UNIVERSITY OF CINCINNATI

Date: 10-Nov-2009

I, Keith J Gaddie , hereby submit this original work as part of the requirements for the degree of: Doctor of Philosophy in Molecular, Cellular & Biochemical Pharmacology It is entitled: Structural Elements that Regulate Interactions between the

Extracellular and Transmembrane Domains of Human

Triphosphate Student Signature: Keith J Gaddie

This work and its defense approved by: Committee Chair: Terence Kirley, PhD Terence Kirley, PhD

Scott Belcher, PhD Scott Belcher, PhD

Paul Rosevear, PhD Paul Rosevear, PhD

John Maggio, PhD John Maggio, PhD

Ronald Millard, PhD Ronald Millard, PhD

11/23/2009 294

Structural Elements that Regulate Interactions between the Extracellular and Transmembrane Domains of Human Nucleoside Triphosphate Diphosphohydrolase 3

A dissertation submitted to the

Division of Research and Advanced Studies of the University of Cincinnati

in partial fulfillment of the requirements for the degree of

DOCTOR OF PHILOSOPHY (Ph.D.)

in the Department of Pharmacology and Cell Biophysics of the College of Medicine

2009

By Keith J. Gaddie

B.S. Fisk University, 2003

Committee Chairperson: Terence L. Kirley, Ph.D.

Abstract

The nucleoside triphosphate diphosphohydrolases (NTPDases) are a family of

constitutively expressed, endogenous nucleotidases, some of which regulate purinergic signaling

by divalent cation-dependent hydrolysis of acting as agonists at purinergic receptors.

Due to the scarcity of specific inhibitors and genetically modified animals, the functions of most

individual NTPDases are poorly understood and still under investigation. However, the

NTPDases have been implicated in many biological and physiological processes, including

secretion, cell adhesion, pain perception, cancer and malignant transformation,

recycling, and platelet aggregation.

Membrane-bound NTPDase3 expressed on the cell surface has a large extracellular

domain, a “linker region”, and a transmembrane domain. Evidence exists for interactions

between the transmembrane domain and the active site lobes that govern the function of rat

NTPDase1. However, the specific mechanism(s) regulating this cross-talk for NTPDases, as well as how the N- and C-terminal transmembrane helices in NTPDase3 interact within and

between monomers to mediate oligomerization and modulation of enzymatic activity is still

unknown.

The roles of the conserved residues of human NTPDase3, located in the “linker

region” that connects the N- and C-terminal transmembrane helices with the extracellular active

site, were examined by proline to alanine substitutions coupled with single cysteine substitutions

strategically placed in the transmembrane domain to serve as cross-linking “sensors” of helical

interactions. Mutation of several proline residues resulted in decreased nucleotidase activities

and some “uncoupled” the effect of ATP binding on TMD movements. The data suggest a role

for proline residues 53 and 481 in the linker region of human NTPDase3 for “coupling”

iii binding and hydrolysis at the active site to movements and/or rearrangements of the transmembrane helices necessary for optimal nucleotide hydrolysis.

To investigate the structural/functional roles of the conserved polar residues in the transmembrane helices of human NTPDase3, each was singly mutated to alanine. All mutants were properly glycosylated and had specific activities similar to wild-type, except Q44A. The

Q44A mutation decreased specific activities by approximately 50% - 70%, and nearly eliminated

Triton X-100 detergent inhibition. The same conserved polar residues were mutated to cysteine, singly and in pairs, to allow a disulfide cross-linking strategy to map potential inter- and intra- molecular hydrogen bond interactions. The results support the centrality of Q44 for the strong inter-molecular interactions driving the association of the N-terminal domains of two NTPDase3 monomers in a dimer, while S39 and T495 may contribute to helical interactions involved in forming higher order oligomers. These results suggest a model for putative hydrogen bond interactions of the conserved polar residues in the transmembrane domain of native, dimeric

NTPDase3 that are important for protein expression, activity, and susceptibility to membrane perturbations.

This dissertation provides additional insights to the structural elements that regulate the enzymatic activity of the NTPDases. With the recent elucidation of the crystal structure of the extracellular portion of rat NTPDase2, the major remaining structural questions regarding the cell membrane NTPDases are how the TM helices interact and how these interactions are coupled to modulation of enzyme activity and NTPDase function. This dissertation addresses these questions for human NTPDase3, and suggests that similar structure-function relationships are relevant to the family of NTPDases controlling purinergic signaling.

iv Acknowledgements

My graduate experience was a big adjustment, a learning process of my strengths and

weaknesses, and a time where I was always pushed to excel to new heights. Luckily, I was

surrounded by talented and accomplished individuals who cared about my scientific future and

were willing to share their expertise as I traveled this challenging journey. First, I would like to

express my sincere gratitude to my mentor and advisor, Dr. Terry Kirley. Dr. Kirley possesses

so many qualities of a great scientist, disciplined, the ability to pay a great attention to detail, and excellent time management skills. In addition, through our casual conversations and some observation, he is an excellent family man.

