Molecular mechanisms regulating

C‐2 activity

Dissertation zur Erlangung des Doktorgrades Dr. rer. nat. der Fakultät für Naturwissenschaften der Universität Ulm

vorgelegt von Anja Jasmin Bühler aus Ulm

Universität Ulm Institut für Pharmakologie and Toxikologie

Ulm 2013

Current Dean of the Faculty of Natural Sciences: Prof. Dr. Joachim Ankerhold

First Supervisor: Prof. Dr. Peter Gierschik Second Supervisor: Prof. Dr. Ralf Marienfeld

Day Doctorate Awarded: 24th March 2014 Table of contents

Table of Contents

List of Figures...... v

Abbreviations and Units...... vii

1 Introduction ...... 1

1.1 Signal transduction...... 1

1.2 C...... 2

1.3 Regulation of Phospholipases C...... 5

1.4 Role of PLC2 in inflammatory and autoimmune diseases ...... 14

1.5 Anaplastic lymphoma kinase ...... 17

1.6 Aim of the work...... 23

2 Materials and Methods...... 25

2.1 Materials...... 25

2.2 Microbiology and Molecular Biology...... 40

2.3 Cell culture and DNA transfection...... 48

2.4 Protein Biochemistry ...... 49

2.5 Measurement of activity...... 53

2.6 Thermodynamic calculations...... 55

2.7 Cluster robustness analysis...... 56

2.8 Statistical analysis...... 56

3 Results ...... 58

3.1 Autoinhibitory regulation of PLC2 ...... 58

3.2 Functional reassembly of PLC2 by separately expressed fragments ...... 65

3.3 Phosphorylation induced activation of PLC2...... 69

3.4 Anaplastic lymphoma kinase ...... 73

3.5 Characterization of PLC2 deletion mutants associated with PLAID ...... 84

4 Discussion ...... 110

4.1 Autoinhibitory and phosphorylation-induced regulation of PLC2 activity ...... 111

iii Table of contents

4.2 Molecular mechanisms underlying the human hereditary disease PLAID...... 123

5 Summary ...... 131

6 References...... 133

Acknowledgements ...... 165

Curriculum Vitae...... 166

Publications ...... 167

Appendix ...... 168

Statutory Declaration ...... 169

Sections 2.1.1, 2.6, 2.7, 3.5.1, 3.5.2, 3.5.3,.3.5.4, 3.5.5, 3.5.6, 3.5.7,.3.5.10, 3.5.11, and 4.2 of this dissertation have been used for a publication and are taken from the corresponding manuscript [26].

iv List of Figures

List of Figures

Figure 1-1: Domain organization of mammalian PLC isozymes ...... 4 Figure 1-2: Oncogenic signalling of NPM-ALK ...... 21

Figure 3-1: Role of 2 specific array on the autoinhibitory regulation of PLC2...... 59 G12V Figure 3-2: Effect of Rac2 and constitutively active Rac2 on the activity of PLC2SH and PLC2SA . 60

Figure 3-3: Inhibitory effect of 2 SH domains on the activity of PLC2SH...... 62

Figure 3-4: Role of the SH2C and SH3 domain in autoinhibitory regulation of PLC2 ...... 64 G12V Figure 3-5: Effect of Rac2 and constitutively active Rac2 on the activity of PLC2SH mutants...... 65

Figure 3-6: Functional reassembly of truncated PLC2 fragments...... 67 H327A/H372A F897Q Figure 3-7: Functional reassembly of truncated PLC2 fragments with PLC2 and PLC2 ..... 68

Figure 3-8: Effect of mimicking phosphorylation of different tyrosine residues on the activity of PLC2 ...... 71 G12V Figure 3-9: Effect of Rac2 and constitutively active Rac2 on the activity of PLC2 phosphomimetic mutants ...... 73 Figure 3-10: Effect of ALK and its fusion proteins NPM-ALK and EML4-ALK on the activity of PLC isozymes, PLC2 phosphomimetic mutant, and PLC2...... 75 Figure 3-11: Effect of the constitutively active ALK mutants ALKF1174L and ALKR1275Q on the activity of PLC isozymes...... 76 Figure 3-12: Association of ALK and NPM-ALK with PLC isozymes in intact cells ...... 78 Figure 3-13: Effect of ALK and NPM-ALK kinase-dead mutants on the activity of PLC isozymes ...... 79 I310T Figure 3-14: Association of NPM-ALK and NPM-ALK with PLC in intact cells ...... 81 Figure 3-15: Effect of ALKY1604F and NPM-ALKY664F on the activity of PLC isozymes...... 82 Y664F Figure 3-16: Association of NPM-ALK with PLC1 ...... 83

Figure 3-17: Specific activation of PLC219 and PLC220-22 at subphysiological temperatures...... 85

Figure 3-18: Q10 analysis of the activities of wild-type PLC2, PLC219, and PLC220-22...... 86 Figure 3-19: Effect of subphysiological temperatures on cellular inositol phospholipid levels ...... 87

Figure 3-20: Effect of cycloheximide on the activity of PLC219 and PLC220-22 at 31 °C ...... 88

Figure 3-21: The activation of PLC219 and PLC220-22 by subphysiological temperatures is reversible ...... 89

Figure 3-22: Time course of the specific activation and inactivation of PLC219 and PLC220-22 by changes to and from a subphysiological temperatur...... 90 Figure 3-23: Effect of deleting exon 19 to 22 of PLCG2 on the enzymatic activity at 37 °C and 31 °C...... 91

Figure 3-24: Effect of the deletion of PLC1 residues corresponding to exon 19 or exon 20-22 of PLCG2 on the activity of PLC1 at subphysiological temperatures ...... 93

Figure 3-25: Effect of subphysiological temperatures on the activities of PLC2PCI, PLC2SA, PLC2SH and PLC2SH2C...... 94

Figure 3-26: Effect of subphysiological temperatures on the activities of PLC2 deletion mutants...... 96

v List of Figures

Ali5 Ali14 Figure 3-27: Effect of subphysiological temperatures on the activities of PLC2 , PLC2 , and Ali5/Ali14 PLC2 ...... 97

Figure 3-28: Summary and clustering of the effects of the PLC2 mutations...... 98 S707Y Figure 3-29: Effect of subphysiological temperatures on the activity of PLC2 ...... 100

Figure 3-30: Specific activity of PLC2 phosphomimetic mutants at 37 °C and 31 °C...... 101

Figure 3-31: Effect of subphysiological temperatures on the activities of PLC2 phosphomimetic mutants. 102 G12V Figure 3-32: Effect of subphysiological temperatures on the Rac2 -mediated activation of PLC2 and

PLC219 ...... 104 G12V Figure 3-33: Cool temperature shifts the requirement of PLC2 for Rac2 -mediated activation to lower G12V cellular PLC2 expression levels without altering the sensitivity to Rac2 ...... 105

Figure 3-34: Role of the spPH domain in the specific activation of PLC219 at subphysiological temperatures...... 107

Figure 3-35: Covalent linkage of the split PH domain halves blocks PLC2 sensitivity to cool temperatures ...... 109 Figure 4-1: Temperature dependence of conformational equilibrium constant...... 127

vi Abbreviations and Units

Abbreviations and Units

°C Degree Celcius A Absorbance aa Amino acid Ali Abnormal limb ALCL Anaplastic large cell lymphoma ALK Anaplastic lymphoma kinase

APLAID Autoinflammation and PLC2-associated antibody deficiency and immune dysregulation approx. Approximately ATP Adenosine triphosphate BCR B cell receptor bp BSA Bovine serum albumin Btk Bruton’s tyrosine kinase BLNK B cell linker (also known as SLP-65) c Concentration C- Carboxy- CD Cluster of differentiation Cdc Cell division cycle cDNA Complementary DNA cf. Confer CNS Central nervous system cpm Counts per minute CSTT Cold stimulation time test CVID Common variable immunodeficiency d Path length Da Dalton DAG Diacylglycerol ddH2O Double-distilled water DLBCL Diffuse large B-cell lymphoma DMEM Dulbecco’s modified Eagle’s medium

vii Abbreviations and Units

DMSO Dimethyl sulfoxide DNA Deoxyribonucleic acid dNTP Deoxyribonucleoside triphosphate DPPC Dipalmitoylphosphatidylcholine DRG Dorsal root ganglia DTT Dithiothreitol  Molar extinction coefficient EDTA Ethylene diamine tetraacetic acid e.g. Example given EGF Epidermal growth factor EML4 Echinoderm microtubule associated protein-like 4 ENU N-ethyl-N-nitrosourea ERK Extracellular-signal regulated kinase FACU Familial atypical cold urticaria FCS Fetal calf serum fp Forward primer g Gram g Gravitational constant Gab2 GRB2-associated-binding protein 2 GAP GTPase-activating protein GDI Guanine nucleotide dissociation inhibitor GDP Guanosine diphosphate GEF Guanine nucleotide exchange factor GMP Guanosine monophosphate GPCR G-protein-coupled receptor Grb2 Growth factor receptor-bound protein 2 GTP Guanosine triphosphate h hour h (as a prefix) Human Ig Immunoglobulin IGF-1 Insulin-like growth factor 1 IMT Inflammatory myofibroblastic tumor IP Immunoprecipitation

IP3 Inositol 1,4,5-trisphosphate

viii Abbreviations and Units

IR Insulin receptor IRS-1 Insulin receptor substrate 1 ITAM Immunoreceptor tyrosine activation motif Itk Interleukine-2-inducible T cell kinase JAK Janus kinase K Conformational equilibrium constant kb Kilobase kcal Kilo calorie l Liter LAT Linker of activated T cells LB Luria-Bertani LTK Leucocyte tyrosine kinase m Milli- M mol/l MAPK Mitogen-activated protein kinase min Minute MK Midkine mol Mole mRNA Messenger ribonucleic acid N- Amino- NFAT Nuclear factor of activated T-cells NK Natural killer NPM Nucleophosmin NMR Nuclear magnetic resonance NSCLC Non-small cell lung cancer nt Nucleotide OD Optical density PAGE Polyacrylamide gel electrophoresis PBS Phosphate-buffered saline PBS-T PBS-Tween p130CAS p130 Crk-associated substrate PCA Perchloric acid PCR Polymerase chain reaction PDGF Platelet-derived growth factor

ix Abbreviations and Units

PH Pleckstrin homology PI3K Phosphoinositide 3-kinase

PIP2 Phosphatidylinositol 4,5-bisphosphate

PIP3 Phosphatidylinositol 3,4,5-trisphosphate PKA Protein kinase A PKB Protein kinase B (also known as Akt) PKC Protein kinase C

PLAID Phospholipase C-2-associated antibody deficiency and immune dysregulation PLC Phosphoinositide-specific phospholipase C PMSF Phenylmethanesulphonylfluoride PP Polypropylene PTB Phosphotyrosine-binding PTN Pleiotrophin RA Ras associating RNA Ribonucleic acid rp Reverse primer rpm Rotations per minute rs Reference SNP cluster (rs#) RT Room temperature RTK Receptor tyrosine kinase s Second SA Specific array SCC Squamous cell carcinoma SDS Sodium dodecyl sulfate SEM Standard error of the mean SH Src homology Shp2 Src homology region 2 domain-containing -1 SNP Single nucleotide polymorphism SOE Splicing by overlap extension spPH Split pleckstrin homology SPR Surface plasmon resonance STAT Signal transducer and activators of transcription STI Soybean trypsin inhibitor

x Abbreviations and Units

TAE Tris-acetate-EDTA TCR T cell receptor TIM Triosephosphateisomerase TK Tyrosine kinase

Tm Melting temperature Tris Trishydroxymethylaminomethane TRP Transient receptor potential U unit UV Ultraviolet v Volume VSMC Vascular smooth muscle cells w Weight WB Western blotting

xi Introduction

1 Introduction

1.1 Signal transduction

Unicellular organisms have to be responsive to physical and chemical changes in their environment in order to ensure their survival. These mechanisms also include the response to the presence of other cells. This communication between cells in turn is essential to organize themselves in a complex community like multicellular organisms. The function of communicating with the environment and also with other cells within an organism is achieved through a number of pathways that receive and process signals originating from the external environment [2]. By this process, referred to as signal transduction, cells are able to convert endocrine signals, e.g. hormones or neurotransmitter, and exocrine signals, e.g. olfactory or acoustic stimuli, for the coordination of physiological functions or cell- cell communication [14]. Depending on the type of the extracellular signal molecule a complementary receptor protein expressed by the target cell, that specifically binds the signal molecule, is required. In general, there are three main types of extracellular signal molecules: hydrophilic molecules, lipophilic molecules, and dissolved gases. Lipophilic signal molecules, e.g. steroid and thyroid hormones are able to diffuse through the plasma membrane and target intracellular receptor proteins which directly regulate the transcription of specific . The dissolved gases, e.g. nitric oxide and carbon monoxide, also diffuse through the plasma membrane of the target cell activating the production of cyclic GMP. However, the vast majority of extracellular signal molecules are hydrophilic, cannot pass through the plasma membrane, and therefore require cell surface receptor proteins. The three largest classes of cell surface receptors are ligand-gated ion channels, enzyme-linked cell-surface receptors, like the epidermal growth factor (EGF) receptor, and seven-transmembrane (7TM) receptors, commonly referred to as G-protein-coupled receptors (GPCRs), which constitute the largest group among the cell surface receptors [2].

A wide variety of extracellular stimuli like hormones, neurotransmitter, chemokines, flavour and odour substances, or even photons are transmitted to intracellular pathways via receptor-mediated activation of phosphoinositide specific phospholipases C (PLCs) [204]. PLC isozymes are major effectors of GPCRs [20, 29, 236], whereas the activation of PLC isozymes is of significant importance in hematopoietic cells, in which the 1 Introduction

stimulation of the T cell receptor (TCR) leads to the activation of PLC1, and the engagement of the B cell receptor (BCR) leads to the stimulation of PLC2 activity [289].

In the specific case of PLC2, there are three major mechanisms to control the activity of the enzyme, namely (i) the phosphorylation-induced activation, (ii) the activation by small GTPases, and (iii) the autoinhibitory regulation. All three mechanisms are outlined in sections 1.3.3, 1.3.2, and 1.3.4, respectively.

1.2 Phospholipases C

PLCs are calcium-dependent enzymes located in the cytoplasm and activated upon stimulation of cell-surface receptors. PLCs control the phosphatidylinositol 4,5- bisphosphate (PIP2) levels of cells by catalysing the phosphodiesteratic cleavage of the membrane phospholipid PIP2 to yield the two second messengers diacylglycerol (DAG) and inositol 1,4,5-trisphosphate (IP3). This catalytic activity of PLCs was first demonstrated in 1975 with impure PLC preparations from Bacillus cereus [171]. IP3 induces the efflux of Ca2+ from the endoplasmatic reticulum into the cytosol [122, 251, 268], whereas DAG is a direct activator of various protein kinases C (PKC) which are in turn regulating several signal cascades involved in cellular processes like development or immune response [186]. Otherwise, PIP2 not only serves as a precursor for intracellular second messenger molecules, but also exerts direct signalling functions. It is involved in processes like the organization of the actin cytoskeleton [229], membrane trafficking [163], the regulation of cytokinesis [113] or cell-cell adhesion by gap junctions [307]. Owing to its regulatory role in diverse important cellular processes, changes in the balance of PIP2 levels are associated with several human diseases [15, 144, 157, 213]. Therefore, a strict regulation of enzymes significantly involved in the homeostasis of PIP2, like PLCs, is of particular importance [70].

1.2.1 Structure, Function, Classification, and Expression

The family of PLCs comprises 13 members which can be further classified due to structural or regulatory analogy. To date cDNAs of 6 different PLC subfamilies were cloned, divided into the PLC-(1-4), -(1,2), -(1,3,4), -,-,and -(1,2) [99, 129, 181,

2 Introduction

225, 254, 255, 313] (Figure 1-1). Although the general similarity between the amino acid sequences of the PLC isoforms is not high, the individual domains exhibit a similarity of 40-50 % [70]. The highly conserved catalytic subdomains X and Y are present in every

PLC subfamily [211]. The crystal structure of PLC1 reveals that both domains are composed of alternating -helices and -strands, which is a typical motif of a triose phosphate (TIM) barrel. Together, they resemble a distorted but closed TIM barrel which represents the of the enzyme [58]. The two halves of the catalytic domains are separated by linkers of different lengths, also called X-Y linker. Exceptions here are the PLC isozymes whose X-Y linker, also referred to as specific array (SA), is composed of a split pleckstrin homology (spPH) domain separated by two Src homology 2 (SH2) domains and one Src homology 3 (SH3) domain. The regulatory function of these X-Y linkers is outlined in section 1.3.4. Other common domains of PLCs are the C2 (calcium dependent lipid-binding) domain and the EF hand motifs. Both domains can bind Ca2+ which is an important co-factor for the catalytic activity of PLC isozymes [58, 182]. The pleckstrin homology (PH) domain is located at the N-terminus of all PLC isozymes, except for PLC, which does not harbour a PH domain at all. Investigations of the PLC1

PH domain demonstrated the high affinity binding of PIP2 to the domain allowing processive hydrolysis of the PLC substrate at the plasma membrane [39, 74, 149, 199, 300]. The N-terminal PH domain of PLC isozymes binds most specifically to membranes containing phosphatidylinositol 3,4,5-trisphosphate (PIP3) [61], whereas the split PH domain of PLC2 was shown to be essential for the activation of the enzyme by RacGTPases (see section 1.3.2 for details) [275].  subunits of heterotrimeric GTP binding (G) proteins bind to the PH domain of PLC2 thereby mediating the activation of the enzyme (see section 1.3.1 for details) [280, 281]. Beside to the so far depicted common structural features of PLCs, the PLC and PLC isoforms harbour additional isozyme specific domains. The PDZ-binding motif (PDZ is an acronym combining the first letters of three proteins - post synaptic density protein (PSD95), Drosophila disc large tumor suppressor (Dlg1), and zonula occludens-1 protein (zo-1) - which were first discovered to share the domain) at the C-terminal extension of PLC isozymes allows the binding of PDZ domain containing proteins potentially involved in the regulation of these PLC isoforms [252, 261]. The known isozyme specific domains of PLC are the Ras-GTPase exchange factor-like domain (RasGEF) at the N-terminus and the two Ras-associating (RA) domains at the C-terminus of the [129].

3 Introduction

Figure 1-1: Domain organization of mammalian PLC isozymes. Common structural features of PLC isozymes are the EF-hand motifs (EF), the catalytic TIM barrel (consisting of the two highly conserved catalytic subdomains X and Y), and the C2-domain (C2). With the exception of PLC, all PLC isozymes harbour a PH domain (PH) at their N-terminus. Isozyme specific domains are the Ras-GTPase exchange factor-like domain (RasGEF) and the two Ras-associating domains (RA) of PLC, as well as the PDZ- binding motif at the C-terminus of PLC. PLC isozymes harbour a unique region, also called specific array, inserted through a flexible loop between the two catalytic domains X and Y. This specific array harbours a second split PH domain separated by two SH2-domains (SH2) and one SH3-domain (SH3).

The tissue distribution of PLC isozymes is often closely linked to their physiological role found in transgenic animals. PLC1, PLC3, and PLC4 are highly expressed in different -/- -/- regions of the brain [210, 253], leading to epilepsy and ataxie in PLC1 and PLC4 null mice, respectively [131]. Mice lacking PLC3 show a decreased opioid sensitivity [297]. Although expressed in the same physiological compartment, these three PLC isoforms seem to have different physiological roles. In contrast, expression of the PLC2 isoform is restricted to the hematopoietic system [30, 120, 193, 237]. The two members of the PLC subfamily possess a rather different expression pattern. PLC1 is expressed ubiquitously, whereas the expression of PLC2, just like PLC2, is limited to cells of the hematopoietic -/- lineage [253]. In accordance with the widespread expression of PLC1, PLC1 mice die at embryonic day 9 [118], whereas PLC2 plays an essential role in the development and function of B cells [81, 277]. PLC1 is abundantly expressed in lung, heart, pancreas,

4 Introduction

skeletal muscle, and kidney [38, 148]. The mRNA of PLC3 is most common in brain, heart, lung, and testis [153]. PLC4 is present at very low concentrations in most tissues compared to other PLCs but shows the strongest expression in testis [146], where it is essential for the acrosome reaction in sperm [71, 72]. The PLC isozyme is mostly expressed in brain, liver, skeletal muscle, and heart [129, 241]. Mice deficient in PLC develop malformations of the heart and display a decreased cardiac function [257, 279]. The expression of both PLC isozymes is found to be neuron-specific in the brain [99, 181, 246], and the expression of PLC is most common in testis and prostate, especially in sperm [225].

1.3 Regulation of Phospholipases C

1.3.1 Regulation by heterotrimeric G proteins

GPCRs constitute the largest group within the family of transmembrane receptors. They were first discovered in the 1970s by Martin Rodbell and Alfred Gilman [154, 160], who received the Nobel prize in 1994 for this pioneering work. Almost 20 years later, in 2012, Robert Lefkowitz and Brian Kobilka received a Nobel prize for their work on GPCRs, again demonstrating the importance of these receptors in today’s medicine. Around 30 % of the currently available drugs target the function of GPCR family members [92, 109]. Within the GPCR signalling, the heterotrimeric G proteins account for the transduction of signals from the receptor to the effector system. G proteins consist of three subunits, designated G, G and G, whereat the two latter subunits form tightly associated G complexes [51]. G proteins are able to bind guanine nucleotides which are of great importance for their function as molecular switches. G proteins can switch between two states: (i) the inactive state where GDP is bound to G, and (ii) the active state where GTP is bound to G. After the binding of GTP, G dissociates from G, and the components transduce the signal to the effector proteins. The termination of the signal is achieved by the intrinsic GTPase catalytic activity of the G subunit, which hydrolyses GTP to GDP, allowing the reassociation of G and G[51, 173]. The switch between the two states of a G protein is tightly regulated by different accessory proteins which can be classified into three main types: GEFs (guanine nucleotide exchange factors) that increase the nucleotide

5 Introduction binding to G; GDIs (guanine nucleotide dissociation inhibitors) that inhibit the dissociation of GDP from G; GAPs (GTPase activating proteins) that accelerate the hydrolysis of GTP bound to G [224, 271].

To date there are 16 different G subunits encoding genes, 5 G subunit genes, and 12 G subunit genes known in human, theoretically allowing almost thousand combinations of heterotrimeric G proteins. However, not every combination is reasonable in signal transduction, and the number of those occurring in tissue and native cells is unproved [51].

The G subunits are classified into four families on the basis of sequence similarity: Gs

(stimulates adenylyl cyclase), Gi (inhibits adenylyl cyclase), Gq/11 (stimulates phospholipase C-), and G12/13 (regulates Rho GTPase signalling through Rho GEF activation) [136, 183, 250]. G subunits were shown to be not only negative regulators of G-dependent signalling but also bind to and activate downstream effector proteins upon stimulation of GPCRs, including adenylyl cyclases [260], ion channels [156, 305], PLC isoforms [308], protein kinases D [111], and phosphoinositide-3 kinase (PI3K) [130]. Furthermore, they are involved in the activation of the extracellular signal-regulated kinase (ERK)/mitogen-activated protein kinases (MAPK) pathway [45] and in chemotaxis [36].

1.3.1.1 PLC

PLC isoforms are activated by both, Gq/11 and G subunits [29, 30, 34, 126, 237]. All four members of the Gq/11 subfamily (Gq, G11, G14, G16) directly activate PLC isoforms, but not PLC and PLC isoforms [236, 262, 273]. It was previously assumed that this isozyme specific effect is mediated by the carboxy-terminal extension of the phospholipase which is uniquely present in PLC isozymes [147, 193, 227, 293]. However, a more recent study by the Harden group revealed that the absence of the carboxy-terminal extension of PLC is not affecting the high affinity binding of PLC to

Gq. The crystal structure of an activated complex of Gq with PLC3 exhibits an in PLC isozymes highly conserved helix-turn helix motif, which follows immediately the , as the Gq-binding interface [274]. The activation of PLC by members of the

Gq/11 subfamily is declining in the order PLC1 ≥ PLC3 >> PLC2, which is in contrast to the order observed for stimulation by G, PLC3 > PLC2 > PLC1 [192]. Despite their common capacity to activate PLC, G subunits target other structures within the

6 Introduction

lipase than Gq. The interaction site between PLC and G was mapped to both, the catalytic domain Y and the PH domain [223, 280, 281], whereas the presence of the catalytic domains X and Y seems to be sufficient for -dimer stimulation [102].

1.3.1.2 PLC2 and PLC

In addition to PLC isozymes, there is also evidence that PLC2 and PLC are direct downstream effectors of GPCRs. PLC2 is directly activated by G but not by G subunits, which was observed in intact cells as well as with purified proteins. This regulation of PLC2 seems to be an isozyme specific effect since there is no proof that

PLC1 is an effector molecule for G proteins. In opposition to PLC, the PH domain of

PLC2 appears to be dispensable for the activation by G [312, 313]. The last known

PLC to be stimulated by G proteins is PLC, which is selectively activated by both, G12 [158] and G [290] subunits. However, the mechanisms of activation by G proteins for both, PLC2 and PLC, have to be clarified yet.

1.3.2 Regulation by Ras superfamily GTPases

The Ras superfamily of small, monomeric GTPases comprises more than 150 members which can be classified into five subfamilies due to structural and functional similarity: Ras, Rho, Arf/Sar, Ran, and Rab [218]. Similar to heterotrimeric G proteins, monomeric GTPases act as molecular switches cycling between an active, GTP-bound state and an inactive, GDP-bound state. This GTPase cycle is controlled by (i) guanine nucleotide exchange factors (GEFs) that activate the GTPase by catalysing the exchange between GDP and GTP, (ii) GTPase activating proteins (GAPs) that inactivate the GTPase by stimulating the intrinsic GTPase activity, and (iii) guanine nucleotide dissociation inhibitors (GDIs) that keep the inactive GTPase in the cytosol and prevent GDP-GTP exchange [110]. Currently, four PLC isozymes are known to be directly regulated by monomeric GTP-binding proteins, namely PLC2 [100, 101], PLC2 [203], PLC[129,

241, 291], and PLC1 [231].

7 Introduction

1.3.2.1 PLC2 In the late 1990s, the Gierschik group demonstrated that bovine neutrophils contain a factor that stimulates PLC2 in a GTP dependent manner even in the absence of the carboxy-terminal region of PLC2 which is known to be essential for the activation of the lipase by Gq subunits [100]. Later on the factor was identified as a complex between a

RhoGTPase, Cdc42 and/or Rac1, and a Rho GDI, LyGDI. Thereby, PLC2 was the first mammalian target shown to be regulated by both, monomeric and heterotrimeric GTP- binding proteins [101]. Other GTPases that directly regulate PLC2 are Rac2 [102] and

Rac3 [238], and stimulation of PLC2 by Rac1 and Rac2 is much more pronounced than that by Cdc42. PLC isozymes are stimulated by Rho GTPases in the ranking of

PLC2 > PLC3 ≥ PLC1, which is quiet different from the activation by -dimers.

However, the regulation of PLC2 by  dimers and Rho GTPases appears to be independent of each other, since the stimulation of the lipase by both regulators is additive [102].

The between Rac and PLC2 was first investigated by Illenberger and coworkers, who replaced the N-terminal PH domain of PLC2 by the analogous region of

PLC1, showing that this chimeric protein is no longer activated by Rho GTPases [102]. The absolute requirement of the PH domain for Rac stimulation was further approved by surface plasmon resonance (SPR) experiments [238] and finally by a crystal structure displaying Rac1 bound to PLC2 [117]. The crystal structure clearly identifies the PH domain of PLC2 as the major effector site for Rac1. Comparison of the crystal structure of

PLC2 unbound and bound to Rac1 revealed that there are no conformational changes upon binding of Rac, indicating that the activation of PLC2 by Rac is solely mediated by localizing and orientating the lipase at membranes [83]. The translocation of PLC2 to the plasma membrane upon Rac stimulation was also demonstrated in intact cells by confocal fluorescence microscopy [103].

1.3.2.2 PLC2

Our group was able to demonstrate that PLC2, but not PLC1, is activated by constitutively active mutants of Rac1, Rac2, and Rac3, but not by Cdc42 or RhoA in intact cells as well as in a cell free system. The activation by Rac2G12V induced the translocation 8 Introduction of the lipase from the soluble to the particulate fraction. This indicates that, like it was shown for PLC2, the recruitment of the lipase to the substrate membrane is part of the activation of PLC2 by Rac GTPases. Furthermore, the mechanism of PLC2 activation by Rac GTPases is independent of tyrosine phosphorylation since replacement of all four tyrosine residues, known to be involved in phosphorylation-induced activation of PLC2, G12V by phenyalanins had no effect on the activation by Rac2 [203]. In contrast to PLC2, the N-terminal PH domain of PLC2 is not required for Rac-mediated stimulation. Instead, both halves of the spPH domain are necessary and sufficient for the isoform-dependent regulation by Rac2. The direct binding of Rac2 to full length PLC2 as well as to the isolated spPH domain was shown in a surface plasmon resonance approach, whereas for full-length PLC1 and the isolated PLC1 spPH domain no binding was observed. Nuclear magnetic resonance (NMR) spectroscopy data revealed that the two halves of the spPH domain form a canonical PH domain architecture with seven -strands and one -helix.

Despite the low sequence similarity (29 %) between the PLC1 and PLC2 spPH domains, the structures of both domains are quite comparable. Given this fact, three amino acid residues (K862, V893, F897), not conserved in PLC1, where identified to be important for

Rac-mediated activation of PLC2. Since mutation of all three residues (K862I, V893Q, F897Q) did not affect the correct folding of the spPH domain, it is likely that these residues directly affect Rac binding [275]. The crystal structure of Rac2 bound to the spPH domain of PLC2 revealed that the interaction of the spPH domain with Rac2 is mediated by the switch I and switch II domains of Rac2 [28].

1.3.2.3 PLC

The PLC isozyme contains two RA domains at the C-terminus, suggesting that it is maybe regulated by Ras. Indeed it was shown that PLC is directly regulated by H-Ras [129, 241] and Rap1a [241], as well as by RhoA, RhoB, and RhoC [291]. Kelley et al. demonstrated that H-Ras binds in a GTP-dependent manner to the RA2 domain, but not to the RA1 domain of PLC and subsequently activates the lipase in intact cells [129]. Upon activation by H-Ras, PLC is translocated from the cytosol to the plasma membrane [241]. RhoA, RhoB, and RhoC but not other Rho family members like Rac and Cdc42, stimulated PLC in intact cells. However, truncated PLC, missing the two RA domains, still responded to activation by Rho. Also, deletion of the RasGEF and the PH domain did not

9 Introduction affect responsiveness of PLC to Rho stimulation. By comparative sequence analysis a region within the catalytic core of PLC was identified, not present in other PLC subfamilies. Mutational analysis revealed that this region is indeed essential for the activation of PLC by RhoA, RhoB, and RhoC [291].

1.3.2.4 PLC1

The small GTPases RalA and RalB bind to and activate PLC1 in vitro as well as in vivo.

The activation is achieved by abrogating the inhibition of PLC by Calmodulin [231].

Furthermore, the RhoGAP p122, was also shown to activate the PLC1 isoform [89].

1.3.3 Regulation by phosphorylation

Phosphorylation of proteins is one of the major regulatory mechanisms in a wide range of cellular processes. Regulation of the activity of PLC enzymes by phosphorylation is known for PLC, PLC and PLC isozymes. The regulation of PLC1 and PLC2 by receptor or non-receptor tyrosine kinases is studied most extensively and will be described more precisely in sections 1.3.3.1 and 1.3.3.2. PLC2 and PLC3 isoforms are serine- phosphorylated by cAMP-dependent kinase (PKA) and/or PKC which was shown to attenuate their activation by heterotrimeric G proteins [155, 302, 303]. ERK phosphorylates PLC1, thereby mediating the mitogenic activity of insulin-like growth factor 1 (IGF-1) [298]. PLC1 binds to PKC via its PH domain, leading to the phosphorylation and subsequent activation of the lipase [68].

1.3.3.1 Regulation of PLC isoforms by receptor tyrosine kinases

Receptor tyrosine kinases (RTK) are cell surface receptors that are of great significance for many fundamental cellular processes like proliferation, differentiation, cell survival, metabolism and cell cycle control. Their molecular architecture is composed of a ligand binding extracellular part, a single transmembrane helix, and a cytoplasmic part containing the protein tyrosine kinase (TK). Binding of an appropriate ligand induces the dimerization of two receptor molecules giving rise to the mutual phosphorylation of the two constituents at specific tyrosine residues. The resulting phosphotyrosines serve as binding sites for 10 Introduction effector molecules that contain SH2 or PTB (phosphotyrosine-binding) domains, such as PLC or PI3K. The recruited proteins are in turn phosphorylated at specific tyrosine residues by the RTK [150, 210]. PLC1 is phosphorylated in intact cells upon stimulation of e.g. epidermal growth factor (EGF) receptor kinase or platelet-derived growth factor (PDGF) receptor kinase on tyrosine residues 771, 783, and 1254 [132, 133, 185, 272]. Mutational analysis revealed that at least phosphorylation of Tyr783 is essential for the activation of PLC1 in response to RTK engagement. The function of phosphorylation on Tyr771 and Tyr1254 are currently unknown [132]. In contrast, the in literature existing in vitro data for the phosphorylation induced activation of PLC1 are controversial [133, 185].

The interaction of PLC1 with activated growth factor receptors, such as the EGF and

PDGF receptor, is mediated by the amino-terminal SH2 domain (SH2N) of PLC1 [35, 119,

206], whereas the C-terminal SH2 domain (SH2C) was shown to associate intramolecularly with phosphorylated Tyr783 [207]. Most interestingly, Tyr771 and Tyr783 are located within the PLC SA region, more precisely in the linker between the SH2C and the SH3 domain. Since the SA is involved in the autoinhibitory regulation of the lipase, Gresset et al. postulated a mechanism for the phosphorylation induced activation of PLC in which tyrosine phosphorylation of the SA leads to large conformational rearrangements, that cause the release of autoinhibition and thus activation of the enzyme [77]. Furthermore, the second PLC isoform, PLC2, is tyrosine phosphorylated and activated by PDGF stimulation in rat fibroblasts overexpressing PLC2 [256]. These results were confirmed in rabbit vascular smooth muscle cells (VSMCs), endogenously expressing PLC2, which represents a more natural surrounding [90].

1.3.3.2 Regulation of PLC isoforms by non‐receptor tyrosine kinases

Non-receptor tyrosine kinases are, in contrast to RTK, cytoplasmic proteins that play important roles in a large number of signal cascades, especially in the immune system, where they are critical constituents in the signal transduction of activated B and T cells. In consequence of the abundant expression of PLC2 in B cells, and the predominant expression of PLC1 in T cells, the activation of these two proteins has been studied most extensively downstream of the B cell receptor (BCR) and the T cell receptor (TCR), respectively.

11 Introduction

The BCR complex consists of an antigen-binding membrane immunoglobulin (mIg) associated with Ig/Ig heterodimers, containing immunoreceptor tyrosine activation motifs (ITAMs). While the mIg is responsible for the binding of antigens, the Ig/Ig heterodimers permit the transmission of the signal to the interior of the B cell. Upon BCR engagement, ITAMs are phosphorylated by the Src family kinase Lyn, triggering the formation of the signalosome composed of tyrosine kinases such as the spleen tyrosine kinase (Syk) and Bruton´s tyrosine kinase (Btk), of adaptor molecules like B cell linker

(BLNK) and cluster of differentiation (CD)19, and of signalling enzymes such as PLC2, Vav, or PI3K. Signals emitted from the signalosome initiate Ca2+ signalling in which

PLC2 was shown to play a critical role [80, 205]. Studies with chicken DT40 B cells deficient in Lyn, Syk or Btk revealed that at least Btk phosphorylates PLC2 upon BCR stimulation [64, 258, 259]. However, it is hard to prove, if PLC2 is a substrate for Lyn and

Syk kinases as well. Btk phosphorylates PLC2 in vitro as well as in intact B cells on tyrosine residues 753, 759, 1197, and 1217, whereas solely Tyr753 and Tyr759 are important for regulating the activity of the lipase [97, 135, 191, 217, 283]. The phosphorylation of PLC2 by Btk is facilitated by means of the adaptor molecule BLNK, which is accountable for bringing both molecules into close proximity. Upon BCR activation BLNK is phosphorylated by Syk kinase, thus providing a docking site for SH2 domains of PLC2 and Btk [67, 107, 140, 217].

Activation of the TCR results in the phosphorylation and subsequent activation of the

PLC1 isoform. The activation and formation of the TCR signaling complex is quite similar to those observed in B cells. Stimulation of the antigen-specific CD3/T-cell receptor complex leads to the phosphorylation of ITAMs, located at the cytoplasmic part of the TCR, by lymphocyte protein tyrosine kinase (Lck). The phosphotyrosines serve as binding sites for the SH2 domains of Zap-70 (-chain associated protein kinase of 70 kDa) which is phosphorylated and activated, and in turn phosphorylates downstream adaptor proteins such as LAT (linker of activated T cells) and SLP-76 (SH2-domain-containing leukocyte protein of 76 kDa). Phosphorylation of LAT and SLP-76 results in the recruitment of several adaptor and signalling proteins to form a multi protein complex, the so called LAT signalosome. Important proteins that assemble this complex, beside LAT and SLP-76, are PLC1, growth factor receptor-bound protein 2 (Grb2), Vav1, and interleukine-2-inducible T cell kinase (Itk). The three major pathways addressed by the 2+ LAT signalosome are Ca , MAPK, and nuclear factor B (NF-B) signalling [25]. PLC1 12 Introduction is phosphorylated and activated in response to TCR engagement and the major tyrosine phosphorylation sites are the same as observed in cells treated with EGF and PDGF (see section 1.3.3.1) [195, 285]. The recruitment of PLC1 by phosphorylated LAT is mediated via the SH2N of PLC1 and represents an essential step in the phosphorylation induced activation of the lipase upon TCR stimulation [232, 247, 310]. The kinases, however, responsible for the phosphorylation of PLC1 in T cells are still unknown, but possible candidates are ZAP-70, Itk, or Lck. It is worth mentioning that PLC1 is also serine phosphorylated in response to both TCR and EGF stimulation. The phosphorylation of serine 1248 is mediated by PKA and PKC and is likely to exert an inhibitory effect on the lipase [8, 134, 194, 195]. Nevertheless, the role of serine phosphorylation of PLC1 is widely unknown so far.

1.3.4 Autoinhibitory regulation

All previously described regulatory mechanisms exert a positive activating influence on PLC isozymes. However, it is known that PLCs exhibit a specific activity (20-30 µmol/min/mg protein) even in the absence of stimulators like G proteins. Due to the fact that the amount of PIP2 in the brain is on average 2-14 nmol/mg protein and the total amount of PLCs in the brain is 2µg/mg protein, the entire PIP2 would be hydrolysed within 2-20 sec, even without receptor stimulation. The consequences would be a needless Ca2+ signal and waste of energy, as the resynthesis of PIP2 out of DAG and inositol requires 5 equivalents of ATP [212]. Since the human organism is directed very economically, there has to be an inhibitory repressor of PLC activity in unstimulated cells. In the 1990s a general model for the inhibition of PLC activity by autoinhibitory elements within the linker between the two halves of the catalytic domain, also called X-Y linker, was proposed. This includes, among others, activation of the enzyme by limited proteolysis, reassembly of separately expressed X and Y polypeptides and the influence of deletion mutants on the activity of the lipase. Limited protease digestion of PLC1 [39, 55], PLC1

[62], and PLC2 [226] within their X-Y linkers as well as the deletion of the SH domain tandem of PLC1 [93] resulted in elevated PLC activity. Furthermore, recombinant expressed SH2SH2SH3 fragments of PLC1 and PLC2 inhibited PLC1, PLC2, PLC1, and PLC1 in a cell free system [91] and the functional reassembly of independently expressed N- and C-terminal halves of PLC1 [94] and PLC2 [309], missing the X-Y 13 Introduction linker, resulted in an increase in basal PLC activity. Taken together, these findings point to a conserved role of the X-Y linker in the autoinhibitory regulation of PLC isozymes. A general model for the autoinhibitory regulation of PLC isozymes was postulated by Hicks et al., where they demonstrated for PLC2, PLC1, and PLC isozymes, that the deletion of a small part of the X-Y linker is sufficient to activate the enzyme [83]. The crystal structure of PLC2 finally revealed the occlusion of its active site by the X-Y linker. Upon engagement of PLC2 by Rac, the lipase is recruited to the plasma membrane where the highly negatively charged X-Y linker is repelled from the also negatively charged plasma membrane. By this mechanism, the access to the active site of PLC2 is opened and the enzyme is able to catalyze the hydrolysis of its substrate [83]. Since PLC isozymes harbour a structurally very distinct, multidomain X-Y linker (for details see section 1.2.1), it is likely that there is a different mechanism for the autoinhibitory regulation than it was shown for the other PLCs. A model for the intramolecular regulation of PLC1 activity was proposed by De Bell et al., in which they state that the SH2C domain inhibits the enzyme in its basal activity and that the availability of this domain is regulated by the N-terminal half of the spPH domain. Furthermore, the SH2N domain seems to participate in the release of the SH2C domain from its inhibitory constraints [48]. It was additionally demonstrated that phosphorylation of PLC1 within its X-Y linker leads to conformational rearrangements of the SH2C domain with respect to the catalytic domain, resulting in the release of autoinhibition [77]. On the other hand, there is the possibility, that the activation by Rac2 and/or the proximity of PLC2 to the plasma membrane induces the liberation of the enzyme from its autoinhibitory constraints [59].

