482-491. Shalin SC, Zirrgiebel U, Honsa KJ, Julien JP, Miller FD, Kaplan DR, Sweatt JD (2004) Neuronal MEK is important for normal fear conditioning in mice. J Neurosci Res 75:760-770. Sharma C, Vomastek T, Tarcsafalvi A, Catling AD, Schaeffer HJ, Eblen ST, Weber MJ (2004) MEK partner 1 (MP1): Regulation of oligomerization in MAP kinase signaling. J Cell Biochem. Shaul YD, Seger R (2006) ERK1c regulates Golgi fragmentation during mitosis. J Cell Biol 172:885-897. Shaul YD, Seger R (2007) The MEK/ERK cascade: from signaling specificity to diverse functions. Biochim Biophys Acta 1773:1213-1226. Shaw P, Greenstein D, Lerch J, Clasen L, Lenroot R, Gogtay N, Evans A, Rapoport J, Giedd J (2006) Intellectual ability and cortical development in children and adolescents. Nature 440:676-679. Shen Q, Wang Y, Dimos JT, Fasano CA, Phoenix TN, Lemischka IR, Ivanova NB, Stifani S, Morrisey EE, Temple S (2006) The timing of cortical neurogenesis is encoded within lineages of individual progenitor cells. Nat Neurosci 9:743-751. Sherr CJ, Roberts JM (2004) Living with or without cyclins and cyclin-dependent kinases. Genes Dev 18:2699-2711. Shi S, Hayashi Y, Esteban JA, Malinow R (2001) Subunit-specific rules governing AMPA receptor trafficking to synapses in hippocampal pyramidal neurons. Cell 105:331-343. Shin DM, Korada S, Raballo R, Shashikant CS, Simeone A, Taylor JR, Vaccarino F (2004) Loss of glutamatergic pyramidal neurons in frontal and temporal cortex resulting from attenuation of FGFR1 signaling is associated with spontaneous hyperactivity in mice. J Neurosci 24:2247-2258.

212 Sicinski P, Donaher JL, Parker SB, Li T, Fazeli A, Gardner H, Haslam SZ, Bronson RT, Elledge SJ, Weinberg RA (1995) Cyclin D1 provides a link between development and oncogenesis in the retina and breast. Cell 82:621-630. Simpson PJ, Moon C, Kleman AM, Connolly E, Ronnett GV (2007) Progressive and inhibitory cell cycle proteins act simultaneously to regulate neurotrophin- mediated proliferation and maturation of neuronal precursors. Cell Cycle 6:1077- 1089. Skarpen E, Flinder LI, Rosseland CM, Orstavik S, Wierod L, Oksvold MP, Skalhegg BS, Huitfeldt HS (2008) MEK1 and MEK2 regulate distinct functions by sorting ERK2 to different intracellular compartments. Faseb J 22:466-476. Slack RS, El-Bizri H, Wong J, Belliveau DJ, Miller FD (1998) A critical temporal requirement for the retinoblastoma protein family during neuronal determination. J Cell Biol 140:1497-1509. Sobin C, Kiley-Brabeck K, Karayiorgou M (2005) Lower prepulse inhibition in children with the 22q11 deletion syndrome. Am J Psychiatry 162:1090-1099. Soriano P (1999) Generalized lacZ expression with the ROSA26 Cre reporter strain. Nat Genet 21:70-71. Stefanovsky V, Langlois F, Gagnon-Kugler T, Rothblum LI, Moss T (2006a) Growth factor signaling regulates elongation of RNA polymerase I transcription in mammals via UBF phosphorylation and r-chromatin remodeling. Mol Cell 21:629-639. Stefanovsky VY, Langlois F, Bazett-Jones D, Pelletier G, Moss T (2006b) ERK modulates DNA bending and enhancesome structure by phosphorylating HMG1- boxes 1 and 2 of the RNA polymerase I transcription factor UBF. Biochemistry 45:3626-3634. Stockhausen MT, Sjolund J, Axelson H (2005) Regulation of the Notch target gene Hes-1 by TGFalpha induced Ras/MAPK signaling in human neuroblastoma cells. Exp Cell Res 310:218-228. Sweatt JD (2001) The neuronal MAP kinase cascade: a biochemical signal integration system subserving synaptic plasticity and memory. J Neurochem 76:1-10. Sweatt JD (2004) Mitogen-activated protein kinases in synaptic plasticity and memory. Curr Opin Neurobiol 14:311-317. Sweatt JD, Weeber EJ, Lombroso PJ (2003) Genetics of childhood disorders: LI. Learning and memory, Part 4: Human cognitive disorders and the ras/ERK/CREB pathway. J Am Acad Child Adolesc Psychiatry 42:741-744. Swillen A, Vogels A, Devriendt K, Fryns JP (2000) Chromosome 22q11 deletion syndrome: update and review of the clinical features, cognitive-behavioral spectrum, and psychiatric complications. Am J Med Genet 97:128-135. Takahashi T, Nowakowski RS, Caviness VS, Jr. (1993) Cell cycle parameters and patterns of nuclear movement in the neocortical proliferative zone of the fetal mouse. J Neurosci 13:820-833. Takahashi T, Nowakowski RS, Caviness VS, Jr. (1995) The cell cycle of the pseudostratified ventricular epithelium of the embryonic murine cerebral wall. J Neurosci 15:6046-6057.