1.4 Role of PLC2 in inflammatory and autoimmune diseases

In the last decade, several mutations in PLC2 had been associated with inflammatory and autoimmune diseases, in mouse models as well as in humans. A genome-wide N-ethyl-N- nitrosourea (ENU) mutagenesis approach was chosen to identify genetic alterations in genes involved in inflammatory and autoimmune diseases in the mouse. Two of the generated mouse strains exhibited point mutations in the PLCG2 , named Ali5 [301]

14 Introduction

(Abnormal limb) and Ali14 [1]. The phenotype of Ali5 mice is characterised by chronic inflammation of the paws and severe eye deformation which is triggered by proliferation of cells in the epithelium and a thinning of the sclera. The Ali5 point mutation is localized in the catalytic subdomain Y, where the aspartic acid on position 993 is substituted by a Ali5 glycine [301]. The biochemical characterisation of PLC2 revealed that the mutation has no impact on the basal activity of the lipase, neither in vitro, nor in intact cells. However, in B the cells Ali5-type mutation caused an increase in sustained Ca2+ entry in response to antigen receptor stimulation. The mechanistic background of this observation is thought to Ali5 be an increased stability and penetrability of PLC2 at the plasma membrane [60, 301]. Ali5 2+ The same effect was observed for platelets, where PLC2 leads to enhanced Ca mobilization, integrin activation and granule secretion and ultimately to a prothrombotic phenotype in mice [56]. The phenotype of Ali14 mice is quite similar to this of Ali5 mice, also suffering, amongst other things, from spontaneous swelling and inflammation of the hind paws. Furthermore, cultured B cells from these mice show an enhanced calcium Ali14 influx. The amino acid residue mutated in PLC2 is located within the N-terminal split Ali5 PH domain, where tyrosine 495 is replaced by a cystein [1]. As it was shown for PLC2 , also the Ali14-type mutation had no impact on the catalytic activity of the enzyme, but their combination resulted in a constitutively active PLC2 enzyme [60]. The importance of a tight regulation of PLC2 activity has become increasingly clear as the loss of autoinhibitory regulation, by either point mutations [311] or larger deletions [190] in the autoinhibitory region of PLC2 was associated with severe inflammatory phenotypes in humans. Just recently Ombrello et al. identified a novel disease pattern characterized by cold-induced urticaria, immunodeficiency, and autoimmunity. Since this unique phenotype could be linked to deletions in the human PLCG2 gene within the autoinhibitory region between the two catalytic subdomains X and Y, the syndrome was designated PLAID

(phospholipase C-2-associated antibody deficiency and immune dysregulation) [190]. PLAID patients suffer, amongst other things, from a previously undescribed form of cold- induced urticaria, the so called familial atypical cold urticaria (FACU) [73]. Cold urticaria, in general, belongs to the physical urticarias triggered by a cold stimulus, provoking whealing and itching of the cold exposed skin [162]. Affected individuals had lifelong symptoms from early childhood on, including pruritis, erythema, and urticaria after cold exposure like cold atmosphere, aquatic activities, handling of cold objects, and ingestion of cold food or beverages. Skin biopsy after a cold challenge revealed an increase in mast cell degranulation compared to healthy controls. However, the processes leading to mast cell 15 Introduction degranulation are poorly understood. In contrast to the other cold urticarial forms, FACU is limited to the exposed skin without a generalized and systemic reaction [73]. Beside this cold-induced unique inflammatory phenotype, PLAID patients also display various other manifestations like recurrent sinopulmonary infections, antibody deficiency, symptomatic autoimmune and allergic disease, and common variable immunodeficiency. Furthermore, the levels of circulating CD19+ B cells, IgA+ and IgG+ class-switched memory B cells, and natural killer cells are diminished in these patients. The characterisation of PLAID B cells exhibited a reduced in vitro class switched antibody production and poor in vitro expansion. As mentioned above, the described phenotype could be linked to two different exon deletions in the PLCG2 gene. In two of the three affected families heterozygous deletions of exon 19 (19) were identified, whereas in the third family exons 20 through 22 (20-22) were deleted on one allele [190]. Both deletions affect the SH domain tandem of PLC2, more precisely the C-terminal SH2 domain and the SH3 domain, which were shown to be important for the autoinhibitory regulation of the enzyme [59]. Indeed, both deletion mutants showed enhanced basal activity in transfected COS-7 cells compared to wild-type PLC2. In spite of this gain in enzymatic activity, downstream signalling in stimulated PLAID B cells was diminished. However, unstimulated B cells from those subjects showed a clear increase in cytosolic Ca2+ levels with decreasing temperatures and mast cells transfected with PLC219 or PLC220-22 displayed spontaneous degranulation at 20 °C but not at 37 °C [190]. Whole exon sequencing of another family revealed a hypermorphic missense mutation in PLCG2 that causes a dominantly inherited autoinflammatory disease with immunodeficiency [311]. Most interestingly, these patients do not suffer from cold urticaria. The described disease is characterized by recurrent blistering skin lesions, bronchiolitis, arthralgia, ocular inflammation, enterocolitis, absence of autoantibodies, and mild immunodeficiency. The detected missense mutation is localized in the C-terminal SH2 domain were serine 707 is replaced by tyrosine. To distinguish the described phenotype, caused by a single amino acid substitution, from

PLAID, caused by genomic deletions in PLCG2, the term APLAID (autoinflammation and

PLC2-associated antibody deficiency and immune dysregulation) was proposed [232]. Zhou et al. [311] and unpublished data from our group showed that the basal activity of S707Y PLC2 is elevated compared to wild-type PLC2. Consequently, but in contrast to

PLAID, PLC2-dependent signalling was elevated in APLAID patients, which was 2+ confirmed by elevated levels of IP3 production, intracellular Ca release, and ERK phosphorylation. However, the assumption that the substitution of serine for tyrosine 16 Introduction creates a novel phosphorylation site, which compromises autoinhibition and thereby leads to an increase in basal activity [311], could be rebuted by unpublished experiments from our group.

1.5 Anaplastic lymphoma kinase

1.5.1 Structure, Expression, and Function

Anaplastic lymphoma kinase (ALK) was originally described in 1994 as an oncogene activated in anaplastic large cell lymphomas (ALCL) with the chromosomal translocation t(2;5) [230], causing the fusion of the kinase domain of ALK to the amino terminal half of nucleophosmin (NPM), generating a fusion gene, NPM-ALK (see section 1.5.3.1) [174]. The molecular characteristics of ALK itself were not elucidated until 1997. ALK displays the classical structural features of a RTK containing an extracellular ligand-binding domain, a membrane spanning domain and the intracellular kinase domain. Based on homology ALK constitutes a subgroup together with LTK (leucocyte tyrosine kinase) within the IR (insulin receptor) superfamily [108, 175] and is highly conserved across species [209]. The human ALK gene encodes a protein of 1620 amino acids with a molecular weight of approx. 180 kDa. Posttranslational modifications such as N-linked glycosylation are generating a mature ALK with a molecular weight of approx. 220 kDa [108, 175]. The expression of murine ALK mRNA was preliminary found in the central and peripheral nervous system during embryonic development but at considerably lower concentrations in the adult nervous system [108, 175]. Its expression is also observed in the developing sensory organs, reproductive organs, skin and stomach [266]. The fading expression of ALK mRNA from developmental to adult stages in mice was also found in the developing central nervous system (CNS) of chicken, where ALK localizes to a subset of DRG (dorsal root ganglia) [98]. A similar expression pattern was also found in a subset of DRG neurons in the rat [50]. Furthermore, ALK mRNA was as well detected in the developing CNS in Drosophila [159]. Mice homozygous for a deletion of ALK have normal appearance and no obvious tissue abnormalities but they revealed an age-dependent increase in basal hippocampal progenitor proliferation and alterations in behavioural tests [16]. The role of ALK in neuronal development is further supported by in vitro studies where it was shown that stimulation of human neuroblastoma (SK-N-SH) cells,

17 Introduction endogenously expressing ALK, by an agonist monoclonal antibody induces activation of ALK which subsequently leads to the activation of the MAP kinase signaling cascade. The activation of this signaling cascade is driving the cells to proliferation and neurite outgrowth [177, 242]. Thus, ALK RTK appears to have an important role in the normal development and function of the central and peripheral nervous system. Nonetheless its physiological role is still enigmatic.

The control of the activity of RTKs can be achieved by expression levels, binding of ligands, or mutations that inhibit or enhance constitutive receptor activity [286]. So far, two activating ligands for ALK were identified, namely pleiotrophin (PTN) [248] and midkine (MK) [249]. Binding of the ligands induces homodimerization of ALK, leading to trans-phosphorylation and activation of the kinase [248]. Both ligands are heparin-binding growth factors and form an own family within the heparin-binding growth factor families. They are involved in processes such as neural development, cell migration and angiogenesis [123, 141, 180]. The expression of both ligands is similar to that of ALK during embryonic development [208]. Just recently it was shown by Allouche that ALK is a novel dependence receptor, meaning that in the absence of ligands the ALK receptor is inactive and its expression is leading to apoptosis, whereas activation of ALK by ligands or as a result of ALK fusion proteins, decreases apoptosis [3, 178].

1.5.2 ALK in disease

Overexpression and constitutive or excessive activation of ALK can lead to the deregulation of cell proliferation and survival resulting in various human cancers. Since the important first description of the ALK fusion protein NPM-ALK in ALCL [174, 230], many other ALK chimeras have been described in diseases like inflammatory myofibroblastic tumors (IMTs) [78], diffuse large B-cell lymphoma (DLBCL) [6], non- small cell lung cancer (NSCLC) [216, 239], esophageal squamous cell carcinoma (SCC) [53, 115], colon, and breast carcinoma [152]. The expression of ALK was also documented in a variety of cancer cell lines where it was shown that overexpression or gain of function mutants of ALK can lead to tumorigenesis (reviewed in [11])

18 Introduction

1.5.2.1 Point mutations, overexpression, and amplification of full‐ length ALK in neuroblastoma

Activating ALK mutations have been reported to be associated with sporadic and familial cases of neuroblastoma [32, 37, 114, 176], which was first described in 2000 [143]. All together there are 12 different residues known to be mutated in neuroblastoma. However, the two master mutation hotspots were found to be F1174L and R1275Q [7]. These mutations were observed in 49 % and 34.7 % of mutated cases, respectively [24]. Both mutations lead to an increase in the catalytic activity of the enzyme and their oncogenic potential was demonstrated by transformation of NIH-3T3 fibroblasts and tumor formation in nude mice [37, 145]. Furthermore, it was shown that both mutations are associated with a defect in N-linked glycosylation of the kinase and that the constitutive activation leads to impaired receptor trafficking resulting in the retainment of the kinase in the endoplasmic reticulum/Golgi compartments [165]. The just recently elucidated crystal structure of ALK in its inactive conformation revealed the autoinhibition of ALK by its C helix. Most interestingly, the majority of the ALK mutations found in neuroblastoma, including F1174L and R1275Q, are located in or near by the area of this autoinhibitory loop. Therefore, these activating mutations are thought to release ALK from its autoinhibitory constraint by allowing the unrestrained mobility of the C helix [145]. Furthermore, these mutations were shown to be associated with constitutive phosphorylation of ALK and its downstream targets [32, 37, 114, 176] and enhanced ATP binding affinity of the kinase [145]. Beside the described gain of function mutations there is also evidence for the amplification of the ALK gene leading to neuroblastoma in about 2 % of neuroblastoma cases [24]. In addition to ALK mutations and amplification of the ALK gene, overexpression of the wild-type ALK protein was also reported to lead to oncogenic activity in neuroblastoma cells. It seems, that a critical threshold of ALK wild-type expression is necessary to achieve this oncogenic activation [198]. However, this mutation- independent overexpression of full-length ALK is still uncharacterized at the molecular level.

19 Introduction

1.5.3 ALK fusion proteins

1.5.3.1 NPM‐ALK

Among the known ALK fusion proteins NPM-ALK is the most studied and the most frequent fusion protein (80 % of translocations), which is found in 40 % to 60 % of ALCL patients [5]. It is created via the recurrent chromosomal translocation t(2;5)(p23;q35) which was found in the late eighties by several groups [12, 63, 124, 164]. Only four years later it was demonstrated that this translocation creates the fusion gene between a so far unknown tyrosine kinase on 2, and the NPM gene on chromosome 5 [174, 230]. Owing to its association with ALCL the kinase was termed ALK and its expression in ALCL led to the clinically distinct entity ALK+ ALCL [201]. NPM-ALK is a 80 kDa fusion protein containing the first 117 amino acid residues of NPM fused to the cytoplasmic part of ALK starting from residue 1058 [174]. Most interestingly, all so far identified ALK fusion proteins contain exactly the same portion of ALK [54]. NPM itself plays multiple roles in cell growth and cell proliferation and is localized mainly at the nucleolus [189]. There it serves as a bidirectional shuttle for proteins between the nucleus and the cytoplasm [19]. The amino terminal half of NPM, which is fused to the kinase domain of ALK, harbours an oligomerization domain and two nuclear export signals [189]. It was shown that the NPM segment leads to the oligomerization of NPM-ALK resulting in the autophosphorylation of the kinase which is absolutely required for the kinase activity of the chimera [17]. Even more important, NPM-ALK lacking the NPM segment is no longer able to transform NIH 3T3 cells, implying that the NPM segment is essential for the oncogenic potential of NPM-ALK [69]. The presence of a dimerization/oligomerization domain is also a common feature of other ALK fusion partners [201]. In contrast to full- length ALK, NPM-ALK is ectopically expressed in lymphoid cells and localized in the cytoplasm and in the nuclear compartments [17].

1.5.3.1.1 Oncogenic signalling of NPM‐ALK

To date four pathways were identified to be activated by NPM-ALK in ALCL, designated as the PLC, PI3K/Akt, Ras/MAPK, and JAK/STAT pathways (Figure 1-2). STAT (signal transducer and activators of transcription) family members are activated by tyrosine phosphorylation which is in particular mediated by Janus Kinases (JAKs). The subsequent dimerization of STATs allows their translocation to the nucleolus and the activation of transcription of genes involved in proliferation, cell survival, and immune response. The

20 Introduction

JAK/STAT pathway is activated by interferon and cytokine receptor signalling [245]. It was shown that ALK+ ALCL exhibit activated STATs and in turn that STATs are required for the transforming activity of NPM-ALK in mouse embryo fibroblasts. In agreement, inactivation of NPM-ALK with small molecule inhibitors leads to a decreased phosphorylation of STATs. This activation of STATs by NPM-ALK was shown to be a result of the interaction between JAKs and NPM-ALK [201].

Figure 1-2: Oncogenic signalling of NPM-ALK. The so far known pathways addressed by NPM-ALK are the PI3K/AKT, the JAK/STAT, the MEK/ERK and the PLC pathway. For further details see main text.

PI3K was identified as a further downstream signalling molecule of NPM-ALK. It was shown that PI3K and its downstream effector, serine/threonine kinase (Akt), are activated by NPM-ALK and that this activation is induced by the recruitment of the C-terminal SH2 domain of PI3K by NPM-ALK. The activation of this antiapoptotic signalling pathway is important for the transforming potential of NPM-ALK, since the inhibition of PI3K or Akt by inhibitors or dominant-negative mutants is leading to growth arrest or even apoptosis of NPM-ALK+ cells [10, 234].

21 Introduction

The third known pathway addressed by NPM-ALK is the MEK/ERK pathway. This pathway is part of a kinase cascade, the MAPK pathway, which is activated by growth factor receptors. Major cellular processes such as differentiation, cell survival, and proliferation are controlled by this signalling cascade [169]. Co-immunoprecipitation studies showed that NPM-ALK interacts with insulin receptor substrate 1 (IRS-1), Shc, and growth-factor-receptor-bound protein 2 (Grb2) [9, 69]. All three mentioned molecules are responsible for the transmission of signals from growth factor receptors to the MEK/ERK pathway [169]. However, NPM-ALK mutants deficient in interacting with IRS-1 and Shc retain their ability to transform NIH 3T3 cells [69]. In line with these results, there is also evidence that NPM-ALK directly interacts with MEK kinases [46, 295].

The most interesting finding for this work was the link between the PLC pathway and NPM-ALK. Bai et al. already showed in 1998 with pull-down assays that NPM-ALK binds to both SH2 domains of PLC, at which the strongest binding was observed with the N-terminal SH2 domain. The interaction of NPM-ALK and PLC leads to the phosphorylation of the lipase and thereby, most presumably, to the activation of the enzyme. Ba/F3 cells overexpressing NPM-ALK also showed an enhanced inositol phosphate production, pointing again to an activation of PLC by NPM-ALK. Furthermore, Bai et al. identified Tyr664 in NPM-ALK to be essential for the complex formation of the kinase and PLC. In contrast to wild-type NPM-ALK, NPM-ALKY664F did not lead to the transformation of Rat-1 fibroblasts. Taken together these results imply that PLC is an important downstream target of NPM-ALK which accounts for the mitogenic potential of the constitutively active kinase [9]. Somehow conflictive are the results published by Dirks et al.. They indeed reported the phosphorylation of PLC in ALK+ ALCL-derived cell lines, but they mentioned that in these cell lines PLC proteins are even phosphorylated in serum-starved ALCL-cells and that there is no direct proof that NPM- ALK provokes this phosphorylation [52]. The possibility that the activation of PLC by NPM-ALK is cell type specific was further approved by Piccinini et al. [202]. Moreover,

PLC1 was identified in a mass spectrometry approach to be a NPM-ALK interacting protein [46]. Beside this potentially direct interaction between NPM-ALK and PLC there is also evidence that NPM-ALK can activate Rac1 GTPase [42], which is in turn a direct activator of PLC2 [203]. Vav3 is one of the GEFs involved in the activation of Rac1 [42]. The importance of Rac activation in NPM-ALK driven disease progression and metastasis was demonstrated by Rac inhibitors [43]. 22 Introduction

1.5.3.2 EML4‐ALK

Lung cancer is the worldwide second frequent cancer after prostate cancer and breast cancer for men and women, respectively. However, the mortality of lung cancer is the highest among the various cancer forms for both sexes [116]. The fusion of ALK to the echinoderm microtubule associated protein-like 4 (EML4) was only recently described in non-small cell lung cancer (NSCLC) [216, 239]. The basis of this fusion is a small inversion within chromosome 2p (Inv (2) (p21p23)) that joins the N-terminal part of EML4 to the exons 20-29 of ALK, representing the intracellular kinase domain. As it was demonstrated for NPM-ALK, the EML4 fusion partner is mediating ligand-independent dimerization of ALK, leading to the constitutive activation of EML4-ALK. The oncogenic potential of EML4-ALK was demonstrated in cell culture systems and in nude mice [239]. EML4-ALK transgenic mice, expressing the fusion protein specifically in lung alveolar epithelial cells, develop adenocarcinoma a few weeks after birth. Administration of small molecule inhibitors of the kinase activity of EML4-ALK can stop the progression of the tumours and even result in a complete disappearance of tumour tissue [240]. Thus, EML4- ALK seems to play a crucial role in the development of NSCLC and could therefore be a promising target for drug development.

1.6 Aim of the work

The regulatory network of PLC2 consists of three main mechanisms, (i) the activation by RhoGTPases, (ii) the phosphorylation-induced activation, and (iii) the autoinhibitory regulation of the lipase. In the last decade it has become abundantly clear that misregulations of the activity of PLC2 can lead to severe phenotypes in mice as well as in human. A first aim of the present doctoral thesis was to further investigate the molecular mechanisms underlying the autoinhibitory and phosphorylation-dependent regulation of

PLC2, and to identify the structural elements involved in these regulatory processes. The association of exon deletions in the PLCG2 gene with a novel disease pattern, named

PLAID (phospholipase C-2-associated antibody deficiency and immune dysregulation), characterized by cold-induced urticaria, immunodeficiency, and autoimmunity was just recently described. A further aim of this thesis was to investigate the consequences of cold 23 Introduction

temperatures on the activity of the PLC2 mutants associated with PLAID and to identifiy structural elements that mediate the temperature sensitivity of the enzyme.

24 Materials and Methods

2 Materials and Methods

2.1 Materials

2.1.1 Chemicals

Acrylamide 4K 30 % (w/v) (37.5:1 mix) AppliChem, Darmstadt Agar Serva, Heidelberg Agarose Sigma, Deisenhofen Albumine bovine fraction V (BSA) Serva, Heidelberg Ammonium formate Fluka, Taufkirchen Ammonium persulphate Serva, Heidelberg Ampicillin, sodium salt Sigma, Deisenhofen Aprotinine (Trasylol®) Bayer, Leverkusen Bacto Trypton Becton Dickinson, Heidelberg Benzamidine Sigma, Deisenhofen Bromphenol blue Serva, Heidelberg Crystal violet Sigma, Deisenhofen Dithiothreitol (DTT) Serva, Heidelberg DMEM Medium (# 41965-039) Gibco Life Technologies, Karlsruhe Deoxynucleoside triphosphate (dNTP) mix MBI Fermentas, St. Leon-Rot Dimethyl sulfoxide (DMSO) Roth, Karlsruhe - Dowex resin (1x8, 100-200 mesh, Cl form) Sigma, Deisenhofen Dulbecco´s PBS (1x) (# H15-002) PAA, Cölbe ECL™ Western Blotting Detection Reagent Amersham, Freiburg ECL Western Blotting Substrate Thermo Scientific, Schwerte Ethylene diamine tetraacetic acid (EDTA), disodium salt Merck, Darmstadt Ethidium bromide (EtBr) AppliChem, Darmstadt Fetal calf serum (FCS, used in mammalian and avian cell culture) Gibco Life technologies, Karlsruhe

25 Materials and Methods

Formic acid Roth, Karlsruhe L-Glutamine PAA, Cölbe Glycerine Roth, Karlsruhe Glycine AppliChem, Darmstadt 4-(2-Hydroxyethyl)-1-piperazineethanesulfonic acid HEPES (for cell culture) PAA, Cölbe High molecular weight protein marker (HMW) Sigma, Deisenhofen Ink Pelikan, Hannover Kanamycin Serva, Heidelberg Lambda DNA/EcoRI+HindIII Marker, 3 MBI Fermentas, St. Leon-Rot Leupeptine Serva, Heidelberg TM Lipofectamine 2000 Transfection reagent Invitrogen, Karlsruhe Lithium chloride Sigma, Deisenhofen Loading dye (6x) MBI Fermentas, St. Leon-Rot jetPRIME® Polyplus Transfection, Peqlab, Erlangen β-Mercaptoethanol Roth, Karlsruhe Milk powder Fluka, Neu-Ulm Milk powder Roth, Karlsruhe OptiMEM I Medium Gibco Life technologies, Karlsruhe PageRulerTM Broad Range Unstained Protein Ladder Thermo Scientific, Schwerte Penicillin/Streptomycin PAA, Cölbe Pepstatin A Serva, Heidelberg Phosphatase Inhibitor Cocktail 2 Sigma, Deisenhofen Phenylmethylsulphonyl fluoride (PMSF) Sigma, Deisenhofen Polyoxyethylene-sorbitan monolaurate (Tween 20) Sigma, Deisenhofen Protein A-Agarose Roche Applied Science, Mannheim Pyronin Y Sigma, Deisenhofen Quicksafe A Scintillation Fluid Zinsser, Frankfurt RTU 60 Developer Adefo, Dietzenbach RTU 60 Fixer Adefo, Dietzenbach

26 Materials and Methods

Sodium chloride AppliChem, Darmstadt Sodiumdodecyl sulfate (SDS) Roth, Karlsruhe Sodium formate Fluka, Neu-Ulm Sodium pyruvate PAA, Cölbe Sodium tetraborate decahydrate (Borax) Sigma, Deisenhofen Soybean trypsin inhibitor (STI) Fluka, Neu-Ulm N,N,N',N'-Tetramethyl-ethane-1,2-diamine (TEMED) Serva, Heidelberg Thimerosal Sigma, Deisenhofen Trishydroxymethylaminomethane (Tris) Roth, Karlsruhe Trypan blue Merck, Darmstadt Trypsin-EDTA solution (# L11-004) PAA, Cölbe Tween 20 Sigma, Deisenhofen Water for chromatography (LiChrosolv®) Merck, Darmstadt Yeast Extract Servabacter Serva, Heidelberg

All chemicals and reagents used but not listed were obtained in p.a. quality from the following companies: Merck (Darmstadt), Riedel-de Haen (Seelze), or Roth (Karlsruhe).

2.1.2 Radiochemicals myo-[2-3H]Inositol with PT6-271 stabilizer PerkinElmer, Rodgau-Jügesheim

2.1.3 Enzymes and Kits

Antarctic Phosphatase New England Biolabs, Frankfurt Big Dye™ Terminator Ready Reaction Mix PE Biosystems, Wertigheim ® Nucleobond AX PC-Kit-100 Macherey-Nagel, Düren ® Nucleospin Kit Macherey-Nagel, Düren QIAquick Gel Extraction Kit Qiagen, Hilden ® QuikChange II XL Site-Directed Mutagenesis Kit Agilent Technologies, Waldbronn

27 Materials and Methods

™ PfuUltra High-Fidelity DNA Polymerase Agilent Technologies, Waldbronn ™ Phusion High-Fidelity DNA Polymerase Thermo Scientific, Schwerte T4 DNA Fermentas, St. Leon-Rot ® Zero Blunt PCR Cloning Kit Invitrogen, Karlsruhe

All restriction were purchased from Fermentas, St. Leon-Rot or New England Biolabs, Frankfurt.

2.1.4 Consumables 6-well-plate Renner, Dannstadt 24-well-plate Renner, Dannstadt AGFA Cronex 5, Medical X-Ray Film AGFA, Mortsel, Belgium Cell culture dish,  100 mm Renner, Dannstadt Cell culture flask, 75 cm2 Renner, Dannstadt Cell scraper 25 cm Sarstedt, Nümbrecht Chromatography paper 3MM Chr Schleicher & Schuell, Dassel CryoPure tube 1.8 ml Sarstedt, Nümbrecht Micro tube (1.5 ml, PP) Sarstedt, Nümbrecht Microfuge® tube Polyallomer BeckmannCoulter, Brea, USA Multiply®-Pro PCR tubes 0.5 ml Sarstedt, Nümbrecht Nitrocellulose membrane disc filter Sartorius, Göttingen (0.45 µm pore size) Petri dish (100 mm) Sarstedt, Nümbrecht Protran® Nitrocellulose transfer membrane Schleicher & Schuell, Dassel Sterile surcical blade BBraun Aesculap, Tuttlingen Scintillation tube (6 ml) Perkin-Elmer, Rodgau-Jügesheim Scintillation tube (20 ml) Perkin-Elmer, Rodgau-Jügesheim Syringe (1 ml, 5 ml, 20 ml) Becton Dickinson, Heidelberg Syringe filter Filtropur S 0.2 (0.22 µm) Merck, Bruchsal Tube (13 ml, 100x16 mm, PP) Sarstedt, Nümbrecht Tube (15 ml, 120x17 mm, PP) Sarstedt, Nümbrecht Tube (50 ml, 114x28 mm, PP) Sarstedt, Nümbrecht

28 Materials and Methods

2.1.5 Bacterial and Mammalian Cells African green monkey kidney (COS-7) cells Deutsche Sammlung von Mikroorganismen und Zellkulturen, Braunschweig Escherichia coli DH5α Clontech, Heidelberg

2.1.6 Plasmids

2.1.6.1 Commercial Plasmids pcDNA3.1(+) Invitrogen, Karlsruhe pcDNA3.1(–) Invitrogen, Karlsruhe pCR®-Blunt Invitrogen, Karlsruhe

2.1.6.2 Recombinant In‐House Plasmids pcDNA3.1(+)-Rac2 C. Walliser (accession number CAG30441.1) G12V pcDNA3.1(+)-Rac2 C. Walliser pcDNA3.1(-)-PLCβ2 B. Möpps (accession number NP_004564) pcDNA3.1(-)-myc-hPLCγ1 C. Walliser (accession number ABB84466) pcDNA3.1(+)-myc-hPLCγ2 C. Walliser (accession number NP_002652) pcDNA3.1(+)-myc-hPLC2PCI L. Becker F897Q pcDNA3.1(+)-myc-hPLC2 C. Walliser  pcDNA3.1(+)-myc-hPLC2 C. Walliser Y733E pcDNA3.1(+)-myc-hPLCγ2 L. Becker Y753E pcDNA3.1(+)-myc-hPLCγ2 L. Becker Y759E pcDNA3.1(+)-myc-hPLCγ2 L. Becker Y753E/Y759E pcDNA3.1(+)-myc-hPLCγ2 L. Becker Y733E/Y753E/Y759E pcDNA3.1(+)-myc-hPLCγ2 L. Becker pcDNA3.1(+)-myc-hPLC2-212 M. Retlich

29 Materials and Methods

pcDNA3.1(+)-myc-hPLC2-221 M. Retlich pcDNA3.1(+)-myc-hPLC2-211 M. Retlich Ali5 pcDNA3.1(+)-myc-hPLC2 J. Haas Ali14 pcDNA3.1(+)-myc-hPLC2 J. Haas Ali5/Ali14 pcDNA3.1(+)-myc-hPLC2 J. Haas S707Y pcDNA3.1(+)-myc-hPLC2 J. Haas

2.1.6.3 Recombinant Plasmids received from other institutes pcDNA3-hALK S. Turner, Cambridge (accession number NM_004304.4) pcDNA3-hALKF1174L S. Turner, Cambridge pcDNA3-hALKR1275Q S. Turner, Cambridge pcDNA3-hNPM-ALK S. Turner, Cambridge (accession number U04946.1) pcDNA3-hEML4-ALK S. Turner, Cambridge (accession number AB663645.1)

2.1.7 Oligonucleotides

All oligonucleotides used for polymerase chain reactions or for sequencing purposes were synthesised by Thermo Fisher Scientific, Ulm or Sigma-Aldrich, Steinheim.

2.1.7.1 Sequencing Primers

2.1.7.1.1 General Plasmid Sequencing Primers

T7 Promotor (for sequencing the 5´ ends of pcDNA3.1 inserts) 5´-TAA TAC GAC TCA CTA TAG GGA GA-3´ 5´-TAA TAC GAC TCA CTA TAG GG-3´ (GATC Biotech, Konstanz)

BGH Reverse Primer (for sequencing the 3´ ends of pcDNA3.1 inserts) 5´-TAG AAG GCA CAG TCG AGG-3´

30 Materials and Methods

Reverse Primer (for sequencing the 5´ ends of pCR®-Blunt inserts) 5´-CAG GAA ACA GCT ATG AC-3´

Universal Primer (for sequencing the 3´ ends of pCR®-Blunt inserts) 5´-GTA AAA CGA CGG CCA G-3´

2.1.7.1.2 Human PLC1 cDNA Sequencing Primers hPLCg1_S1 5´-TCT ACT CCA AGA AGT CGC-3´ hPLCg1_S2 5´-GAT ACA TTG CAG GCA CCC AC-3´ hPLCg1_S3 5´-ATG TAC AGC GCC CAG AAG-3´ hPLCg1_S4 5´-GAC ACC ATG AAC AAC CCT C-3´ hPLCg1_S5 5´-CAG AGA AAC ATG GCC CAA TAC-3´ hPLCg1_S6 5´-ACC AGC AGC AAG ATC TAC-3´ hPLCg1_S6b 5´-CAA CGA GGA TGA GGA GGA G-3´ hPLCg1_S7 5´-CCC CAA GTT CTT CTT GAC AGA C-3´ hPLCg1_S8 5´-CAG ACA GTG ATG CTA GGG AAC-3´ hPLCg1_S9 5´-CCA GAA TGT GGA GAA GCA AG-3´ hPLCg1_S10 5´-GTG AAA AAG ATC CGT GAA G-3´ hPLCg1_S11 5´-TTG TGG CCC TCA ACT TCC-3´ hPLCg1_S12 5´-GAG TTT GTG GTG GAC AAT GG-3´

2.1.7.1.3 Human PLC2 cDNA Sequencing Primers PLCg2_S1 5´-GGG CTT CTT GGA TAT CAT GG-3´ PLCg2_S2 5´-CTC CGA GAG TTG AAG ACC ATC-3´ PLCg2_S3 5´-TGA ACA AAG TCC GTG AGC-3´ PLCg2_S4b 5´-CAT TGA ACT GGA CTG CTG-3´ PLCg2_S5 5´-ATG TCA ACA TGG AGG ACA AG-3´ PLCg2_S5b 5´-CCA GCT GCG GGA GAA GAT C-3´ PLCg2_S6 5´-AGA CCT TCC CCA ATG ACT ACA C-3´ PLCg2_S7 5´-ATG CTG ATG AGG ATT CCC-3´ PLCg2_S8 5´-GAT CCC AGT GAA ATC AAT CCG-3´ PLCg2_S9c 5´-AGT TTG CCA CAG ACA GGG TG-3´ PLCg2_S10 5´-CTG ACA GCA TCA TCA GAC-3´ PLCg2_S11 5´-TCA AGG TTC TCG GTG CTC-3´ 31 Materials and Methods

2.1.7.1.4 Human ALK cDNA Sequencing Primers

ALK_S1 5´-TGC ATG ACC TCA GGA ACC-3´ ALK_S2 5´-GTG CTA GTG GAG AAC AAA AC-3´ ALK_S3 5´-AGA GAG ACT GGA GAA TAA C-3´ For sequencing the whole ALK cDNA the sequencing primers NPM-ALK_S2 to NPM-ALK_S7 were additionally used.

2.1.7.1.5 Human NPM‐ALK cDNA Sequencing Primers

NPM-ALK_S1 5´-AAG GAT GAG TTG CAC ATT G-3´ NPM-ALK_S2 5´-TGA GTA CAA GCT GAG CAA G-3´ NPM-ALK_S3 5´-CAG GAC GAA CTG GAT TTC-3´ NPM-ALK_S4 5´-GGA AAA CCA CTT CAT CCA C-3´ NPM-ALK_S5 5´-TGG ATA TAT GCC ATA CCC-3´ NPM-ALK_S6 5´-AGG AAG AGA AAG TGC CTG-3´ NPM-ALK_S7 5´-TAC GGC TCC TGG TTT ACA G-3´

2.1.7.2 PCR Primers

2.1.7.2.1 Primers for the amplification of PLC2 SH domains

All Primers contain an AvrII (CCTAGG) restriction site.

SH2N (aa 515-638)

AB063 (fp) 5´-GTCCTAGGGGAGGAAGTGCCCC-3´

AB080 (rp) 5´-CCTAGGGGGGTTGGGCACAG-3´

SH2C (aa 637-747)

AB066 (fp) 5´-GTCCTAGGGAACCCCAACCCC-3´

AB081 (rp) 5´-GGCCTAGGTCTTTCCATATTGTAGCGC-3´

SH3 (aa 767-839)

AB077 (fp) 5´-GTCCTAGGGCCGTCCATGCC-3´

AB048 (rp) 5´-GGGCCTAGGAATAATCTGCTTTTCTAGCTC-3´

32 Materials and Methods

YY-SH3 (aa 746-839)

AB078 (fp) 5´-GGGCCTAGGGAGAGATATAAACTCCCTCTAC-3´

AB048 (rp) Sequence see SH3

SH2N-SH2C (aa 515-747)

AB063 (fp) 5´-GTCCTAGGGGAGGAAGTGCCCC-3´

AB081 (rp) Sequence see SH2C

SH2C-SH3 (aa 637-839)

AB079 (fp) 5´-GTCCTAGGGAACCCCAACCCC-3´

AB048 (rp) Sequence see SH3

2.1.7.2.2 Primers for the generation of the two truncated PLC2 halves,

PLC2‐X and PLC2‐Y All Primers contain an XbaI (TCTAGA) restriction site.

PLC2-X (aa 1-514)

AB038 (fp) 5´-AGTTCTAGAGCCGCCATGTCCACCACGGT-3´

AB091 (rp) 5´-GGTCTAGACTACAGGTCCTCCTCGGAGATCAGCTTCTGC TCCTCCATAGTCTGTTCAATGTCATCAC-3´ contains a myc-tag

PLC2-Y (aa 842-1265)

AB092 (fp) 5´-GCTCTAGAGCCGCCATGGACAATCCCTTAGGGTCTCTTTG-3´

AB037 (rp) 5´-CCCTCTAGACTACTACAGGTCCTCCTCGG-3´

33 Materials and Methods

2.1.7.2.3 Mutagenesis Primers

PLC2SH with artificial linker containing an AvrII restriction site (GCCCTAGG) and introducing five additional residues in positions 515-519 of the protein (ALG_SH domains_PR) (plasmid for cloning)

AB073 (fp) 5´-GAACAGACTATGGAGGCCCTAGGGAAGACAATCCCTTAGG-3´

AB074 (rp) 5´-CCTAAGGGATTGTCTTCCCTAGGGCCTCCATAGTCTGTTC-3´

PLC2SH with artificial linker containing an AvrII restriction site (GCCCTAGG) and introducing five additional residues in positions 515-519 of the protein (...ALGPR...) (control plasmid)

AB075 (fp) 5´-GAACAGACTATGGAGGCCCTAGGGCCTAGGGAAGACAATCCC TTAGG-3´

AB076 (rp) 5´-CCTAAGGGATTGTCTTCCCTAGGCCCTAGGGCCTCCATAGTC TGTTC-3´

Y1197E PLC2 (T3589G/C3591G)

AB087 (fp) 5´-GAGCGAAGAGGAACTTGAGTCCTCCTGTCGCCAGC-3´

AB088 (rp) 5´-GCTGGCGACAGGAGGACTCAAGTTCCTCTTCGCTC-3´

Y1217E PLC2 (T3650G/T3652G)

AB089 (fp) 5´-CAACCAGCTCTTTCTGGAGGACACACACCAGAACTTG-3´

AB090 (rp) 5´-CAAGTTCTGGTGTGTGTCCTCCAGAAAGAGCTGGTTG-3´

ALKI1250T (T3749C) and NPM-ALKI310T (T929C)

AB094 (fp) 5´-CACTTCATCCACCGAGACACTGCTGCCAGAAACTGCCTC-3´

AB095 (rp) 5´-GAGGCAGTTTCTGGCAGCAGTGTCTCGGTGGATGAAGTG-3´

PLC219 (646-685), PLC2-21219, PLC2-22119, PLC2-21119

AB096 (fp) 5´-CCACGAGTCCAAGCCGGCTAGGGGCAAGGTAAAGC-3´

AB097 (rp) 5´-GCTTTACCTTGCCCCTAGCCGGCTTGGACTCGTGG-3´

34 Materials and Methods

ALKY1604F (A4811T) and NPM-ALKY664F (A1991T)

AB100 (fp) 5´-CCTGGAGCTGGTCATTTCGAGGATACCATTCTG-3´

AB101 (rp) 5´-CAGAATGGTATCCTCGAAATGACCAGCTCCAGG-3´

PLC1668-707 (PLC1“19“)

AB102 (fp) 5´-CGCCCACGAGAGCAAAGAGGCTGAGGGCAAGATCAAGC-3´

AB103 (rp) 5´-GCTTGATCTTGCCCTCAGCCTCTTTGCTCTCGTGGGCG-3´

PLC219-22 (template PLC220-22)

AB108 (fp) 5´-CCCCACGAGTCCAAGCCGTGGAAAGGAGACTATGGAACCAG-3´

AB109 (rp) 5´-CCTGGTTCCATAGTCTCCTTTCCACGGCTTGGACTCGTGGGG-3´

W899A PLC219 (template PLC219) AB110 (fp) 5´-GGAGCTCTTTGAGGCGTTTCAGAGCATCCGAG-3´

AB111 (rp) 5´-CTCGGATGCTCTGAAACGCCTCAAAGAGCTCC-3´

Y509A/F510A PLC2-21119 (template PLC2-21119) AB112 (fp) 5´-GAATGGTATCCCCACGCCGCTGTTCTGACCAGCAG-3´

AB113 (rp) 5´-CTGCTGGTCAGAACAGCGGCGTGGGGATACCATTC-3´

2.1.7.2.4 SOE‐Primers

PLC2SA (476-908)

AB038 (fp A) 5´-AGTTCTAGAGCCGCCATGTCCACCACGGT-3´ contains a XbaI restriction site

AB031 (rp A) 5´-GTACTTCATGTTGTTCTCCTTGGTGTCAATCTTGT GTTCGTCCTTCTTGTCCTCCATGTTGACA-3´

AB030 (fp B) 5´-TGTCAACATGGAGGACAAGAAGGACGAACACAAG ATTGACACCAAGGAGAACAACATGAAGTAC-3´

AB037 (rp B) 5´-CCCTCTAGACTACTACAGGTCCTCCTCGG-3´ contains a XbaI restriction site

35 Materials and Methods

PLC2SH (515-840)

AB038 (fp A) Sequence see PLC2SA

AB033 (rp A) 5´-CTCTGCAAAGAGACCCTAAGGGATTGTCTTCCTCCAT AGTCTGTTCAATGTCATCACTGAAGG-3´

AB032 (fp B) 5´-CCTTCAGTGATGACATTGAACAGACTATGGAGGAAGA CAATCCCTTAGGGTCTCTTTGCAGAG-3´

AB037 (rp B) Sequence see PLC2SA

PLC2SH2C (639-766)

AB038 (fp A) Sequence see PLC2SA

AB067 (rp A) 5´-CTCTGAGGCATGGACGGGGGGTTGGGCACAGGG-3´

AB068 (fp B) 5´-CCCTGTGCCCAACCCCCCGTCCATGCCTCAGAG-3´

AB037 (rp B) Sequence see PLC2SA

PLC220-22 (686-806)

AB038 (fp A) Sequence see PLC2SA

AB098 (rp A) 5´-GCTGGATCCTGGTTCCATAGTCTCCTTTCCACCTGAA GGTGATGGCATAGGAGTCGCTCC-3´

AB099 (fp B) 5´-GGAGCGACTCCTATGCCATCACCTTCAGGTGGAAAG GAGACTATGGAACCAGGATCCAGC-3´

AB037 (rp B) Sequence see PLC2SA

PLC1708-828 (PLC1”20-22”)

PLCg1_Sense 5´-TTTGGATCCGCCGCCATGGCGGGCGCCGCGTC-3´ (fp A) contains a BamHI (GGATCC) restriction site (C. Walliser)

AB105 (rp A) 5´-CTTCCCTCCGTAGTCCCCTCGCCACCGGAAAGAGATGG CATATGAGTTGGGTTC-3´

AB107 (fp B) 5´-GAACCCAACTCATATGCCATCTCTTTCCGGTGGCGAGGG GACTACGGAGGGAAG-3´

36 Materials and Methods

AB106 (rp B) 5´-CCGAATTCCTACTACAGGTCCTCCTCGGAGATC-3´

contains an EcoRI (GAATTC) restriction site

2.1.8 Antibodies

2.1.8.1 Primary Antibodies

Myc-Tag (# 2276, 9B11) (mouse monoclonal IgG) Cell Signaling Technology, Danvers, MA, USA

PLC2 (# sc-407, Q-20) (rabbit polyclonal IgG) Santa Cruz, Heidelberg

PLC2 (404) (rabbit polyclonal IgG) Peter J. Parker, London, UK ALK (# 3791, 31F12) (WB) (mouse monoclonal IgG) Cell Signaling Technology, Danvers, MA, USA ALK (# 3333 ,C26G7) (IP) (rabbit monoclonal IgG) Cell Signaling Technology, Danvers, MA, USA

2.1.8.2 Secondary Antibodies

Anti-Mouse IgG-Peroxidase (# A5278) Sigma, Deisenhofen (goat IgG) Anti-Rabbit IgG-Peroxidase (#A6154) Sigma, Deisenhofen (goat IgG)

2.1.9 Instruments ABI Prism™ 310 Genetic Analyzer Perkin-Elmer, Weiterstadt Axiovert 25 Microscope Zeiss, Oberkochen Bench centrifuge 5430 Eppendorf, Hamburg Bench centrifuge 5415 D Eppendorf, Hamburg Bench centrifuge 5417 R Eppendorf, Hamburg Electrophoresis chamber (DNA) MBT, Giessen 37 Materials and Methods

Electrophoresis chamber (big) (SDS-PAGE) Amersham Pharmacia, Freiburg Electrophoresis Power Supply EPS 301 Amersham, Freiburg Electrophoresis Power Supply EPS 600 Amersham, Freiburg Elix® 5 Millipore, Eschborn Heater TechneDRI-BLOCK® DB3A Labtech, Burkhadtshausen HERA Safe Laminar Flow Workbench Heraeus, Fellbach Immunoblot Transfer Chamber Renner GmbH, Darmstadt Incubator CB 210 Binder, Tuttlingen Incubator Certomat H Braun, Melsungen J2-HS Centrifuge Beckman Instruments, München Liquid Scintillator Analyzer, Tricarb 288TR Packard, Canberra Magnetic hot plate stirrer RCTbasic IKA, Staufen Milli-Q® Millipore, Eschborn Optima™ TLX Ultracentrifuge Beckman Instruments, München Orbital shaker Unimax 2010 Heidolph, Kehlheim Overhead shaker Reax 2 Heidolph, Kehlheim pH meter 765 calimatic Knick, Berlin PTC 200 Peltier Thermal Cycler MJ Research, Oldendorf Rocking platform shaker Duomax 1030 Heidolph, Kehlheim Shaker Certomat R Braun, Melsungen Shaker Thermomixer comfort Eppendorf, Hamburg Shaker Thermomixer compact Eppendorf, Hamburg Shaker Titramax 100 Heidolph, Kehlheim Spectrophotometer GeneQuant Pro Amersham, Freiburg UV transilluminator IL 200M (312 nm) Bachofen, Reutlingen Videodocumentation system CF8/1 DX Kappa, Gleichen Vortex mixer Reax control Heidolph, Kelheim Vortex Whirlmix 34524 Cenco Instrumenten, Breda, Netherlands

2.1.10 Construction of individual incubation chambers

Chambers allowing the temperature-controlled incubation of individual 24-well tissue culture plates (92424; TPP, Switzerland) were custom assembled using 15 x 12 x 23 cm

38 Materials and Methods

Styrofoam containers (Schaumaplast, Reilingen, Germany) equipped with circuits made up of one Velleman VM148 thermostat control module (190655-62) and one Dallas DS18S20-55 temperature sensors with digital output (176168-62) to control two serially connected heating foils (532878-62; all from Conrad Electronic, http://www.conrad.de) to be placed on either side of the tissue culture plate during incubation. To allow for additional external, analogue control of the temperature, a 10 mm whole was drilled into the wall of the container and the tissue culture plate to allow insertion of a thermometer into one well. Power supply to up to 8 individual chambers was through a TDK-Lambda LS75-12 AC/DC converter unit (511823-62; Conrad). A circuit diagram can be found in the appendix.

2.1.11 Software

Adobe® Photoshop® CS3 Extended 10.0 Adobe Systems GmbH, München Chromas version 2.33 Technelysium Pty Ltd, Tewantin, Australia Clustal Omega Sievers et al. [233] GraphPad Prism® 4 GraphPad Software, Inc., San Diego, USA Micrografx Designer 9.0 Micrografx Inc., München Microsoft Office 2003 Microsoft Corp., Washington, USA Microsoft Office 2010 Microsoft Corp., Washington, USA NEBcutter V 2.0 Vincze et al. [267] OMIGA 2.0 Oxford Molecular Ltd. 1996-1999, Oxford, UK PPSP: prediction of PK-specifc phosphorylation site Xue et al. [299] Zotero 4.0.11 Center for history and new media, George Mason University, Fairfax, USA, http://zotero.org

39 Materials and Methods

2.2 Microbiology and Molecular Biology

2.2.1 Preparation of Luria‐Bertani (LB) Medium and Agar Plates

E. coli bacteria were grown in sterile Luria-Bertani (LB) medium. LB medium was autoclaved and then stored at room temperature (RT). For the selection of bacteria harbouring a plasmid with the antibiotic resistance gene the medium was supplemented with the appropriate antibiotic. Ampicillin and Kanamycin were prepared as stock solutions and diluted 1:1000 in LB medium. LB agar plates were prepared by adding 1.5 % (w/v) agar to the LB medium before autoclaving. The solution was then allowed to cool to hand-warm temperature and after adding the appropriate antibiotics the solution was poured in sterile Petri dishes. After solidifying at RT the plates were stored at 4 °C.

LB-Medium: 1 % (w/v) Bacto Trypton 0.5 % (w/v) Yeast Extract Servabacter 1 % (w/v) NaCl

Ampicillin stock solution: 50 mg/ml in ddH2O, sterile filtered through a nitrocellulose membrane disc filter (0.22 μm pore size)

Kanamycin stock solution: 30 mg/ml in ddH2O, sterile filtered through a nitrocellulose membrane disc filter (0.22 μm pore size)

2.2.2 Preparation of Competent Bacteria

The Inoue method [104] was used to prepare competent bacteria. A single colony of E. coli DH5 bacteria streaked on an agar plate without antibiotics was picked with a sterile tip and introduced into a sterile tube containing 25 ml of LB medium. The bacterial suspension was grown for 6-8 h in an incubator (37 °C, 200 rpm). Of this preculture, 2 ml, 4 ml, and 6 ml were added to 250 ml of SOB medium in a sterile flask, and allowed to grow overnight (20 °C, 200 rpm) to an optical density at 600 nm (OD600) of 0.55. The culture reaching this density first was centrifuged at 2500 x g at 4 °C for 10 min. The pellet was resuspended in 80 ml of ice-cold Inoue buffer. The suspension was again centrifuged at 2500 x g at 4 °C for 10 min and the pellet was resuspended in 20 ml of ice-cold Inoue buffer. The suspension was supplemented with 1.5 ml DMSO and incubated in an ice 40 Materials and Methods

water bath for 10 min, after which 50 μl aliquots were snap-frozen in liquid N2 and the competent bacteria stored at -80 °C.

Inoue buffer: 55 mM MnCl2

15 mM CaCl2 250 mM KCl 10 mM PIPES/KOH (0.5 M, pH 6.7)

SOB medium: 2 % (w/v) Bacto Trypton 0.5 % (w/v) Yeast Extract Servabacter 10 mM NaCl 2.5 mM KCl

10 mM MgCl2

10 mM MgSO4, pH 7.0

2.2.3 Transformation of competent bacteria

An aliquot of competent bacteria was thawed on ice (50 µl for transformation of bacteria with a plasmid from a ligation reaction, 25 µl for re-transformation of bacteria with a plasmid). The appropriate plasmid (10 ng in a volume of ≥ 2µl), ligation reaction (2 µl) or PCR reaction (2 µl) was added and the bacterial suspension was incubated on ice for 30 min, followed by a heat-shocking for 90 sec at 42 °C in a water bath. Immediately after the heat shock the bacteria were placed on ice for 2 min and then supplemented with 500 µl of prewarmed (37 °C) LB medium, and incubated at 37 °C for 1 h on a rotary shaker (220 rpm). Fifty µl of a re-transformation reaction or 400 µl of the suspension was then plated on a LB agar plate containing appropriate antibiotics and incubated at 37 °C over night.

2.2.4 Plasmid DNA preparation

For small or medium scale preparations of plasmid DNA the Nucleospin® Plasmid Kit and the Nucleo Bond® PC 100 Kit (both from Macherey-Nagel) were used, respectively. Therefore LB medium (3 ml for small scale and 50 ml for medium scale preparations) containing the appropriate antibiotics were inoculated with single colonies picked from the respective LB plates and grown over night in an incubator (37 °C, 200 rpm). The

41 Materials and Methods preparations were processed according to the manufacturer’s protocols. DNA was eluted in ddH2O and stored at -20°C.

2.2.5 Photometric Quantification of DNA

Nucleic acids absorb UV light with an absorption peak at 260 nm. This feature allows the photometric measurement of DNA concentrations. To relate the absorption at 260 nm with the DNA concentration the Beer-Lambert law A = ε × d × c, with A, absorbance, ε, molar extinction coefficient, d, path length, c, concentration, was used. The average molar -1 -1 extinction coefficient for double-stranded DNA is ~50 (μg/ml) cm at 260 nm. The purity of the DNA solution was determined by measuring the ratio of the absorbance at 260 nm and 280 nm (OD260/280). Pure DNA solutions resulted in values between 1.8 and 2.0. Lower values indicate contaminations with proteins or phenol.

2.2.6 Agarose Gel Electrophoresis

Agarose gel electrophoresis is a method used to separate DNA molecules by size. The negatively charged nucleic acid molecules migrate through an agarose matrix when an electric field is applied. Short DNA molecules migrate faster and therefore farther than long ones. The Agarose gel was prepared by dissolving 1 % (w/v) agarose in TAE buffer by boiling the solution in a microwave. The solution was allowed to cool to ~ 50 °C before pouring into a gel rack. The DNA samples were supplemented with DNA sample buffer at the rate of 5:1 (v/v) and loaded into the agarose gel wells. A voltage of 100 V was applied for 1 hour. To visualize the DNA the agarose gel was stained for 10 min in a TAE bath containing ~ 0.0003 % (w/v) ethidium bromide. After a short destaining (~ 1 min) in H2O the stained DNA was visualized on a UV transilluminator and images obtained were printed out for documentation.

1 x TAE buffer: 40 mM Tris/HCl, pH 8.0 1 mM EDTA 0.1 % (v/v) acetic acid 6 x DNA sample buffer: 60 % (v/v) Glycerin 0.25 % (w/v) bromphenol blue 60 mM EDTA, pH 8.0

42 Materials and Methods

To avoid a possible damage of DNA by UV light which may impair reactions like the ligation of DNA fragments, a more gentle staining method was applied. DNA stained with crystal violet is already visible at ambient light which renders irradiation of DNA with UV light unnecessary. Therefore 30 μl of crystal violet solution (10 mg/ml (w/v)) were added to 30 ml of dissolved, cooled agarose before pouring the gel. The DNA samples were mixed with crystal violet sample buffer instead of DNA sample buffer.

6 x crystal violet sample buffer: 30 % (v/v) glycerine 0.3 mg/ml crystal violet

2.2.7 Gel Purification of DNA Fragments

® DNA fragments to be ligated were extracted from agarose gels using the QIAquick Gel Extraction Kit (Qiagen) according to the manufacturer’s protocol except that the DNA was eluted in 30 µl of ddH2O.

2.2.8 DNA Restriction Digest

Restriction endonucleases cut double-stranded DNA at specific recognition sites which turns them into an important tool to analyse or prepare DNA fragments, e.g. for ligation. For analytical or preparative DNA restriction digests, the following reactions were prepared:

analytical preparative Plasmid DNA 1 µg 5 µg

Specific buffer (10x) 1,5 µl 5 µl

Restriction 5 U 25 U ad H O 15 µl 50 µl 2

The reactions were incubated for 1h at the temperature specific for the used restriction endonuclease. Digests involving more than one enzyme were performed according to the manufacturer’s suggestions either as one reaction or with an intermittent step involving the isolation of digested DNA by agarose gel electrophoresis and gel extraction.

43 Materials and Methods

2.2.9 Dephosphorylation of DNA Fragments

To prevent re-ligation of DNA plasmids cut at a single restriction site in a later ligation reaction the linearized plasmid DNA had to be dephosphorylated. Antarctic phosphatase (NEB) was used according to the manufacturer’s protocol to remove the 5' phosphate groups of DNA molecules and thereby preventing re-ligation. Products from preparative restriction digests were directly used for dephosphorylation after agarose gel electrophoresis and gel purification of the desired DNA fragments. After heat inactivation of antarctic phosphatase according to the manufacturer’s protocol the DNA molecules were directly used for ligation reaction, omitting any purification step.

2.2.10 DNA Ligation

DNA Ligation is used to generate recombinant plasmid DNA by inserting DNA fragments into plasmid vectors. For this purpose T4 DNA ligase, which forms covalent phosphodiester bonds in an ATP-dependent manner between the 3' hydroxyl group and the 5' phosphate group in double-stranded DNA or RNA molecules, was used. For ligation of two DNA fragments the following reaction was prepared:

25 ng Linear vector DNA 5:1 Insert DNA molar ratio over vector 1,5 µl 10 x T4 DNA ligase buffer 0,5 µl T4 DNA ligase (5 U/µl)

15 µl ad H2O

The ligation reaction was incubated at 22 °C for 1 hour before transformation of competent bacteria. The leftovers of this reaction were incubated at 16 °C over night which allows a second transformation in case the first transformation failed.

2.2.11 Polymerase Chain Reaction (PCR)

Hardly any invention has changed biological sciences in a way polymerase chain reaction (PCR) did [179]. By means of this method it is possible to amplify marginal amounts of DNA very rapid, which is of great interest for molecular biology as well as for clinical 44 Materials and Methods approaches. Kary Mullis, the innovator of this method, received the Nobel prize in chemistry in 1993 for this pioneering work. To specifically amplify defined DNA sequences DNA oligonucleotides, highly complementary to the DNA region targeted for amplification, so called primers have to be designed. The following 50 µl standard PCR reaction and the cycling parameters were used, unless specified otherwise:

1 µl DNA template (50 ng) 1 µl dNTP mix (10 mM each dNTP) 10 µl 5 x Phusion® HF buffer 1 µl Forward primer (100 µM) 1 µl Reverse primer (100 µM) 0.5 µl Phusion® high-fidelity DNA polymerase

50 µl ad H2O

Table 2-1: Standard PCR cycling conditions

2-step 3-step

Step Temperature Time Temperature Time Cycles

Initial denaturation 98 °C 2 min 98 °C 2 min 1

Denaturation 98 °C 20 s 98 °C 20 s

Primer annealing - - X °C 20 s 30

Extension 72 °C 30 s/kb 72 °C 30 s/kb

Final extension 72 °C 7 min 72 °C 7 min 1

Melting temperatures (Tm) of the primers were calculated with the nearest-neighbour method [23]. The annealing temperature was calculated dependent on the melting temperature (Tm) of the primers. Annealing was always performed at + 3°C of the lower

Tm primer, except for primers with Tm ≥ 72 °C where the 2-step PCR protocol was used.

45 Materials and Methods

The PCR products were separated by agarose gel electrophoresis and the desired bands were extracted from the gel and subcloned into the vector pCR-Blunt.

2.2.12 Site‐directed Mutagenesis

The QuikChange II XL Site-Directed Mutagenesis Kit (Agilent Technologies) was used to insert point mutations or deletions into plasmid DNA. This method relies on the fact that DNA produced in a PCR reaction is resistant against digestion by DpnI endonuclease since DpnI endonuclease specifically digests methylated and hemi-methylated DNA. The whole plasmid is amplified in vitro by the help of oligonucleotides containing the desired mutation, followed by digest of the template DNA by DpnI. The generated mutated plasmid DNA is directly introduced into competent bacteria. For introducing point mutations or deletions into plasmid DNA the reaction was performed according to the manufacturer’s protocol except that 100 ng template DNA was used. For the deletion of multiple amino acids annealing of the primers was performed at lower temperatures. To generate the PLC, PLC119, PLC19-22 constructs an annealing temperature of 55 °C was used. In both cases transformation of competent DH5 bacteria was performed as described in section 2.2.3.

2.2.13 Splicing by Overlap Extension (SOE)

Splicing by overlap extension is a technique used to recombine two DNA fragments at precise junctions and thereby e.g. deleting large fragments within a gene [95]. Therefore two DNA fragments containing overlapping sequences are produced in two separate PCR reactions, PCR A and PCR B. In a third PCR reaction, PCR C, these PCR products are recombined. Thereby the overlapping sequence of both DNA fragments constitutes an origin for DNA polymerase like primers do in a standard PCR reaction. Extension of this overlap effectively „splices“ the original molecules together. SOE was used to generate the

PLC2SA, PLC2SH, PLC2SH2C, PLC220-22, and PLC120-22 constructs. For PCR A and PCR B the standard PCR protocol was used (Table 2-1). The resulting PCR products were separated by agarose gel electrophoresis and the particular bands were extracted from the gel. The reaction mix for PCR C was prepared as in the standard PCR reaction adding the same amount of the two extracted products from PCR A and PCR B (usually 1µl of the higher concentrated PCR product and the appropriate higher volume of

46 Materials and Methods the lower concentrated PCR product) instead of regular primers and template. The cycling conditions are specified in Table 2-2 and Table 2-3. After 10 cycles 1 μl each of forward primer A and reverse primer B were added to the reaction mix and cycling continued for additional 30 cycles.

Table 2-2: PCR cycling conditions for splicing by overlap extension

Step Temperature Time Cycles

Initial denaturation 98 °C 2 min 1

Denaturation 98 °C 20 s

Primer annealing X °C X s 10

Extension 72 °C 30 s/kb

Addition of forward primer A and reverse Primer B

Denaturation 98 °C 20 s

Primer annealing X °C 20 s 30

Extension 72 °C X s

Final extension 72 °C 7 min 1

Table 2-3: Annealing temperatures and extension times for splicing by overlap extension

Annealing Construct Extension time temperature

PLC2SA 68 °C 1 min 30 s

PLC2SH 68 °C 1 min 30 s

PLC2SH2C 70 °C 1 min 30 s

PLC220-22 68 °C 2 min

PLC120-22 72 °C 3 min

The PCR products were separated by agarose gel electrophoresis and the desired bands were extracted from the gel and subcloned into the vector pCR-Blunt.

47 Materials and Methods

2.2.14 DNA Sequencing

DNA sequences were either determined by GATC Biotech Konstanz or by the use of ABI PRISM® BigDyeTM Terminator Cycle Sequencing Ready Reaction Kits, Version 2.0, Applied Biosystems.

2.3 Cell culture and DNA transfection

2.3.1 Cell counting

For counting, cells were diluted 1:4 in trypan blue and counted in a Neubauer counting chamber. Viable cells in all four big squares were counted and the number of cells per ml was determined by multiplying the number of counted cells with 104.

Trypan blue solution: 0.5 % (w/v) trypan blue 0.85 % (w/v) NaCl

2.3.2 COS‐7 cell culture

The COS-7 cell line is a fibroblast-like cell line derived from the kidney of African green monkeys. The COS-7 cells are cultured in culture dishes (Ø 100 mm) in adherent culture at

37 °C in a humidified atmosphere of 90 % air and 10 % CO2 in 10 ml of Dulbecco’s modified Eagle’s medium (DMEM) supplemented with 10 % (v/v) fetal calf serum (FCS), 2 mM glutamine, 100 U/ml penicillin, and 100 µg/ml streptomycin. Before reaching confluence (every three or four days) the cells were passaged. Therefore the old medium was aspirated and the cells were washed with 5 ml of Dulbecco’s PBS (PAA). Thereafter the cells were incubated with 2.5 ml trypsin/EDTA at 37 °C for ~10 min to detach them from the cell culture dish. After the incubation the cells were rinsed from the cell culture dish with 5 ml of medium by pipetting up and down. The resulting single cell suspension was centrifuged (300 x g, 5 min, RT) and the pellet was resuspended in an appropriate volume of medium to dilute by a factor of eight for subculture.

48 Materials and Methods

2.3.3 Transient transfection of COS‐7 cells

To transiently transfect COS-7 cells, the cells were seeded 24 h prior transfection in culture dishes (Ø 100 mm), 6-well plates, or 24-well plates at a density of 2.5 x 106, 0.4 x 106, or 0.075 x 106 cells per well in a volume of 10 ml, 2 ml/well, or 0.5 ml/well, respectively. At least 1 h before transfection the medium was replaced in each case with 10 ml, 2 ml/well, or 0.5 ml/well of fresh medium. For transfection LipofectamineTM 2000 (Invitrogen) or jetPRIME® (Polyplus transfectionTM) was used according the manufacturer’s manual. The total amount of DNA was kept constant within the samples of one assay by adding empty pcDNA3.1(+) vector. After transfection with LipofectamineTM 2000 the cells were incubated for 24 h in the same medium. In case of transfection with jetPRIME® the medium was changed 4 h after transfection.

2.4 Protein Biochemistry

2.4.1 Protein Quantification

The determination of protein concentrations in crude cell extracts was performed using the Bradford method [21]. This method is based on the binding of the dye Coomassie Brilliant Blue to proteins and the resulting shift of the absorption peak of the dye from 465 nm to 595 nm. A standard curve was prepared by adding water to 0, 10, 20, 30, 40, and 50 μl of the protein standard (0.34 mg/ml bovine IgG) to a total volume of 800 μl. Protein samples were diluted and an appropriate volume (20 to 40 µl) was added to separate tubes and supplemented with water to a total volume of 800 µl. After addition of 200 µl Bradford reagent the tubes were vortexed and the absorbance was measured at 595 nm. The protein concentration was determined by linear regression of the standard curve by means of the computer program PG Prot. All values were determined by double or triple determination.

Bradford reagent: 0.04 % (w/v) Coomassie Brilliant Blue G 250 21 % (v/v) ethanol 42.5 % (v/v) phosphoric acid

49 Materials and Methods

2.4.2 Sodium Dodecyl Sulfate Polyacrylamide Gel Electrophoresis (SDS‐PAGE)

SDS-PAGE is a method according to Laemmli [142] in which denatured proteins are separated according to their Stokes radius, and thus according to their molecular mass. The separation is achieved through the molecular sieve effect of a porous polyacrylamide matrix. The running gel was poured between two glass plates, covered with water and allowed to polymerize, so that a vertical gel is formed. In order to achieve a sharper band separation, a stacking gel was poured on top of the running gel after removing the water. Each of the protein samples were mixed with an equal volume of (+)-sample buffer and boiled for 5 min at 95 °C. The samples were loaded and the gel was run in tank buffer at 45 mA. Empty lanes were loaded with a mix of equal volumes of (-)-sample buffer and water.

Table 2-4: SDS-PAGE gel preparation

Running gel Stacking gel

7 % 10 % 6%

Monomer solution 7 ml 10 ml 2 ml

Running buffer 7.5 ml 7.5 ml -

Stacking buffer - - 2.5 ml

H2O 15 ml 12 ml 5.3 ml

10 % (w/v) SDS 300 µl 300 µl 100 µl

10% (w/v) APS 150 µl 150 µl 50 µl

TEMED 10 µl 10 µl 5 µl

Monomer solution: 30 % (w/v) acrylamide 4K solution mix 37.5:1 (292.2 g/l acrylamide 4K, 7.8 g/l bisacrylamide 4K) Running buffer: 1.5 M Tris/HCl, pH 8.8

90.75 g Tris were dissolved in 400 ml ddH2O and pH was adjusted to 8.8 with approximately 16 ml of conc. HCl, filled

up with ddH2O to 500 ml and the solution was filtered through a 0.45 µm cellulose nitrate filter.

50 Materials and Methods

Stacking buffer: 0.5 M Tris/HCl, pH 6.8

12 g Tris were dissolved in 160 ml ddH2O and pH was adjusted to 6.8 with approximately 9 ml of conc. HCl, filled

up with ddH2O to 200 ml and the solution was filtered through a 0.45 µm cellulose nitrate filter. 2 x (+)-sample buffer: 125 mM Tris/HCl, pH 6.8 4 % (w/v) SDS 20 % (v/v) glycerol 10 % (v/v) β-mercaptoethanol 0.002 % (m/v) pyronineY 2 x (-)-sample buffer: 125 mM Tris/HCl pH 6.8 4 % (w/v) SDS 20 % (v/v) glycerol 10 % (v/v) β-mercaptoethanol Tank buffer: 25 mM Tris/HCl, pH 8.6 190 mM glycine 0.1 % (w/v) SDS

2.4.3 Western Blotting of SDS‐Polyacrylamide Gels

In the Western blotting method proteins are transferred form a SDS gel to a nitrocellulose membrane. Therefore the SDS gel and a piece of nitrocellulose membrane (Protran®) were soaked in transfer buffer and sandwiched between two sponge layers and two layers of blot paper in a blot chamber filled with transfer buffer. The proteins were then electroblotted with a current of 125 mA for 15 h for PLCs or 18 h for larger proteins like ALK. After blotting, the marker lane was separately stained with ink (3.3 % (v/v) blue ink, 0.3 % (v/v) Tween-20).

For the immunodetection of the transferred proteins the membrane was first blocked with 5 % (w/v) dried milk in PBS-T for 1 h. After blocking the membrane was incubated with the primary antibody diluted in PBS-T for 2 h. After washing the membrane three times for 7 min in PBS-T the membrane was incubated with the horseradish peroxidase-labelled secondary antibody diluted 1:1000 in PBS-T for 30 min. After washing as described above

51 Materials and Methods

and briefly rinsing the membrane with H2O the immunoreactive proteins were detected using the ECL Western blotting detection system according to the manufacturer’s instructions. The chemiluminescence was visualized by blackening and developing of an x- ray film. All incubation steps were performed at room temperature on a vertical shaker.

Transfer buffer: 25 mM Tris/HCl, pH 8.6 190 mM glycine 20 % (v/v) methanol

PBS-T: 80 mM Na2HPO4

20 mM NaH2PO4, pH 7.5 100 mM NaCl 0.5 % (v/v) Tween-20

2.4.4 Immunoprecipitation (IP)

COS-7 cells were grown on culture dishes (Ø 100 mm) and transiently transfected as described in section 2.3.3. Twenty-four h after transfection the medium was replaced with 10 ml of fresh medium. After another 20 h of incubation the cells were washed once with 5 ml of Dulbeccos’s PBS (PAA), scraped into 500 µl of ice cold lysis buffer and transferred into 1.5 ml centrifuge tubes (Beckman coulter). After 30 min of incubation on ice the lysates were precleared by ultracentrifugation (100,000 x g, 4 °C, 1 h). To couple antibodies reactive against the c-myc epitope or ALK to Protein A-Agarose beads (Roche), 30 µl of beads were washed once with 50 µl of 1 x PBS and equilibrated in 15 µl of 1 x PBS (50 % (v/v) slurry). The equilibrated beads were mixed with 500 µl of antibody stock diluted 1:1000 in 1 x PBS. The mixture was incubated by end-over-end rotation at 4 °C for 1 h and afterwards washed two times with 500 µl of lysis buffer. The precleared lysate was added to the antibody-coated Protein A-Agarose beads and the samples were incubated for 2 h at 4 °C by end-over end rotation. The beads were collected by centrifugation, washed three times with 500 µl of lysis buffer, resuspendend in 60 µl of 1 x SDS-PAGE sample buffer and boiled at 95 °C for 5 min. The interaction between the proteins was analyzed by western blot using antibodies reactive against the c-myc epitope and ALK.

52 Materials and Methods

Lysis buffer (according to [9]): 10 mM Tris/HCl, pH 7.5 130 mM NaCl 5 mM EDTA 0.5 % (v/v) Triton X-100 0.1 % (w/v) BSA 1 mM PMSF 10 µM Leupeptine 2 µM Pepstatin A 2 µg/ml Soybean trypsin inhibitor (STI) 1 µg/ml Aprotinin 3 mM Benzamidine 1 mM DTT 10 µl/ml Phosphatase Inhibitor Cocktail 2 (Sigma)

1 x PBS: 80 mM Na2HPO4

20 mM NaH2PO4 100 mM NaCl, pH 7.5

2.5 Measurement of phospholipase C activity

2.5.1 Analysis of Inositol Phosphate Formation in Intact COS‐7 Cells

Cells were seeded in 24-well plates and transfected as described in section 2.3.3. Twenty- four hours after transfection the medium was removed by aspiration and the cells were washed once with 300 µl/well of Dulbecco’s PBS and the cells were incubated for 20 h in 200 µl per well of DMEM with supplements as specified in section 2.3.2, 2.5 µCi/ml myo- [2-3H]inositol, and 10 mM LiCl [188]. In case the cells were incubated in the absence of

10 % CO2 atmosphere the medium was additionally supplemented with 25 mM Hepes and 2 mM sodium pyruvate to maintain pH [243]. The cells were then lysed with 200 µl/well of 10 mM ice-cold formic acid at 4 °C. After 30 min the lysis was stopped by transferring the supernatants from the wells into reaction tubes filled with 300 µl of 10 mM NH4OH 53 Materials and Methods

[188]. To remove cell debris the samples were centrifuged at 20,000 x g for 5 min at 4 °C. The supernatant was loaded onto a column containing 0.25 ml of Dowex® 1x 8-200 ion exchange resin that had been converted to the formate form and equilibrated with H2O as described [31]. The columns were washed once with 3 ml of H2O and then twice with 3.5 ml of 60 mM sodium formate and 5 mM sodium tetraborate, and inositol phosphates were eluted with 3 ml of 1 M ammonium formate and 100 mM formic acid [294]. The eluate was supplemented with 15 ml of Quicksafe A liquid scintillator (Zinsser Analytic) and the radioactivity was quantified by liquid scintillation counting. The columns were reused after regeneration as described [31].

2.5.2 Lipid extraction

To investigate the uptake of myo-[2-3H]inositol and the subsequent conversion of this 3 reagent into [ H]PIP2 in COS-7 cells when incubating the cells at different sub- physiological temperatures the lipid extraction protocol [60] was performed. COS-7 cells were grown on 6-well plates with a density of 0.25 x 106 cells/well and transiently transfected with empty pcDNA3.1(+) vector as described in section 2.3.3. Twenty-four hours after transfection the medium was replaced with 1.5 ml/well of fresh medium supplemented with 2.5 µCi/ml myo-[2-3H]inositol, and 10 mM LiCl. In order to maintain pH in the absence of 10 % CO2 atmosphere the medium was additionally supplemented with 25 mM Hepes and 2 mM sodium pyruvate [243]. After another 20 h of incubation in individual incubation chambers (see section 2.1.10) at different temperatures (25 °C, 27 °C, 29 °C, 31 °C, 33 °C, 35 °C, and 37 °C) the cells were washed once with 1.5 ml of Dulbecco’s PBS and lysed with 1.2 ml of 4.5 % (v/v) perchloric acid (PCA) for 30 min at 4 °C. To determine the amount of myo-[2-3H]inositol taken up by the cells 10 µl aliquots of the labling mix before and after incubation of the cells at different temperatures were transferred into a scintillation vial. Three ml of Quicksafe A liquid scintillator (Zinsser Analytic) were added and total radioactivity was determined by liquid scintillation counting. After lysis the cells were scraped into PCA, transferred into a 1.5 ml reaction tube and centrifuged at 3700 x g for 20 min at 4 °C. The PCA pellets were resuspended in 100 µl of water and 375 µl of chloroform:methanol:HCl, 100:200:15 (ratio differs from [60]) were added. The tubes were vortexed and 125 µl of chloroform plus 125 µl of 0.1 M HCl were added. The mixtures were vortexed and centrifuged at 700 x g for 10 min at RT.

54 Materials and Methods

Fifty µl of the chlorophormic phase (lower one), which contains the phospholipids was subjected to liquid scintillation counting as described above.

2.6 Thermodynamic calculations

According to [85], the 10-degree temperature coefficient, Q10 , of an enzyme reaction can be calculated for an arbitrary temperature intervall T from

T 10 QT  (Q10 ) (1)

Ai Using T  Ti  Tref and QT  , where Ti and Tref are the individual and an Aref arbitrarily chosen reference temperatures and Ai and Aref are the individual and the reference activities, this equation can be linearized to

 A  log  i   0.1(T  T )log (Q ) (2) 10   i ref 10 10  Aref 

 A  log  i   0.1log (Q )T  0.1log (Q )T (3) 10   10 10 i 10 10 ref  Aref 

 A  Upon plotting the T vs. the log i  , the Q value(s) can be calculated from the i 10   10  Aref  slopes of the linear portions of the resultant graphs.

The molar heat capacity CP of PLC219 and PLC220-22 was estimated according to the protocol developed by Clapham and Miller [41] for the thermodynamic analysis of temperature-sensitive transient receptor potential (TRP) channels, using the following working relationship:

55 Materials and Methods

 ln K(T )  S0 R  CP 1 T0 T  ln(T0 T) R (4) where K(T) is the equilibrium constant of a protein in equilibrium between two

 conformations, S0 the standard state entropy, R is the gas constant, CP is the molar

 heat capacity, and T0 is the temperature of minimum K. To estimate S0 , we chose the

o  minimal K determined experimentally at 39 C, 0.003, as K 0 , and set H 0 to zero, using equation 9 of [41]:

 S0  ln K 0 R (5)

CP and T0 were then estimated by non-linear least squares curve fitting to equation (4) and K(T )  PA (T ) (1 PA )(T ) , where PA is the probability of PLC2 be in the fully active conformation A, using the global curve fitting procedure contained in GraphPad Prism, version 4.03 (GraphPad Software, San Diego California USA). To this end, the extra sum of squares F-test was used to determine whether the best-fit values of the two parameters differed between the two data sets. The simpler model was selected unless had a P value less than 0.05.

2.7 Cluster robustness analysis

An exhaustive cluster robustness analysis was performed using model-based clustering of infinite mixture multivariate Gaussians for k = 2-5 clusters. First, all leave-out-one subsets of the dataset were clustered. Then all cluster assignments were compared pair-wise using the maximum cluster assignment (MCA) index [139]. Finally, the MCA indices were compared to random baseline clustering using the Wilcoxon rank sum test. From this comparison, k = 3 (corresponding to three clusters) was chosen as the one with the largest mean difference in the MCA index (Figure 3-28).

2.8 Statistical analysis

The experiments described in this thesis were replicated at least three times. The results of representative experiments are shown in the form of mean ± SEM of triplicates. The

56 Materials and Methods significance of differences between means ± standard errors was assessed by using the unpaired t test with two-tailed P values contained in GraphPadInStat®, version 3.10 (www.graphpad.com). To account for propagation of errors, the standard errors of quantities corresponding to sums or differences of measured quantities were calculated as the sums in quadrature of the absolute standarderrors of the measured quantities [170]. Two effects are said to be synergistic, if the effect of two components tested in combination is statistically significantly higher than the sum of the individual effects [76].

57 Results

3 Results

3.1 Autoinhibitory regulation of PLC2

Besides the mechanisms of activating PLC isozymes by upstream regulators, like Rho GTPases (see section 1.3.2) or receptor and non-receptor tyrosine kinases (see section 1.3.3), there is also abundant evidence in the literature that PLCs are autoinhibited in their basal state by structural elements within their X/Y linker. Protease cleavage of PLC1,

PLC1, and PLC2 within the X/Y linker region resulted in an increase in basal PLC activity [39, 55, 62, 226]. Furthermore, the deletion of the SH2SH2SH3 tandem in PLC1 led to an enhanced PLC activity [93]. The analysis of the crystal structure of PLC2 and subsequent biochemical analyses revealed the occlusion of the active site of the enzyme by its X/Y linker [83]. Data from our group show, that also PLC2 is regulated by autoinhibitory elements within its specific array [59, 275]. All these findings point to a conserved role of the X/Y linker in autoinhibitory regulation of PLC isozymes. In the last years it became increasingly clear, that the loss of the autoinhibitory regulation of PLC2 can lead to severe phenotypes in mouse models as well as in humans [1, 56, 190, 301, 311]. Therefore it is of great importance to get a better understanding of the molecular mechanisms that lie behind this kind of intramolecular regulation.

3.1.1 PLC2 is autoinhibited by its specific array.

To analyse the role of the PLC2 specific array in autoinhibition of the enzyme, two deletion mutants were established, in which either the whole specific array (PLC2SA, aa 476-908) or only the SH2SH2SH3 tandem (PLC2SH, aa 515-840) was deleted. In intact, transiently transfected COS-7 cells both mutants showed a marked increase in basal activity compared to wild-type PLC2 (Figure 3-1, left panel), whereas the expression of all three proteins was similar (Figure 3-1, inset). These data suggest that PLC2 is subject to autoinhibitory regulation by its specific array. In order to further characterize and quantify the enhanced basal activities of the deletion mutants, COS-7 cells were transfected with increasing amounts of plasmid DNA encoding wild-type PLC2, PLC2SA, and

PLC2SH (Figure 3-1, right panel). The activity of both mutants increased with 58 Results increasing amounts of vector used for transfection. Specifically, transfected cells showed a 40-fold increase in PLC activity, for both deletion mutants, compared to cells expressing wild-type PLC2 when 50 ng each of the encoding vectors were used for transfection. An octamer peptide, referred to as the phospholipase C inhibitory (PCI) peptide, is located at the C-terminal end of the SH2C domain (aa 726-733) and was reported to be inhibitory when added to PLC enzymes in vitro and in intact cells [87, 91]. Furthermore, the deletion of this octamer was shown to be sufficient to overcome autoinhibition of the PLC2 isozyme [59]. However, the basal activity of the PLC2PCI mutant was considerably lower compared to PLC2SA and PLC2SH (Figure 3-1, right panel). COS-7 cells transfected with 50 ng vector encoding PLC2PCI only showed a 3.7-fold increase in basal activity compared to wild-type PLC2. This implies that deletion of the PCI region is not sufficient to fully negotiate autoinhibition.

Figure 3-1: Role of 2 specific array on the autoinhibitory regulation of PLC2. Left panel, COS-7 cells were transfected as indicated at the abscissa with 500 ng of either empty vector (Co.), vector encoding PLC2

(PLC2), PLC2SA (SA), or PLC2SH (SH). Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. Inset, COS-7 cells were transfected as described above with empty vector (lane 1), vector encoding PLC2 (lane 2), PLC2SA (lane 3), or PLC2SH (lane 4). Cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope. Right panel, COS-7 cells were transfected with 500 ng per well of vector encoding PLC2 (PLC2, □) or increasing amounts

59 Results

(1 ng, 2 ng, 5 ng, 10 ng, 20 ng, and 50 ng) of vector encoding PLC2SA (SA, ▼), PLC2SH (SH, ), or PLC2PCI (PCI, ○). The total amount of DNA was maintained constant at 500 ng for each transfection by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined.

3.1.2 The spPH domain but not the SH domain tandem is required

for Rac2‐mediated activation of PLC2.

Experiments from our group showed that the spPH domain of PLC2 is both necessary and sufficient for the Rac2-mediated activation of the enzyme [275]. To examine the effect of

Rac2 on the deletion mutants PLC2SH and PLC2SA, COS-7 cells were cotransfected with decreasing amounts of DNA encoding PLC2SH and PLC2SA together with Rac2-DNA or DNA encoding constitutively active Rac2G12V-DNA (Figure 3-2). In line with the previously obtained results, PLC2SA lacking both halves of the spPH domain is G12V not activated by constitutively active Rac2 , whereas PLC2SH, still harbouring the G12V spPH domain, is clearly activated by Rac2 . Surprisingly PLC2SH was also activated by wild-type Rac2, even though to a lower extend than by Rac2G12V. Furthermore, it is obvious that the SH domain tandem is not necessary for the Rac2-mediated activation of

PLC2 and that the autoinhibitory regulation and the activation by Rho GTPases seem to be independent of each other.

G12V Figure 3-2: Effect of Rac2 and constitutively active Rac2 on the activity of PLC2SH and

PLC2SA. COS-7 cells were cotransfected as indicated at the abscissa with either 500 ng of empty vector

(Co.), 500 ng of vector encoding PLC2 (wt), or decreasing amounts (10 ng, 5 ng, 2 ng, 1 ng, and 0.5 ng) of 60 Results

vector encoding PLC2SH (SH, left panel) or PLC2SA (SA, right panel), together with 25 ng of either empty vector (Control), vector encoding Rac2 (Rac2), or vector encoding Rac2G12V (Rac2G12V). The total amount of DNA was maintained constant at 525 ng by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined.

3.1.3 The C‐terminal SH2 domain and the SH3 domain in combination, but not alone, exhibit the strongest

autoinhibitory effect on PLC2.

The results obtained so far identified the SH domain tandem as the autoinhibitory element inhibiting PLC2 in its basal activity. Since the SH domain tandem is a multi-domain complex, consisting of two SH2 domains and one SH3 domain, it is possible that the single domains contribute with diverse strength to the inhibitory regulation of the lipase. To assess this hypothesis, single SH domains or combinations of two SH domains were inserted into PLC2SH, to identify the domains executing the strongest inhibitory effect on PLC2SH. The insertion of the 2 SH domain sequences into the PLC2SH cDNA was accomplished by introducing a linker containing an AvrII restriction site in the deletion site of PLC2SH. The SH domains (SH2N, aa 515-638; SH2C, aa 637-747; SH3, aa 767-839; SH2N-SH2C, aa 515-747; SH2C-SH3, aa 637-839) were amplified with primers containing an AvrII restriction site on either end and inserted into PLC2SH by restriction and ligation. The linker introduced five additional residues in positions 515-519 of the protein. However, there were no functional differences between the PLC2SH mutants with and without these residues (data not shown). The system was further verified by introducing the whole SH domain tandem in PLC2SH, thereby generating a reconstituted

PLC2 (PLC2SH+SH). The reconstituted lipase behaved similar to wild-type PLC2, in terms of the basal activity and the stimulation by Rac2G12V (Figure 3-3A, upper panel). Furthermore, the expression of both enzymes was similar (Figure 3-3A, lower panel). The insertion of each SH domain alone reduced the basal activity of PLC2SH by 46.7 % for the insertion of SH2N, by 32.9 % for the insertion of SH2C, and by 39.4 % for the insertion of the SH3 domain (Figure 3-3B, upper panel). Since important regulatory tyrosine phosphorylation sites (Y753 and Y759) are located in the linker between the SH2C and the

SH3 domain, an extended version of the SH3 domain (YY-SH3), including the SH2C-SH3 linker (aa 746-766), was also checked for its ability to inhibit PLC2SH. As shown in the 61 Results upper panel of Figure 3-3B, the additional insertion of the linker containing the tyrosine phosphorylation sites did not cause a further inhibition (24 % inhibition of PLC2SH activity) compared to the insertion of SH3 alone. The same results were observed by the simultaneous insertion of two SH domains, at which SH2N together with SH2C and SH2N together with SH3 caused a 33.8 % and 57.3 % inhibition of PLC2SH, respectively.

However, the strongest inhibitory effect was detected when the SH2C domain together with the SH3 domain were introduced into PLC2SH. Both domains in combination reduced the basal activity of the deletion mutant by 87.4 % in this experiment, indicating that both domains in combination but not alone are the major components inhibiting PLC2 in its basal activity. But it should also be noted, that the other domains also account, albeit with lower intensity than SH2C-SH3, for the autoinhibitory regulation of the lipase. This was also confirmed by the observation, that only the insertion of the whole SH domain tandem was able to fully inhibit PLC2SH.

Figure 3-3: Inhibitory effect of 2 SH domains on the activity of PLC2SH. A, upper panel, COS-7 cells were cotransfected as indicated at the abscissa with either 500 ng of empty vector (Co.), 500 ng of vector encoding PLC2 (wt), or 500 ng of vector encoding PLC2SH+SH (SH+SH), or PLC2SH (SH) together with 25 ng of either empty vector (Control), Rac2 (Rac2), or Rac2G12V (Rac2G12V). The total amount

62 Results of DNA was maintained constant at 525 ng by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. A, lower panel, COS-7 cells were transfected as described above and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope. B, upper panel, COS-7 cells were transfected as indicated at the abscissa with 50 ng of either empty vector (Co.), vector encoding PLC2 (wt), PLC2SH (SH), PLC2SH+SH2N (+SH2N), PLC2SH+SH2C

(+SH2C), PLC2SH+SH3 (+SH3), PLC2SH+YY-SH3 (+YY-SH3), PLC2SH+SH2NSH2C (+SH2N-

SH2C), PLC2SH+SH2CSH3 (+SH2C-SH3), or PLC2SH+SH2N-SH3 (+SH2N-SH3). The total amount of DNA was maintained constant at 500 ng by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. B, lower panel, COS-7 cells were transfected with 500 ng of the encoding vectors described above and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope.

The amounts of encoding vectors used for transfection in Figure 3-3B had to be quite small (50 ng of plasmid DNA/well) to clearly see the differences in the activities of the mutants. Due to the small amounts of protein expressed, we were not able to check for expression differences via immunoblotting in this case. To get at least a rough idea of the expression, COS-7 cells were transfected with 500 ng of encoding vectors and the expression was demonstrated via immunoblotting as shown in the lower panel of Figure 3-3B. The expression levels of the insertion mutants were similar or even higher than the expression level observed for PLC2SH in this experiment, indicating that the lower activities of the insertion mutants compared to PLC2SH are not due to expression differences.

The role of the SH2C and SH3 domain in the autoinhibitory regulation of PLC2 was further approved by a more detailed analysis of the activity of the PLC2SH+SH2C,

PLC2SH+SH3, and PLC2SH+SH2C-SH3 mutants. Therefore, COS-7 cells were transfected with increasing amounts of the encoding vectors. The activity of all three mutants increased with increasing amounts of vector used for transfection (Figure 3-4A).

However, cells transfected with PLC2SH+SH2C-SH3 showed an on average about 6.8-fold lower activity than the mutants in which each domain alone was inserted. The expression levels of all three mutants were similar, with the tendency of an even higher expression of the PLC2SH+SH2C-SH3 mutant compared to the other two mutants (Figure 3-4B). 63 Results

Figure 3-4: Role of the SH2C and SH3 domain in autoinhibitory regulation of PLC2. A, COS-7 cells were transfected with 500 ng of vector encoding PLC2 (PLC2, □) or increasing amounts (5 ng, 15 ng,

50 ng, 150 ng, and 500 ng) of vector encoding PLC2SH+SH2C (SH+SH2C, ▼), PLC2SH+SH3

(SH+SH3, ▲), or PLC2SH+SH2C-SH3 (SH+SH2C-SH3, ○). The total amount of DNA was maintained constant at 500 ng in each transfection by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected as described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope.