213 Takahashi T, Goto T, Miyama S, Nowakowski RS, Caviness VS, Jr. (1999) Sequence of neuron origin and neocortical laminar fate: relation to cell cycle of origin in the developing murine cerebral wall. J Neurosci 19:10357-10371. Takuwa N, Takuwa Y (1997) Ras activity late in G1 phase required for p27kip1 downregulation, passage through the restriction point, and entry into S phase in growth factor-stimulated NIH 3T3 fibroblasts. Mol Cell Biol 17:5348-5358. Tarui T, Takahashi T, Nowakowski RS, Hayes NL, Bhide PG, Caviness VS (2005) Overexpression of p27 Kip 1, probability of cell cycle exit, and laminar destination of neocortical neurons. Cereb Cortex 15:1343-1355. Teixeira A, Chaverot N, Strosberg AD, Cazaubon S (2000) Differential regulation of cyclin D1 and D3 expression in the control of astrocyte proliferation induced by endothelin-1. J Neurochem 74:1034-1040. Thomas KL, Hunt SP (1993) The regional distribution of extracellularly regulated kinase- 1 and -2 messenger RNA in the adult rat central nervous system. Neuroscience 56:741-757. Thomas GM, Huganir RL (2004) MAPK cascade signalling and synaptic plasticity. Nat Rev Neurosci 5:173-183. Thomson RE, Pellicano F, Iwata T (2007) Fibroblast growth factor receptor 3 kinase domain mutation increases cortical progenitor proliferation via mitogen-activated protein kinase activation. J Neurochem 100:1565-1578. Toma JG, El-Bizri H, Barnabe-Heider F, Aloyz R, Miller FD (2000) Evidence that helix- loop-helix proteins collaborate with retinoblastoma tumor suppressor protein to regulate cortical neurogenesis. J Neurosci 20:7648-7656. Ussar S, Voss T (2004) MEK1 and MEK2, different regulators of the G1/S transition. J Biol Chem 279:43861-43869. Vaccarino FM, Schwartz ML, Raballo R, Nilsen J, Rhee J, Zhou M, Doetschman T, Coffin JD, Wyland JJ, Hung YT (1999) Changes in cerebral cortex size are governed by fibroblast growth factor during embryogenesis. Nat Neurosci 2:246- 253. Valjent E, Corvol JC, Pages C, Besson MJ, Maldonado R, Caboche J (2000) Involvement of the extracellular signal-regulated kinase cascade for cocaine-rewarding properties. J Neurosci 20:8701-8709. Vantaggiato C, Formentini I, Bondanza A, Bonini C, Naldini L, Brambilla R (2006) ERK1 and ERK2 mitogen-activated protein kinases affect Ras-dependent cell signaling differentially. J Biol 5:14. Vazdarjanova A, McNaughton BL, Barnes CA, Worley PF, Guzowski JF (2002) Experience-dependent coincident expression of the effector immediate-early genes arc and Homer 1a in hippocampal and neocortical neuronal networks. J Neurosci 22:10067-10071. Villanueva J, Yung Y, Walker JL, Assoian RK (2007) ERK activity and G1 phase progression: identifying dispensable versus essential activities and primary versus secondary targets. Mol Biol Cell 18:1457-1463. Vogels A, Verhoeven WM, Tuinier S, DeVriendt K, Swillen A, Curfs LM, Frijns JP (2002) The psychopathological phenotype of velo-cardio-facial syndrome. Ann Genet 45:89-95.