3.1.4 PLC2SH mutants are more sensitive to stimulation by Rac2.

The activation of the PLC2SH insertions mutants, described above, by Rac2 was assessed by cotransfection of COS-7 cells with the encoding vectors for wild-type Rac2 or constitutively active Rac2G12V. To detect a possible stimulation of the mutants by Rac2G12V, it was important to adjust a low basal activity of the mutants. This was obtained by using only 1 ng of lipase encoding vector for transfection. As shown in Figure 3-5,

PLC2SH as well as all of the insertion mutants are responsive to the stimulation by G12V G12V Rac2 . In contrast, wild-type PLC2 was only activated by Rac2 when 500 ng instead of 1 ng of encoding vector was used for transfection. This demonstrates the increased susceptibility of PLC2SH and its insertion mutants to the activation by Rho GTPases like Rac2. 64 Results

G12V Figure 3-5: Effect of Rac2 and constitutively active Rac2 on the activity of PLC2SH mutants. COS-7 cells were cotransfected as indicated at the abscissa with either 500 ng of empty vector (Co.), 500 ng or 1 ng of vector encoding PLC2 (wt), or 1 ng of vector encoding PLC2SH (SH), PLC2SH+SH2N

(+SH2N), PLC2SH+SH2C (+SH2C), PLC2SH+SH3 (+SH3), PLC2SH+YY-SH3 (+YY-SH3),

PLC2SH+SH2NSH2C (+SH2NSH2C), PLC2SH+SH2CSH3 (+SH2CSH3), or PLC2SHSH2N-SH3

(SH2N-SH3) together with 25 ng of either empty vector (Control), vector encoding Rac2 (Rac2), or vector encoding Rac2G12V (Rac2G12V). The total amount of DNA was maintained constant at 525 ng by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined.

3.2 Functional reassembly of PLC2 by separately expressed fragments

The functional reassembly of PLC isozymes by separately expressed halves was already demonstrated for PLC1 [94] and PLC2 isozymes [309]. Furthermore, coexpression of the N- and C-terminal halves of these enzymes led to an enhanced PLC activity compared to the holoenzymes, again demonstrating the importance of the X-Y linkers in autoinhibitory regulation of the . To further exploit the utility of assembling PLC2 fragments for functional analysis, we constructed two truncated versions of PLC2. One represents the 65 Results

N-terminal part of the enzyme from the N-terminus up to the end of spPHN (PLC2-X, aa 1-514) whereas the other truncated PLC2 starts at the beginning of spPHC and ends at the C-terminus of the enzyme (PLC2-Y, aa 842-1265). The assembly of these two parts results in a PLC2 variant lacking the SH2SH2SH3 tandem (Figure 3-6A). As expected, both mutants did not show any activity in intact cells by themselves, since both lack important catalytic residues (Figure 3-6B, left panel). However, coexpression of the two halves led to catalytic activity, which was markedly (approximately 20-fold) enhanced compared to wild-type PLC2, whereas the expression of the proteins was similar (Figure

3-6C, left panel). These results indicate (i) that the two halves of PLC2 reassemble to an intact enzyme which is able to catalyze the hydrolysis of PIP2 and (ii) that the absence of the SH2SH2SH3 tandem relieves the enzyme from autoinhibitory control.

Experiments from our group showed that both spPH domains are necessary for the Rac2- mediated activation of PLC2 [275]. Furthermore, the three-dimensional structure of the

PLC1 spPH domain revealed that the two halves of the spPH domain fold into a canonical PH domain and that the SH2SH2SH3 domain tandem is not required for the correct folding of the PH domain [288]. This led us to the assumption that the reassembly of the two

PLC2 fragments not only results in the formation of an intact catalytic domain, but potentially also in the formation of an intact PH domain. Indeed, coexpression of the G12V PLC2-X and PLC2-Y together with constitutively active Rac2 led to an about 4.7-fold increase in PLC activity compared to the control activity, whereas both halves alone were not stimulated by Rac2G12V (Figure 3-6B, right panel). Intriguingly, the effect of stimulation by wild-type Rac2, which was previously observed for PLC2SH (Figure

3-2), was also detected for the reassembled PLC2 enzyme.

66 Results

Figure 3-6: Functional reassembly of truncated PLC2 fragments. A, Domain organisation of PLC2 and the truncated PLC2 mutants PLC2-X and PLC2-Y. For details see main text. B, left panel, COS-7 cells were transfected as indicated at the abscissa with 500 ng of vector encoding PLC2 (wt), truncated PLC2 mutants PLC2-X (X) and PLC2-Y (Y) or cotransfected with 250 ng of vector encoding PLC2-X together with 250 ng of vector encoding PLC2-Y (X + Y). Right panel, COS-7 cells were cotransfected as indicated at the abscissa with 500 ng of vector encoding PLC2 (wt), truncated PLC2 mutants PLC2-X (X) and PLC2-Y

(Y), or 250 ng of vector encoding PLC2-X and 75 ng of vector encoding PLC2-Y (X+Y) in combination, together with either empty vector (Control) or 25 ng of vector encoding wild-type Rac2 (Rac2) or constitutively active Rac2G12V (Rac2G12V). The total amount of DNA was maintained constant at 525 ng by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. C, cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope.

The formation of the intact catalytic domain as well as the PH domain out of separately expressed halves was further investigated by the use of two distinct and well characterized

PLC2 mutants. The replacement of histidines 327 and 372, which are located in the

67 Results

H327A/H372A catalytic domain X, by alanins (PLC2 ) abrogates the catalytic activity of the enzyme [235]. As expected, no stimulation by Rac2G12V was observable in intact cells expressing the catalytically inactive PLC2 mutant (Figure 3-7). However, coexpression together with PLC2-X, harbouring an intact X domain, produced a PLC2 enzyme that was catalytically active, which was indicated by the sensitivity to Rac2-mediated activation. A complementary experiment regarding the spPH domain was carried out using F897Q the PLC2 mutant, which is insusceptible to stimulation by Rac2, but still exhibits its catalytic activity [275]. Intriguingly, the Rac2 sensitivity was regained by coexpression with PLC2-Y. The presented results impressively demonstrate the in vivo association of full-length PLC2 mutants with the appropriate PLC2 halves, which can replace the respective mutational inactivated domains. It is also striking that no elevated basal activity was observed for coexpression of the truncated PLC2 halves with the full-length mutant enzyme, indicating that the presence of the SH domain tandem restores the autoinhibitory control of the lipase.

H327A/H372A F897Q Figure 3-7: Functional reassembly of truncated PLC2 fragments with PLC2 and PLC2 . A, COS-7 cells were cotransfected as indicated at the abscissa with 250 ng of either empty vector (Co.), 68 Results

H327A/H372A vector encoding PLC2 (wt), truncated PLC2 mutants PLC2-X (X) and PLC2-Y (Y), PLC2 F897Q (H327A/H372A), or PLC2 (F897Q) together with 25 ng of either empty vector (Control) or vector encoding wild-type Rac2 (Rac2) or constitutively active Rac2G12V (Rac2G12V). The total amount of DNA was maintained constant at 525 ng by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected as described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope.

3.3 Phosphorylation induced activation of PLC2

PLC2 is activated by receptor and non-receptor protein tyrosine kinases, which plays an important role in B-cell receptor signalling. Upon receptor stimulation, PLC2 gets phosphorylated at tyrosine residues 753 and 759 [135], which are located in the SH domain tandem, more precisely in the linker between the C-terminal SH2 domain (SH2C) and the SH3 domain (Figure 3-8A). In addition to these two phosphorylation sites, Watanabe et al. identified two additional phosphorylation sites in the C-terminal region of PLC2, Y1197 and Y1217 [283]. However, the regulatory impact of phosphorylating PLC2 on these residues still remains ambiguous.

3.3.1 Tyrosine 733 plays a crucial role in the phosphorylation‐

induced activation of PLC2.

Since protein tyrosine phosphorylation in cells is a dynamic, rapidly reversible event and the modification itself is inherently labile, we established phosphomimetic mutants of

PLC2 by replacing the appropriate tyrosine residues by glutamic acid, which carries a terminal carboxyl group mimicking the negative charge of the phosphate. These phosphomimetic mutants of PLC2 can then be used for biochemical analysis both in intact cells and in crude cell extracts. To examine the effect of mimicking phosphorylation of different tyrosine residues in PLC2, COS-7 cells were transfected with the encoding vectors and inositol phosphate formation was determined. Figure 3-8B shows that

69 Results replacing tyrosine residues 753 or 759 by glutamic acid had no effect on the activity of

PLC2 when compared to the activity of the wild-type enzyme. Also, the double Y753E/Y759E phosphomimetic mutant PLC2 (2YE) did not show any differences in basal activity. All further mutants were generated in the setting of the 2YE mutant, as indicated at the abscissa in Figure 3-8B. Combining 2YE with Y1197E resulted in a 12.2-fold stimulation of PLC2 activity compared to the wild-type enzyme, whereas the triple phosphomimetic mutant 2YE/Y1217E only showed a slight activation (4.7 fold). Mimicking phosphorylation of all four well known phosphorylation sites

(Y753E/Y759E/Y1197E/Y1217E) stimulated PLC2 about 13-fold, which is comparable as it was observed for the triple phosphomimetic 2YE/Y1197E. Interestingly, Rikova et al. recently identified a novel phosphorylation site in PLC2 which was found in non-small cell lung cancer (NSCLC) tissue [216]. This novel phosphorylation site, Y733, lies immediately upstream of the two established phosphorylation sites, Y753 and Y759 in an area within the SH2C domain, which was shown to be an important regulatory element for the autoinhibitory regulation of PLC2. This element, also called PCI (phospholipase C inhibiting peptide), consists of 8 amino acids (726YRKMRLRY733), and its deletion results in the loss of autoinhibition, thereby generating a constitutively active PLC2 enzyme [59]. Furthermore, the PCI peptide was reported to be inhibitory when added to PLC enzymes in vitro [91] as well as in intact cells [87]. We therefore extended our set of phosphomimetic

PLC2 mutants by including the mutation Y733E. Like the previously tested phosphomimetic PLC2 mutations, the mutation Y733E by itself had no effect on the activity of PLC2 and the combination with 2YE had an effect on the basal activity that is comparable to the activity of the triple phosphomimetic mutant 2YE/Y1197E. However, mimicking phosphorylation of Y733 together with the both so far most actively mutants 2YE/Y1197E and 2YE/Y1197E/Y1217E gave raise to the most pronounced increase in phospholipase activity, 81-fold and 85.9- fold compared to PLC2, respectively. These results showed for the first time the activation of PLC2 using a phosphomimetic approach and furthermore demonstrated arrestingly the constitutive activation of PLC2 by phosphorylation of tyrosine 733.

The expression levels of the most active phosphomimetic mutants 2YE/Y733E/Y1197E and the quintuple mutant 2YE/Y733E/Y1197E/Y1217E are elevated compared to the wild- type enzyme (Figure 3-8C), but not in an intensity that could explain the 81-fold and

70 Results

85.9-fold activation of both mutants, respectively, meaning, that the obtained PLC activities are not due to the observed differences in expression of the lipases.

Figure 3-8: Effect of mimicking phosphorylation of different tyrosine residues on the activity of PLC2.

A, Domain organisation and localisation of regulatory tyrosine phosphorylation sites in PLC2. For details see main text. B, COS-7 cells were transfected as indicated at the abscissa with 500 ng of empty vector (control Y733E Y753E activity), vector encoding PLC2 (PLC2) or PLC2 phosphomimetic mutants PLC2 (733), PLC2 Y759E Y753E/Y759E Y733E/Y753E/Y759E (753), PLC2 (759), PLC2 (753/759), PLC2 (733/753/759), Y753E/Y759E/Y1197E Y753E/Y759E/Y1217E Y753E/Y759E/Y1197E/Y1217E PLC2 (753/759/1197), PLC2 (753/759/1217), PLC2 Y733E/Y753E/Y759E/Y1197E Y733E/Y753E/Y759E/Y1197E/Y1217E (753/759/1197/1217), PLC2 (733/753/759/1197), PLC2 (733/753/759/1197/1217). Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. C, COS-7 cells were transfected as described above, and cells from one well were lysed in 100

71 Results

µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope.

3.3.2 Activation of PLC2 mediated by Rac2 is independent of the phosphorylation induced activation.

Our group was previously able to show that PLC2, but not PLC1, is activated in intact cells by active mutants of the Rho GTPases Rac1, Rac2, and Rac3, but not by Cdc42 and RhoA, and that this activation is mediated by the two halves of the spPH domain [203,

275]. However, the activation of PLC2 by phosphorylation or Rho GTPases seems to be independent of each other. The mechanisms of both activating processes and their relation to each other still remain unclear. To examine the effect of Rac2 on PLC2, already activated by phosphorylation, all phosphomimetic mutants were expressed in COS-7 cells with or without Rac2 or constitutively active Rac2G12V (Figure 3-9A). Coexpression of G12V Rac2 and PLC2 increased PLC2 activity about 32-fold (Figure 3-9A, left panel) and 16-fold (Figure 3-9A, right panel) in the experiments shown. All phosphomimetic mutants are activated by constitutively active Rac2G12V in a manner that is comparable or even G12V stronger compared to the Rac2 -mediated activation of PLC2. Phosphomimetic mutants that showed enhanced phospholipase activity are also activated by wild-type Rac2, as it was already observed for other constitutively active PLC2 mutants, e.g. PLC2SH (Figure 3-2). The expression levels of the lipases alone, as well as together with Rac2 or Rac2G12V were similar (Figure 3-9B), indicating, that the observed stimulations by Rac2G12V as well as wild-type Rac2 are not due to expression differences.

72 Results

G12V Figure 3-9: Effect of Rac2 and constitutively active Rac2 on the activity of PLC2 phosphomimetic mutants. A, COS-7 cells were cotransfected as indicated at the abscissa with either 500 ng of empty vector Y733E Y753E (Co.), 500 ng of vector encoding PLC2 (wt), PLC2 phosphomimetic mutants PLC2 (733), PLC2 Y759E Y753E/Y759E Y733E/Y753E/Y759E (753), PLC2 (759), PLC2 (753/759), PLC2 (733/753/759), Y753E/Y759E/Y1197E Y753E/Y759E/Y1217E Y753E/Y759E/Y1197E/Y1217E PLC2 (753/759/1197), PLC2 (753/759/1217), PLC2 Y733E/Y753E/Y759E/Y1197E (753/759/1197/1217), or 100 ng of vector encoding PLC2 (733/753/759/1197) and Y733E/Y753E/Y759E/Y1197E/Y1217E PLC2 (733/753/759/1197/1217) together with 25 ng of either empty vector (Control), vector encoding Rac2 (Rac2), or vector encoding Rac2G12V (Rac2G12V). The total amount of DNA was maintained constant at 525 ng by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected as described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope.

3.4 Anaplastic lymphoma kinase

The interesting finding, that tyrosin 733 seems to play a crucial role in the phosphorylation-induced activation of PLC2 (see section 3.3.1), raised the question as to which protein kinase is involved in phosphorylating PLC2 at this important regulatory tyrosine residue. In silico prediction of protein phosphorylation sites and the correponding 73 Results kinases using the PPSP algorithm (prediction of protein kinase specific phosphorylation sites) [299], identified anaplastic lymphoma kinase (ALK) as a protein kinase potentially phosphorylating PLC2 on position Y733. Interestingly, (i) a novel, constitutively active ALK fusion protein, EML4-ALK, was identified in NSCLC, in the same phosphoproteomic approach that also detected Y733 as a novel phosphorylation site in

PLC2 [216], and (ii) Bai et al. [9] showed that another activated ALK fusion protein

(NPM-ALK) interacts with the SH2 domains of PLC2.

3.4.1 Isozyme specific activation of PLC isozymes by ALK and NPM‐ALK

In order to investigate the effect of ALK and its fusion proteins NPM-ALK and EML4- ALK on the activity of different PLC isozymes, COS-7 cells were transfected with the encoding vectors and inositol phosphate formation was measured (Figure 3-10A).

Coexpression of ALK and PLC1 or PLC2 resulted in a 17-fold and 5-fold stimulation of phospholipase activity, respectively. Surprisingly, NPM-ALK stimulated PLC1 (5.7-fold) but failed to stimulate PLC2 in transfected COS-7 cells. The ALK fusion protein EML4- ALK, recently discovered in NSCLC patients, had neither an impact on the activity of

PLC1, nor on the activity of PLC2. In contrast to the PLC isozymes, coexpression of

PLC2 together with NPM-ALK, or EML4-ALK did not lead to an enhanced, or in the case of ALK only resulted in a slightly enhanced inositol phosphate formation compared to the control activity. The initial hypothesis, that ALK, and maybe also its fusion proteins, phosphorylate PLC2 on tyrosine residue 733 and thereby activate the lipase, was refuted Y733E/Y753E/Y759E/Y1197E/Y1217E 5YE as the quintuple phosphomimetic mutant PLC2 (PLC2 ) was markedly activated by ALK (9.5-fold). As it had been observed before for wild-type 5YE PLC2, NPM-ALK and EML4-ALK failed to activate PLC2 . The same holds true for all other PLC2 phosphomimetic mutants (data not shown). These results present the identification of novel tyrosine kinases activating PLC isozymes at least in transfected COS-7 cells.

As shown in Figure 3-10B, coexpression of PLC isozymes with ALK or ALK fusion proteins, led to an obvious increase in the expression of the lipases compared to cells transfected with lipases only (Control). However, this fact is also true for PLC2,

74 Results

coexpression of PLCs together with EML4-ALK, as well as for coexpression of PLC2 together with NPM-ALK, which all together did not show an enhanced inositol phosphate formation. Therefore, the stimulation of PLC1 by ALK and NPM-ALK, and the 5YE stimulation of PLC2 and PLC2 by ALK, is likely not to be due to the enhanced protein levels of the PLC isozymes.

Figure 3-10: Effect of ALK and its fusion proteins NPM-ALK and EML4-ALK on the activity of PLC isozymes, PLC2 phosphomimetic mutant, and PLC2. A, COS-7 cells were cotransfected as indicated at the abscissa with either 500 ng of empty vector (Co.), 300 ng of vector encoding PLC1 (PLC1), 500 ng of Y733E/Y753E/Y759E/Y1197E/Y1217E 5YE vector encoding PLC2 (PLC2), 150 ng of vector encoding PLC2 (PLC2 ),

300 ng of vector encoding PLC2 (PLC2) together with 250 ng of either empty vector (Control), vector encoding ALK (ALK), NPM-ALK (NPM-ALK), or EML4-ALK (EML4-ALK). The total amount of DNA was maintained constant at 750 ng by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected as described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope for PLC1, PLC2, and 5YE PLC2 , an antibody reactive against PLC2 for PLC2, and an antibody reactive against ALK for ALK, NPM-ALK, and EML4-ALK. 75 Results

3.4.2 Wild‐type ALK as well as constitutively active ALK mutants are able to activate PLC isozymes.

ALKF1174L and ALKR1275Q were found to be mutated in sporadic and familial cases of neuroblastoma, creating a constitutively active ALK, and thereby gaining their oncogenic R1275Q potential [7, 37, 145]. In Figure 3-11A it is depicted that ALK activates PLC1 to a similar extend as wild-type ALK (both about 9-fold compared to control activity), whereas F1174L ALK exhibits the strongest stimulatory effect on PLC1 (14.9-fold). Both ALK variants as well as the wild-type enzyme had more or less comparable effects on the activity of PLC2. We observed a 4.6- and 4.7-fold stimulation for ALK and its F1174L R1275Q variant, whereas ALK only led to a 3.2-fold stimulation of PLC2 in this experiment.

On the other hand, PLC2 is also only weakly stimulated by the constitutively active mutants of ALK (1.7-fold compared to control activity). The expression levels of ALK and its variants were similar, whereas the expression levels of all lipases were again upregulated in the presence of ALK, as it had been elucidated before (Figure 3-11B).

Figure 3-11: Effect of the constitutively active ALK mutants ALKF1174L and ALKR1275Q on the activity of PLC isozymes. A, COS-7 cells were cotransfected as indicated at the abscissa with either 500 ng of empty

76 Results

vector (Co.), 300 ng of vector encoding PLC1 (PLC1), 500 ng of vector encoding PLC2 (PLC2), or 300 ng of vector encoding PLC2 (PLC2) together with 250 ng of either empty vector (Control), vector encoding ALK (ALK), vector encoding ALKF1174L (F1174L), or vector encoding ALKR1275Q (R1275Q). The total amount of DNA was maintained constant at 750 ng by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2,5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected as described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope for PLC1 and PLC2, an antibody reactive against PLC2 for PLC2, and an antibody reactive against ALK for ALK, ALKF1174L, and ALKR1275Q.

3.4.3 NPM‐ALK associates with PLC1 but not with PLC2.

To further study the observed activation of PLC isozymes by ALK and NPM-ALK, immunoprecipitation studies were performed. Therefore, COS-7 cells were cotransfected with cDNAs encoding PLC1 or PLC2 together with vectors encoding ALK or NPM-ALK and immunoprecipitations using an antibody reactive against the c-myc epitope of PLC isozymes and an antibody reactive against ALK and NPM-ALK were performed. PLC1 and PLC2 were efficiently precipitated using an anti-myc antibody (Figure 3-12A and C, left panels). There was no coprecipitation observed for ALK neither with PLC1 nor with

PLC2 (Figure 3-12A, right panel). However, the inverse experiment in which ALK was precipitated with an anti-ALK antibody, a coprecipitation of both PLC isozymes was detectable (Figure 3-12B). These results are in line with the observed activation of PLC1 and PLC2 by ALK (section 3.4.2). NPM-ALK clearly coprecipitated with PLC1 whereas there was only a weak association observable between NPM-ALK and PLC2 (Figure 3-12C). The same results were obtained by precipitating NPM-ALK with an anti-ALK antibody. There was again a strong association observeable for NPM-ALK and PLC1 whereas the detectable signal for coprecipitation of PLC2 was again only very negligible (Figure 3-12D), which is in accordance with the results observed for the isozyme specific activation of PLC isozymes by NPM-ALK in transfected COS-7 cells (section 3.4.1).

77 Results

Figure 3-12: Association of ALK and NPM-ALK with PLC isozymes in intact cells. COS-7 cells were transfected with 15 µg per well of either empty vector (Control), 10 µg per well of vector encoding PLC1

(PLC1) and PLC2 (PLC2), or 5 µg per well of vector encoding ALK (ALK) or NPM-ALK (NPM-ALK), or cotransfected with 10 µg per well of vector encoding PLC1 or PLC2 together with 5 µg per well of vector encodingALK or NPM-ALK. The total amount of DNA was maintained constant at 15 µg in each transfection by adding empty vector. Forty-eight hours after transfection the cells were harvested and immunoprecipitation was performed with an antibody reactive against the myc-tags of PLC1 and PLC2 (A and C) or with an antibody reactive against ALK and NPM-ALK (B and D). The samples were then subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope for PLC1 and PLC2, and an antibody reactive against ALK for ALK and NPM-ALK. IP: immunoprecipitation, WB: Western blotting

78 Results

3.4.4 The activation of PLC isozymes is mediated by the kinase activity of ALK and NPM‐ALK.

To proof, that the activation of PLC isozymes by ALK and NPM-ALK is a direct effect of the kinase activity of ALK and its fusion protein NPM-ALK, kinase-dead mutants of ALK (ALKI1250T) [228] and NPM-ALK (NPM-ALKI310T) were established and tested for their I1250T ability to activate PLC1 and PLC2 in transfected COS-7 cells. ALK was originally detected in neuroblastoma and was shown to be a kinase-dead RTK [228]. The kinase dead NPM-ALK mutant, NPM-ALKI310T, was established by substituting the corresponding isoleucine in the ALK part of NPM-ALK. As shown before, PLC2 is activated by ALK, whereas the kinase-dead mutant of ALK failed to activate PLC2 (Figure 3-13A).

Furthermore, the activation of PLC1 by ALK as well as by NPM-ALK was abrogated by introducing point mutations into the kinases, leading to kinase-dead RTKs. Therefore, the activation of PLC isozymes by ALK and NPM-ALK is mediated by the kinase activity of the enzymes and is therefore likely to be due to the phosphorylation of PLC2 tyrosine residues, which are not identified yet.

Figure 3-13: Effect of ALK and NPM-ALK kinase-dead mutants on the activity of PLC isozymes. A, COS-7 cells were cotransfected as indicated at the abscissa with 500 ng of either empty vector (Co.), 300 ng 79 Results

of vector encoding PLC1 (PLC1), 500 ng of vector encoding PLC2 (PLC2), or 150 ng of vector encoding

PLC2 (PLC2) together with 250 ng of either empty vector (Control), vector encoding ALK (ALK), vector encoding ALKI1250T (I1250T), vector encoding NPM-ALK (NPM-ALK), or vector encoding NPM-ALKI310T (I310T). The total amount of DNA was maintained constant at 750 ng by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2,5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected as described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope for PLC1 and PLC2, an antibody reactive against PLC2 for PLC2, and an antibody reactive against ALK for ALK, ALKI1250T, NPM-ALK, and NPM-ALKI310T.

It is noticeable that the expression of the kinase-dead mutants is obviously lower than the expression of the wild-type enzymes, and that furthermore the expression of the lipases also decreased when coexpressed together with the kinase-dead mutants. This effect is apparently also observable with PLC2 and PLC2 that are not activated by NPM-ALK or ALK and NPM-ALK, respectively (Figure 3-13B). To further investigate if the binding of

NPM-ALK to PLC1 is also dependent on the kinase activity of the fusion protein immunoprecipitation studies were performed. As shown in the left panel of Figure 3-14 wild-type NPM-ALK binds to PLC1 whereas the binding of the kinase-dead mutant was markedly reduced. The possibilty that the reduced association of NPM-ALKI310T is due to a lower expression could be excluded by western blot analysis of the whole cell lysates used for immunoprecipitation, where NPM-ALK and the kinase-dead mutant showed a similar expression (Figure 3-14, right panel). The results presented here indicate that the phosphorylation-induced actvation as well as the association of PLC1 and NPM-ALK are dependent on an active kinase domain.

80 Results

I310T Figure 3-14: Association of NPM-ALK and NPM-ALK with PLC in intact cells. COS-7 cells were transfected with 15 µg per well of either empty vector (Control), 10 µg per well of vector encoding PLC1

(PLC1) and PLC2 (PLC2), or 5 µg per well of vector encoding ALK (ALK) or NPM-ALK (wt) or NPM- I310T ALK (I310T), or cotransfected with 10 µg per well of vector encoding PLC1 or PLC2 together with 5 µg per well of vector encoding NPM-ALK or NPM-ALKI310T. The total amount of DNA was maintained constant at 15 µg in each transfection by adding empty vector. Left panel, forty-eight hours after transfection the cells were harvested and immunoprecipitation was performed with an antibody reactive against the myc- tag of PLC1. The samples were then subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope for PLC1, and an antibody reactive against ALK for NPM-ALK and NPM-ALKI310T. Right panel, 10 µl of the whole-cell lysates used for immunoprecipitation were subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope for I310T PLC1, and an antibody reactive against ALK for NPM-ALK and NPM-ALK . IP: immunoprecipitation, WB: Western blotting

3.4.5 Tyrosine residues 664 in NPM‐ALK and 1604 in ALK are not essential for the complex formation with PLC isozymes in transfected COS‐7 cells.

The tyrosine phosphorylation site Y664 in NPM-ALK was found to be essential for the Y664F complex formation between the kinase and PLC2. It was shown that NPM-ALK no longer forms complexes with or activates PLC in transfected Ba/F3 cells [9]. The PLC binding site is located at the C-terminus of NPM-ALK and therefore is also present in the full-length ALK enzyme. It is hence likely that Y1604 in ALK also utilizes PLC isozymes 81 Results for signal transduction. The two mutants NPM-ALKY664F and ALKY1604F were constructed and tested for their ability to activate PLC1 and PLC2 (Figure 3-15). Surprisingly, and in contrast to the results presented by Bai et al. [9], NPM-ALKY664F, as well as ALKY1604F, activated PLC2 or both isozymes, PLC2 and PLC1, in intact COS-7 cells to an extend that was only slightly diminished compared to the wild-type enzymes (Figure 3-15A), while the expression levels of the kinases and the lipases were similar (Figure 3-15B).

Figure 3-15: Effect of ALKY1604F and NPM-ALKY664F on the activity of PLC isozymes. A, COS-7 cells were cotransfected as indicated at the abscissa with 500 ng of either empty vector (Co.), 300 ng of vector encoding PLC1 (PLC1), or 500 ng of vector encoding PLC2 (PLC2) together with 250 ng of either empty vector (Control), vector encoding ALK (ALK), vector encoding ALKY1604F (Y1604F), vector encoding NPM- ALK (NPM-ALK), or vector encoding NPM-ALKY664F (Y664F). The total amount of DNA was maintained constant at 750 ng by adding empty vector. Twenty four hours after transfection, the cells were incubated for 20 h in the presence of myo-[2-3H]inositol (2,5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected as described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and

82 Results

immunoblotting was performed using an antibody reactive against the c-myc epitope for PLC1 and PLC2 and an antibody reactive against ALK for ALK, ALKY1604F, NPM-ALK, and NPM-ALKY664F.

The obtained results where further strengthened by demonstrating the association of PLC1 Y664F and NPM-ALK . COS-7 cells were cotransfected with DNA encoding PLC1 and DNA Y664F encoding NPM-ALK or NPM-ALK , and PLC1 was selectively immunoprecipitated using an antibody reactive against the c-myc epitope. A clear coprecipitation was observed for NPM-ALK as well as for its Y664F variant (Figure 3-16).

Y664F Figure 3-16: Association of NPM-ALK with PLC1. COS-7 cells were transfected with 15 µg per well of either empty vector (Control), 10 µg per well of vector encoding PLC1 (PLC1), or 5 µg per well of vector encoding NPM-ALK (wt) or NPM-ALKY664F (Y664F), or cotransfected with 10 µg per well of vector Y664F encoding PLC1 together with 5 µg per well of vector encoding NPM-ALK (wt) or NPM-ALK (Y664F). The total amount of DNA was maintained constant at 15 µg in each transfection by adding empty vector. Forty-eight hours after transfection the cells were harvested and immunoprecipitation was performed with an antibody reactive against the myc-tag of PLC1. The samples were then subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope for PLC1, and an antibody reactive against ALK for NPM-ALK. IP: immunoprecipitation, WB: Western blotting

83 Results

3.5 Characterization of PLC2 deletion mutants associated with PLAID

Genomic deletions in the PLCG2 gene were associated with a novel disease pattern named

PLAID (phospholipase C-2-associated antibody deficiency and immune dysregulation) including amongst other symptoms cold-induced urticaria, immunodeficiency, and autoimmunity [190]. However, the impact of cold temperatures on the activity of the

PLC2 deletion mutants and the mechanistical implications leading to the described phenotype are unknown so far.

3.5.1 PLC219 and PLC220‐22 are specifically activated at subphysiological temperatures.

Variants of human PLC2 carrying deletions corresponding to those described by Ombrello et al. [190] and designated PLC219 and PLC220-22, lacking the amino acids encoded by exons 19 and 20-22 of the human PLCG2 gene, were expressed in transiently transfected COS-7 cells and maintained for 24 h at 37 °C. Radiolabeling of the cells with [3H]inositol was then performed for 20 h at temperatures ranging from 39 °C to 25 °C, followed by measurement of inositol phosphate formation. Figure 3-17A shows that expression of the two PLC2 deletion mutants at 37 °C caused slight increases in inositol phosphate formation in comparison to expression of the wild-type enzyme. Consistent with earlier results [190], these increases were approximately 2.8- and 3.6-fold for PLC219 and PLC220-22, respectively, relative to the increase over basal activity observed for wild-type PLC2. Much more strikingly, however, there was a marked stimulation of inositol phosphate formation when cells expressing either deletion mutant were incubated at subphysiologic temperatures. In both cases, there was a biphasic stimulatory response with declining temperatures, with a maximum at 31 °C and a gradual reduction upon further cooling to 25 °C. There was only a modest monophasic decrease in inositol phosphate formation in cells expressing wild-type PLC2 and in control cells when the temperature was reduced from 39 °C to 25 °C. Strikingly, at 31 °C, the increase in inositol phosphate formation in cells expressing PLC219 and PLC220-22 was enhanced approx. 480 ± 91- and 430 ± 84-fold (means ± SEM) relative to the increase over basal 84 Results

inositol phosphate formation observed in cells expressing wild-type PLC2. Figure 3-17B shows that the expression of wild-type PLC2 decreased with decreasing incubation temperature, while the expression of both PLC219 and PLC220-22 took slight increases at intermediate temperatures ranging from 35 °C to 31 °C and from 35 °C to

25 °C, respectively. While reduced expression of wild-type PLC2 at lower temperatures may explain, at least in part, the monophasic decrease in inositol phosphate formation by this enzyme, the limited magnitude of the changes observed in Figure 3-17B for PLC219 and PLC220-22 argues against a critical role of fluctuating enzyme expression in the marked changes of inositol phosphate formation evident in Figure 3-17A.

Figure 3-17: Specific activation of PLC219 and PLC220-22 at subphysiological temperatures. COS-7 cells were transfected with 500 ng each per well of either empty vector (), or vector encoding either wild-type PLC2 (●), PLC219 (▼), or PLC220-22 (▲). A, twenty-four hours after transfection, the cells were incubated for a further 20 h in individual incubation chambers at the temperatures indicated at the abscissa, in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. The levels of inositol phosphate formation at 37 °C are shown in expanded scale on the right vertical axis (open symbols). The values correspond to the means ± standard error of triplicate determinations. The statistical significance of the difference between the increases in inositol phosphate formation over basal activity (□) caused by expression of wild-type PLC2 (○), PLC219 (), and PLC220-22 () at 37 °C was determined using the unpaired t test with two-tailed P values: wild-type

PLC2 vs. PLC219, P = 0.0616; wild-type PLC2 vs. PLC220-22, P = 0.0231 (not shown). B, COS-7 cells were transfected as described above and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against PLC2.

85 Results

The determination of the 10 °C temperature coefficient (Q10) value, widely used to characterize the regulation of transient receptor potential (TRP) channels by temperature

[40], for wild-type PLC2, PLC219, and PLC220-22 is shown in Figure 3-18. While only a single linear component with a Q10 value of 4.6 (corresponding to a 4.6-fold decrease in activity per 10 °C) was evident for wild-type PLC2 between 39 °C and 25 °C, both PLC219 and PLC220-22 displayed two separate phases of opposite signs and markedly distinct Q10 values. Specifically, while Q10 was as high as 6745 for the marked increase in activity between 39 °C and 33 °C, it was lower by more than three orders of magnitude and not different, by sign and by magnitude, by global curve fitting analysis from the Q10 value describing the decrease in activity of the wild-type enzyme, 4.6, between 31 °C and 25 °C. Thus, it appears that the major functional consequence of the two deletions is their increased activity upon cooling from 37 °C to approximately 31 °C. The decrease in activity observed at temperatures below 31 °C appears to be a property inherently present in the remaining elements of wild-type PLC2.

Figure 3-18: Q10 analysis of the activities of wild-type PLC2, PLC219, and PLC220-22. The individual temperatures T were plotted against  A  , with the maximum activity of PLC 19 at 31 °C i log   2 10    Aref  chosen as the reference activity Aref and reference temperature, Tref, respectively. The data of the linear components was analysed by non-linear least square curve fitting to a polynominal first order (straight line) equation. The slopes of the curves of PLC219 and PLC220-22 were not significantly different by clobal curve fitting using shared parameters from each other and, between 25 °C and 31 °C, from slope obtained for wild-type PLC2 (P = 0.9356 and P = 0.2121, respectively). 86 Results

The influence of a decrease in incubation temperature on the abundance of inositol phospholipids in transfected COS-7 cells is shown in Figure 3-19. Of note, PLC is capable of hydrolyzing PtdIns, PtdIns(4)P, and PtdIns(4,5)P2 [146, 220, 221]. There was a monophasic loss of inositol phospholipids with decreasing temperature to approximately 50 % and 8 % at 31 °C and 25 °C, respectively, of the level observed at 37 °C. While the loss between 31 °C and 25 °C may explain some of the decline in inositol phosphate formation by PLC219 and PLC220-22 at these low temperatures (cf. Figure 3-17), the loss between 37 °C and 31 °C would be expected to counteract the increased inositol phosphate formation by the two deletion mutants in this range of temperatures, if the amount of radiolabeled inositol phospholipids available for hydrolysis ever becomes limiting in transfected cells. Thus, the degree of stimulation of the two enzymes between 37 °C and 31 °C may actually be higher than evident in Figure 3-17A.

Figure 3-19: Effect of subphysiological temperatures on cellular inositol phospholipid levels. COS-7 cells were transfected with 2 µg/well of empty vector. Twenty-four hours after transfection, the cells were incubated for a further 20 h in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl at the temperatures indicated at the abscissa. The amount of [3H]inositol present in the culture medium after radiolabeling of the cells (; right vertical axis) and the cellular formation of inositol phospholipids (●; left vertical axis) was then determined as described under Materials and Methods.

87 Results

3.5.2 The activation of PLC219 and PLC220‐22 occurs posttranslationally and is reversible.

The next experiment was designed to examine whether stimulation of PLC219 and

PLC220-22 by cool temperatures occurs during or after mRNA translation. To this end, cells expressing either wild-type PLC2, PLC219, or PLC220-22 were treated with 100 µg/ml cycloheximide during [3H]inositol radiolabeling at either 37 °C or 31 °C. Figure 3-20B, shows that addition of cycloheximide caused a halt in the approximately two-fold increase in expression of both wild-type and deletion mutant PLC2 at both 37 °C and 31 °C, indicating that cycloheximide was in fact effective as a protein biosynthesis inhibitor in this experiment. As shown in Figure 3-20A, however, cycloheximide had only a minor effect on the increased formation of inositol phosphates by the two deletion enzymes at 31 °C versus 37 °C. These results clearly indicate that the stimulatory effect of subphysiologic temperatures on deletion mutant PLC2 activity may occur at the posttranslational stage.

Figure 3-20: Effect of cycloheximide on the activity of PLC219 and PLC220-22 at 31 °C. COS-7 cells were transfected as indicated at the abscissa with 500 ng of vector encoding PLC2 (Wt), PLC219

(19), or PLC220-22 (20-22). A, twenty-four hours after transfection, the cells were incubated for a further 20 h at 31 °C or 37 °C in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml), 10 mM LiCl, and as indicated in the absence or presence of 100 µg/ml cycloheximide (CHX), and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected as described 88 Results above.Twenty four hours after transfection cells from one well were lysed in 100 µl of SDS-PAGE sample buffer (Co., lane 1) or incubated for another 20 h in individual incubation chambers at 37 °C (lanes 2 and 3) or 31 °C (lanes 4 and 5), as indicated in the absence (lanes 2 and 4) or presence (lanes 3 and 5) of 100 µg/ml cycloheximide (CHX). Cells from one well were then lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope of PLC2, PLC219 and PLC220-22.

To examine whether the stimulatory effects of the 19 and 20-22 deletions on PLC2 activity were reversible, wild-type PLC2, PLC219 and PLC220-22 were incubated for the last four hours of the transfection protocol, i.e. in the absence of radiolabeled inositol phospholipid substrate, at either 37 °C or 31 °C and then radiolabeled with [3H]inositol at one of the two temperatures. Figure 3-21, right panel, shows that the incubation protocol had no effect on the expression of the three enzymes. As shown in the corresponding left panel, preincubation of the three enzymes at 31 °C had no effect on their ability to promote inositol phosphate formation. Only when cells were incubated at 31 °C in the presence of [3H]inositol, enhanced inositol phosphate formation was evident by the two deletion mutants, in contrast to wild-type PLC2.

Figure 3-21: The activation of PLC219 and PLC220-22 by subphysiological temperatures is reversible. Left panel, COS-7 cells were transfected with 500 ng each per well of either vector encoding wild-type PLC2 (Wt), PLC219 (19), or PLC220-22 (20-22). Left panel, twenty hours after transfection the cells were pre-incubated for four hours in individual incubation chambers at 31 °C (31→31;

89 Results

31→37) or 37 °C (37→37). The cells were then incubated for another 20 h in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl at 31 °C (31→31) or 37 °C (31→37;37→37) and the levels of inositol phosphates were then determined. Right panel, COS-7 cells were transfected and incubated as described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope of PLC2, PLC219, and PLC219-22.

Figure 3-22 shows a time-course of the changes of inositol phosphate formation by wild- type and mutant PLC2 upon changes in incubation temperature from 37 °C to 31 °C and vice versa. Both deletion mutants were activated with no apparent lag time with cooling to 31 °C. However, deactivation of both enzymes by warming to 37 °C required a lag time of approximately 10 min to come into effect. Taken together, the results suggest that activation of PLC219 and PLC220-22 by cool temperatures is a rapid and slowly, but fully reversible process.

Figure 3-22: Time course of the specific activation and inactivation of PLC219 and PLC220-22 by changes to and from a subphysiological temperature. Left panel, COS-7 cells were transfected with 500 ng each per well of either vector encoding wild-type PLC2 (Wt, ●), PLC219 (19, ▲), or PLC220-22 (20-22, ▼). Twenty-four hours after transfection, the cells were incubated for a further 20 h at 37 °C, in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. From the time point 0 min on (37→31 °C) the cells were incubated in individual incubation chambers at 31 °C, and at the time point 40 min (31→37 °C) the cells were placed back in the incubator at 37 °C for another 30 min. The levels of inositol phosphate formation of cells incubated at 37 °C for the whole time are shown as open symbols. Right panel, COS-7 cells were transfected as described above, and cells

90 Results from one well were lysed in 100 µl of SDS-PAGE sample buffer at the indicated time points. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope of PLC2, PLC219, and PLC219-22.

3.5.3 The deletion of PLCG2 exons 19 and 20‐22 synergize to

promote PLC2 activation.

The fact that either of the two PLC2mutants, PLC219 and PLC220-22, were activated by cool temperatures to a similar degree raised the question as to the effect of a combined deletion of the residues encoded by exons 19 through 22. Figure 3-23 shows that the compound deletion mutant, designated PLC219-22, displayed a markedly

(approximately 24-fold) enhanced activity in comparison to PLC219 and PLC220-22 already upon incubation of the cells at 37 °C.The stimulatory effect of cool temperatures was assayed at two expression levels of the deletion mutants. In both cases, the stimulatory effect of cooling from 37 °C to 31 °C was markedly higher for PLC219-22 than for both

PLC219 and PLC220-22. At both temperatures, 37 °C and 31 °C, the deletions of

PLCG2 exons 19 and 20-22 clearly synergized to promote activation of PLC2.

Figure 3-23: Effect of deleting exon 19 to 22 of PLCG2 on the enzymatic activity at 37 °C and 31 °C. Left panel, COS-7 cells were transfected as indicated at the abscissa with 500 ng each per well of either

91 Results

vector encoding wild-type PLC2 (Wt), or 500 ng, 20 ng or 10 ng of vector encoding PLC219 (19),

PLC220-22 (20-22), or PLC219-22 (19-22). The total amount of DNA was maintained constant at 500 ng by adding empty vector. Twenty-four hours after transfection, the cells were incubated for a further 20 h at 31 °C or 37 °C in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. Right panel, COS-7 cells were transfected with 500 ng of vector encoding wild-type PLC2 (Wt), PLC219 (19), PLC220-22 (20-22), or PLC219-22 (19-22). Twenty-four hours after transfection, the cells were incubated for a further 20 h at 37 °C. Cells from one well were then lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope of

PLC2, PLC219, PLC220-22 and PLC219-22.

3.5.4 The temperature induced activation mediated by exon

deletions is an isozyme specific effect for PLC2.

The structural organization of PLC2 between its two catalytic subdomains X and Y is very similar to that of its close relative PLC1 [289]. To address the question, whether deletion of residues in PLC1 corresponding to those encoded by exons 19 and 20-22 of PLCG2 has similar functional consequences, the relevant mutants of human PLC1 were produced. Since the human PLCG1 and PLCG2 genes exhibit distinct patterns of genomic organization, the deletions in PLC1 do not correspond to specific exons in PLCG1. Hence, the mutants are designated PLC1"19" and PLC1"20-22". Figure 3-24A shows that both PLC1"19" and PLC1"20-22", in contrast to their PLC2 counterparts, were already highly constitutively active at 37 °C (54- and 37-fold increases in inositol phosphate formation, respectively, in comparison to the increase caused by wild-type

PLC1 relative to control). Furthermore, only minor changes in their activity were observed upon reduction of the incubation temperature from 37 °C to 27 °C. Within the same range of temperatures, there were only minor changes in inositol phosphate formation by wild- type PLC1. Figure 3-24B shows that there were only subtle, if any changes in protein expression of the PLC enzymes with decreasing temperatures, except for PLC1"20-22" and PLC220-22, which showed increased expression. These changes notwithstanding, the results shown in Figure 3-24 clearly indicate that the temperature sensitivities of PLC1 and PLC2 carrying deletions corresponding to exons 19 and 20-22 of PLCG2 are distinct.

92 Results

Figure 3-24: Effect of the deletion of PLC1 residues corresponding to exon 19 or exon 20-22 of PLCG2 on the activity of PLC1 at subphysiological temperatures. A, COS-7 cells were transfected with 500 ng each per well of vector encoding either wild-type PLC1 (1Wt, ●), PLC1 (1”19”, ▼),

PLC1 (1”20-22”, ▲), PLC219(219, ), or PLC220-22(220-22, ). Twenty-four hours after transfection, the cells were incubated for a further 20 h in individual incubation chambers at the temperatures indicated at the abscissa, in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl, and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected as described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope of

PLC1, PLC1668-707, PLC1708-828, PLC219, and PLC219-22.

3.5.5 Stimulation of basal PLC2 activity by deletions within the SH domain tandem does not correlate with activation by cool temperatures.

The region encoded by exons 20 to 22 of PLCG2 contains an octapeptide (YRKMRLRY) at the end of SH2C that is absolutely conserved in PLC1 and has previously been shown to mediate inhibition of PLC2 in trans and in cis [59, 88, 91]. To determine whether deletion of the octamer, designated PCI (for phospholipase C inhibitor), is sufficient to mediate sensitivity of PLC2 to cool temperatures, the deletion mutant PLC2PCI was expressed in COS-7 cells and functionally compared to PLC220-22 and wild-type PLC2. At the 93 Results same time, three other mutants carrying deletions known to promote constitutive activity of

PLC2 at 37 °C, PLC2SA, PLC2SH, and PLC2SH2C, were characterized. Figure

3-25 shows that PLC2PCI shared most features of its temperature sensitivity with

PLC220-22 and, by extension, with PLC219 (not shown): only a very slight (approximately 2.5-fold) enhancement of its activity at 37 °C, a marked activation upon lowering the incubation temperature from 37 °C to 31 °C, and a decrease in activity at temperatures below 31 °C. In contrast, PLC2SA, PLC2SH, and PLC2SH2C, although exhibiting constitutive activity at 37 °C as reported earlier for both PLC1 and

PLC2 [59, 77], showed an only modest, further increase in activity with cooling, which was monophasic for PLC2SH2C and only vaguely biphasic for PLC2SA and

PLC2SH.

Figure 3-25: Effect of subphysiological temperatures on the activities of PLC2PCI, PLC2SA,

PLC2SH and PLC2SH2C. A, COS-7 cells were transfected with 50 ng each per well of vector encoding either wild-type PLC2 (●), PLC220-22 (▲), or PLC2PCI () (left panel), or with 10 ng each per well of vector encoding either wild-type PLC2SA (○), PLC2SH (□), or PLC2SH2C (◊) (right panel). Twenty- 94 Results four hours after transfection, the cells were incubated for 20 h in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl at the temperatures indicated at the abscissa, and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected as described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope of

PLC2 and its deletion mutants.

A more comprehensive comparison of the effects of the various SH tandem constituents, alone or in combination, on PLC2 constitutive activity and sensitivity to cool temperatures is shown in Figure 3-26. As described before for PLC1 and PLC2 [59, 77], deletion of

SH2N did not markedly affect basal activity at 37 °C, while deletion of SH2C, caused an approximately 2.6-fold increase. Unlike observed previously, however, deletion of SH3 also caused an increase in basal activity, which was even slightly higher (approximately

4.1-fold) than that monitored for deletion of SH2C, but lower than that for deletion of all three SH domains (approximately 7.5-fold). The increases for the variants lacking two SH domains, PLC2SH2NSH2C, PLC2SH2CSH3, and PLC2SH2NSH3, were approximately 3.6-, 1.9-, and 2.4-fold, respectively. While deletion of SH2N did not change the temperature sensitivity of PLC2 relative to the wild-type enzyme and deletion of all three SH domains resulted in a PLC2 mutant largely insensitive to cool temperature, all other deletion mutants showed a clear sensitivity. Maximal sensitivity (approximately

6.1-fold) was observed for PLC2SH2CSH3, followed, in that order, by SH2NSH3,

SH2NSH2C, SH2C,and SH3.

95 Results

Figure 3-26: Effect of subphysiological temperatures on the activities of PLC2 deletion mutants. COS- 7 cells were transfected with 500 ng each per well of empty vector (Co.) and vector encoding wild-type

PLC2 (Wt), or with 10 ng each per well of vector encoding PLC2SH2N (SH2N), PLC2SH2C (SH2C),

PLC2SH3 (SH3), PLC2SH2NSH2C (SH2NSH2C), PLC2SH2CSH3 (SH2CSH3), PLC2SH2NSH3

(SH2NSH3), or PLC2SH (SH). Twenty-four hours after transfection, the cells were incubated for 20 h in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl at 31 °C or 37 °C, and the levels of inositol phosphates were then determined.

3.5.6 The hypermorphic mutations Ali5 and Ali14 in combination

induce temperature‐mediated activation of PLC2.

The PLC2 point mutants identified in the two mouse disease models Ali5 and Ali14, D993G and Y495C, respectively, have previously been shown to cause constitutive activation of PLC2 at 37 °C [1, 56, 60, 301]. In one of the models, Ali5, autoinflammatory dermatitis commenced in the superficial layers of the paws and ears, i.e. in cool body regions [301]. It has been noted that this distribution resembles those of cutaneous granulomatous lesions in patients with PLAID [190]. We therefore examined and Ali5, compared the effect of cool temperatures on the activity of the mutants PLC2 Ali14 Ali5/Ali14 PLC2 , and the compound mutant PLC2 in intact cells. Figure 3-27A, left panel, shows that all three mutants exhibit moderately enhanced activity at 37 °C Ali5/Ali14 Ali5 Ali14 (PLC2 > PLC2 > PLC2 ) as decribed before [60]. Importantly, however, the three mutants displayed markedly different sensitivities to cool temperatures. Thus, while 96 Results

Ali5/Ali14 the response of PLC2 to decreasing temperatures closely resembled the pattern Ali14 observed for PLC220-22 (Figure 3-27A, center panel), PLC2 showed only a minor, monophasic increase in activity, which was opposite to the decrease observed for wild-type Ali5 PLC2 (Figure 3-27A, right panel). PLC2 displayed an intermediate phenotype, nonetheless showing a 3.9-fold activation upon cooling from 37 °C to 31 °C in comparison to wild-type PLC2.

Ali5 Ali14 Figure 3-27: Effect of subphysiological temperatures on the activities of PLC2 , PLC2 , and Ali5/Ali14 PLC2 . A, left panel, COS-7 cells were transfected with 500 ng each per well of either empty vector D993G Y495C (Co.), or vector encoding either wild-type PLC2 (PLC2), PLC2 (Ali5), PLC2 (Ali14), or Y495C/D993G PLC2 (Ali5/14). Twenty-four hours after transfection, the cells were incubated for 20 h in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl at 31°C or 37°C, and the levels of inositol phosphates were then determined. Center and right panel, COS-7 cells were transfected with 500 ng each per well of vector encoding either wild-type PLC2 (PLC2, ○), D993G Y495C Y495C/D993G PLC220-22 (20-22, ▲), PLC2 (Ali5, ), PLC2 (Ali14, ), or PLC2 (Ali5/14, ●). Twenty-four hours after transfection, the cells were incubated for 20 h in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl at the temperatures indicated at the abscissa, and the levels of inositol phosphates were then determined. B, left, center, and right panel, COS-7 cells were transfected as described above, and cells from one well were lysed in 100 µl of SDS-PAGE

97 Results sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an Ali5 Ali14 Ali5/Ali14 antibody reactive against the c-myc epitope of PLC2, PLC2 , PLC2 , PLC2 , and PLC220-22.

3.5.7 The temperature‐induced activation of PLC219 and

PLC220‐22 is distinct from a loss of autoinhibition.

Figure 3-28 shows a summary of the effects of the PLC2 mutations analyzed so far, in which the constitutive activities of the mutants at 37 °C were plotted against their activation by cool temperature (31 °C). Analyzing the results by cluster robustness analysis [139], it was found that the functional phenotypes fell into three clusters: one characterized by high constitutive activity at 37 °C (SA, SH, SH2C, SH3, and Ali5), one by high activation by cold (19, 20-22, PCI, and 19-22), and a third, intermediate one, consisting of mutants that could not be clearly categorized into one of the other two. These results suggest that the activation pattern of the two PLAID mutants, PLC219 and

PLC220-22, is distinct form a mere loss of autoinhibition, and indicate, by extension, that it is based on a specific molecular mechanism(s).

Figure 3-28: Summary and clustering of the effects of the PLC2 mutations. The ratios of the activities of the mutants and the activity of the wild-type enzyme at 37 °C corresponding to constitutive activity, were 98 Results

plotted against the ratios of the mutant activities at 31 °C and 37 °C, representing PLC2 activation by cool temperatures. Since some activities determined in cells expressing wild-type or variant PLC2 were similar to background inositol phosphate formation in the absence of exogeneous PLC2, the former values were used to calculate the ratios throughout. The values shown on both axes thus represent minimum estimates. The results of ensemble clustering (majority vote) for the leave-one-out clusterings are indicated by three different colours.

3.5.8 A hypermorphic missense mutation in PLC2 associated with APLAID is only limited activated by cool temperatures.

A hypermorphic missense mutation in PLCG2 was associated with a dominantly inherited autoinflammatory disease with immunodeficiency, termed APLAID (autoinflammation and PLC2-associated antibody deficiency and immune dysregulation). In contrast to PLAID, APLAID patients do not suffer from cold-induced urticaria. The detected missense mutation is located in the SH2C domain, resulting in a replacement of serine 707 by a S707Y tyrosine residue [311]. Figure 3-29A shows that the expression of PLC2 at 37 °C leads to an about 16.2-fold increase in inositol phosphate formation compared to cells expressing wild-type PLC2, which is in accordance with previously published data [311]. S707Y Despite the fact that APLAID patients do not suffer from cold urticaria, PLC2 displayed an intermediate sensitivity to cold temperatures as shown in Figure 3-29A. However, the fold activation upon cooling from 37 °C to 31 °C is about 20.6 magnitudes lower than the one observed for PLC220-22 in this experiment. The expression levels of the enzymes were similar and did not change significantly with decreasing temperatures (Figure 3-29B).

99 Results

S707Y Figure 3-29: Effect of subphysiological temperatures on the activity of PLC2 . A, COS-7 cells were transfected with 500 ng each per well of vector encoding either wild-type PLC2 (Wt, ○), PLC220-22 (20-22, ▲), or PLCS707Y (S707Y, ). Twenty-four hours after transfection, the cells were incubated for 20 h in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl at the temperatures indicated at the abscissa, and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected and incubated at the temperatures described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope.

3.5.9 Phosphorylation of Y733 plays a crucial role in the

temperature‐mediated activation of PLC2.

The effect of decreasing temperatures on the phosphorylation-induced activation of PLC2 was investigated by using the phosphomimetic mutants that were already characterized in section 3.3. As shown in Figure 3-30A the two phosphomimetic mutants, Y733E/Y753E/Y759E/Y1197E Y733E/Y753E/Y759E/Y1197E/Y1217E PLC2 and PLC2 that showed the highest basal activity at 37 °C are, in contrast to all other phosphomimetic mutants, further activated at 31 °C (about 3.4-fold and 3.9-fold compared to their activity at 37 °C, respectively). The expression levels of the mutants at 31 °C did not significantly differ from those at 37 °C (Figure 3-30B).

100 Results

Figure 3-30: Specific activity of PLC2 phosphomimetic mutants at 37 °C and 31 °C. A, COS-7 cells were transfected as indicated at the abscissa with 500 ng of empty vector (Co.), vector encoding PLC2 (Wt) Y733E Y753E Y759E Y753E/Y759E or PLC2 phosphomimetic mutants PLC2 (733), PLC2 (753), PLC2 (759), PLC2 Y733E/Y753E/Y759E Y753E/Y759E/Y1197E Y753E/Y759E/Y1217E (753/759), PLC2 (733/753/759), PLC2 (753/759/1197), PLC2 Y753E/Y759E/Y1197E/Y1217E Y733E/Y753E/Y759E/Y1197E (753/759/1217), PLC2 (753/759/1197/1217), PLC2 Y733E/Y753E/Y759E/Y1197E/Y1217E (733/753/759/1197), PLC2 (733/753/759/1197/1217). Twenty-four hours after transfection, the cells were incubated for 20 h in individual incubation chambers in the presence of myo-[2- 3H]inositol (2.5 µCi/ml) and 10 mM LiCl at 31°C or 37°C, and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected and incubated at the temperatures described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope.

101 Results

The critical role of phosphorylation of Y733 for the activation of PLC2 at subphysiological temperatures was further characterized as shown in Figure 3-31A. The quadruple phosphomimetic mutant (Y753E/Y759E/Y1197E/Y1217E) in which phosphorylation of all four well known phosphorylation sites was mimicked and the PLC2 mutant in which phosphorylation of Y733 alone was mimicked showed no or only a minor monophasic response to decreasing temperatures. In contrast, the combination of both mutants, resulting in the quintuple phosphomimetic mutant displayed a clear sensitivity to cool temperatures that also showed the characteristic biphasic stimulatory response with decreasing temperatures that was observed before for PLC220-22. Nevertheless, the fold activation upon cooling from 37 °C to 31 °C is about 29 magnitudes lower than the one observed for PLC220-22 in this experiment. The expression levels of PLC220-22 and the quintuple phosphomimetic mutant did not change significantly with decreasing temperatures (Figure 3-31B).

Figure 3-31: Effect of subphysiological temperatures on the activities of PLC2 phosphomimetic mutants. A, COS-7 cells were transfected with 500 ng each per well of vector encoding either wild-type Y733E Y753E/Y759E/Y1197E/Y1217E PLC2 (2Wt, □), PLC220-22 (20-22, ), PLC2 (Y733E, ), PLC2 (4YtoE, Y733E/Y753E/Y759E/Y1197E/Y1217E ), or PLC2 (5YtoE, ● ). Twenty-four hours after transfection, the cells were incubated for 20 h in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl at the temperatures indicated at the abscissa, and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected and incubated at the temperatures described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope.

102 Results

3.5.10 The activation of PLC2 by cool temperatures and RacGTPases is competitive.

Beside the activation by tyrosine phosphorylation, PLC2 can also be activated by RacGTPases. The experiment shown in Figure 3-32 was designed to determine the effect of decreasing temperatures on the ability of constitutively active Rac2G12V to activate G12V PLC219 in comparison to wild-type PLC2. At 37 °C, Rac2 caused similar enhancements of the activity of both enzymes (Figure 3-32A, left panel). Lowering the incubation temperature, however, took a very different effect on Rac2G12V-mediated activation of the two enzymes. While only minor changes were observed for wild-type G12V PLC2, there was a progressive loss of the stimulatory effect of Rac2 on PLC219, ranging from about 27.2-fold at 37 °C, to about 1.3-fold at 31 °C, and no effect at 27 °C. This loss is unlikely to be due to exhaustion of the inositol phospholipids substrate at lower temperatures, since PLC219 is well capable of producing even higher levels of inositol phosphates when expressed at higher density (Figure 3-32A, right panel). It thus follows that activation of the PLC219 deletion mutant studied herein by cool temperatures is competitive with its activation by Rac2G12V.

103 Results

G12V Figure 3-32: Effect of subphysiological temperatures on the Rac2 -mediated activation of PLC2 and PLC219. A, Left panel, COS-7 cells were cotransfected with 500 ng each per well of vector encoding either wild-type PLC2 (, □) or 100 ng of vector encoding PLC219 (●, ○) together with 25 ng per well of either empty vector (open symbols) or vector encoding Rac2G12V (closed symbols). Twenty-four hours after transfection, the cells were incubated for 20 h in individual incubation chambers in the presence of myo-[2- 3H]inositol (2.5 µCi/ml) and 10 mM LiCl at the temperatures indicated at the abscissa, and the levels of inositol phosphates were then determined. Right panel, COS-7 cells were transfected with either 500 ng per well of vector encoding wild-type PLC2 (□) or increasing amounts (10 ng, 20 ng, 50 ng, 100 ng, 200 ng, 500 ng) of vector encoding PLC219 (○).The total amount of DNA was maintained constant at 500 ng by adding empty vector. Twenty-four hours after transfection, the cells were incubated for 20 h in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl at 31°C. Inositol phosphate formation was then determined as described under Materials and Methods. B, COS-7 cells were transfected and incubated at the temperatures described above, and cells from one well were lysed in 100 µl of SDS- PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope.

The experiment shown in Figure 3-33 examines the influence of the PLC219 expression level on the stimulatory effect of Rac2G12V. Two markedly different patterns were 104 Results

G12V observed. At 37 °C, stimulation of PLC219 by Rac2 was lost upon lowering the enzyme´s cellular expression level. In contrast, at 31 oC, enzyme stimulation was markedly enhanced with decreasing expression (Figure 3-33, left panel). There was no difference in G12V the sensitivity of PLC219 to Rac2 under these two conditions (Figure 3-33, right G12V panel).Given that Rac2 mediates translocation of PLC2 from the cytosol to the membrane fraction, i.e. the site of the enzyme substrate [168, 203], these findings suggest that enhanced membrane translocation by Rac2G12V is only effective at low cellular expression of PLC219 at 31 °C. According to this model, at higher expression levels,

PLC219 would already be present in the membrane compartment at concentrations sufficient to account for high inositol phosphate formation and, hence, less sensitive to further stimulation by activated Rac.

G12V Figure 3-33: Cool temperature shifts the requirement of PLC2 for Rac2 -mediated activation to G12V lower cellular PLC2 expression levels without altering the enzymes sensitivity to Rac2 . Left panel, COS-7 cells were cotransfected with increasing amounts (10 ng, 20 ng, 50 ng, or 100 ng) of vector encoding G12V PLC219 together with 25 ng per well of either empty vector (, □) or vector encoding Rac2 (●,○).The total amount of DNA was maintained constant at 500 ng by adding empty vector. Twenty-four hours after transfection, the cells were incubated for 20 h in individual incubation chambers in the presence of myo-[2-3H]inositol (2,5 µCi/ml) and 10 mM LiCl at 31 °C (closed symbols) or 37 °C (open symbols). Inositol phosphate formation was then determined as described under Materials and Methods. Right panel, COS-7 cells were transfected with either 100 ng () or 10 ng () per well of vector encoding PLC219 or 20 ng G12V per well of vector encoding Rac2 (□, ○) or cotransfected with 100 ng () or 10 ng (●) of vector encoding

PLC219 together with increasing amounts (0.5 ng, 1 ng, 2 ng, 5 ng, 10 ng, or 20 ng) of vector encoding Rac2G12V.Twenty-four hours after transfection, the cells were incubated for 20 h in individual incubation

105 Results chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl at 31 °C (●,○,) or 37 °C

(,□,). Inositol phosphate formation was then determined as described under Materials and Methods.

3.5.11 The cold‐induced activation of PLAID PLC2 mutants is mediated by the split PH domain.

The activation of PLC2 by Rac is mediated by the internal, split PH domain of the enzyme [28, 275]. The interdependence of the stimulatory effects of Rac2G12V and cooling prompted us to determine the role of this domain in PLC2 deletion mutant activation by cooling. Replacement in spPH of PLC2 of a tryptophan residue conserved in all PH domains [196] by alanine has previously been shown to result in a loss of PLC2 stimulation by Rac, but not by tyrosine phosphorylation [59]. Figure 3-34 shows that the W899A replacement completely abrogates the response of the enzyme to cool temperature activation. Furthermore, replacement of one or the other half of the split PH domain of

PLC2 by the corresponding portions of PLC1, previously shown to block activation of the mutant enzymes by Rac, but to take no effect on their catalytic activity [275], also eliminated enzyme activation by cool temperatures. In contrast, replacement of both PH domains halves, which does not convey Rac sensitivity to PLC2 [275], fully rescued the response of PLC2 to cooling. These results indicate that a functional split PH domain, not necessarily the one genuinely contained in PLC2, is required for the sensitivity of PLC2 to cool temperatures.

Two residues of the N-terminal half of the PLC1 spPH domain, Y509 and F510, have previously been shown to be important for the ability of spPHN to cooperate with SH2C to maintain PLC1 in an inactive conformation [48, 60] and to reside, by NMR mapping, at the interface of spPH and SH2C [27]. To examine the role of the interaction in mediating the cool temperature response of PLC2, the two residues were substituted by alanine residues in PLC219 carrying the entire spPH of PLC1 (PLC2-11-19), thus generating the mutant PLC2-11-19-Y509A/F510A. Consistent with earlier findings [48, 60], there was a slight increase in basal activity of the latter mutant in comparison to PLC2-11-19.

Remarkably, however, and in striking contrast to PLC2-11-19, there was no response of the mutant to a decrease in incubation temperature from 37 °C to 31 °C. These results

106 Results

suggest that stimulation of the PLAID PLC2 deletion mutants by cool temperatures is dependent on the integrity of the site on spPHN that normally interacts with SH2C. Hence, it is likely that a functional interaction between this site and a structural element of the SH domain tandem retained in the deletion mutants mediates their cool temperature sensitivity.

Figure 3-34: Role of the spPH domain in the specific activation of PLC219 at subphysiological temperatures. A, COS-7 cells were transfected with 500 ng each per well of vector encoding either wild- W899A type PLC2 (Wt), PLC219 (19), PLC219 (W899A19), PLC2-21219 (12-19), PLC2-22119 Y509A/F510A (21-19), PLC2-21119 (11-19), PLC2-211 (Wt-11), or PLC2-21119 (11-19-Y509A/F510A). Twenty-four hours after transfection, the cells were incubated for 20 h in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl at 31 °C or 37 °C, and and the levels of inositol phosphates were then determined. B, COS-7 cells were transfected and incubated as described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-

PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope of PLC2 and PLC2 mutants.

107 Results

The observation that PLC2SH, a mutant lacking the entire SH tandem, displays enhanced basal activity, but is largely insensitive to activation by cold is remarkable, because this mutant lacks all constituents thought to mediate autoinhibition of the enzyme (Figure 3-35A). On the one hand, this finding is very well consistent with and confirms the notion that activation by cold and loss of autoinhibition by SH tandem constituents are independent processes. On the other hand, it is rather puzzling because cold activation appears to be both dependent on and independent of specific constituents of the SH tandem (cf. Figure 3-25, Figure 3-26, and Figure 3-35A). However, given the importance of the split PH domain for cold activation of PLC2 emerging from the experiments shown in Figures 3-31, 3-32, and 3-33, the possibility remained that activation by cold temperatures is sterically restricted in the mutant PLC2SH by the (artificial) covalent linkage between the two halves of the split PH domain. To examine this possibility, we took advantage of previous findings suggesting that PLC isozymes can be expressed as two independent polypeptide fragments containing either the catalytic subdomain X or Y [94, 309] (section

3.2). Thus, PLC2SH was expressed either as a single polypeptide and as two separate chains, encompassing either the sequence from the amino terminus up to and including spPHN or the sequence beginning with and including spPHC and extending to the carboxyl- terminus (cf. Figure 3-6). The cells were then analyzed and compared for inositol phosphate formation at 37 °C and 31 °C. Figure 3-35B shows that expression of the

N-terminal or the C-terminal half of PLC2 did not lead to enhanced inositol phosphate formation. In marked contrast, when both halves were coexpressed, basal inositol phosphate formation at 37 °C was markedly enhanced. More importantly, however, decreasing the incubation temperature to 31 °C caused a substantial, approximately 6.8-fold enhancement of PLC activity, in striking contrast to cells expressing the SH deletion mutant of PLC2, where no change in activity was observed. These results strongly suggest that the structural element(s) mediating cold activation of PLAID PLC2 mutants reside outside the SH tandem and that its responsiveness to cool temperatures is blocked by covalent linkage between the two constituents of the PLC2 split PH domain.

108 Results

Figure 3-35: Covalent linkage of the split PH domain halves blocks PLC2 sensitivity to cool temperatures. A, Schematic representation of the deletions within the SH domain tandem. A linear sketch of the region between residues 515 and 839 of human PLC2 is shown. The positions of the two repeats detected within the two SH2 domains (KDGTFLVR/RDGAFLIR and GRVQHCRI/GKVKHCRI; SH2N/SH2C, identical residues underlined) and of the various deletions used in this study are indicated. B, COS-7 cells were transfected with 500 ng each per well of either empty vector (Co.) or vector encoding wild-type PLC2

(PLC2), 10 ng of vector encoding PLC2SH (SH), 250 ng of vector encoding either PLC2-X (X) or

PLC2-Y (Y), or 250 ng each of both of the latter two vectors (X+Y).Twenty-four hours after transfection, the cells were incubated for 20 h in individual incubation chambers in the presence of myo-[2-3H]inositol (2.5 µCi/ml) and 10 mM LiCl at 31 °C or 37 °C, and and the levels of inositol phosphates were then determined. C, COS-7 cells were transfected and incubated as described above, and cells from one well were lysed in 100 µl of SDS-PAGE sample buffer. An aliquot was subjected to SDS-PAGE, and immunoblotting was performed using an antibody reactive against the c-myc epitope of PLC2 and PLC2 mutants.

109 Discussion

4 Discussion

PLCs are signaling molecules that catalyze the hydrolysis of phosphatidylinositol 4,5- bisphosphate to produce the Ca2+-mobilizing second messenger inositol 1,4,5-trisphosphate and the protein kinase C-activating second messenger diacylglycerol [171, 186, 251]. The mammalian PLCs are divided into six subfamilies, designated , , , , , and  [99, 129, 181, 225, 254, 255, 313]. Within the PLC subfamily, the two PLC isoforms share a number of features that are distinct from those of the other PLC subfamilies. The most striking difference is the insertion of additional domains between the catalytic subdomains X and Y. This specific array (SA) contains a second, split pleckstrin homology (spPH) domain, consisting of two halves separated by two Src homology 2 (SH2) domains and one

Src homology 3 (SH3) domain. Activation of PLC2 upon B cell antigen receptor (BCR) stimulation has been implicated to be a critical step in the BCR-mediated Ca2+ signaling

[81]. Therefore, it is important to understand the mechanisms of how the activity of PLC2 is stimulated and inhibited.

The need for a stringent regulation of PLC activity in unstimulated cells becomes apparent on closer examination of the balance of energy of a cell. A constitutively active PLC isozyme would imply a tremendous waste of energy, needed for the resynthesis of the PLC substrate PIP2 out of DAG and inositol (5 equivalents of ATP) [212]. Furthermore, the 2+ depletion of PIP2 and the persisting Ca signalling may lead to regulatory feedback mechanisms that cause insensitivity of the cells to extracellular stimuli. The physiological consequences of constitutive activation in the case of PLC2 were reported to be severe defects in immune regulation leading to autoimmunity and immune dysregulation in mouse models as well as in humans [1, 190, 301, 311]. Therefore, a better knowledge of the mechanisms of autoinhibitory regulation of the PLC2 isozyme is of exceptional importance.

110 Discussion

4.1 Autoinhibitory and phosphorylation‐induced

regulation of PLC2 activity

The involvement of PLC X-Y linkers in the intramolecular regulation were already revealed in the 1990s by several different experimental approaches, all of which have one common ground: the deletion of either parts of or the whole X-Y linkers or the disruption of the structure of these linkers. The deletion of PLC1 aa 517-883 encompassing the SH2SH2SH3 tandem and, unknown to the authors at that time, parts of the spPH domains

[93], as well as the coexpression of amino and carboxy terminal halves of PLC1 [94] and

PLC2 [309] to assemble enzymes lacking the SH domains and parts of the spPH domains in case of PLC1 or the X-Y linker in case of PLC2, created constitutively active PLC isozymes. Similarly, protease cleavage of PLC1 [39, 55], PLC2 [226], and PLC1 [62] within the X-Y linker region, disrupting the structure of this region, also resulted in an increase in basal PLC activity. Furthermore, PLC1 and PLC1 were inhibited in trans by peptides corresponding to sequences within the X-Y linkers [91]. All described approaches finally led to one common result, namely the loss of autoinhibitory regulation of the appropriate PLC isozymes. The `hinged´ lid model was proposed as a general model for the autoinhibitory regulation of PLC isozymes [77]. This model is based on biochemical analysis [83] as well as on the crystal structure of PLC2 in complex with Rac1 [117] and describes the occlusion of the active site of PLC isozymes by their X-Y linkers. Upon activation, PLC isozymes are recruited to and orientated at plasma membranes at which the negatively charged X-Y linker is repelled from the negatively charged membrane, revealing the active site and allowing substrate entry and processing. Since the X-Y linker of PLC isozymes is, in contrast to all other PLC isozymes, composed of multiple domains and does not exhibit large accumulations of negatively charged side groups, it is likely that these PLCs are subject to different autoinhibitory mechanisms.

The results of this work show that the deletion of either the whole specific array (SA) or just the SH2SH2SH3 (SH) domain tandem resulted in a marked increase in basal activity of the enzymes (both about 40-fold compared to wild-type), suggesting that PLC2 is negatively regulated by its specific array (Figure 3-1). Since PLC2SA and PLC2SH did not show any obvious differences in PLC activity in transfected COS-7 cells, the most simple reasoning for this result would be that the spPH domains, present in PLC2SH but 111 Discussion

missing in PLC2SA are not involved in autoinhibitory regulation of the enzyme and that the deletion of the SH domain tandem is sufficient to overcome autoinhibition. However, the effect of the direct deletion of the spPH domains on the basal activity of PLC2 is discussed contrary in literature. The Katan group [59] reported that upon deletion of both halves of the spPH domain the basal activity of PLC2 increased whereas the Sondek group [77] stated that there is no change in the activity of the enzyme compared to wild-type

PLC2. The contradictory effects possibly arise from the different deletion strategies that were utilized by the two groups. While the Katan group directly deleted aa 471-513 and aa 841-913, the Sondek group replaced aa 464-515 and aa 852-921 by a G-S-G-S linker. With regard to the direct deletion of the spPH domains, there is no clear evidence for their involvement in PLC2 autoinhibitory regulation. However, results from our group could show that chimeric PLC1 mutants in whom the spPH domains were replaced by those of

PLC2 display an enhanced phospholipase activity in intact cells as well as in vitro. This was in particular the case when the N-terminal spPH (spPHN) domain was swapped, indicating a direct involvement of the spPH domains in the intramolecular regulation of

PLC1. Interestingly, the counterpart chimeric PLC2 mutants were not susceptible to enhanced basal activity [275]. Possibly, PLC2 underlies a more stringent intramolecular regulation than PLC1, or the spPH domain is not involved at all in this particular regulatory mechanism. Further evidence for the contribution of the spPHN domain in the regulation of PLC1 provided a study in which the SH2C domain was found to be involved in autoinhibitory regulation of PLC1 [48]. It was suggested by the authors that the spPHN domain (in particular residues Y509 and F510) regulates the accessibility of the SH2C domain, and thereby plays an ancillary role in the regulation of PLC1 activity. However, it appears that the identified residues in the spPH domain are not sufficient to fully overcome autoinhibition, but possibly drive the lipase to a higher susceptibility for the activation by upstream activators like Rho GTPases, tyrosine or non-tyrosine kinases [48]. This hypothesis was further supported by a point mutation in PLC2 (Ali14, Y495C) found in a murine model and corresponding to Y509 in PLC1. Although the Ali14-type mutation did not affect basal activity of PLC2, the lipase exhibited enhanced stimulation by Rac and an increased response to EGF stimulation in transfected cells [60]. Finally, NMR spectra revealed that the Ali14 mutation appears to be located at the interface of the spPH and

SH2C domains [27].

112 Discussion

In addition to the depicted role of the spPH domains in autoinhibitory regulation, both domains are of crucial importance for the Rac2-mediated activation of PLC2. Biochemical analysis, NMR spectra, as well as the crystal structure of the isolated spPH domain bound to Rac2 clearly demonstrated that Rac2 binds to PLC2 as well as to the isolated spPH domain [28, 275]. Taking these results into account it is to be expected that the PLC2SA mutant lacking the spPH domains is no longer activated by Rac2. This hypothesis was confirmed experimentally. Furthermore, the PLC2SH mutant, still harbouring spPH domains, was additionally stimulated by Rac2G12V (Figure 3-2), indicating (i) that the two halves of the spPH domain fold into a functional PH domain and (ii) that the SH domain tandem in not required for the activation of PLC2 by Rac2. Moreover, the results argue that the autoinhibitory regulation and the activation by Rho GTPases rely on distinct mechanisms that partially implicate common domains, in particular the spPH domain.

The PLC SH domain tandem is composed of two SH2 domains and one SH3 domain, at which each domain possesses particular functions. SH3 domains are known to interact with proteins containing proline rich sequences [306], whereas the SH2N domain is critical for binding to receptor tyrosine kinases or adaptor molecules like BLNK, that mediate translocation of the lipase to the plasma membrane [35, 49, 106, 107, 119, 206]. The function of the SH2C domain appears to be more directed to the intramolecular regulation of PLC isozymes. It was demonstrated for PLC1 that the deletion or the functional interruption of the SH2C domain leads to constitutive activation of the lipase [48, 77].

Likewise, the deletion of the SH2C domain of PLC2 also resulted in an increase in basal activity. In contrast, point mutations that specifically affect the function of the SH2 domains were not sufficient to enhance the basal activity of the enzyme. Also, the deletion of SH2N or SH3 had no impact on PLC activity [59]. Based on these results, one would expect that the insertion of the SH2C domain into PLC2SH should be adequate to reduce the elevated basal activity of the deletion mutant. Nevertheless, neither the insertion of the

SH2C domain nor the insertion of the SH2N or SH3 domains gave rise to a fully inhibition of the activity of PLC2SH (Figure 3-3). The observed inhibition for each of the SH domains ranged between 33 % and 47 %, indicating that each of the domains contribute equally to the autoinhibitory regulation of PLC2. Likewise, the insertion of the SH2C domain together with the SH2N domain, and the insertion of the SH2N domain together with the SH3 domain were not sufficient to reduce the basal activity of PLC2SH to the level of the wild-type enzyme. Intriguingly, the SH2C domain in combination with the SH3 113 Discussion domain exhibited the strongest inhibitory effect on the deletion mutant and inhibited the basal activity of PLC2SH by almost 88 % (Figure 3-3). A more detailed and comparative analysis of the activities of PLC2SH+SH2C, PLC2SH+SH3 and

PLC2SH+SH2C-SH3 revealed that both mutants in whom either the SH2C or the SH3 domain was inserted showed an about 6.8-fold higher activity than the mutant in which both domains in combination were inserted (Figure 3-4). The linker between the SH2C and the SH3 domain contains two important regulatory tyrosine phosphorylation sites (Y753 and Y759). Phosphorylation of PLC1 at the corresponding tyrosine residues mediates high intramolecular interactions with the SH2C domain resulting in conformational changes that release the enzyme from its autoinhibitory constraint [77]. However, insertion of the SH3 domain together with the SH2C-SH3 linker exhibited no further inhibitory effect on the activity of PLC2SH compared to the insertion of the SH3 domain alone (Figure 3-3).

Taken together, the results imply that the SH2C domain together with the SH3 domain and the SH2C-SH3 linker are the major elements involved in the autoinhibitory regulation of

PLC2. However, on the basis of the performed experiments, the role of the SH2C-SH3 domain linker is still undefined. There is the possibility that the SH2C domain together with the SH2C-SH3 linker in combination with phosphorylation of Y753 and Y759 contribute to autoinhbitory regulation, like it was described for PLC1 [77]. However, the direct deletion of the PLC2 SH2C domain, like it is described in literature [59], is not sufficient to address this hypothesis. To test the possibility that the observed inhibition of

PLC2SH activity by introducing the SH2C and SH3 domain is dependent on both domains or probably only on the SH2C domain together with the linker region, it would be adequate as a future experiment to insert the SH2C domain together with the linker region, containing Y753 and Y759, into PLC2SH.

As previously demonstrated for PLC2SH, also the appropriate insertion mutants are further activated by Rac2G12V (Figure 3-5). Impressively, only very small amounts of vector (1ng/well) encoding for the insertion mutants as well as for PLC2SH were needed for transfection to obtain an observable stimulation by Rac2G12V. In contrast, no stimulation of wild-type PLC2 was detectable under these conditions. These results point to an enhanced susceptibility of the desinhibited PLC2 deletion mutants to Rac2-mediated activation. The possibility that the binding of Rac2 leads to conformational changes in the three dimensional structure of PLC2, releasing the lipase from its autoinhibitory

114 Discussion constraints and thereby achieving a stimulated PLC activity was excluded by using FRET (fluorescence resonance energy transfer) analysis. At least in this kind of experiments,

Rac2 binding did not provoke major conformational rearrangements in the PLC2 structure

[59]. The same results were observed for PLC2, where the comparison of the crystal structure of PLC2 alone and in complex with Rac1 revealed that there are no structural changes upon binding of Rac1 [83, 117]. The proximity to membranes was shown to be essential for the Rac2-mediated activation of PLC2 [59]. However, the role of membrane proximity in the release of autoinhibition of PLC2, as it was described for PLC2 [83], has to be further clarified.

The functional reassembly of PLC1 [94] or PLC2 [309] out of independently expressed amino and carboxy terminal halves was initially used to study the formation of the catalytic site or the functional contribution of different domains to enzyme function, respectively. The physically association between the two halves of the catalytic domain was furthermore demonstrated by limited proteolysis of PLC1 within the SH domain tandem resulting in enhanced PLC activity [62]. A common outcome of these studies was (i) that both enzymes can reassemble from separately expressed halves and form a catalytically active enzyme and (ii) that in both cases the specific activation of the enzyme was elevated compared to the holoenzyme if the X-Y linker was missing. The same results were obtained in this work for independently expressed halves of PLC2, in which the

N-terminal part of the enzyme, reaching from the N-terminus up to the end of spPHN

(PLC2-X, aa 1-514), and the C-terminal part, starting at the beginning of spPHC and ending at the C-terminus of the enzyme (PLC2-Y, aa 842-1265), reassembled to a catalytically active enzyme that showed an elevated basal activity (about 20-fold) compared to the holo-PLC2 enzyme (Figure 3-6). This showed once again the role of the

SH domain tandem as an intrinsic negative regulator of PLC2 activity and furthermore demonstrates the formation of an intact catalytic TIM-barrel by the two halves, X and Y, of the catalytic domain. The complex between PLC2-X and PLC2-Y was additionally activated by Rac2G12V (Figure 3-6) suggesting the formation of an intact spPH domain, which is necessary for the Rac2-mediated activation of PLC2 [275]. This hypothesis is further supported by the three dimensional structure of the PLC1 spPH domain, exhibiting that the structure of the spPH domain is not changed by the insertion of the SH domain tandem [288]. The formation of intact domains out of two halves might be driven by at

115 Discussion least two distinct reasons: (i) both halves contain residues that are directly involved in the catalytic activity in case of the catalytic domains, or in the Rac2-mediated activation in case of the spPH domain, or (ii) only one of the halves contains the described residues and the other halve imparts the proper folding of the domain. In case of the PLC1 spPH domain it was shown that the isolated halves of the spPH domains remained unfolded and that correct folding was only achieved upon interaction with the complementary halve

[288]. Based on NMR data and the crystal structure of the 2 spPH domain bound to GTP-loaded Rac2G12V, the interaction site of Rac2 was mapped to the C-terminal halve of the spPH domain, in which residues K862, V893 and F897 were found to be important for the activation of PLC2 by Rac2 [28, 275]. Taken together, it is most likely that the C-terminal halve of the spPH domain mediates the interaction with Rac2 and the N-terminal halve is necessary to achieve the correct folding of the spPH domain.

Impressively, PLC2-X and PLC2-Y also formed complexes with full-length PLC2 enzymes in transfected COS-7 cells. This was demonstrated by coexpression of H327A/H372A PLC2 , in which two important residues in the catalytic domain X were replaced by alanin resulting in a catalytically inactive enzyme [235], together with PLC2-X, harbouring a functional X domain. The complementary experiment regarding the spPH F897Q domain was performed using PLC2 , which is insensitive to stimulation by Rac2 due to the mutation of a residue in the C-terminal spPH domain important for the interaction with Rac2 [275]. The results shown in Figure 3-7 arrestingly demonstrate that the X domain of PLC2-X can functionally restore the nonfunctional X domain of H327A/H372A PLC2 resulting in a catalytically active enzyme. The identical findings were F897Q observed for PLC2 in which the spPHC domain of PLC2-Y redressed the ability of the enzyme to be responsive to the activation by Rac2G12V. Another interesting observation emerging from this experiment is that no enhanced basal activity was observed upon coexpression of full-length PLC2 mutants and the truncated PLC2 halves. This is probably due to the presence of the SH domain tandem, which was shown to exert an intrinsic inhibitory effect on PLC2 (section 3.1.1). These results suggest that the SH domain tandem has not to be covalently linked between the two halves of the spPH domain to exert its inhibitory effect. Recombinant SH2SH2SH3 domain regions of PLC1 and

PLC2 inhibited phosphoinositide hydrolysis induced by PLC1, PLC1, PLC2, and PLC1 in vitro [91], supporting the above described hypothesis.