214 Vomastek T, Schaeffer HJ, Tarcsafalvi A, Smolkin ME, Bissonette EA, Weber MJ (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 101:6981-6986. Voss AK, Gruss P, Thomas T (2003) The guanine nucleotide exchange factor C3G is necessary for the formation of focal adhesions and vascular maturation. Development 130:355-367. Voss AK, Krebs DL, Thomas T (2006) C3G regulates the size of the cerebral cortex neural precursor population. Embo J 25:3652-3663. Walz R, Roesler R, Quevedo J, Sant'Anna MK, Madruga M, Rodrigues C, Gottfried C, Medina JH, Izquierdo I (2000) Time-dependent impairment of inhibitory avoidance retention in rats by posttraining infusion of a mitogen-activated protein kinase kinase inhibitor into cortical and limbic structures. Neurobiol Learn Mem 73:11-20. Walz R, Roesler R, Barros DM, de Souza MM, Rodrigues C, Sant'Anna MK, Quevedo J, Choi HK, Neto WP, DeDavid e Silva TL, Medina JH, Izquierdo I (1999) Effects of post-training infusions of a mitogen-activated protein kinase kinase inhibitor into the hippocampus or entorhinal cortex on short- and long-term retention of inhibitory avoidance. Behav Pharmacol 10:723-730. Weeber EJ, Sweatt JD (2002) Molecular neurobiology of human cognition. Neuron 33:845-848. Weiss LA, Shen Y, Korn JM, Arking DE, Miller DT, Fossdal R, Saemundsen E, Stefansson H, Ferreira MA, Green T, Platt OS, Ruderfer DM, Walsh CA, Altshuler D, Chakravarti A, Tanzi RE, Stefansson K, Santangelo SL, Gusella JF, Sklar P, Wu BL, Daly MJ (2008) Association between microdeletion and microduplication at 16p11.2 and autism. N Engl J Med 358:667-675Wilkinson MG, Millar JB (2000) Control of the eukaryotic cell cycle by MAP kinase signaling pathways. Faseb J 14:2147-2157. Wiese S, Pei G, Karch C, Troppmair J, Holtmann B, Rapp UR, Sendtner M (2001) Specific function of B-Raf in mediating survival of embryonic motoneurons and sensory neurons. Nat Neurosci 4:137-142. Willard FS, Crouch MF (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 13:653-664. Williams RW, Herrup K (1988) The control of neuron number. Annu Rev Neurosci 11:423-453. Witelson SF, Beresh H, Kigar DL (2006) Intelligence and brain size in 100 postmortem brains: sex, lateralization and age factors. Brain 129:386-398. Wixler V, Smola U, Schuler M, Rapp U (1996) Differential regulation of Raf isozymes by growth versus differentiation inducing factors in PC12 pheochromocytoma cells. FEBS Lett 385:131-137. Wojnowski L, Stancato LF, Zimmer AM, Hahn H, Beck TW, Larner AC, Rapp UR, Zimmer A (1998) Craf-1 protein kinase is essential for mouse development. Mech Dev 76:141-149. Wu GY, Deisseroth K, Tsien RW (2001) Spaced stimuli stabilize MAPK pathway activation and its effects on dendritic morphology. Nat Neurosci 4:151-158.

215 Yamamoto S, Yoshino I, Shimazaki T, Murohashi M, Hevner RF, Lax I, Okano H, Shibuya M, Schlessinger J, Gotoh N (2005) Essential role of Shp2-binding sites on FRS2alpha for corticogenesis and for FGF2-dependent proliferation of neural progenitor cells. Proc Natl Acad Sci U S A 102:15983-15988. Yamamoto T, Ebisuya M, Ashida F, Okamoto K, Yonehara S, Nishida E (2006) Continuous ERK activation downregulates antiproliferative genes throughout G1 phase to allow cell-cycle progression. Curr Biol 16:1171-1182. Yao Y, Li W, Wu J, Germann UA, Su MS, Kuida K, Boucher DM (2003) Extracellular signal-regulated kinase 2 is necessary for mesoderm differentiation. Proc Natl Acad Sci U S A 100:12759-12764. Yntema HG, van den Helm B, Kissing J, van Duijnhoven G, Poppelaars F, Chelly J, Moraine C, Fryns JP, Hamel BC, Heilbronner H, Pander HJ, Brunner HG, Ropers HH, Cremers FP, van Bokhoven H (1999) A novel ribosomal S6-kinase (RSK4; RPS6KA6) is commonly deleted in patients with complex X-linked mental retardation. Genomics 62:332-343. York RD, Yao H, Dillon T, Ellig CL, Eckert SP, McCleskey EW, Stork PJ (1998) Rap1 mediates sustained MAP kinase activation induced by nerve growth factor. Nature 392:622-626. Yuan LL, Adams JP, Swank M, Sweatt JD, Johnston D (2002) Protein kinase modulation of dendritic K+ channels in hippocampus involves a mitogen-activated protein kinase pathway. J Neurosci 22:4860-4868. Yuan PX, Huang LD, Jiang YM, Gutkind JS, Manji HK, Chen G (2001) The mood stabilizer valproic acid activates mitogen-activated protein kinases and promotes neurite growth. J Biol Chem 276:31674-31683. Yung Y, Yao Z, Hanoch T, Seger R (2000) ERK1b, a 46-kDa ERK isoform that is differentially regulated by MEK. J Biol Chem 275:15799-15808. Zheng W, Nowakowski RS, Vaccarino FM (2004) Fibroblast growth factor 2 is required for maintaining the neural stem cell pool in the mouse brain subventricular zone. Dev Neurosci 26:181-196. Zhong J, Li X, McNamee C, Chen AP, Baccarini M, Snider WD (2007) Raf kinase signaling functions in sensory neuron differentiation and axon growth in vivo. Nat Neurosci 10:598-607. Zhuo L, Theis M, Alvarez-Maya I, Brenner M, Willecke K, Messing A (2001) hGFAP- cre transgenic mice for manipulation of glial and neuronal function in vivo. Genesis 31:85-94.

216