116 Discussion

Phosphorylation of proteins is an important mechanism for activating or inhibiting enzymes. This reversible posttranslational modification is furthermore crucial for the assembly of multiprotein complexes. Antigen-mediated activation of the BCR entails the phosphorylation and activation of PLC2 at specific tyrosine residues located in the SH domain tandem, more precisely in the linker between the SH2C and SH3 domain (Y753 and Y759), and at the C-terminal end of the lipase (Y1197 and Y1217) [135, 283]. The replacement of Y753 and Y759 by phenylalanine residues revealed that both tyrosine residues are important for PLC2 signalling function in B cells [217]. However, the regulatory role of phosphorylation of Y1197 and Y1217 remains questionable and also the mechanistically impact of phosphorylation of different tyrosine residues on the activity of

PLC2 remains an open question. In order to study the direct impact of tyrosine phosphorylation on the activity of PLC2, we replaced all four known tyrosine phosphorylation sites alone or in combination, by glutamic acids, generating PLC2 phosphomimetic mutants. An acidic substitution, like glutamic acid or aspartic acid, can mimic the effects of tyrosine, serine or threonine phosphorylation by its negative charge. This is a widely used approach to study the effects of phosphorylation on the activity of proteins [47, 125, 200, 296]. The replacement of Y753 and Y759 alone or in combination by glutamic acid had no influence on the activity of PLC2. This could indicate that phosphorylation of Y753 and Y759 is not sufficient to activate PLC2 and that the phosphorylation sites on the C-terminus of the enzyme are additionally required to achieve activation of the lipase. In deed, mimicking phosphorylation of Y753 and Y759 together with Y1197 or Y1217 resulted in a 12.2-fold and 4.7-fold activation of PLC2 compared to the wild-type enzyme, respectively. The quadruple phosphomimetic mutant displayed an about 13-fold enhanced activity compared to wild-type PLC2 (Figure 3-8). At least three of the four common phosphorylation sites had to be phosphorylated to exert an activating effect on the lipase. This is consistent with the observation that chicken DT40 B cells expressing PLC2 mutants in which single tyrosine residues were substituted by phenylalanine exhibited a diminished Ca2+ mobilization whereas the expression of the quadruple YF mutant almost completely abolished Ca2+ signalling [283]. The results obtained indicate that mimicking tyrosine phosphorylation by substitution of the appropriate tyrosine residues with glutamic acid is a valuable tool to study the effect of tyrosine phosphorylation on the activity of at least Phospholipases C. It is worth mentioning that the chemical structure of phosphorylated tyrosine differs strongly from that of glutamate. Both of them are negatively charged, but glutamate is singly charged 117 Discussion whereas phosphorylated tyrosine is, nominally, doubly charged at physiological pH [44]. Furthermore, glutamate is much shorter than tyrosine and it totally lacks the aromatic ring.

However, in the case of PLC2 it seems that the negative charge of glutamate alone is sufficient to mimic the phosphorylation of tyrosine residues, since the impact of these substitutions led to changes in the activity of PLC2 that are in line with similar results obtained from other experimental approaches, as described above. To fully exclude the possibility that the absence of the aromatic ring of tyrosine itself has an influence on the activity of the enzyme it would be adequate to substitute the appropriate tyrosine residues for alanine.

A phosphoproteomic approach targeted on the survey of tyrosine kinase activity in lung cancer identified a novel phosphorylation site in PLC2, Y733, which is located at the

C-terminus of the SH2C domain and thus lies immediately upstream of the two established phosphorylation sites Y753 and Y759 [216]. Most interestingly, the SH2C domain per se and in particular an octapeptide, the PCI peptide (726YRKMRLRY733), including Y733, were shown to be involved in the autoinhibitory regulation of PLC2 [59, 87, 91].

Furthermore, phosphorylation induced activation of PLC1 was shown to be coupled to the release of autoinhibition of the enzyme [77]. Based on this knowledge it is most likely that phosphorylation of tyrosine residue 733 is a major contributive component in a PLC2 activating mechanism that potentially leads to tumorigenesis. As previously observed for Y753 and Y759, mimicking phosphorylation of Y733 alone had no impact on the activity of PLC2. However, mimicking phosphorylation of Y733, Y753 and Y759 in combination elevated PLC activity about 11-fold which is comparable to the measured activities of the Y753E/Y759E/Y1197E triple and quadruple phosphomimetic mutants PLC2 and Y753E/Y759E/Y1197E/Y1217E PLC2 (Figure 3-8). Despite the obvious difference in the location of Y1197 and Y733 in respect to Y753 and Y759, they seem to exert the same effect on the activity of PLC2. There are at least two possible explanations for this surveillance, (i) the three dimensional structure of PLC2 also brings Y1197 and Y1217 in direct spatial proximity to Y753 and Y759, or (ii) distinct mechanisms lead to the activation of PLC2 depending on the pattern of phosphorylated tyrosines. Intriguingly, mimicking phosphorylation of Y733 in the context of the two most actively phosphomimetic mutants Y753E/Y759E/Y1197E Y753E/Y759E/Y1197E/Y1217E PLC2 and PLC2 , led to a marked activation of PLC activity, 81-fold and 85.9- fold compared to PLC2, respectively (Figure 3-8). Thus,

118 Discussion

phosphorylation of Y733 seems to execute an exceeding effect on the activation of PLC2, which may play an important role in the pathogenesis of lung cancer.

All established phosphomimetic mutants are additionally activated by constitutively active Rac2G12V (Figure 3-9), suggesting that tyrosine phosphorylation-dependent activation and the Rac2-mediated activation of PLC2 are independent processes that both, separately as well as in combination, fulfil the requirements to activate PLC2. In line with the described results, mutations in the binding interface of Rac2 and PLC2 did not affect the phosphorylation induced activation of PLC2 downstream of the EGF receptor [28].

Similarly, abolishing the activation of PLC2 by Rac had no effect on the stimulation of the lipase by G subunits [117]. Strikingly, all phosphomimetic mutants that displayed an elevated activity compared to the wild-type enzyme also showed an increased basal activity in cells expressing wild-type Rac2 (Figure 3-9). By culturing the transfected COS-7 cells in cell culture medium containing serum, a small proportion of endogenously expressed Rac1 as well as over-expressed Rac2 exists in its activated GTP bound form. However, this small population of activated Rac2 is not sufficient to activate wild-type

PLC2 but seems to be adequate to induce a further activation of the constitutively active phosphomimetic mutants. This leads to the assumption that those mutants are more susceptible to Rac2-mediated activation. It seems that the intrinsic inhibition of PLC2 by autoinhibitory elements is a general barrier that has to be overcome to fully activate the lipase. This appears to be also the case for the activation of PLC2 by upstream regulators like Rho GTPases or receptor and non-receptor tyrosine kinases. If the intrinsic barrier, or in other words, the autoinhibitory regulation is weakened through certain events like phosphorylation of regulatory tyrosine residues or deletion of SH domains, it is easier for a second regulator, e.g. Rac2, to exert its activating effect on PLC2.

The apparently striking effect of the phosphorylation of tyrosine residue 733 on the activity of PLC2 and the potentially pathophysiological effects of this constitutive activation of the lipase regarding tumorigenesis incited the search for the kinase phosphorylating PLC2 on Y733. Using the PPSP algorithm, several kinases were suggested to be able to phosphorylate this specific tyrosine residue in PLC2 [299]. However, ALK seemed to be the most promising hit for two distinct reasons: (i) NPM-ALK, an ALK fusion protein, was shown to directly bind to SH2 domains of PLC, in which Y733 is located, resulting in

119 Discussion phosphorylation and subsequent activation of the lipase [9] and (ii) furthermore, a novel ALK fusion protein, EML4-ALK, was found in the same phosphoproteomic approach that also discovered Y733 as a novel phosphorylation site in PLC2 [216]. ALK displays the classical structural features of a receptor tyrosine kinase and belongs to the insulin receptor superfamily [16, 108]. NPM-ALK is an oncogenic constitutively active tyrosine kinase associated with ALCL and is created via the recurrent chromosomal translocation t(2;5)(p23;q35), which fuses the C-terminal half of ALK to the N-terminal half of nucleophosmin (NPM) [12, 63, 124, 164]. A further constitutively active ALK fusion protein, EML4-ALK, was shown to be associated with NSCLC [216, 239]. Full length- ALK, as well as constitutively active ALK mutants, ALKF1175L and ALKR1275Q, found to be mutated in sporadic and familial cases of neuroblastoma [7, 37, 145], activated PLC1 and

PLC2 in transfected COS-7 cells (Figure 3-11). The distinct stimulating impact of wild- type ALK, which is not constitutively active, on PLC activity is most likely due to the overexpression of the kinase. Constitutive phosphorylation and activation of ALK was also observed in human NBL (naval biosciences laboratory) cell lines that displayed a high expression of ALK, leading to mutation-independent activation of the kinase [198].

The proof of an interaction between ALK and the PLC isozymes was assessed by using a coimmunoprecipitation approach (Figure 3-12). PLC1 as well as PLC2 coprecipitated with ALK, indicating that the observed activation of both PLC isozymes is based on an association of the lipases with the kinase. More direct approaches to characterize the interaction of two proteins would be surface plasmon resonance (SPR) spectroscopy or fluorescence resonance energy transfer (FRET) analysis. However, there is a need of recombinant expressed and purified proteins for both of the two methods. So far, the production of recombinant soluble ALK protein failed using baculovirus infected Sf9 insect cells (data not shown).

As shown in Figure 3-10, ALK also activated the quintuple phosphomimetic mutant, in which all 5 known tyrosine phosphorylation sites were replaced by glutamic acid. This observation refuted the initial hypothesis that ALK is a potential kinase phosphorylating

PLC2 on tyrosine residue 733. The same results were obtained for all other phosphomimetic mutants as well (data not shown). Therefore, Y733, Y753, Y759, Y1197, and Y1217 seem to be no or not the only phosphorylation sites in PLC2 that are targeted by ALK. The identification of the ALK-specific phosphorylation sites could be conducted 120 Discussion

by a mass spectromic approach in the future. PLC2 was only marginal stimulated by ALK and by the constitutively active ALK mutants (Figure 3-10 and Figure 3-11). Thus ALK seems to be no interaction partner of PLC2. Furthermore, the expression of PLC2 is restricted to cells of the haematopoietic system [30, 120, 192, 237], whereas expression of full-length ALK is absent in cell lines of the haematopoietic system [52]. This would also mean, that the observed association of ALK and PLC2, which is also only expressed in haematopoietic cells [253], has probably no physiological impact.

Most interestingly, the activation of PLC isozymes by NPM-ALK appears to be isozyme specific. PLC1 is clearly activated by NPM-ALK whereas PLC2 did not show any change in PLC activity by coexpression with NPM-ALK (Figure 3-10). NPM-ALK is associated with ALCL which is a type of non-Hodgkin lymphoma involving aberrant T cells. The predominantly expressed PLC isoform in these cells is PLC1 [253], emphasizing the isozyme selective activation by NPM-ALK. Furthermore, PLC1 was identified to associate with NPM-ALK in Karpas299 cells by coimmunoprecipitation and subsequently tandem mass spectrometry (MS/MS) [46]. In this work, the association of NPM-ALK and

PLC1 was demonstrated by coimmunoprecipitation experiments, whereas no or only a very weak interaction was detectable for PLC2 (Figure 3-12). The results further confirm the observed isozyme specific activation of PLC isozymes by NPM-ALK in transfected COS-7 cells.

EML4-ALK did not exert any effect on the activity of PLC1, PLC2 or PLC2 in transfected COS-7 cells (Figure 3-10). This may indicate that EML4-ALK does not regulate the PLC pathway. Pharmacologic inhibition of EML4-ALK by small molecule tyrosine kinase inhibitors led to the downregulation of the Ras/MEK/ERK and PI3K/Akt pathways. In contrast, no or only a minor downregulation of STAT3 phosphorylation was observed in the same experimental approach [137]. So far, there is no indication in literature about a link between the PLC pathway and EML4-ALK. However, the results obtained here could possibly imply that EML4-ALK exhibits its mitogenic activity through other downstream signalling pathways than the PLC pathway.

The ability of NPM-ALK to activate PLC1 was found to be due to the phosphorylation of the lipase. Tyrosine phosphorylation of PLC was exclusively detectable in BaF3

121 Discussion lymphocyte cells that stably overexpressed NPM-ALK. Furthermore, this phosphorylation of PLC led to enhanced inositol phosphate levels, corroborating the activation of PLC

[9]. In order to ensure that the observed activation of PLC1 and/or PLC2 by ALK and NPM-ALK in transfected COS-7 cells is also dependent on the phosphorylation of the lipases, kinase-dead mutants of both kinases were tested for their ability to activate PLC isozymes. In line with the results described in literature, kinase-dead mutants of ALK as well as of NPM-ALK failed to activate PLC1 and/or PLC2 (Figure 3-13). These results imply that the presumably phosphorylation-induced activation of PLC isozymes by ALK and/or NPM-ALK and at least the binding of NPM-ALK to PLC1 (Figure 3-14) is strictly dependent on the intrinsic kinase activity of ALK. It was also demonstrated for other NPM-ALK interacting molecules like pC130 CAS (p130 Crk-associate substrate), Grb2, JAK3, Shp2, or Gab2, that the interaction is dependent on the catalytic activity of NPM- ALK, as there was no association with NPM-ALK kinase dead mutants [4, 215, 269, 282, 304]

Phosphopeptide competition experiments identified tyrosine residue 664 in NPM-ALK as the binding site for PLC isozymes. Furthermore, Y664 was shown to be an autophosphorylation site in vivo that mediates the complex formation between NPM-ALK and PLC. Expression of NPM-ALK in Rat-1 fibroblasts and Ba/F3 lymphocyte cells was shown to lead to transformation of these cells, whereas NPM-ALKY664F lost the ability to transform Rat-1 fibroblasts as well as Ba/F3 cells. Thus, the activation of PLC seems to be essential for the transduction of a mitogenic signal by NPM-ALK [9]. In contrast to the described observations, the results of this work show that NPM-ALKY664F was able to activate PLC1 in transfected cells to a level that is comparable to wild-type NPM-ALK (Figure 3-15). Furthermore, a clear association of both enzymes in transfected COS-7 cells was demonstrated by coimmunoprecipitation (Figure 3-16). The same results were observed for full-length ALK in which the corresponding tyrosine was replaced by Y1604F phenylalanine. ALK behaved as wild-type ALK in regard to the activation of PLC1 and PLC2 (Figure 3-15). In line with the described results, Bischof et al. reported that a truncated NPM-ALK mutant lacking the C-terminal 154 amino acids, including Y664, still exhibited its transforming ability on Fr3T3 fibroblasts [17]. It might be possible that the observed effects are dependent on the used cell-types.

122 Discussion

4.2 Molecular mechanisms underlying the human hereditary disease PLAID

The results presented in this work show that deletion of residues encoded by exons 19 and

20-22 from PLC2 causes both a mild constitutive activation of inositol phosphate formation at 37 °C and a marked, several-hundred-fold enhancement of this activity in response to very small decreases in temperature, e.g. to 31 °C, upon expression of the mutant enzymes in intact cells. These functional responses are qualitatively and quantitatively similar for both types of mutants. While the former changes are consistent with those reported earlier for transfected cells, the latter are much higher than increases in 2+ intracellular calcium [Ca ]i observed when peripheral blood B cells of PLAID patients are exposed to cold. In the latter case, differences to normal B cells were only observed at temperatures below 29 °C, not exceeding a maximum of an about two-fold increase at 21 °C [190]. The results shown here suggest that the primary defect of the mutant enzymes, i.e. sensitivity to stimulation by cool temperatures, emerges at temperatures only a few degrees below the normal body temperature. These findings are in agreement with clinical findings on patients with a deletion of exons 20-22, where only very subtle cooling, such as the one caused by a tear rolling down the cheek of an affected family member at room temperature caused symptoms within one minute [73]. Of note, even at an ambient, comfortable temperature of 27 °C, the temperature in many cutaneous and subcutaneous regions is already somewhat lower than the esophageal core temperature of 37 °C [284], such that only minimal further temperature decreases by evaporative heat loss may suffice to cause maximal PLC2 activation. The considerable decrease of the stimulatory effect observed at temperatures below 31 °C may explain why PLAID patients are typically negative in the cold stimulation time test (CSTT), involving cooling with ice, and upon cold-water immersion [73, 190]. The time course of PLC2 deletion mutant activation and its reversible nature closely matches the clinical observation on PLAID patients of an immediate onset of inflammatory symptoms upon evaporative skin cooling and a somewhat slower resolution of the symptoms developing over the course of

30 minutes of rewarming at room temperature [73]. Mouse PLC2 carrying a gain-of- Ali5 function mutation, D993G, in the catalytic region, PLC2 , and causing spontaneous autoinflammation and autoimmunity shows a similar, albeit less dramatic response to cold temperatures [301]. This suggests that some of the lesions noticed in the affected animals,

123 Discussion such as inflammation in superficial skin layers, may in fact be caused by further activation Ali5 S707Y of PLC2 by cool temperatures. Intriguingly, PLC2 , found in APLAID patients, that do not suffer from cold urticaria [311], is further activated by cool temperatures, even though to a much lower extend than it was observed for PLC219 and PLC220-22. APLAID is a dominantly inherited autoinflammatory disease with immunodeficiency which is characterized by recurrent blistering skin lesions, bronchiolitis, arthralgia, ocular inflammation, enterocolitis, absence of autoantibodies, and mild immunodeficiency [311]. In contrast to PLAID, unstimulated B cells from APLAID patients show an enhanced intracellular Ca2+ release, which may lead to B cell anergy, as discussed below. Hence, an effect of this could be a desensitization of B cells or even mast cells to a further response to S707Y e.g. cool temperatures. Due to the only moderate response of PLC2 to cool temperatures it is also possible that APLAID patients only suffer from a very mild form of cold urticaria which was possibly not assessed in the phenotypical analysis of the patients so far. Similarly, also phosphomimetic mutants of PLC2 showed a further activation by Ali5 cool temperatures, with an intensity that is comparable to those observed for PLC2 and S707Y PLC2 . These results suggest that a hyperphosphorylation of PLC2 can result in a sensitivity of the enzyme to cool temperatures.

Three mechanisms should be considered to explain the sensitivity of the PLC2 mutants to cold temperatures observed in this study [270]. First, other cold-sensitive molecules endogenously present in COS-7 cells might indirectly cause activation of the mutants, e.g. via a soluble mediator or via altered protein-protein-interaction patterns. Second, the mutants may be specifically enabled to sense temperature-mediated changes in the physical properties of the plasma membrane phospholipid bilayer containing the enzymes´substrate(s). The third possibility is that a cool temperature sensor is intrinsically present in the PLC2 mutants. We do not think, however, that the first possibility is a likely one. Thus, among the known cold-temperature-sensitive molecules, TRPA1, TRPM8, and TRPC5, the former two responded to cold temperatures within lower temperature ranges [41] and endogenous expression of the latter two was not evident in COS-7 cells [127, 161, 214, 264]. While we cannot at present formally exclude the second mechanism, we note that the phase transition temperature of DPPC (Dipalmitoylphosphatidylcholine) (Tm) in artificial phospholipid membrane vesicles is 41 °C [105] and thus outside the range of temperatures mediating activation of the PLC2 mutants. However, the behavior of native

124 Discussion plasma membranes may be different. Experiments in cell-free systems using soluble PLC substrate (e.g. [96]) will allow addressing this question in the future.

If the structural element mediating the response to cool temperatures resides on the enzyme itself, the simplest interpretation of the results would be that removal of the portions encoded by exons 19 or 20-22 causes a loss of an autoinhibition, since the two regions overlap with regions identified before [27, 59, 77, 79] or in this study (see section 3.1) to be autoinhibitory at 37 °C. Comparison of the functional consequences of deletion and point mutations within the SH tandem on constitutive activity at 37 °C vs. stimulation by cool temperatures, e.g. 31 °C, shows that the mutants can be grouped into three distinct categories and that the two functional consequences are not necessarily correlated with each other (Figure 3-28). While this would not formally exclude a loss of autoinhibition, acquisition of an autostimulatory mechanism sensitive to decreasing temperatures appears as an alternative explanation. At first glance, a structural element residing in the SH domain tandem of PLC2 would appear a likely candidate for triggering such a stimulation. However, the deletion experiments shown in Figure 3-26 revealed that none of the constituents of the entire SH tandem is specifically required for enzyme activation by cool temperatures and that even a deletion mutant lacking both of the repeats shared between the two SH2 domains (PLC2SH2NSH2C) (Figure 3-26, Figure 3-35A) is activated by incubation at 31 °C.

Thus, the possibility remains that temperature-sensitivity resides in the split PH domain and that this sensitivity is specifically kept in check by either the intact, native SH tandem, by its SH2C-SH3 portion, or by a covalent linkage between the two split PH domain halves. The fact that the cool-temperature-sensitivity is lost when non-homonymous split

PH domain halves are present in PLC2, when the tertiary structure of the PH domain is disrupted by the W899A mutation [60], or when two residues of its N-terminal half, Y509 and F510, are replaced by alanine residues, is consistent with this view. The latter residues have recently been mapped, by NMR titration analysis, to the interface of spPH and SH2C and suggested to be involved in SH2C-mediated autoinhibition [27]. The fact that removal of the two residues is also effective in the absence of almost the entire SH2C domain in

PLC219 is difficult to reconcile with an important role of a spPH-SH2C interaction in cool-temperature regulation of mutant PLC2. However, chemical shift perturbations observed in NMR experiments probing protein-protein-interactions may imply structural

125 Discussion reorientation of residues upon binding rather than direct involvement in the interaction surface [187]. This and the weak interaction between the two domains observed in [77] is consistent with additional functions of these two spPH domain residues. Interestingly, Ali14 Y509 of PLC1 corresponds to Y495 of PLC2, which is mutated to C495 in PLC2 and Ali14/Ali5 the double mutant PLC2 . The considerable stimulatory effect of the Ali14 mutation Ali5 on the stimulation of PLC2 by cool temperature (cf. Figure 3-27, left panel) suggests there may loss- and gain-of function mutations in this locus with regard to cool- temperature sensitivity of PLC2.

The thermodynamic changes of PLC219 and PLC220-22 induced by cool temperatures exceed those reported for other temperature-regulated proteins, such as thermoTRPs, where Q10 values between approximately 10 and > 100 have been reported [41]. However, higher Q10 values are not without precedence in the literature. For example, the step of heat damage to the development of the posterior crossvein of Drosophila that is most sensitive to increasing temperature showed a Q1 of 1.8, corresponding to a Q10 of about 360 [86, 172]. Clapham and Miller recently developed a thermodynamic framework for understanding temperature sensing by transient receptor potential (thermoTRP) channels [41]. It is based on the requirement, under the laws of thermodynamics, of unusually large

 conformational standard-state enthalpies, H 0 , in the opening of temperature-sensitive TRP channels. The authors suggest that thermoTRP channel gating is accompanied by large changes in molar heat capacity, CP , and postulate that hot- and cold-sensing TRPs

 operate by identical conformational changes. For the cold-activated TRP channels, H 0 is

-1 -1  between -40 kcal mol and -112 kcal mol . According to H 0  20ln Q10 , the enthalpies

-1 are even higher, approximately -176 kcal mol , for PLC219 and PLC220-22. The -1 -1 molar heat capacity of tempTRPs, CP , is in the order of 3-5 kcal mol K . Since high heat capacity increases accompany the transfer of nonpolar compounds from small molecules into water, these values have been taken by Clapham and Miller to suggest that on the order of 10 to 20 non-polar side chains are unburied from the thrust of each subunit of the tetrameric thermoTRPs upon activation. Figure 4-1 shows that CP is even higher, -1 -1 approximately 35 kcal mol K , for PLC219 and PLC220. If the thermodynamic framework developed for the thermoTRPs apply to the deletion mutants of PLC2, this would suggest exposure of an even higher number of hydrophobic residues to the aqueous solvent upon cool-temperature-mediated enzyme activation. 126 Discussion

Figure 4-1: Temperature dependence of conformational equilibrium constant. The effect of incubation temperature on the conformational equilibrium constant K is shown for PLC219 (▲) and

PLC220-22 (▼). The data was fit to equation (4) (section 2.6) by non-linear least squares curve fitting to estimate CP and T0. The two parameters were not significantly different from each other by clobal curve fitting using shared parameters for the curves of PLC219 and PLC220-22 (P = 0.1746). The region of the curve that has not been accessed experimentally is drawn by a dot-dashed line. The relationships between -1 -1 -1 -1 temperature and K proposed by [41] for thermoTRPs for CP = 3 kcal mol K and CP = 5 kcal mol K , using K0 = 0.01 and T0 = 25 °C, are shown in comparison. The dashed line marks the half-activarion level of the equilibrium. The positions of T0 are marked by dotted lines.

The marked sensitivity of the two PLAID PLC2 mutants to cool temperatures may explain the gain-of-function symptoms observed in all PLAID patients such as cold urticaria [73,

190]. They are likely to be caused by cold-temperature-mediated activation of PLC2 in cutaneous mast cells and monocytes/macrophages, where the enzyme plays important roles in FcR- and FcR-mediated inflammatory skin reactions [287]. Symptomatic allergic and autoimmune disease, occurring in 56 % and 26 % of PLAID patients and potentially precipitated by enhanced basal (37 °C) or cool-temperature-mediated activation of PLC2 in mast and other immune cells, may also fall into this category. The latter may include B cells and dendritic cells. Loss-of-function symptoms, such as antibody deficiency, recurrent sinopulmonary infection, and symptoms resembling certain forms of common variable immunodeficiency [222], detected in 75 %, 44 %, and 11 %, respectively, of PLAID patients, are more difficult to explain. In this context, it is interesting to note that 127 Discussion constant antigen receptor occupancy and signaling, including an increased basal 2+ 2+ basal concentration of intracellular Ca ([Ca ]i ) is required for the maintenance of B cell anergy, which is characterized by the persistence in the periphery of B cells unresponsive to immunogen [13, 75, 82, 314]. The latter change is thought to downmodulate the expression of surface IgM, but not IgD BCRs [314], most likely by activating ERK and 2+ NFAT pathways [18, 219], resulting a dampened [Ca ]i response to surface IgM ligation. 2+ A defect leading to impaired BCR-induced increases in [Ca ]i has been described for type Ia patients with common variable immunodeficiency (Freiburg classification) and shown to be associated with a reduction in IgD-IgM- CD27+ class-switched memory B cells, hypogammaglobulinemia, and autoimmune dysregulation [65], all of which are observed in PLAID patients. Defective Ca2+ signaling leading to reduced memory B cell subsets has also been observed in pediatric patients with CVID disorders, with B cells displaying a normal surface phenotype, but disrupted BCR-CD20 dissociation [265]. Importantly, in normal individuals there is a continuum of anergy or unresponsiveness to anti-IgM stimulation in the mature B-cell compartment [314]. In normal B cells, anergy is rapidly reversed after dissociation of self antigen, e.g. by using hapten competition, and these cells 2+ basal regain antigen responsiveness [75]. In PLAID B cells, [Ca ]i would be expected to be elevated, either chronically due to constitutively enhanced PLC2 mutant activity at 37 °C or intermittently during passage through cool body regions of the patient, causing reversible reductions of B cell responsiveness. Some of the trans-inhibitory mechanisms observed in anergic B cells, inhibiting the signaling of G-protein-coupled chemokine receptors [22], may be operative even in other immune cells, such as neutrophils, potentially contributing to the loss-of-immune-response phenotype observed in PLAID patients.

An interesting question raised by the dramatic functional consequence of the PLCG2 deletions observed in PLAID patients is whether the genomic regions encompassing exons 19 through 22 are subject to alternative processing of pre-messenger RNA [128, 278], leading to exclusion of residues encoded by exons 19 and/or 20-22 from the mature PLC2 protein. This appears as an important issue, since almost 90 % of human genes undergo alternative splicing with a minor isoform frequency of 15 % or more, with variations, intraindividually, between tissues and between individuals [278]. Pertinent to this issue, variations in position -15 from the exon 19 splice acceptor site have been detected in about one third of the examined samples and suggested to be potentially involved in altering the

128 Discussion efficiency of splicing at this site [276]. Very recently, alternative RNA splicing has been shown to be pervasive across immune system lineages [57]. Intriguingly, one of the two + changes observed for PLC2-encoding mRNA in mouse CD19 B cells is skipping of exons 20-22. The marked similarity between the genomic organization of the mouse Plcg2 and the human PLCG2 genes raises the possibility that this alteration may also exist in humans to convey cool-temperature-sensitivity to PLC2 in certain cell types and/or particular individuals. Four single nucleotide polymorphisms (SNPs) in the PLCG2 gene associated with myocardial infarction (rs4072683), stroke (rs4889428), left ventricular hypertrophy (rs1143689), and arterial diseases (rs13331678) can be found in the NCBI (National Center for Biotechnology Information) Gene data base. The SNPs were identified in the course of the Cardiovascular Health Study (CHS) [66] and the NHLBI Family Heart Study (FamHS) [84]. The SNPs associated with myocardial infarction, stroke and arterial diseases are located in the introns between exon 1 and 2, 9 and 10, and 20 and 21 of the PLCG2 gene, respectively. The impacts of these SNPs on splicing or other regulatory mechanisms which affect the transcription of the gene are unknown so far. The fourth SNP, associated with hypertrophy of the left ventricle, is located in exon 15 and represents a synonymous variant of the coding amino acid.

Cold environmental temperatures are associated with a number of human disease states, including, e.g., exercise-induced asthma and acute cardiovascular events [166, 263]. While the pathophysiology of both states is certainly complex and still incompletely understood, both are associated with cell types functionally dependent on PLC2-mediated signaling. Thus, substantial evidence suggests that exercise-induced asthma is effected by the release of mediators from mast cells and other inflammatory cells of the airways [33, 197, 244]. During quiet breathing of room air, the temperatures in the upper trachea and subsegmental bronchi were approximately 32.0 °C and 35.5 °C, respectively [167]. With increasing ventilation, these temperatures fell to approximately 29.2 °C and 32.9 °C, respectively, with normal room temperature and to 20.6 °C and 31.6 °C with cold air, respectively. Thus, under the latter conditions, even in subsegmental bronchi, temperatures are well within the range of those promoting maximal activation of the PLC2 deletion mutants studied here. Although TRPM8 has been implicated in mediating the response of asthmatic subjects to cold temperatures by induces mucin hypersecretion from bronchial epithelial cells, a much lower temperature of 18 °C has been used in these studies as a cold stimulus for extended time periods [151]. Since the 1920s, numerous studies have suggested that

129 Discussion acute coronary events are more frequent during the cold winter season [292], especially in conjunction with physical exercise, such as snow shoveling [112, 121, 184]. Although an increase in myocardial oxygen demand, due to a cold-induced increase in peripheral vascular resistance, may be causative in these situations, coronary vasoconstriction or the absence of normal coronary vasodilation during exercise could also be involved [121]. Kounis syndrome is the concurrence of acute coronary syndromes with conditions associated with mast cell activation [138]. While the syndrome is usually connected to hypersensitivity reactions, activation of mast cells by cold temperatures, either at pulmonary or extrapulmonary sites such as the skin and subcutaneous tissues may also play a role. In this context, it is important to note that the hyperactive PLC2 variant harboring the Ali5 mutation has been shown to cause platelet hyperreactivity and a prothrombotic phenotype in mice [56]. Furthermore, PLC2 has long been known to be the predominant PLC subtype, together with PLC1, in proliferating smooth muscle cells, suggesting that the enzyme may play a particular role in vascular lesions [90].

130 Summary

5 Summary

The generation of the two second messengers inositol 1,4,5-trisphosphate (IP3) and diacylglycerol (DAG) by inositol-phospholipid-specific phospholipases C (PLCs) from its substrate phosphatidylinositol 4,5-bisphosphate (PIP2) constitutes one of the main signaling responses to extracellular stimuli. The produced second messengers IP3 and DAG finally promote the release of Ca2+ from intracellular stores and the activation of protein kinase C, respectively. There are six subfamilies of mammalian PLCs, designated , , , , , and . Among these subfamilies, the PLC isozymes take a particular position because of their structural organization, which reveals a unique linker inserted between the two halves of the catalytic subdomains X and Y. This specific array (SA) contains a second, split pleckstrin homology (spPH) domain, consisting of two halves separated by two Src homology 2 (SH2) domains and one Src homology 3 (SH3) domain. The activity of the PLC isozymes is regulated by several inter- and intramolecular mechanisms, constituting a complex regulatory network. Deregulated PLC activity is associated with several severe immunological phenotypes in mice and in human, and with several types of cancer, which demonstrates the absolute necessity to understand the regulation of these important signalling molecules.

While the activation of PLC2 by Rac GTPases is fairly well understood and characterized on the molecular level, the autoinhibitory regulation and the phosphorylation-induced activation of the lipase still remain undefined. In this work, it is shown that PLC2 is autoinhibited by its specific array, and that the main elements involved in this autoinhibitory regulation are the C-terminal SH2 (SH2C) domain together with the SH3 domain. This was investigated by deleting the SH2SH2SH3 domain tandem of PLC2 resulting in a constitutively active enzyme, PLC2SH, in transfected COS-7 cells. The insertion of SH2C domain together with the SH3 domain inhibited the activity of

PLC2SH efficiently, whereas the insertion of the N-terminal SH2 (SH2N) domain and other combinations of SH domains only exhibited a minor inhibitory effect on the constitutively active deletion mutant.

131 Summary

Most interestingly, two regulatory phosphorylation sites, Y753 and Y759, are located in the identified autoinhibitory region, more precisely in the linker between the SH2C and the

SH3 domain. To investigate the impact of phosphorylation on the activity of PLC2, phosphomimetic mutants were established by replacing the appropriate tyrosine residues by glutamic acid. In the present thesis it was demonstrated that mimicking phosphorylation of PLC2 is an appropriate tool to study the effect of phosphorylation on the activity of the lipase. Furthermore, it was demonstrated that all four regulatory phosphorylation sites known in PLC2 are essential for the activation of the enzyme. Most interestingly, mimicking phosphorylation on Y733, which is a novel PLC2 phosphorylation site identified in non-small cell lung cancer (NSCLC) had the most striking effect on the activity of the lipase pointing to a role of PLC2 in the tumorigenesis of NSCLC and prompting the question to the kinase phosphorylating PLC2 on Y733. On the basis of an in silico analysis anaplastic lymphoma kinase (ALK) was predicted to phosphorylate Y733. However, the results of this study show that Y733 is no or not the only site targeted by ALK. Nonetheless, it was clearly demonstrated that ALK activates and associates with

PLC1 as well as with PLC2 in transfected COS-7 cells and that this activation is due to a phosphorylation of the lipases on a so far unknown tyrosine residue. Furthermore, it was shown that NPM-ALK, an ALK fusion protein, activates and binds to PLC1 but not to

PLC2 and that the activation as well as the association of both molecules depends on the kinase activity of NPM-ALK.

Quite recently, deletions of exon 19 and exon 20-22 in the PLCG2 gene were shown to cause a novel disease pattern, designated PLAID (phospholipase C-2-associated antibody deficiency and immune dysregulation) and characterized by immunodeficiency, autoimmunity and cold urticaria. In this work it was demonstrated that the two PLC2 deletion mutants, PLC219 and PLC220-22, which were shown to be associated with PLAID, are specifically activated by cool temperatures. This was investigated by incubation of transfected COS-7 cells at different temperatures and the subsequently measurement of inositol phosphate formation. The activation of the PLAID mutants by cool temperatures has been shown to be a reversible process that competes with the Rac2- mediated activation of the lipase. Furthermore, the mechanism underlying the described sensitivity of the mutants to cool temperatures has been shown to be dependent on the integrity of the spPH domain which is distinct from the mechanism by which autoinhibition of PLC2 is regulated. 132 References

6 References

[1] Abe, K. et al. 2011. A novel N-ethyl-N-nitrosourea-induced mutation in phospholipase Cγ2 causes inflammatory arthritis, metabolic defects, and male infertility in vitro in a murine model. Arthritis and rheumatism. 63, 5 (May. 2011), 1301–1311. [2] Alberts, B., Johnson, A., Walter, P., Lewis, J., Raff, M. and Roberts, K. 2007. Molecular Biology of the Cell. Taylor & Francis. [3] Allouche, M. 2007. ALK is a Novel Dependence Receptor: Potential Implications in Development and Cancer. Cell Cycle. 6, 13 (Jul. 2007), 1533–1538. [4] Ambrogio, C. et al. 2005. p130Cas mediates the transforming properties of the anaplastic lymphoma kinase. Blood. 106, 12 (Dec. 2005), 3907–3916. [5] Amin, H.M. and Lai, R. 2007. Pathobiology of ALK+ anaplastic large-cell lymphoma. Blood. 110, 7 (Oct. 2007), 2259–2267. [6] Arber, D.A., Sun, L.H. and Weiss, L.M. 1996. Detection of the t(2;5)(p23;q35) chromosomal translocation in large B-cell lymphomas other than anaplastic large cell lymphoma. Human pathology. 27, 6 (Jun. 1996), 590–594. [7] Azarova, A.M., Gautam, G. and George, R.E. 2011. Emerging importance of ALK in neuroblastoma. Seminars in Cancer Biology. 21, 4 (Oktober. 2011), 267–275. [8] Bae, S.S., Choi, J.H., Oh, Y.-S., Yun, S.U., Ryu, S.H. and Suh, P.-G. 2002. Regulation of phospholipase C-gamma1 by protein kinase A-dependent phosphorylation. Advances in enzyme regulation. 42, (2002), 195–211. [9] Bai, R.-Y., Dieter, P., Peschel, C., Morris, S.W. and Duyster, J. 1998. Nucleophosmin-Anaplastic Lymphoma Kinase of Large-Cell Anaplastic Lymphoma Is a Constitutively Active Tyrosine Kinase That Utilizes Phospholipase C-γ To Mediate Its Mitogenicity. Molecular and Cellular Biology. 18, 12 (Dec. 1998), 6951– 6961. [10] Bai, R.-Y., Ouyang, T., Miething, C., Morris, S.W., Peschel, C. and Duyster, J. 2000. Nucleophosmin–anaplastic lymphoma kinase associated with anaplastic large-cell lymphoma activates the phosphatidylinositol 3-kinase/Akt antiapoptotic signaling pathway. Blood. 96, 13 (Dec. 2000), 4319–4327. [11] Barreca, A., Lasorsa, E., Riera, L., Machiorlatti, R., Piva, R., Ponzoni, M., Kwee, I., Bertoni, F., Piccaluga, P.P., Pileri, S.A. and Inghirami, G. 2011. Anaplastic

133 References

lymphoma kinase in human cancer. Journal of Molecular Endocrinology. 47, 1 (Aug. 2011), R11–R23. [12] Le Beau, M.M., Bitter, M.A., Larson, R.A., Doane, L.A., Ellis, E.D., Franklin, W.A., Rubin, C.M., Kadin, M.E. and Vardiman, J.W. 1989. The t(2;5)(p23;q35): a recurring chromosomal abnormality in Ki-1-positive anaplastic large cell lymphoma. Leukemia. 3, 12 (Dec. 1989), 866–870. [13] Benschop, R.J., Aviszus, K., Zhang, X., Manser, T., Cambier, J.C. and Wysocki, L.J. 2001. Activation and anergy in bone marrow B cells of a novel immunoglobulin transgenic mouse that is both hapten specific and autoreactive. Immunity. 14, 1 (Jan. 2001), 33–43. [14] Berg, J.M., Tymoczko, J.L. and Stryer, L. 2007. Stryer Biochemie. Spektrum Akademischer Verlag. [15] Berman, D.E., Dall’Armi, C., Voronov, S.V., McIntire, L.B.J., Zhang, H., Moore, A.Z., Staniszewski, A., Arancio, O., Kim, T.-W. and Di Paolo, G. 2008. Oligomeric amyloid-beta peptide disrupts phosphatidylinositol-4,5-bisphosphate metabolism. Nature neuroscience. 11, 5 (May. 2008), 547–554. [16] Bilsland, J.G., Wheeldon, A., Mead, A., Znamenskiy, P., Almond, S., Waters, K.A., Thakur, M., Beaumont, V., Bonnert, T.P., Heavens, R., Whiting, P., McAllister, G. and Munoz-Sanjuan, I. 2007. Behavioral and Neurochemical Alterations in Mice Deficient in Anaplastic Lymphoma Kinase Suggest Therapeutic Potential for Psychiatric Indications. Neuropsychopharmacology. 33, 3 (online Mai. 2007), 685– 700. [17] Bischof, D., Pulford, K., Mason, D.Y. and Morris, S.W. 1997. Role of the nucleophosmin (NPM) portion of the non-Hodgkin’s lymphoma-associated NPM- anaplastic lymphoma kinase fusion protein in oncogenesis. Molecular and Cellular Biology. 17, 4 (Apr. 1997), 2312–2325. [18] Borde, M., Barrington, R.A., Heissmeyer, V., Carroll, M.C. and Rao, A. 2006. Transcriptional basis of lymphocyte tolerance. Immunological reviews. 210, (Apr. 2006), 105–119. [19] Borer, R.A., Lehner, C.F., Eppenberger, H.M. and Nigg, E.A. 1989. Major nucleolar proteins shuttle between nucleus and cytoplasm. Cell. 56, 3 (Feb. 1989), 379–390. [20] Boyer, J.L., Waldo, G.L. and Harden, T.K. 1992. Beta gamma-subunit activation of G-protein-regulated phospholipase C. Journal of Biological Chemistry. 267, 35 (Dec. 1992), 25451–25456.

134 References

[21] Bradford, M.M. 1976. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Analytical biochemistry. 72, (May. 1976), 248–254. [22] Brauweiler, A., Merrell, K., Gauld, S.B. and Cambier, J.C. 2007. Cutting Edge: Acute and chronic exposure of immature B cells to antigen leads to impaired homing and SHIP1-dependent reduction in stromal cell-derived factor-1 responsiveness. Journal of immunology (Baltimore, Md.: 1950). 178, 6 (Mar. 2007), 3353–3357. [23] Breslauer, K.J., Frank, R., Blöcker, H. and Marky, L.A. 1986. Predicting DNA duplex stability from the base sequence. Proceedings of the National Academy of Sciences of the United States of America. 83, 11 (Jun. 1986), 3746–3750. [24] Brouwer, S.D. et al. 2010. Meta-analysis of Neuroblastomas Reveals a Skewed ALK Mutation Spectrum in Tumors with MYCN Amplification. Clinical Cancer Research. 16, 17 (Sep. 2010), 4353–4362. [25] Brownlie, R.J. and Zamoyska, R. 2013. T cell receptor signalling networks: branched, diversified and bounded. Nature reviews. Immunology. 13, 4 (Apr. 2013), 257–269. [26] Buehler, A., Walliser, C., Haas, J., Kraus, J.M., Kestler, H., Milner, J.D. and Gierschik, P. 2013. Molecular mechanisms underlying the human hereditary disease PLAID, phosholipase C-gamma2-associated antibody deficiency and immune dysregulation. publication in preparation. (2013). [27] Bunney, T.D., Esposito, D., Mas-Droux, C., Lamber, E., Baxendale, R.W., Martins, M., Cole, A., Svergun, D., Driscoll, P.C. and Katan, M. 2012. Structural and functional integration of the PLCγ interaction domains critical for regulatory mechanisms and signaling deregulation. Structure (London, England: 1993). 20, 12 (Dec. 2012), 2062–2075. [28] Bunney, T.D., Opaleye, O., Roe, S.M., Vatter, P., Baxendale, R.W., Walliser, C., Everett, K.L., Josephs, M.B., Christow, C., Rodrigues-Lima, F., Gierschik, P., Pearl, L.H. and Katan, M. 2009. Structural insights into formation of an active signaling complex between Rac and phospholipase C gamma 2. Molecular cell. 34, 2 (Apr. 2009), 223–233. [29] Camps, M., Carozzi, A., Schnabel, P., Scheer, A., Parker, P.J. and Gierschik, P. 1992. Isozyme-selective stimulation of phospholipase C-β2 by G protein βγ-subunits. Nature. 360, 6405 (print Dezember. 1992), 684–686. [30] Camps, M., Hou, C., Sidiropoulos, D., Stock, J.B., Jakobs, K.H. and Gierschik, P. 1992. Stimulation of phospholipase C by guanine-nucleotide-binding protein beta

135 References

gamma subunits. European journal of biochemistry / FEBS. 206, 3 (Jun. 1992), 821– 831. [31] Camps, M., Hou, C.F., Jakobs, K.H. and Gierschik, P. 1990. Guanosine 5’-[gamma- thio]triphosphate-stimulated hydrolysis of phosphatidylinositol 4,5-bisphosphate in HL-60 granulocytes. Evidence that the guanine nucleotide acts by relieving phospholipase C from an inhibitory constraint. The Biochemical journal. 271, 3 (Nov. 1990), 743–748. [32] Carén, H., Abel, F., Kogner, P. and Martinsson, T. 2008. High incidence of DNA mutations and gene amplifications of the ALK gene in advanced sporadic neuroblastoma tumours. Biochemical Journal. 416, 2 (Dec. 2008), 153. [33] Carlsen, K.-H. and Carlsen, K.C.L. 2002. Exercise-induced asthma. Paediatric respiratory reviews. 3, 2 (Jun. 2002), 154–160. [34] Carozzi, A., Camps, M., Gierschik, P. and Parker, P.J. 1993. Activation of phosphatidylinositol lipid-specific phospholipase C-beta 3 by G-protein beta gamma subunits. FEBS letters. 315, 3 (Jan. 1993), 340–342. [35] Chattopadhyay, A., Vecchi, M., Ji, Q. s, Mernaugh, R. and Carpenter, G. 1999. The role of individual SH2 domains in mediating association of phospholipase C-gamma1 with the activated EGF receptor. The Journal of biological chemistry. 274, 37 (Sep. 1999), 26091–26097. [36] Chen, S., Spiegelberg, B.D., Lin, F., Dell, E.J. and Hamm, H.E. 2004. Interaction of Gbetagamma with RACK1 and other WD40 repeat proteins. Journal of molecular and cellular cardiology. 37, 2 (Aug. 2004), 399–406. [37] Chen, Y., Takita, J., Choi, Y.L., Kato, M., Ohira, M., Sanada, M., Wang, L., Soda, M., Kikuchi, A., Igarashi, T., Nakagawara, A., Hayashi, Y., Mano, H. and Ogawa, S. 2008. Oncogenic mutations of ALK kinase in neuroblastoma. Nature. 455, 7215 (print Oktober. 2008), 971–974. [38] Cheng, H.-F., Jiang, M.-J., Chen, C.-L., Liu, S.-M., Wong, L.-P., Lomasney, J.W. and King, K. 1995. Cloning and Identification of Amino Acid Residues of Human Phospholipase Cδ1 Essential for Catalysis. Journal of Biological Chemistry. 270, 10 (Mar. 1995), 5495–5505. [39] Cifuentes, M.E., Honkanen, L. and Rebecchi, M.J. 1993. Proteolytic fragments of phosphoinositide-specific phospholipase C-delta 1. Catalytic and membrane binding properties. The Journal of biological chemistry. 268, 16 (Jun. 1993), 11586–11593.

136 References

[40] Clapham, D.E. 2003. TRP channels as cellular sensors. Nature. 426, 6966 (Dec. 2003), 517–524. [41] Clapham, D.E. and Miller, C. 2011. A thermodynamic framework for understanding temperature sensing by transient receptor potential (TRP) channels. Proceedings of the National Academy of Sciences of the United States of America. 108, 49 (Dec. 2011), 19492–19497. [42] Colomba, A., Courilleau, D., Ramel, D., Billadeau, D.D., Espinos, E., Delsol, G., Payrastre, B. and Gaits-Iacovoni, F. 2007. Activation of Rac1 and the exchange factor Vav3 are involved in NPM-ALK signaling in anaplastic large cell lymphomas. Oncogene. 27, 19 (Nov. 2007), 2728–2736. [43] Colomba, A., Giuriato, S., Dejean, E., Thornber, K., Delsol, G., Tronchere, H., Meggetto, F., Payrastre, B. and Gaits-Iacovoni, F. 2011. Inhibition of Rac controls NPM-ALK-dependent lymphoma development and dissemination. Blood Cancer Journal. 1, 6 (Jun. 2011), e21. [44] Cooper, J.A., Sefton, B.M. and Hunter, T. 1983. Detection and quantification of phosphotyrosine in proteins. Methods in enzymology. 99, (1983), 387–402. [45] Crespo, P., Xu, N., Simonds, W.F. and Gutkind, J.S. 1994. Ras-dependent activation of MAP kinase pathway mediated by G-protein beta gamma subunits. Nature. 369, 6479 (Jun. 1994), 418–420. [46] Crockett, D.K., Lin, Z., Elenitoba-Johnson, K.S.J. and Lim, M.S. 2004. Identification of NPM-ALK interacting proteins by tandem mass spectrometry. Oncogene. 23, 15 (Apr. 2004), 2617–2629. [47] Cutler, R.E., Jr, Stephens, R.M., Saracino, M.R. and Morrison, D.K. 1998. Autoregulation of the Raf-1 serine/threonine kinase. Proceedings of the National Academy of Sciences of the United States of America. 95, 16 (Aug. 1998), 9214– 9219. [48] DeBell, K., Graham, L., Reischl, I., Serrano, C., Bonvini, E. and Rellahan, B. 2007. Intramolecular regulation of phospholipase C-gamma1 by its C-terminal Src homology 2 domain. Molecular and cellular biology. 27, 3 (Feb. 2007), 854–863. [49] DeBell, K.E., Stoica, B.A., Verí, M.C., Di Baldassarre, A., Miscia, S., Graham, L.J., Rellahan, B.L., Ishiai, M., Kurosaki, T. and Bonvini, E. 1999. Functional independence and interdependence of the Src homology domains of phospholipase C- gamma1 in B-cell receptor signal transduction. Molecular and cellular biology. 19, 11 (Nov. 1999), 7388–7398.

137 References

[50] Degoutin, J., Brunet-de Carvalho, N., Cifuentes-Diaz, C. and Vigny, M. 2009. ALK (Anaplastic Lymphoma Kinase) expression in DRG neurons and its involvement in neuron–Schwann cells interaction. European Journal of Neuroscience. 29, 2 (2009), 275–286. [51] Denis, C., Sauliere, A., Galandrin, S., Senard, J.-M. and Gales, C. 2012. Probing Heterotrimeric G Protein Activation: Applications to Biased Ligands. Current Pharmaceutical Design. 18, 2 (Jan. 2012), 128–144. [52] Dirks, W.G., Fähnrich, S., Lis, Y., Becker, E., MacLeod, R.A.F. and Drexler, H.G. 2002. Expression and functional analysis of the anaplastic lymphoma kinase (ALK) gene in tumor cell lines. International Journal of Cancer. 100, 1 (2002), 49–56. [53] Du, X.-L., Hu, H., Lin, D.-C., Xia, S.-H., Shen, X.-M., Zhang, Y., Luo, M.-L., Feng, Y.-B., Cai, Y., Xu, X., Han, Y.-L., Zhan, Q.-M. and Wang, M.-R. 2007. Proteomic profiling of proteins dysregulted in Chinese esophageal squamous cell carcinoma. Journal of molecular medicine (Berlin, Germany). 85, 8 (Aug. 2007), 863–875. [54] Duyster, J., Bai, R.-Y. and Morris, S.W. 2001. Translocations involving anaplastic lymphoma kinase (ALK). Oncogene. 40, 20 (Sep. 2001), 5623–5637. [55] Ellis, M.V., Carne, A. and Katan, M. 1993. Structural requirements of phosphatidylinositol-specific phospholipase C delta 1 for enzyme activity. European journal of biochemistry / FEBS. 213, 1 (Apr. 1993), 339–347. [56] Elvers, M., Pozgaj, R., Pleines, I., May, F., Kuijpers, M.J.E., Heemskerk, J.M.W., Yu, P. and Nieswandt, B. 2010. Platelet hyperreactivity and a prothrombotic phenotype in mice with a gain-of-function mutation in phospholipase Cgamma2. Journal of thrombosis and haemostasis: JTH. 8, 6 (Jun. 2010), 1353–1363. [57] Ergun, A., Doran, G., Costello, J.C., Paik, H.H., Collins, J.J., Mathis, D., Benoist, C. and ImmGen Consortium 2013. Differential splicing across immune system lineages. Proceedings of the National Academy of Sciences of the United States of America. 110, 35 (Aug. 2013), 14324–14329. [58] Essen, L.-O., Perisic, O., Cheung, R., Katan, M. and Williams, R.L. 1996. Crystal structure of a mammalian phosphoinositide-specific phospholipase Cδ. Nature. 380, 6575 (Apr. 1996), 595–602. [59] Everett, K.L., Buehler, A., Bunney, T.D., Margineanu, A., Baxendale, R.W., Vatter, P., Retlich, M., Walliser, C., Manning, H.B., Neil, M.A.A., Dunsby, C., French, P.M.W., Gierschik, P. and Katan, M. 2011. Membrane Environment Exerts an

138 References

Important Influence on Rac-Mediated Activation of Phospholipase C 2. Molecular and Cellular Biology. 31, 6 (Jan. 2011), 1240–1251. [60] Everett, K.L., Bunney, T.D., Yoon, Y., Rodrigues-Lima, F., Harris, R., Driscoll, P.C., Abe, K., Fuchs, H., de Angelis, M.H., Yu, P., Cho, W. and Katan, M. 2009. Characterization of phospholipase C gamma enzymes with gain-of-function mutations. The Journal of biological chemistry. 284, 34 (Aug. 2009), 23083–23093. [61] Falasca, M., Logan, S.K., Lehto, V.P., Baccante, G., Lemmon, M.A. and Schlessinger, J. 1998. Activation of phospholipase Cγ by PI 3-kinase-induced PH domain-mediated membrane targeting. The EMBO Journal. 17, 2 (Jan. 1998), 414– 422. [62] Fernald, A.W., Jones, G.A. and Carpenter, G. 1994. Limited proteolysis of phospholipase C-gamma 1 indicates stable association of X and Y domains with enhanced catalytic activity. The Biochemical journal. 302 ( Pt 2), (Sep. 1994), 503– 509. [63] Fischer, P., Nacheva, E., Mason, D.Y., Sherrington, P.D., Hoyle, C., Hayhoe, F.G. and Karpas, A. 1988. A Ki-1 (CD30)-positive human cell line (Karpas 299) established from a high-grade non-Hodgkin’s lymphoma, showing a 2;5 translocation and rearrangement of the T-cell receptor beta-chain gene. Blood. 72, 1 (Jul. 1988), 234–240. [64] Fluckiger, A.C., Li, Z., Kato, R.M., Wahl, M.I., Ochs, H.D., Longnecker, R., Kinet, J.P., Witte, O.N., Scharenberg, A.M. and Rawlings, D.J. 1998. Btk/Tec kinases regulate sustained increases in intracellular Ca2+ following B-cell receptor activation. The EMBO journal. 17, 7 (Apr. 1998), 1973–1985. [65] Foerster, C., Voelxen, N., Rakhmanov, M., Keller, B., Gutenberger, S., Goldacker, S., Thiel, J., Feske, S., Peter, H.-H. and Warnatz, K. 2010. B cell receptor-mediated calcium signaling is impaired in B lymphocytes of type Ia patients with common variable immunodeficiency. Journal of immunology (Baltimore, Md.: 1950). 184, 12 (Jun. 2010), 7305–7313. [66] Fried, L.P., Borhani, N.O., Enright, P., Furberg, C.D., Gardin, J.M., Kronmal, R.A., Kuller, L.H., Manolio, T.A., Mittelmark, M.B. and Newman, A. 1991. The Cardiovascular Health Study: design and rationale. Annals of epidemiology. 1, 3 (Feb. 1991), 263–276. [67] Fu, C., Turck, C.W., Kurosaki, T. and Chan, A.C. 1998. BLNK: a central linker protein in B cell activation. Immunity. 9, 1 (Jul. 1998), 93–103.

139 References

[68] Fujii, M., Yi, K.S., Kim, M.J., Ha, S.H., Ryu, S.H., Suh, P.-G. and Yagisawa, H. 2009. Phosphorylation of phospholipase C-delta 1 regulates its enzymatic activity. Journal of cellular biochemistry. 108, 3 (Oct. 2009), 638–650. [69] Fujimoto, J., Shiota, M., Iwahara, T., Seki, N., Satoh, H., Mori, S. and Yamamoto, T. 1996. Characterization of the transforming activity of p80, a hyperphosphorylated protein in a Ki-1 lymphoma cell line with chromosomal translocation t(2;5). Proceedings of the National Academy of Sciences. 93, 9 (Apr. 1996), 4181–4186. [70] Fukami, K., Inanobe, S., Kanemaru, K. and Nakamura, Y. 2010. Phospholipase C is a key enzyme regulating intracellular calcium and modulating the phosphoinositide balance. Progress in Lipid Research. 49, 4 (Oktober. 2010), 429–437. [71] Fukami, K., Nakao, K., Inoue, T., Kataoka, Y., Kurokawa, M., Fissore, R.A., Nakamura, K., Katsuki, M., Mikoshiba, K., Yoshida, N. and Takenawa, T. 2001. Requirement of Phospholipase Cδ4 for the Zona Pellucida-Induced Acrosome Reaction. Science. 292, 5518 (May. 2001), 920–923. [72] Fukami, K., Yoshida, M., Inoue, T., Kurokawa, M., Fissore, R.A., Yoshida, N., Mikoshiba, K. and Takenawa, T. 2003. Phospholipase Cδ4 is required for Ca2+ mobilization essential for acrosome reaction in sperm. The Journal of Cell Biology. 161, 1 (Apr. 2003), 79–88. [73] Gandhi, C., Healy, C., Wanderer, A.A. and Hoffman, H.M. 2009. Familial atypical cold urticaria: description of a new hereditary disease. The Journal of allergy and clinical immunology. 124, 6 (Dec. 2009), 1245–1250. [74] Garcia, P., Gupta, R., Shah, S., Morris, A.J., Rudge, S.A., Scarlata, S., Petrova, V., McLaughlin, S. and Rebecchi, M.J. 1995. The pleckstrin homology domain of phospholipase C-delta 1 binds with high affinity to phosphatidylinositol 4,5- bisphosphate in bilayer membranes. Biochemistry. 34, 49 (Dec. 1995), 16228–16234. [75] Gauld, S.B., Benschop, R.J., Merrell, K.T. and Cambier, J.C. 2005. Maintenance of B cell anergy requires constant antigen receptor occupancy and signaling. Nature immunology. 6, 11 (Nov. 2005), 1160–1167. [76] Greco, W.R., Unkelbach, H.D., Poch, G., Suhnel, J., Kundi, M. and Bodeker, W. 1992. Consensus on concepts and terminology for interaction assessment: The Saaresilkä agreement. Archives of Complex Environmental Studies. 4 (1992), 65–69. [77] Gresset, A., Hicks, S.N., Harden, T.K. and Sondek, J. 2010. Mechanism of phosphorylation-induced activation of phospholipase C-gamma isozymes. The Journal of biological chemistry. 285, 46 (Nov. 2010), 35836–35847.

140 References

[78] Griffin, C.A., Hawkins, A.L., Dvorak, C., Henkle, C., Ellingham, T. and Perlman, E.J. 1999. Recurrent Involvement of 2p23 in Inflammatory Myofibroblastic Tumors. Cancer Research. 59, 12 (Jun. 1999), 2776–2780. [79] Hajicek, N., Charpentier, T.H., Rush, J.R., Harden, T.K. and Sondek, J. 2013. Autoinhibition and Phosphorylation-Induced Activation of Phospholipase C-γ Isozymes. Biochemistry. (Jul. 2013). [80] Harwood, N.E. and Batista, F.D. 2010. Early events in B cell activation. Annual review of immunology. 28, (2010), 185–210. [81] Hashimoto, A., Takeda, K., Inaba, M., Sekimata, M., Kaisho, T., Ikehara, S., Homma, Y., Akira, S. and Kurosaki, T. 2000. Cutting Edge: Essential Role of Phospholipase C-γ2 in B Cell Development and Function. The Journal of Immunology. 165, 4 (Aug. 2000), 1738–1742. [82] Healy, J.I., Dolmetsch, R.E., Timmerman, L.A., Cyster, J.G., Thomas, M.L., Crabtree, G.R., Lewis, R.S. and Goodnow, C.C. 1997. Different nuclear signals are activated by the B cell receptor during positive versus negative signaling. Immunity. 6, 4 (Apr. 1997), 419–428. [83] Hicks, S.N., Jezyk, M.R., Gershburg, S., Seifert, J.P., Harden, T.K. and Sondek, J. 2008. General and versatile autoinhibition of PLC isozymes. Molecular cell. 31, 3 (Aug. 2008), 383–394. [84] Higgins, M., Province, M., Heiss, G., Eckfeldt, J., Ellison, R.C., Folsom, A.R., Rao, D.C., Sprafka, J.M. and Williams, R. 1996. NHLBI Family Heart Study: objectives and design. American journal of epidemiology. 143, 12 (Jun. 1996), 1219–1228. [85] Hille, B. 2001. Ion channels of excitable membranes. Sinauer. [86] Hille, B. and Milkman, R. 1966. A Quantitative Description of Some Temperature Effects on Drosophila. Biological Bulletin. 131, 2 (Oct. 1966), 346. [87] Homma, M.K., Yamasaki, M., Ohmi, S. and Homma, Y. 1997. Inhibition of phosphoinositide hydrolysis and cell growth of Swiss 3T3 cells by myristoylated phospholipase C inhibitor peptides. Journal of biochemistry. 122, 4 (Oct. 1997), 738– 742. [88] Homma, Y. 1993. Suppression of Membrane Phospholipase C Activity by Synthetic Autoinhibitor Peptides. Methods. 5, 3 (Dec. 1993), 229–232. [89] Homma, Y. and Emori, Y. 1995. A dual functional signal mediator showing RhoGAP and phospholipase C-delta stimulating activities. The EMBO journal. 14, 2 (Jan. 1995), 286–291.

141 References

[90] Homma, Y., Sakamoto, H., Tsunoda, M., Aoki, M., Takenawa, T. and Ooyama, T. 1993. Evidence for involvement of phospholipase C-gamma 2 in signal transduction of platelet-derived growth factor in vascular smooth-muscle cells. The Biochemical journal. 290 ( Pt 3), (Mar. 1993), 649–653. [91] Homma, Y. and Takenawa, T. 1992. Inhibitory effect of src homology (SH) 2/SH3 fragments of phospholipase C-gamma on the catalytic activity of phospholipase C isoforms. Identification of a novel phospholipase C inhibitor region. The Journal of biological chemistry. 267, 30 (Oct. 1992), 21844–21849. [92] Hopkins, A.L. and Groom, C.R. 2002. The druggable genome. Nature reviews. Drug discovery. 1, 9 (Sep. 2002), 727–730. [93] Horstman, D.A., Chattopadhyay, A. and Carpenter, G. 1999. The influence of deletion mutations on phospholipase C-gamma 1 activity. Archives of biochemistry and biophysics. 361, 1 (Jan. 1999), 149–155. [94] Horstman, D.A., DeStefano, K. and Carpenter, G. 1996. Enhanced phospholipase C- gamma1 activity produced by association of independently expressed X and Y domain polypeptides. Proceedings of the National Academy of Sciences of the United States of America. 93, 15 (Jul. 1996), 7518–7521. [95] Horton, R.M., Hunt, H.D., Ho, S.N., Pullen, J.K. and Pease, L.R. 1989. Engineering hybrid genes without the use of restriction enzymes: gene splicing by overlap extension. Gene. 77, 1 (Apr. 1989), 61–68. [96] Huang, W., Hicks, S.N., Sondek, J. and Zhang, Q. 2011. A fluorogenic, small molecule reporter for mammalian phospholipase C isozymes. ACS chemical biology. 6, 3 (Mar. 2011), 223–228. [97] Humphries, L.A., Dangelmaier, C., Sommer, K., Kipp, K., Kato, R.M., Griffith, N., Bakman, I., Turk, C.W., Daniel, J.L. and Rawlings, D.J. 2004. Tec kinases mediate sustained calcium influx via site-specific tyrosine phosphorylation of the phospholipase Cgamma Src homology 2-Src homology 3 linker. The Journal of biological chemistry. 279, 36 (Sep. 2004), 37651–37661. [98] Hurley, S.P., Clary, D.O., Copie, V. and Lefcort, F. 2006. Anaplastic lymphoma kinase is dynamically expressed on subsets of motor neurons and in the peripheral nervous system. The Journal of Comparative Neurology. 495, 2 (2006), 202–212. [99] Hwang, J.-I., Oh, Y.-S., Shin, K.-J., Kim, H., Ryu, S.H. and Suh, P.-G. 2005. Molecular cloning and characterization of a novel phospholipase C, PLC-η. Biochemical Journal. 389, 1 (Jul. 2005), 181.

142 References

[100] Illenberger, D., Schwald, F. and Gierschik, P. 1997. Characterization and purification from bovine neutrophils of a soluble guanine-nucleotide-binding protein that mediates isozyme-specific stimulation of phospholipase C beta2. European journal of biochemistry / FEBS. 246, 1 (May. 1997), 71–77. [101] Illenberger, D., Schwald, F., Pimmer, D., Binder, W., Maier, G., Dietrich, A. and Gierschik, P. 1998. Stimulation of phospholipase C-beta2 by the Rho GTPases Cdc42Hs and Rac1. The EMBO journal. 17, 21 (Nov. 1998), 6241–6249. [102] Illenberger, D., Walliser, C., Nurnberg, B., Diaz Lorente, M. and Gierschik, P. 2003. Specificity and structural requirements of phospholipase C-beta stimulation by Rho GTPases versus G protein beta gamma dimers. The Journal of biological chemistry. 278, 5 (Jan. 2003), 3006–3014. [103] Illenberger, D., Walliser, C., Strobel, J., Gutman, O., Niv, H., Gaidzik, V., Kloog, Y., Gierschik, P. and Henis, Y.I. 2003. Rac2 regulation of phospholipase C-beta 2 activity and mode of membrane interactions in intact cells. The Journal of biological chemistry. 278, 10 (Mar. 2003), 8645–8652. [104] Inoue, H., Nojima, H. and Okayama, H. 1990. High efficiency transformation of Escherichia coli with plasmids. Gene. 96, 1 (1990), 23–28. [105] Inoue, T., Fukushima, K. and Shimozawa, R. 1990. Surfactant partition between bulk water and DPPC vesicle membrane: solid-gel vs. liquid-crystalline membrane. Chemistry and physics of lipids. 52, 2 (Jan. 1990), 157–161. [106] Irvin, B.J., Williams, B.L., Nilson, A.E., Maynor, H.O. and Abraham, R.T. 2000. Pleiotropic contributions of phospholipase C-gamma1 (PLC-gamma1) to T-cell antigen receptor-mediated signaling: reconstitution studies of a PLC-gamma1- deficient Jurkat T-cell line. Molecular and cellular biology. 20, 24 (Dec. 2000), 9149–9161. [107] Ishiai, M., Sugawara, H., Kurosaki, M. and Kurosaki, T. 1999. Cutting edge: association of phospholipase C-gamma 2 Src homology 2 domains with BLNK is critical for B cell antigen receptor signaling. Journal of immunology (Baltimore, Md.: 1950). 163, 4 (Aug. 1999), 1746–1749. [108] Iwahara, T., Fujimoto, J., Wen, D., Cupples, R., Bucay, N., Arakawa, T., Mori, S., Ratzkin, B. and Yamamoto, T. 1997. Molecular characterization of ALK, a receptor tyrosine kinase expressed specifically in the nervous system. Oncogene. 14, 4 (Jan. 1997), 439–449.

143 References

[109] Jacoby, E., Bouhelal, R., Gerspacher, M. and Seuwen, K. 2006. The 7 TM G- protein-coupled receptor target family. ChemMedChem. 1, 8 (Aug. 2006), 761–782. [110] Jaffe, A.B. and Hall, A. 2005. Rho GTPases: biochemistry and biology. Annual review of cell and developmental biology. 21, (2005), 247–269. [111] Jamora, C., Yamanouye, N., Van Lint, J., Laudenslager, J., Vandenheede, J.R., Faulkner, D.J. and Malhotra, V. 1999. Gbetagamma-mediated regulation of Golgi organization is through the direct activation of protein kinase D. Cell. 98, 1 (Jul. 1999), 59–68. [112] Janardhanan, R., Henry, Z., Hur, D.J., Lin, C.M., Lopez, D., Reagan, P.M., Rudnick, S.R., Koshko, T.J. and Keeley, E.C. 2010. The snow-shoveler’s ST elevation myocardial infarction. The American journal of cardiology. 106, 4 (Aug. 2010), 596–600. [113] Janetopoulos, C. and Devreotes, P. 2006. Phosphoinositide signaling plays a key role in cytokinesis. The Journal of cell biology. 174, 4 (Aug. 2006), 485–490. [114] Janoueix-Lerosey, I. et al. 2008. Somatic and germline activating mutations of the ALK kinase receptor in neuroblastoma. Nature. 455, 7215 (print Oktober. 2008), 967–970. [115] Jazii, F.-R., Najafi, Z., Malekzadeh, R., Conrads, T.-P., Ziaee, A.-A., Abnet, C., Yazdznbod, M., Karkhane, A.-A. and Salekdeh, G.-H. 2006. Identification of squamous cell carcinoma associated proteins by proteomics and loss of beta tropomyosin expression in esophageal cancer. World journal of gastroenterology: WJG. 12, 44 (Nov. 2006), 7104–7112. [116] Jemal, A., Siegel, R., Ward, E., Hao, Y., Xu, J., Murray, T. and Thun, M.J. 2008. Cancer Statistics, 2008. CA: A Cancer Journal for Clinicians. 58, 2 (Jan. 2008), 71– 96. [117] Jezyk, M.R., Snyder, J.T., Gershberg, S., Worthylake, D.K., Harden, T.K. and Sondek, J. 2006. Crystal structure of Rac1 bound to its effector phospholipase C- beta2. Nature structural & molecular biology. 13, 12 (Dec. 2006), 1135–1140. [118] Ji, Q., Winnier, G.E., Niswender, K.D., Horstman, D., Wisdom, R., Magnuson, M.A. and Carpenter, G. 1997. Essential role of the tyrosine kinase substrate phospholipase C-γ1 in mammalian growth and development. Proceedings of the National Academy of Sciences. 94, 7 (Apr. 1997), 2999–3003. [119] Ji, Q.S., Chattopadhyay, A., Vecchi, M. and Carpenter, G. 1999. Physiological requirement for both SH2 domains for phospholipase C-gamma1 function and

144 References

interaction with platelet-derived growth factor receptors. Molecular and cellular biology. 19, 7 (Jul. 1999), 4961–4970. [120] Jiang, H., Kuang, Y., Wu, Y., Xie, W., Simon, M.I. and Wu, D. 1997. Roles of phospholipase C β2 in chemoattractant-elicited responses. Proceedings of the National Academy of Sciences. 94, 15 (Jul. 1997), 7971–7975. [121] Juneau, M., Johnstone, M., Dempsey, E. and Waters, D.D. 1989. Exercise-induced myocardial ischemia in a cold environment. Effect of antianginal medications. Circulation. 79, 5 (May. 1989), 1015–1020. [122] Kadamur, G. and Ross, E.M. 2013. Mammalian phospholipase C. Annual review of physiology. 75, (2013), 127–154. [123] Kadomatsu, K. and Muramatsu, T. 2004. Midkine and pleiotrophin in neural development and cancer. Cancer Letters. 204, 2 (Feb. 2004), 127–143. [124] Kaneko, Y., Frizzera, G., Edamura, S., Maseki, N., Sakurai, M., Komada, Y., Sakurai, M., Tanaka, H., Sasaki, M. and Suchi, T. 1989. A novel translocation, t(2;5)(p23;q35), in childhood phagocytic large T- cell lymphoma mimicking malignant histiocytosis. Blood. 73, 3 (Feb. 1989), 806–813. [125] Kassenbrock, C.K. and Anderson, S.M. 2004. Regulation of ubiquitin protein ligase activity in c-Cbl by phosphorylation-induced conformational change and constitutive activation by tyrosine to glutamate point mutations. The Journal of biological chemistry. 279, 27 (Jul. 2004), 28017–28027. [126] Katz, A., Wu, D. and Simon, M.I. 1992. Subunits beta gamma of heterotrimeric G protein activate beta 2 isoform of phospholipase C. Nature. 360, 6405 (Dec. 1992), 686–689. [127] Kawasaki, B.T., Liao, Y. and Birnbaumer, L. 2006. Role of Src in C3 transient receptor potential channel function and evidence for a heterogeneous makeup of receptor- and store-operated Ca2+ entry channels. Proceedings of the National Academy of Sciences of the United States of America. 103, 2 (Jan. 2006), 335–340. [128] Kelemen, O., Convertini, P., Zhang, Z., Wen, Y., Shen, M., Falaleeva, M. and Stamm, S. 2013. Function of alternative splicing. Gene. 514, 1 (Feb. 2013), 1–30. [129] Kelley, G.G., Reks, S.E., Ondrako, J.M. and Smrcka, A.V. 2001. Phospholipase Cϵ: a novel Ras effector. The EMBO Journal. 20, 4 (Feb. 2001), 743–754. [130] Kerchner, K.R., Clay, R.L., McCleery, G., Watson, N., McIntire, W.E., Myung, C.- S. and Garrison, J.C. 2004. Differential sensitivity of phosphatidylinositol 3-kinase

145 References

p110gamma to isoforms of G protein betagamma dimers. The Journal of biological chemistry. 279, 43 (Oct. 2004), 44554–44562. [131] Kim, D., Jun, K.S., Lee, S.B., Kang, N.-G., Min, D.S., Kim, Y.-H., Ryu, S.H., Suh, P.-G. and Shin, H.-S. 1997. Phospholipase C isozymes selectively couple to specific neurotransmitter receptors. Nature. 389, 6648 (Sep. 1997), 290–293. [132] Kim, H.K., Kim, J.W., Zilberstein, A., Margolis, B., Kim, J.G., Schlessinger, J. and Rhee, S.G. 1991. PDGF stimulation of inositol phospholipid hydrolysis requires PLC- gamma 1 phosphorylation on tyrosine residues 783 and 1254. Cell. 65, 3 (May. 1991), 435–441. [133] Kim, J.W., Sim, S.S., Kim, U.H., Nishibe, S., Wahl, M.I., Carpenter, G. and Rhee, S.G. 1990. Tyrosine residues in bovine phospholipase C-gamma phosphorylated by the epidermal growth factor receptor in vitro. The Journal of biological chemistry. 265, 7 (Mar. 1990), 3940–3943. [134] Kim, U.H., Kim, J.W. and Rhee, S.G. 1989. Phosphorylation of phospholipase C- gamma by cAMP-dependent protein kinase. The Journal of biological chemistry. 264, 34 (Dec. 1989), 20167–20170. [135] Kim, Y.J., Sekiya, F., Poulin, B., Bae, Y.S. and Rhee, S.G. 2004. Mechanism of B- cell receptor-induced phosphorylation and activation of phospholipase C-gamma2. Molecular and cellular biology. 24, 22 (Nov. 2004), 9986–9999. [136] Kimple, A.J., Bosch, D.E., Giguère, P.M. and Siderovski, D.P. 2011. Regulators of G-protein signaling and their Gα substrates: promises and challenges in their use as drug discovery targets. Pharmacological reviews. 63, 3 (Sep. 2011), 728–749. [137] Koivunen, J.P. et al. 2008. EML4-ALK Fusion Gene and Efficacy of an ALK Kinase Inhibitor in Lung Cancer. Clinical Cancer Research. 14, 13 (Jul. 2008), 4275– 4283. [138] Kounis, N.G., Mazarakis, A., Tsigkas, G., Giannopoulos, S. and Goudevenos, J. 2011. Kounis syndrome: a new twist on an old disease. Future cardiology. 7, 6 (Nov. 2011), 805–824. [139] Kraus, J.M. and Kestler, H.A. 2010. A highly efficient multi-core algorithm for clustering extremely large datasets. BMC bioinformatics. 11, (2010), 169. [140] Kurosaki, T. and Tsukada, S. 2000. BLNK: connecting Syk and Btk to calcium signals. Immunity. 12, 1 (Jan. 2000), 1–5.

146 References

[141] Kurtz, A., Schulte, A.M. and Wellstein, A. 1995. Pleiotrophin and midkine in normal development and tumor biology. Critical reviews in oncogenesis. 6, 2 (1995), 151–177. [142] Laemmli, U.K. 1970. Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature. 227, 5259 (Aug. 1970), 680–685. [143] Lamant, L., Pulford, K., Bischof, D., Morris, S.W., Mason, D.Y., Delsol, G. and Mariamé, B. 2000. Expression of the ALK Tyrosine Kinase Gene in Neuroblastoma. The American Journal of Pathology. 156, 5 (Mai. 2000), 1711–1721. [144] Landman, N., Jeong, S.Y., Shin, S.Y., Voronov, S.V., Serban, G., Kang, M.S., Park, M.K., Di Paolo, G., Chung, S. and Kim, T.-W. 2006. Presenilin mutations linked to familial Alzheimer’s disease cause an imbalance in phosphatidylinositol 4,5-bisphosphate metabolism. Proceedings of the National Academy of Sciences of the United States of America. 103, 51 (Dec. 2006), 19524–19529. [145] Lee, C.C., Jia, Y., Li, N., Sun, X., Ng, K., Ambing, E., Gao, M., Hua, S., Chen, C., Kim, S., Michellys, P., Lesley, S.A., Harris, J.L. and Spraggon, G. 2010. Crystal structure of the ALK (anaplastic lymphoma kinase) catalytic domain. Biochemical Journal. 430, 3 (Sep. 2010), 425–437. [146] Lee, S.B. and Rhee, S.G. 1996. Molecular Cloning, Splice Variants, Expression, and Purification of Phospholipase C-4. Journal of Biological Chemistry. 271, 1 (Jan. 1996), 25–31. [147] Lee, S.B., Shin, S.H., Hepler, J.R., Gilman, A.G. and Rhee, S.G. 1993. Activation of phospholipase C-beta 2 mutants by G protein alpha q and beta gamma subunits. The Journal of biological chemistry. 268, 34 (Dec. 1993), 25952–25957. [148] Lee, W.K., Kim, J.K., Seo, M.-S., Cha, J.-H., Lee, K.J., Rha, H.K., Min, D.S., Jo, Y.-H. and Lee, K.-H. 1999. Molecular Cloning and Expression Analysis of a Mouse Phospholipase C-δ1. Biochemical and Biophysical Research Communications. 261, 2 (Aug. 1999), 393–399. [149] Lemmon, M.A., Ferguson, K.M., O’Brien, R., Sigler, P.B. and Schlessinger, J. 1995. Specific and high-affinity binding of inositol phosphates to an isolated pleckstrin homology domain. Proceedings of the National Academy of Sciences of the United States of America. 92, 23 (Nov. 1995), 10472–10476. [150] Lemmon, M.A. and Schlessinger, J. 2010. Cell signaling by receptor-tyrosine kinases. Cell. 141, 7 (Jun. 2010), 1117–1134.

147 References

[151] Li, M., Li, Q., Yang, G., Kolosov, V.P., Perelman, J.M. and Zhou, X.D. 2011. Cold temperature induces mucin hypersecretion from normal human bronchial epithelial cells in vitro through a transient receptor potential melastatin 8 (TRPM8)-mediated mechanism. The Journal of allergy and clinical immunology. 128, 3 (Sep. 2011), 626–634.e1–5. [152] Lin, E., Li, L., Guan, Y., Soriano, R., Rivers, C.S., Mohan, S., Pandita, A., Tang, J. and Modrusan, Z. 2009. Exon Array Profiling Detects EML4-ALK Fusion in Breast, Colorectal, and Non–Small Cell Lung Cancers. Molecular Cancer Research. 7, 9 (Sep. 2009), 1466–1476. [153] Lin, F.-G., Cheng, H.-F., Lee, I.-F., Kao, H.-J., Loh, S.-H. and Lee, W.-H. 2001. Downregulation of Phospholipase C δ3 by cAMP and Calcium. Biochemical and Biophysical Research Communications. 286, 2 (Aug. 2001), 274–280. [154] Lin, M.C., Nicosia, S., Lad, P.M. and Rodbell, M. 1977. Effects of GTP on binding of (3H) glucagon to receptors in rat hepatic plasma membranes. The Journal of biological chemistry. 252, 8 (Apr. 1977), 2790–2792. [155] Liu, M. and Simon, M.I. 1996. Regulation by cAMP-dependent protein kinease of a G-protein-mediated phospholipase C. Nature. 382, 6586 (Jul. 1996), 83–87. [156] Logothetis, D.E., Kurachi, Y., Galper, J., Neer, E.J. and Clapham, D.E. 1987. The βγ subunits of GTP-binding proteins activate the muscarinic K+ channel in heart. Nature. 325, 6102 (Jan. 1987), 321–326. [157] Lopes, C.M.B., Zhang, H., Rohacs, T., Jin, T., Yang, J. and Logothetis, D.E. 2002. Alterations in conserved Kir channel-PIP2 interactions underlie channelopathies. Neuron. 34, 6 (Jun. 2002), 933–944. [158] Lopez, I., Mak, E.C., Ding, J., Hamm, H.E. and Lomasney, J.W. 2001. A novel bifunctional phospholipase c that is regulated by Galpha 12 and stimulates the Ras/mitogen-activated protein kinase pathway. The Journal of biological chemistry. 276, 4 (Jan. 2001), 2758–2765. [159] Lorén, C.E., Scully, A., Grabbe, C., Edeen, P.T., Thomas, J., McKeown, M., Hunter, T. and Palmer, R.H. 2001. Identification and characterization of DAlk: a novel Drosophila melanogaster RTK which drives ERK activation in vivo. Genes to Cells. 6, 6 (2001), 531–544. [160] Maguire, M.E., Van Arsdale, P.M. and Gilman, A.G. 1976. An agonist-specific effect of guanine nucleotides on binding to the beta adrenergic receptor. Molecular pharmacology. 12, 2 (Mar. 1976), 335–339.

148 References

[161] Mahieu, F., Owsianik, G., Verbert, L., Janssens, A., De Smedt, H., Nilius, B. and Voets, T. 2007. TRPM8-independent menthol-induced Ca2+ release from endoplasmic reticulum and Golgi. The Journal of biological chemistry. 282, 5 (Feb. 2007), 3325–3336. [162] Malcom W Greaves and Sabroe, R.A. 1998. Allergy and the skin: Urticaria. British Medical Journal. 516, 11 (Apr. 1998), 1147–1150. [163] Martin, T.F.. 2001. PI(4,5)P2 regulation of surface membrane traffic. Current Opinion in Cell Biology. 13, 4 (Aug. 2001), 493–499. [164] Mason, D.Y., Bastard, C., Rimokh, R., Dastugue, N., Huret, J.L., Kristoffersson, U., Magaud, J.P., Nezelof, C., Tilly, H. and Vannier, J.P. 1990. CD30-positive large cell lymphomas (’Ki-1 lymphoma’) are associated with a chromosomal translocation involving 5q35. British journal of haematology. 74, 2 (Feb. 1990), 161–168. [165] Mazot, P., Cazes, A., Boutterin, M.C., Figueiredo, A., Raynal, V., Combaret, V., Hallberg, B., Palmer, R.H., Delattre, O., Janoueix-Lerosey, I. and Vigny, M. 2011. The constitutive activity of the ALK mutated at positions F1174 or R1275 impairs receptor trafficking. Oncogene. 30, 17 (Apr. 2011), 2017–2025. [166] McFadden, E.R., Jr and Ingram, R.H., Jr 1983. Exercise-induced airway obstruction. Annual review of physiology. 45, (1983), 453–463. [167] McFadden, E.R., Jr, Pichurko, B.M., Bowman, H.F., Ingenito, E., Burns, S., Dowling, N. and Solway, J. 1985. Thermal mapping of the airways in humans. Journal of applied physiology (Bethesda, Md.: 1985). 58, 2 (Feb. 1985), 564–570. [168] McLaughlin, S., Wang, J., Gambhir, A. and Murray, D. 2002. PIP(2) and proteins: interactions, organization, and information flow. Annual review of biophysics and biomolecular structure. 31, (2002), 151–175. [169] Meister, M., Tomasovic, A., Banning, A. and Tikkanen, R. 2013. Mitogen- Activated Protein (MAP) Kinase Scaffolding Proteins: A Recount. International Journal of Molecular Sciences. 14, 3 (Mar. 2013), 4854–4884. [170] Meyer, S.L. 1975. Data analysis for scientists and engineers. Wiley. [171] Michell, R.H. and Allan, D. 1975. Inositol cyclic phosphate as a product of phosphatidylinositol breakdown by phospholipase C (Bacillus cereus). FEBS Letters. 53, 3 (Mai. 1975), 302–304. [172] Milkman, R. and Hille, B. 1966. Analysis of Some Temperature Effects on Drosophila pupae. Biological Bulletin. 131, 2 (Oct. 1966), 331.

149 References

[173] Milligan, G. and Kostenis, E. 2006. Heterotrimeric G-proteins: a short history. British journal of pharmacology. 147 Suppl 1, (Jan. 2006), S46–55. [174] Morris, S.W., Kirstein, M.N., Valentine, M.B., Dittmer, K.G., Shapiro, D.N., Saltman, D.L. and Look, A.T. 1994. Fusion of a kinase gene, ALK, to a nucleolar protein gene, NPM, in non-Hodgkin’s lymphoma. Science (New York, N.Y.). 263, 5151 (Mar. 1994), 1281–1284. [175] Morris, S.W., Naeve, C., Mathew, P., James, P.L., Kirstein, M.N., Cui, X. and Witte, D.P. 1997. ALK, the chromosome 2 gene locus altered by the t(2;5) in non- Hodgkin’s lymphoma, encodes a novel neural receptor tyrosine kinase that is highly related to leukocyte tyrosine kinase (LTK). Oncogene. 14, 18 (May. 1997), 2175– 2188. [176] Mosse, Y.P. et al. 2008. Identification of ALK as the Major Familial Neuroblastoma Predisposition Gene. Nature. 455, 7215 (Oct. 2008), 930–935. [177] Motegi, A., Fujimoto, J., Kotani, M., Sakuraba, H. and Yamamoto, T. 2004. ALK receptor tyrosine kinase promotes cell growth and neurite outgrowth. Journal of Cell Science. 117, 15 (Jul. 2004), 3319–3329. [178] Mourali, J., Bénard, A., Lourenço, F.C., Monnet, C., Greenland, C., Moog-Lutz, C., Racaud-Sultan, C., Gonzalez-Dunia, D., Vigny, M., Mehlen, P., Delsol, G. and Allouche, M. 2006. Anaplastic Lymphoma Kinase Is a Dependence Receptor Whose Proapoptotic Functions Are Activated by Caspase Cleavage. Molecular and Cellular Biology. 26, 16 (Aug. 2006), 6209–6222. [179] Mullis, K., Faloona, F., Scharf, S., Saiki, R., Horn, G. and Erlich, H. 1986. Specific enzymatic amplification of DNA in vitro: the polymerase chain reaction. Cold Spring Harbor symposia on quantitative biology. 51 Pt 1, (1986), 263–273. [180] Muramatsu, T. 2002. Midkine and Pleiotrophin: Two Related Proteins Involved in Development, Survival, Inflammation and Tumorigenesis. Journal of Biochemistry. 132, 3 (Sep. 2002), 359–371. [181] Nakahara, M., Shimozawa, M., Nakamura, Y., Irino, Y., Morita, M., Kudo, Y. and Fukami, K. 2005. A Novel Phospholipase C, PLCη2, Is a Neuron-specific Isozyme. Journal of Biological Chemistry. 280, 32 (Aug. 2005), 29128–29134. [182] Nakashima, S., Banno, Y., Watanabe, T., Nakamura, Y., Mizutani, T., Sakai, H., Zhao, Y., Sugimoto, Y. and Nozawa, Y. 1995. Deletion and site-directed mutagenesis of EF-hand domain of phospholipase C-delta 1: effects on its activity. Biochemical and biophysical research communications. 211, 2 (Jun. 1995), 365–369.

150 References

[183] Neves, S.R., Ram, P.T. and Iyengar, R. 2002. G protein pathways. Science (New York, N.Y.). 296, 5573 (May. 2002), 1636–1639. [184] Nichols, R.B., McIntyre, W.F., Chan, S., Scogstad-Stubbs, D., Hopman, W.M. and Baranchuk, A. 2012. Snow-shoveling and the risk of acute coronary syndromes. Clinical research in cardiology: official journal of the German Cardiac Society. 101, 1 (Jan. 2012), 11–15. [185] Nishibe, S., Wahl, M.I., Hernández-Sotomayor, S.M., Tonks, N.K., Rhee, S.G. and Carpenter, G. 1990. Increase of the catalytic activity of phospholipase C-gamma 1 by tyrosine phosphorylation. Science (New York, N.Y.). 250, 4985 (Nov. 1990), 1253– 1256. [186] Nishizuka, Y. 1984. The role of protein kinase C in cell surface signal transduction and tumour promotion. Nature. 308, 5961 (Apr. 1984), 693–698. [187] O’Connell, M.R., Gamsjaeger, R. and Mackay, J.P. 2009. The structural analysis of protein-protein interactions by NMR spectroscopy. Proteomics. 9, 23 (Dec. 2009), 5224–5232. [188] Offermanns, S. and Simon, M.I. 1995. G alpha 15 and G alpha 16 couple a wide variety of receptors to phospholipase C. The Journal of biological chemistry. 270, 25 (Jun. 1995), 15175–15180. [189] Okuwaki, M. 2008. The Structure and Functions of NPM1/Nucleophsmin/B23, a Multifunctional Nucleolar Acidic Protein. Journal of Biochemistry. 143, 4 (Apr. 2008), 441–448. [190] Ombrello, M.J. et al. 2012. Cold urticaria, immunodeficiency, and autoimmunity related to PLCG2 deletions. The New England journal of medicine. 366, 4 (Jan. 2012), 330–338. [191] Ozdener, F., Dangelmaier, C., Ashby, B., Kunapuli, S.P. and Daniel, J.L. 2002. Activation of phospholipase Cgamma2 by tyrosine phosphorylation. Molecular pharmacology. 62, 3 (Sep. 2002), 672–679. [192] Park, D., Jhon, D.Y., Lee, C.W., Lee, K.H. and Rhee, S.G. 1993. Activation of phospholipase C isozymes by G protein beta gamma subunits. The Journal of biological chemistry. 268, 7 (Mar. 1993), 4573–4576. [193] Park, D., Jhon, D.Y., Lee, C.W., Ryu, S.H. and Rhee, S.G. 1993. Removal of the carboxyl-terminal region of phospholipase C-beta 1 by calpain abolishes activation by G alpha q. The Journal of biological chemistry. 268, 5 (Feb. 1993), 3710–3714.

151 References

[194] Park, D.J., Min, H.K. and Rhee, S.G. 1992. Inhibition of CD3-linked phospholipase C by phorbol ester and by cAMP is associated with decreased phosphotyrosine and increased phosphoserine contents of PLC-gamma 1. The Journal of biological chemistry. 267, 3 (Jan. 1992), 1496–1501. [195] Park, D.J., Rho, H.W. and Rhee, S.G. 1991. CD3 stimulation causes phosphorylation of phospholipase C-gamma 1 on serine and tyrosine residues in a human T-cell line. Proceedings of the National Academy of Sciences of the United States of America. 88, 12 (Jun. 1991), 5453–5456. [196] Parker, P.J., Hemmings, B.A. and Gierschik, P. 1994. PH domains and phospholipases--a meaningful relationship? Trends in biochemical sciences. 19, 2 (Feb. 1994), 54–55. [197] Parsons, J.P., Hallstrand, T.S., Mastronarde, J.G., Kaminsky, D.A., Rundell, K.W., Hull, J.H., Storms, W.W., Weiler, J.M., Cheek, F.M., Wilson, K.C., Anderson, S.D. and American Thoracic Society Subcommittee on Exercise-induced Bronchoconstriction 2013. An official American Thoracic Society clinical practice guideline: exercise-induced bronchoconstriction. American journal of respiratory and critical care medicine. 187, 9 (May. 2013), 1016–1027. [198] Passoni, L. et al. 2009. Mutation-Independent Anaplastic Lymphoma Kinase Overexpression in Poor Prognosis Neuroblastoma Patients. Cancer Research. 69, 18 (Sep. 2009), 7338–7346. [199] Paterson, H.F., Savopoulos, J.W., Perisic, O., Cheung, R., Ellis, M.V., Williams, R.L. and Katan, M. 1995. Phospholipase C delta 1 requires a pleckstrin homology domain for interaction with the plasma membrane. The Biochemical journal. 312 ( Pt 3), (Dec. 1995), 661–666. [200] Pearlman, S.M., Serber, Z. and Ferrell, J.E., Jr 2011. A mechanism for the evolution of phosphorylation sites. Cell. 147, 4 (Nov. 2011), 934–946. [201] Pearson, J.D., Lee, J.K.H., Bacani, J.T.C., Lai, R. and Ingham, R.J. 2012. NPM- ALK: The Prototypic Member of a Family of Oncogenic Fusion Tyrosine Kinases. Journal of Signal Transduction. 2012, (2012), 1–14. [202] Piccinini, G., Bacchiocchi, R., Serresi, M., Vivani, C., Rossetti, S., Gennaretti, C., Carbonari, D. and Fazioli, F. 2002. A Ligand-inducible Epidermal Growth Factor Receptor/Anaplastic Lymphoma Kinase Chimera Promotes Mitogenesis and Transforming Properties in 3T3 Cells. Journal of Biological Chemistry. 277, 25 (Jun. 2002), 22231–22239.

152 References

[203] Piechulek, T., Rehlen, T., Walliser, C., Vatter, P., Moepps, B. and Gierschik, P. 2005. Isozyme-specific Stimulation of Phospholipase C-γ2 by Rac GTPases. Journal of Biological Chemistry. 280, 47 (Nov. 2005), 38923–38931. [204] Pierce, K.L., Premont, R.T. and Lefkowitz, R.J. 2002. Seven-transmembrane receptors. Nature reviews. Molecular cell biology. 3, 9 (Sep. 2002), 639–650. [205] Dal Porto, J.M., Gauld, S.B., Merrell, K.T., Mills, D., Pugh-Bernard, A.E. and Cambier, J. 2004. B cell antigen receptor signaling 101. Molecular immunology. 41, 6-7 (Jul. 2004), 599–613. [206] Poulin, B., Sekiya, F. and Rhee, S.G. 2000. Differential roles of the Src homology 2 domains of phospholipase C-gamma1 (PLC-gamma1) in platelet-derived growth factor-induced activation of PLC-gamma1 in intact cells. The Journal of biological chemistry. 275, 9 (Mar. 2000), 6411–6416. [207] Poulin, B., Sekiya, F. and Rhee, S.G. 2005. Intramolecular interaction between phosphorylated tyrosine-783 and the C-terminal Src homology 2 domain activates phospholipase C-gamma1. Proceedings of the National Academy of Sciences of the United States of America. 102, 12 (Mar. 2005), 4276–4281. [208] Powers, C., Aigner, A., Stoica, G.E., McDonnell, K. and Wellstein, A. 2002. Pleiotrophin Signaling through Anaplastic Lymphoma Kinase Is Rate-limiting for Glioblastoma Growth. Journal of Biological Chemistry. 277, 16 (Apr. 2002), 14153– 14158. [209] Pulford, K., Morris, S. w. and Turturro, F. 2004. Anaplastic lymphoma kinase proteins in growth control and cancer. Journal of Cellular Physiology. 199, 3 (2004), 330–358. [210] Rebecchi, M.J. and Pentyala, S.N. 2000. Structure, function, and control of phosphoinositide-specific phospholipase C. Physiological reviews. 80, 4 (Oct. 2000), 1291–1335. [211] Rhee, S.G. and Choi, K.D. 1992. Regulation of inositol phospholipid-specific phospholipase C isozymes. Journal of Biological Chemistry. 267, 18 (Jun. 1992), 12393–12396. [212] Rhee, S.G., Suh, P.G., Ryu, S.H. and Lee, S.Y. 1989. Studies of inositol phospholipid-specific phospholipase C. Science (New York, N.Y.). 244, 4904 (May. 1989), 546–550. [213] Van Rheenen, J., Song, X., van Roosmalen, W., Cammer, M., Chen, X., Desmarais, V., Yip, S.-C., Backer, J.M., Eddy, R.J. and Condeelis, J.S. 2007. EGF-induced PIP2

153 References

hydrolysis releases and activates cofilin locally in carcinoma cells. The Journal of cell biology. 179, 6 (Dec. 2007), 1247–1259. [214] Riccio, A., Medhurst, A.D., Mattei, C., Kelsell, R.E., Calver, A.R., Randall, A.D., Benham, C.D. and Pangalos, M.N. 2002. mRNA distribution analysis of human TRPC family in CNS and peripheral tissues. Brain research. Molecular brain research. 109, 1-2 (Dec. 2002), 95–104. [215] Riera, L., Lasorsa, E., Ambrogio, C., Surrenti, N., Voena, C. and Chiarle, R. 2010. Involvement of Grb2 adaptor protein in nucleophosmin-anaplastic lymphoma kinase (NPM-ALK)-mediated signaling and anaplastic large cell lymphoma growth. The Journal of biological chemistry. 285, 34 (Aug. 2010), 26441–26450. [216] Rikova, K. et al. 2007. Global Survey of Phosphotyrosine Signaling Identifies Oncogenic Kinases in Lung Cancer. Cell. 131, 6 (Dezember. 2007), 1190–1203. [217] Rodriguez, R., Matsuda, M., Perisic, O., Bravo, J., Paul, A., Jones, N.P., Light, Y., Swann, K., Williams, R.L. and Katan, M. 2001. Tyrosine residues in phospholipase Cgamma 2 essential for the enzyme function in B-cell signaling. The Journal of biological chemistry. 276, 51 (Dec. 2001), 47982–47992. [218] Rojas, A.M., Fuentes, G., Rausell, A. and Valencia, A. 2012. The Ras : evolutionary tree and role of conserved amino acids. The Journal of cell biology. 196, 2 (Jan. 2012), 189–201. [219] Rui, L., Vinuesa, C.G., Blasioli, J. and Goodnow, C.C. 2003. Resistance to CpG DNA-induced autoimmunity through tolerogenic B cell antigen receptor ERK signaling. Nature immunology. 4, 6 (Jun. 2003), 594–600. [220] Ryu, S.H., Cho, K.S., Lee, K.Y., Suh, P.G. and Rhee, S.G. 1987. Purification and characterization of two immunologically distinct phosphoinositide-specific phospholipases C from bovine brain. The Journal of biological chemistry. 262, 26 (Sep. 1987), 12511–12518. [221] Ryu, S.H., Suh, P.G., Cho, K.S., Lee, K.Y. and Rhee, S.G. 1987. Bovine brain cytosol contains three immunologically distinct forms of inositolphospholipid- specific phospholipase C. Proceedings of the National Academy of Sciences of the United States of America. 84, 19 (Oct. 1987), 6649–6653. [222] Salzer, U., Warnatz, K. and Peter, H.H. 2012. Common variable immunodeficiency - an update. Arthritis research & therapy. 14, 5 (Sep. 2012), 223.

154 References

[223] Sankaran, B., Osterhout, J., Wu, D. and Smrcka, A.V. 1998. Identification of a structural element in phospholipase C beta2 that interacts with G protein betagamma subunits. The Journal of biological chemistry. 273, 12 (Mar. 1998), 7148–7154. [224] Sato, M., Blumer, J.B., Simon, V. and Lanier, S.M. 2006. Accessory proteins for G proteins: partners in signaling. Annual review of pharmacology and toxicology. 46, (2006), 151–187. [225] Saunders, C.M., Larman, M.G., Parrington, J., Cox, L.J., Royse, J., Blayney, L.M., Swann, K. and Lai, F.A. 2002. PLCζ: a sperm-specific trigger of Ca2+ oscillations in eggs and embryo development. Development. 129, 15 (Aug. 2002), 3533–3544. [226] Schnabel, P. and Camps, M. 1998. Activation of a phospholipase Cbeta2 deletion mutant by limited proteolysis. The Biochemical journal. 330 ( Pt 1), (Feb. 1998), 461–468. [227] Schnabel, P., Camps, M., Carozzi, A., Parker, P.J. and Gierschik, P. 1993. Mutational analysis of phospholipase C-beta 2. Identification of regions required for membrane association and stimulation by guanine-nucleotide-binding protein beta gamma subunits. European journal of biochemistry / FEBS. 217, 3 (Nov. 1993), 1109–1115. [228] Schönherr, C., Ruuth, K., Eriksson, T., Yamazaki, Y., Ottmann, C., Combaret, V., Vigny, M., Kamaraj, S., Palmer, R.H. and Hallberg, B. 2011. The neuroblastoma ALK(I1250T) mutation is a kinase-dead RTK in vitro and in vivo. Translational oncology. 4, 4 (Aug. 2011), 258–265. [229] Sechi, A.S. and Wehland, J. 2000. The actin cytoskeleton and plasma membrane connection: PtdIns(4,5)P(2) influences cytoskeletal protein activity at the plasma membrane. Journal of Cell Science. 113, 21 (Nov. 2000), 3685–3695. [230] Shiota, M., Fujimoto, J., Takenaga, M., Satoh, H., Ichinohasama, R., Abe, M., Nakano, M., Yamamoto, T. and Mori, S. 1994. Diagnosis of t(2;5)(p23;q35)- associated Ki-1 lymphoma with immunohistochemistry. Blood. 84, 11 (Dec. 1994), 3648–3652. [231] Sidhu, R.S., Clough, R.R. and Bhullar, R.P. 2005. Regulation of phospholipase C- delta1 through direct interactions with the small GTPase Ral and calmodulin. The Journal of biological chemistry. 280, 23 (Jun. 2005), 21933–21941. [232] Sieh, M., Batzer, A., Schlessinger, J. and Weiss, A. 1994. GRB2 and phospholipase C-gamma 1 associate with a 36- to 38-kilodalton phosphotyrosine protein after T-cell receptor stimulation. Molecular and cellular biology. 14, 7 (Jul. 1994), 4435–4442.

155 References

[233] Sievers, F., Wilm, A., Dineen, D., Gibson, T.J., Karplus, K., Li, W., Lopez, R., McWilliam, H., Remmert, M., Söding, J., Thompson, J.D. and Higgins, D.G. 2011. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Molecular systems biology. 7, (2011), 539. [234] Slupianek, A., Nieborowska-Skorska, M., Hoser, G., Morrione, A., Majewski, M., Xue, L., Morris, S.W., Wasik, M.A. and Skorski, T. 2001. Role of Phosphatidylinositol 3-Kinase-Akt Pathway in Nucleophosmin/Anaplastic Lymphoma Kinase-mediated Lymphomagenesis. Cancer Research. 61, 5 (Mar. 2001), 2194–2199. [235] Smith, M.R., Liu, Y.L., Matthews, N.T., Rhee, S.G., Sung, W.K. and Kung, H.F. 1994. Phospholipase C-gamma 1 can induce DNA synthesis by a mechanism independent of its lipase activity. Proceedings of the National Academy of Sciences of the United States of America. 91, 14 (Jul. 1994), 6554–6558. [236] Smrcka, A.V., Hepler, J.R., Brown, K.O. and Sternweis, P.C. 1991. Regulation of polyphosphoinositide-specific phospholipase C activity by purified Gq. Science (New York, N.Y.). 251, 4995 (Feb. 1991), 804–807. [237] Smrcka, A.V. and Sternweis, P.C. 1993. Regulation of purified subtypes of phosphatidylinositol-specific phospholipase C beta by G protein alpha and beta gamma subunits. The Journal of biological chemistry. 268, 13 (May. 1993), 9667– 9674. [238] Snyder, J.T., Singer, A.U., Wing, M.R., Harden, T.K. and Sondek, J. 2003. The pleckstrin homology domain of phospholipase C-beta2 as an effector site for Rac. The Journal of biological chemistry. 278, 23 (Jun. 2003), 21099–21104. [239] Soda, M. et al. 2007. Identification of the transforming EML4–ALK fusion gene in non-small-cell lung cancer. Nature. 448, 7153 (Aug. 2007), 561–566. [240] Soda, M., Takada, S., Takeuchi, K., Choi, Y.L., Enomoto, M., Ueno, T., Haruta, H., Hamada, T., Yamashita, Y., Ishikawa, Y., Sugiyama, Y. and Mano, H. 2008. A mouse model for EML4-ALK-positive lung cancer. Proceedings of the National Academy of Sciences. 105, 50 (Dec. 2008), 19893–19897. [241] Song, C., Hu, C.-D., Masago, M., Kariya, K., Yamawaki-Kataoka, Y., Shibatohge, M., Wu, D., Satoh, T. and Kataoka, T. 2001. Regulation of a Novel Human Phospholipase C, PLCε, through Membrane Targeting by Ras. Journal of Biological Chemistry. 276, 4 (Jan. 2001), 2752–2757.

156 References

[242] Souttou, B., Carvalho, N.B.-D., Raulais, D. and Vigny, M. 2001. Activation of Anaplastic Lymphoma Kinase Receptor Tyrosine Kinase Induces Neuronal Differentiation through the Mitogen-activated Protein Kinase Pathway. Journal of Biological Chemistry. 276, 12 (Mar. 2001), 9526–9531. [243] Spierenburg, G.T., Oerlemans, F.T., van Laarhoven, J.P. and de Bruyn, C.H. 1984. Phototoxicity of N-2-hydroxyethylpiperazine-N’-2-ethanesulfonic acid-buffered culture media for human leukemic cell lines. Cancer research. 44, 5 (May. 1984), 2253–2254. [244] Spooner, C.H., Spooner, G.R. and Rowe, B.H. 2003. Mast-cell stabilising agents to prevent exercise-induced bronchoconstriction. The Cochrane database of systematic reviews. 4 (2003), CD002307. [245] Stark, G.R. and Darnell, J.E. 2012. The JAK-STAT Pathway at Twenty. Immunity. 36, 4 (Apr. 2012), 503–514. [246] Stewart, A.J., Mukherjee, J., Roberts, S.J., Lester, D. and Farquharson, C. 2005. Identification of a novel class of mammalian phosphoinositol-specific phospholipase C enzymes. International Journal of Molecular Medicine. 15, 1 (Jan. 2005), 117. [247] Stoica, B., DeBell, K.E., Graham, L., Rellahan, B.L., Alava, M.A., Laborda, J. and Bonvini, E. 1998. The amino-terminal Src homology 2 domain of phospholipase C gamma 1 is essential for TCR-induced tyrosine phosphorylation of phospholipase C gamma 1. Journal of immunology (Baltimore, Md.: 1950). 160, 3 (Feb. 1998), 1059– 1066. [248] Stoica, G.E., Kuo, A., Aigner, A., Sunitha, I., Souttou, B., Malerczyk, C., Caughey, D.J., Wen, D., Karavanov, A., Riegel, A.T. and Wellstein, A. 2001. Identification of Anaplastic Lymphoma Kinase as a Receptor for the Growth Factor Pleiotrophin. Journal of Biological Chemistry. 276, 20 (May. 2001), 16772–16779. [249] Stoica, G.E., Kuo, A., Powers, C., Bowden, E.T., Sale, E.B., Riegel, A.T. and Wellstein, A. 2002. Midkine Binds to Anaplastic Lymphoma Kinase (ALK) and Acts as a Growth Factor for Different Cell Types. Journal of Biological Chemistry. 277, 39 (Sep. 2002), 35990–35998. [250] Strathmann, M.P. and Simon, M.I. 1991. G alpha 12 and G alpha 13 subunits define a fourth class of G protein alpha subunits. Proceedings of the National Academy of Sciences of the United States of America. 88, 13 (Jul. 1991), 5582–5586.

157 References

[251] Streb, H., Irvine, R.F., Berridge, M.J. and Schulz, I. 1983. Release of Ca2+ from a nonmitochondrial intracellular store in pancreatic acinar cells by inositol-1,4,5- trisphosphate. Nature. 306, 5938 (Nov. 1983), 67–69. [252] Suh, P.G., Hwang, J.I., Ryu, S.H., Donowitz, M. and Kim, J.H. 2001. The roles of PDZ-containing proteins in PLC-beta-mediated signaling. Biochemical and biophysical research communications. 288, 1 (Oct. 2001), 1–7. [253] Suh, P.-G., Park, J.-I., Manzoli, L., Cocco, L., Peak, J.C., Katan, M., Fukami, K., Kataoka, T., Yun, S. and Ryu, S.H. 2008. Multiple roles of phosphoinositide-specific phospholipase C isozymes. BMB reports. 41, 6 (Jun. 2008), 415–434. [254] Suh, P.-G., Ryu, S.H., Moon, K.H., Suh, H.W. and Goo Rhee, S. 1988. Cloning and sequence of multiple forms of phospholipase C. Cell. 54, 2 (Jul. 1988), 161–169. [255] Suh, P.G., Ryu, S.H., Moon, K.H., Suh, H.W. and Rhee, S.G. 1988. Inositol phospholipid-specific phospholipase C: complete cDNA and protein sequences and to tyrosine kinase-related oncogene products. Proceedings of the National Academy of Sciences. 85, 15 (Aug. 1988), 5419–5423. [256] Sultzman, L., Ellis, C., Lin, L.L., Pawson, T. and Knopf, J. 1991. Platelet-derived growth factor increases the in vivo activity of phospholipase C-gamma 1 and phospholipase C-gamma 2. Molecular and cellular biology. 11, 4 (Apr. 1991), 2018– 2025. [257] Tadano, M. et al. 2005. Congenital Semilunar Valvulogenesis Defect in Mice Deficient in Phospholipase Cε. Molecular and Cellular Biology. 25, 6 (Mar. 2005), 2191–2199. [258] Takata, M. and Kurosaki, T. 1996. A role for Bruton’s tyrosine kinase in B cell antigen receptor-mediated activation of phospholipase C-gamma 2. The Journal of experimental medicine. 184, 1 (Jul. 1996), 31–40. [259] Takata, M., Sabe, H., Hata, A., Inazu, T., Homma, Y., Nukada, T., Yamamura, H. and Kurosaki, T. 1994. Tyrosine kinases Lyn and Syk regulate B cell receptor- coupled Ca2+ mobilization through distinct pathways. The EMBO journal. 13, 6 (Mar. 1994), 1341–1349. [260] Tang, W.J. and Gilman, A.G. 1991. Type-specific regulation of adenylyl cyclase by G protein beta gamma subunits. Science (New York, N.Y.). 254, 5037 (Dec. 1991), 1500–1503. [261] Tang, Y., Tang, J., Chen, Z., Trost, C., Flockerzi, V., Li, M., Ramesh, V. and Zhu, M.X. 2000. Association of Mammalian Trp4 and Phospholipase C Isozymes with a

158 References

PDZ Domain-containing Protein, NHERF. Journal of Biological Chemistry. 275, 48 (Dec. 2000), 37559–37564. [262] Taylor, S.J., Smith, J.A. and Exton, J.H. 1990. Purification from bovine liver membranes of a guanine nucleotide-dependent activator of phosphoinositide-specific phospholipase C. Immunologic identification as a novel G-protein alpha subunit. The Journal of biological chemistry. 265, 28 (Oct. 1990), 17150–17156. [263] Thompson, P.D. et al. 2007. Exercise and acute cardiovascular events placing the risks into perspective: a scientific statement from the American Heart Association Council on Nutrition, Physical Activity, and Metabolism and the Council on Clinical Cardiology. Circulation. 115, 17 (May. 2007), 2358–2368. [264] Varnai, P., Thyagarajan, B., Rohacs, T. and Balla, T. 2006. Rapidly inducible changes in phosphatidylinositol 4,5-bisphosphate levels influence multiple regulatory functions of the lipid in intact living cells. The Journal of cell biology. 175, 3 (Nov. 2006), 377–382. [265] Van de Ven, A.A.J.M., Compeer, E.B., Bloem, A.C., van de Corput, L., van Gijn, M., van Montfrans, J.M. and Boes, M. 2012. Defective calcium signaling and disrupted CD20-B-cell receptor dissociation in patients with common variable immunodeficiency disorders. The Journal of allergy and clinical immunology. 129, 3 (Mar. 2012), 755–761.e7. [266] Vernersson, E., Khoo, N.K.S., Henriksson, M.L., Roos, G., Palmer, R.H. and Hallberg, B. 2006. Characterization of the expression of the ALK receptor tyrosine kinase in mice. Patterns. 6, 5 (Jun. 2006), 448–461. [267] Vincze, T., Posfai, J. and Roberts, R.J. 2003. NEBcutter: A program to cleave DNA with restriction enzymes. Nucleic acids research. 31, 13 (Jul. 2003), 3688–3691. [268] Vines, C.M. 2012. Phospholipase C. Advances in experimental medicine and biology. 740, (2012), 235–254. [269] Voena, C., Conte, C., Ambrogio, C., Boeri Erba, E., Boccalatte, F., Mohammed, S., Jensen, O.N., Palestro, G., Inghirami, G. and Chiarle, R. 2007. The tyrosine phosphatase Shp2 interacts with NPM-ALK and regulates anaplastic lymphoma cell growth and migration. Cancer research. 67, 9 (May. 2007), 4278–4286. [270] Voets, T., Droogmans, G., Wissenbach, U., Janssens, A., Flockerzi, V. and Nilius, B. 2004. The principle of temperature-dependent gating in cold- and heat-sensitive TRP channels. Nature. 430, 7001 (Aug. 2004), 748–754.

159 References

[271] De Vries, L., Zheng, B., Fischer, T., Elenko, E. and Farquhar, M.G. 2000. The regulator of G protein signaling family. Annual review of pharmacology and toxicology. 40, (2000), 235–271. [272] Wahl, M.I., Nishibe, S., Kim, J.W., Kim, H., Rhee, S.G. and Carpenter, G. 1990. Identification of two epidermal growth factor-sensitive tyrosine phosphorylation sites of phospholipase C-gamma in intact HSC-1 cells. The Journal of biological chemistry. 265, 7 (Mar. 1990), 3944–3948. [273] Waldo, G.L., Boyer, J.L., Morris, A.J. and Harden, T.K. 1991. Purification of an AlF4- and G-protein beta gamma-subunit-regulated phospholipase C-activating protein. The Journal of biological chemistry. 266, 22 (Aug. 1991), 14217–14225. [274] Waldo, G.L., Ricks, T.K., Hicks, S.N., Cheever, M.L., Kawano, T., Tsuboi, K., Wang, X., Montell, C., Kozasa, T., Sondek, J. and Harden, T.K. 2010. Kinetic scaffolding mediated by a phospholipase C-beta and Gq signaling complex. Science (New York, N.Y.). 330, 6006 (Nov. 2010), 974–980. [275] Walliser, C., Retlich, M., Harris, R., Everett, K.L., Josephs, M.B., Vatter, P., Esposito, D., Driscoll, P.C., Katan, M., Gierschik, P. and Bunney, T.D. 2008. Rac Regulates Its Effector Phospholipase Cγ2 through Interaction with a Split Pleckstrin Homology Domain. Journal of Biological Chemistry. 283, 44 (Oct. 2008), 30351– 30362. [276] Wang, D., Boylin, E.C., Minegishi, Y., Wen, R., Smith, C.I., Ihle, J.N. and Conley, M.E. 2001. Variations in the human phospholipase Cgamma2 gene in patients with B-cell defects of unknown etiology. Immunogenetics. 53, 7 (Sep. 2001), 550–556. [277] Wang, D., Feng, J., Wen, R., Marine, J.-C., Sangster, M.Y., Parganas, E., Hoffmeyer, A., Jackson, C.W., Cleveland, J.L., Murray, P.J. and Ihle, J.N. 2000. Phospholipase Cγ2 Is Essential in the Functions of B Cell and Several Fc Receptors. Immunity. 13, 1 (Jul. 2000), 25–35. [278] Wang, E.T., Sandberg, R., Luo, S., Khrebtukova, I., Zhang, L., Mayr, C., Kingsmore, S.F., Schroth, G.P. and Burge, C.B. 2008. Alternative isoform regulation in human tissue transcriptomes. Nature. 456, 7221 (Nov. 2008), 470–476. [279] Wang, H., Oestreich, E.A., Maekawa, N., Bullard, T.A., Vikstrom, K.L., Dirksen, R.T., Kelley, G.G., Blaxall, B.C. and Smrcka, A.V. 2005. Phospholipase C ε Modulates β-Adrenergic Receptor– Dependent Cardiac Contraction and Inhibits Cardiac Hypertrophy. Circulation Research. 97, 12 (Dec. 2005), 1305–1313.

160 References

[280] Wang, T., Dowal, L., El-Maghrabi, M.R., Rebecchi, M. and Scarlata, S. 2000. The pleckstrin homology domain of phospholipase C-beta(2) links the binding of gbetagamma to activation of the catalytic core. The Journal of biological chemistry. 275, 11 (Mar. 2000), 7466–7469. [281] Wang, T., Pentyala, S., Rebecchi, M.J. and Scarlata, S. 1999. Differential association of the pleckstrin homology domains of phospholipases C-beta 1, C-beta 2, and C-delta 1 with lipid bilayers and the beta gamma subunits of heterotrimeric G proteins. Biochemistry. 38, 5 (Feb. 1999), 1517–1524. [282] Wang, Y.-W., Tu, P.-H., Lin, K.-T., Lin, S.-C., Ko, J.-Y. and Jou, Y.-S. 2011. Identification of oncogenic point mutations and hyperphosphorylation of anaplastic lymphoma kinase in lung cancer. Neoplasia (New York, N.Y.). 13, 8 (Aug. 2011), 704–715. [283] Watanabe, D., Hashimoto, S., Ishiai, M., Matsushita, M., Baba, Y., Kishimoto, T., Kurosaki, T. and Tsukada, S. 2001. Four tyrosine residues in phospholipase C-gamma 2, identified as Btk-dependent phosphorylation sites, are required for B cell antigen receptor-coupled calcium signaling. The Journal of biological chemistry. 276, 42 (Oct. 2001), 38595–38601. [284] Webb, P. 1992. Temperatures of skin, subcutaneous tissue, muscle and core in resting men in cold, comfortable and hot conditions. European journal of applied physiology and occupational physiology. 64, 5 (1992), 471–476. [285] Weiss, A., Koretzky, G., Schatzman, R.C. and Kadlecek, T. 1991. Functional activation of the T-cell antigen receptor induces tyrosine phosphorylation of phospholipase C-gamma 1. Proceedings of the National Academy of Sciences of the United States of America. 88, 13 (Jul. 1991), 5484–5488. [286] Wellstein, A. 2012. ALK receptor activation, ligands and therapeutic targeting in glioblastoma and in other cancers. Frontiers in Oncology. 2, (Dec. 2012). [287] Wen, R., Jou, S.-T., Chen, Y., Hoffmeyer, A. and Wang, D. 2002. Phospholipase C gamma 2 is essential for specific functions of Fc epsilon R and Fc gamma R. Journal of immunology (Baltimore, Md.: 1950). 169, 12 (Dec. 2002), 6743–6752. [288] Wen, W., Yan, J. and Zhang, M. 2006. Structural characterization of the split pleckstrin homology domain in phospholipase C-gamma1 and its interaction with TRPC3. The Journal of biological chemistry. 281, 17 (Apr. 2006), 12060–12068.

161 References

[289] Wilde, J.I. and Watson, S.P. 2001. Regulation of phospholipase C gamma isoforms in haematopoietic cells: why one, not the other? Cellular signalling. 13, 10 (Oct. 2001), 691–701. [290] Wing, M.R., Houston, D., Kelley, G.G., Der, C.J., Siderovski, D.P. and Harden, T.K. 2001. Activation of phospholipase C-epsilon by heterotrimeric G protein betagamma-subunits. The Journal of biological chemistry. 276, 51 (Dec. 2001), 48257–48261. [291] Wing, M.R., Snyder, J.T., Sondek, J. and Harden, T.K. 2003. Direct activation of phospholipase C-epsilon by Rho. The Journal of biological chemistry. 278, 42 (Oct. 2003), 41253–41258. [292] Wolff, L. and White, P.D. 1926. Acute Coronary Occlusion. The Boston Medical and Surgical Journal. 195, 1 (Jul. 1926), 13–25. [293] Wu, D., Jiang, H., Katz, A. and Simon, M.I. 1993. Identification of critical regions on phospholipase C-beta 1 required for activation by G-proteins. The Journal of biological chemistry. 268, 5 (Feb. 1993), 3704–3709. [294] Wu, D.Q., Lee, C.H., Rhee, S.G. and Simon, M.I. 1992. Activation of phospholipase C by the alpha subunits of the Gq and G11 proteins in transfected Cos- 7 cells. The Journal of biological chemistry. 267, 3 (Jan. 1992), 1811–1817. [295] Wu, F., Wang, P., Young, L.C., Lai, R. and Li, L. 2009. Proteome-Wide Identification of Novel Binding Partners to the Oncogenic Fusion Gene Protein, NPM-ALK, using Tandem Affinity Purification and Mass Spectrometry. The American Journal of Pathology. 174, 2 (Feb. 2009), 361–370. [296] Wu, Y., Spencer, S.D. and Lasky, L.A. 1998. Tyrosine phosphorylation regulates the SH3-mediated binding of the Wiskott-Aldrich syndrome protein to PSTPIP, a cytoskeletal-associated protein. The Journal of biological chemistry. 273, 10 (Mar. 1998), 5765–5770. [297] Xie, W., Samoriski, G.M., McLaughlin, J.P., Romoser, V.A., Smrcka, A., Hinkle, P.M., Bidlack, J.M., Gross, R.A., Jiang, H. and Wu, D. 1999. Genetic alteration of phospholipase C β3 expression modulates behavioral and cellular responses to μ opioids. Proceedings of the National Academy of Sciences. 96, 18 (Aug. 1999), 10385–10390. [298] Xu, A., Suh, P.G., Marmy-Conus, N., Pearson, R.B., Seok, O.Y., Cocco, L. and Gilmour, R.S. 2001. Phosphorylation of nuclear phospholipase C beta1 by

162 References

extracellular signal-regulated kinase mediates the mitogenic action of insulin-like growth factor I. Molecular and cellular biology. 21, 9 (May. 2001), 2981–2990. [299] Xue, Y., Li, A., Wang, L., Feng, H. and Yao, X. 2006. PPSP: prediction of PK- specific phosphorylation site with Bayesian decision theory. BMC bioinformatics. 7, (2006), 163. [300] Yagisawa, H., Sakuma, K., Paterson, H.F., Cheung, R., Allen, V., Hirata, H., Watanabe, Y., Hirata, M., Williams, R.L. and Katan, M. 1998. Replacements of Single Basic Amino Acids in the Pleckstrin Homology Domain of Phospholipase C- δ1 Alter the Ligand Binding, Phospholipase Activity, and Interaction with the Plasma Membrane. Journal of Biological Chemistry. 273, 1 (Jan. 1998), 417–424. [301] Yu, P. et al. 2005. Autoimmunity and inflammation due to a gain-of-function mutation in phospholipase C gamma 2 that specifically increases external Ca2+ entry. Immunity. 22, 4 (Apr. 2005), 451–465. [302] Yue, C., Dodge, K.L., Weber, G. and Sanborn, B.M. 1998. Phosphorylation of serine 1105 by protein kinase A inhibits phospholipase Cbeta3 stimulation by Galphaq. The Journal of biological chemistry. 273, 29 (Jul. 1998), 18023–18027. [303] Yue, C., Ku, C.Y., Liu, M., Simon, M.I. and Sanborn, B.M. 2000. Molecular mechanism of the inhibition of phospholipase C beta 3 by protein kinase C. The Journal of biological chemistry. 275, 39 (Sep. 2000), 30220–30225. [304] Zamo, A., Chiarle, R., Piva, R., Howes, J., Fan, Y., Chilosi, M., Levy, D.E. and Inghirami, G. 2002. Anaplastic lymphoma kinase (ALK) activates Stat3 and protects hematopoietic cells from cell death. Oncogene. 21, 7 (Feb. 2002), 1038–1047. [305] Zamponi, G.W., Bourinet, E., Nelson, D., Nargeot, J. and Snutch, T.P. 1997. Crosstalk between G proteins and protein kinase C mediated by the calcium channel alpha1 subunit. Nature. 385, 6615 (Jan. 1997), 442–446. [306] Zarrinpar, A., Park, S.-H. and Lim, W.A. 2003. Optimization of specificity in a cellular protein interaction network by negative selection. Nature. 426, 6967 (Dec. 2003), 676–680. [307] Van Zeijl, L., Ponsioen, B., Giepmans, B.N.G., Ariaens, A., Postma, F.R., Várnai, P., Balla, T., Divecha, N., Jalink, K. and Moolenaar, W.H. 2007. Regulation of connexin43 gap junctional communication by phosphatidylinositol 4,5-bisphosphate. The Journal of cell biology. 177, 5 (Jun. 2007), 881–891. [308] Zhang, S., Coso, O.A., Collins, R., Gutkind, J.S. and Simonds, W.F. 1996. A C- terminal mutant of the G protein beta subunit deficient in the activation of

163 References

phospholipase C-beta. The Journal of biological chemistry. 271, 33 (Aug. 1996), 20208–20212. [309] Zhang, W. and Neer, E.J. 2001. Reassembly of phospholipase C-beta2 from separated domains: analysis of basal and G protein-stimulated activities. The Journal of biological chemistry. 276, 4 (Jan. 2001), 2503–2508. [310] Zhang, W., Sloan-Lancaster, J., Kitchen, J., Trible, R.P. and Samelson, L.E. 1998. LAT: the ZAP-70 tyrosine kinase substrate that links T cell receptor to cellular activation. Cell. 92, 1 (Jan. 1998), 83–92. [311] Zhou, Q. et al. 2012. A hypermorphic missense mutation in PLCG2, encoding phospholipase Cγ2, causes a dominantly inherited autoinflammatory disease with immunodeficiency. American journal of human genetics. 91, 4 (Oct. 2012), 713–720. [312] Zhou, Y., Sondek, J. and Harden, T.K. 2008. Activation of Human Phospholipase C-η2 by Gβγ†. Biochemistry. 47, 15 (Apr. 2008), 4410–4417. [313] Zhou, Y., Wing, M.R., Sondek, J. and Harden, T.K. 2005. Molecular cloning and characterization of PLC-η2. Biochemical Journal. 391, 3 (Nov. 2005), 667. [314] Zikherman, J., Parameswaran, R. and Weiss, A. 2012. Endogenous antigen tunes the responsiveness of naive B cells but not T cells. Nature. 489, 7414 (Sep. 2012), 160–164.

164 Acknowledgements

Acknowledgements

Though only my name appears on the cover of this dissertation, a great many people have contributed to its production. I owe my gratitude to all those people who have made this dissertation possible.

Foremost, I would like to express my sincerest gratitude to my supervisor and promoter Prof. Dr. Peter Gierschik, for his constant guidance, support, motivation and untiring help during the course of my PhD.

I also appreciate my advisory committee, Prof. Dr. Ralf Marienfeld and Prof. Dr. Reinhard Wetzker for all their guidance and valuable comments on my research, especially through my intermediate examinations.

Thank you also to PD Dr. Hans Kestler and Johann M. Kraus for doing the cluster analysis.

I owe particular thanks to Dr. Claudia Walliser who always supported me in any possible way. With her in-depth expertise in the scientific field she always provided me efficient assistance and support in experimental as well as in theoretical aspects. Thank you Claudia!

In my daily work I have been blessed with a friendly and cheerful group of colleagues. Thanks for the pleasant working atmosphere go to Norbert Zanker, Patricia Schilling, Zinna Rasonabe, Angela Mansard, Torben Langer, Jennifer Haas, Maren Lillich, Dr. Barbara Möpps, Dr. Carolin König, Dr. Petra Vatter, Carolin Weiss, Susanne Gierschik, Vincenzina Palmieri and all the other members of the institute.

Special thanks go to Norbert, who always provided me with valuable information and input for discussions beyond the scientific field ;)

A special thank also goes to my brother Armin who developed and built up the individual incubation chambers. Without his technical support an outstanding and important part of this thesis would not have been possible.

At the end, I want to say that this dissertation would not be possible with the great support of my parents, my sister Carina and my brother Armin.

One of the most important persons in my life, Christoph, supported and encouraged me all the time. Thank you!!!

165 Curriculum Vitae

Curriculum Vitae

Personal Name Anja Jasmin Bühler Date of Birth June 10th 1984 Place of Birth Ulm Nationality German

Education since August 2009 Ph. D. thesis under the supervision of Prof. Dr. Peter Gierschik

October 2007 - Master of Science (M.Sc.) program in biochemistry at Ulm May 2009 University; grade: 1.4; Master Thesis at the Institute of Pharmacology and Toxicology under the supervision of Prof. Dr. Peter Gierschik: “Molecular mechanism of autoinhibitory regulation

of phosphoinositide-specific phospholipase C-γ2” (grade 1.0).

October 2004- Bachelor of Science (B.Sc.) program in biochemistry at Ulm October 2007 University; grade: 2.2; Bachelor Thesis at the Institute of Pharmacology and Toxicology under the supervision of Prof. Dr.

Peter Gierschik: “Analysis of the regulatory PHN-SH2-SH2-SH3-

PHC-Domain of phosphoinositide-specific phospholipase C-γ2” (grade 1.0).

2001-2004 Abitur Ernährungswissenschaftliches Gymnasium Valckenburgschule Ulm

1995-2001 Mittlere Reife Ulrich von Ensingen Realschule Ulm

Languages English Good in writing, reading and speaking Spanish Basic knowledge

166 Publications

Publications

Everett, K.L., Buehler, A., Bunney, T.D., Margineanu, A., Baxendale, R.W., Vatter, P., Retlich, M., Walliser, C., Manning, H.B., Neil, M.A.A., Dunsby, C., French, P.M.W., Gierschik, P., and Katan, M. (2011) Membrane Environment Exerts an Important Influence on Rac-Mediated Activation of Phospholipase C-2. Molecular and Cellular Biology. 31; 6: 1240-1251

Gierschik, P., Buehler, A., Walliser, C. (2012) Activated PLCγ breaking loose. Structure. 5, 20: 1989-1990

Buehler, A., Walliser, C., Haas, J., Kraus, J.M., Filingeri, D., Havenith G., Kestler, H.A., Milner, J.D., and Gierschik, P. (2013) Cool-temperature-mediated activation of human hereditary PLAID variant phospholipase C-2. submitted for publication

167 Appendix

Appendix

Circuit diagram of individual incubation chambers

168 Statutory Declaration

Statutory Declaration

I hereby declare that I wrote the present dissertation with the topic:

Molecular mechanisms regulating phospholipase C-2 activity independently and used no other aids than those cited. In each individual case, I have clearly identified the source of the passages that are taken word for word or paraphrased from other works.

I also hereby declare that I have carried out my scientific work according to the principles of good scientific practice in accordance with the current „Satzung der Universität Ulm zur Sicherung guter wissenschaftlicher Praxis“ [Rules of the University of Ulm for Assuring Good Scientific Practice].

Ulm, 12.12.13

______Anja Bühler

169