Dr. Kirley was very instrumental in developing my skills as a research scientist and he always encouraged me to work and think in an independent manner, but more importantly, to

pay attention to the small details. This has allowed me to grow as a scientist and to think critically about my data. I am truly indebted to Dr. Kirley for all his hard work, support, and guidance. In addition, I owe thanks to my committee members, Drs. Scott Belcher, John

Maggio, Ronald Millard, and Paul Rosevear, for providing constructive criticism and insight over the years. Their input and suggestions were instrumental in shaping the progression of my research projects.

I would also like to thank all the members of the Kirley laboratory for their assistance and guidance throughout the years. When I first joined the laboratory, Dr. Patrick Crawford was very instrumental in teaching me the basic laboratory protocols and worked closely with me on his project regarding the NTPDase3 splice variant, NTPDase3β. Dr. Vasily Ivanenkov instructed and motivated me to interpret my data from several points of view. In addition, he discovered that Tween 20 stimulated NTPDase3 activity and I applied in my dissertation

v research. I would also like to thank Dr. Min Yang and Clifford Cookman for all of their assistance during their time in the laboratory.

I would like to thank everybody in the department of Pharmacology, especially Nancy

Thyberg, Damita Jo Harris, and Carol Ross for always handling the seen and unseen obstacles of graduate school from a paperwork point of view.

Finally, I am so thankful for having a supportive family (Godsister’s Stephanie Ford and

Janelle Henson) and friends (Linda Smith and Dr. Bradley Sheares) who never gave up on me along the way. I would not have successfully completed my studies if it were not for my mom

(Pamela J. Gaddie) and dad’s (Keith Taylor) unconditional love, my aunts’ (Brenda, Dee Dee,

LaVada, and Margo) unwavering support, as well as my best friend’s (Derrick G. Demmons) and roommate’s (Mark A. Edmonson) encouraging words.

vi Table of Contents

Page

Abstract iii

Acknowledgements v

Table of contents vii

List of figures and tables xiii

List of abbreviations xvii

Chapter I: Introduction 1

1. Evolutionary Appearance of 1

1.1. Nucleotide structure 2

1.2. Nucleotide synthesis 2

1.2.1. De novo nucleotide synthesis 2

1.2.2. De novo nucleotide synthesis 4

2. The Concept of Purinergic Signaling 8

3. ATP: The Neurotransmitter 12

4. Purinergic Receptors 15

4.1. P1 (Adenosine) Receptors 18

4.2. P2X Receptors 18

4.3. P2Y Receptors 19

5. Functional Significance of Signaling 20

6. Ecto-nucleotidases 21

6.1. Alkaline phosphatase (AP) 21

6.2. Ecto-5’-nucleotidase/CD73 23

vii 6.3. Ecto-nucleotide phosphodiesterase/pyrophosphatase (E-NPP) 23

6.4. Nucleoside Triphosphate Diphosphohydrolase (NTPDase) 24

7. Molecular Cloning, Sequencing, Expression, and Characterization of the NTPDases 26

7.1. The membrane-bound, cell surface NTPDases (NTPDase1-3 and 8) 26

7.2. The intraluminal NTPDases (NTPDase4-7) 27

8. The Biochemical and Structural Characterization of the NTPDases 30

8.1. Enzymology of the NTPDases 30

8.2. Inhibitors of the NTPDases 31

8.3. Site-directed mutagenesis and computational comparative protein modeling of the

membrane-bound, cell surface NTPDases 36

8.3.1. The structural features of the extracellular domain of the NTPDases 38

8.3.2. The structural features of the transmembrane domain of the NTPDases 41

9. Functional Significance of the NTPDases 44

10. Dissertation Scope and Objectives 46

Chapter II: Methods and Materials 49

1. General Methods and Materials 49

1.1. Purification of oligonucleotides for site-directed mutagenesis 49

1.2. Site-directed mutagenesis of NTPDase3 50

1.3. Bacterial transformation and inoculation of PCR products 50

1.4. Isolation of DNA from XL1-Blue Supercompetent Cells 51

1.5. Restriction enzyme digestion of DNA 52

1.6. Transient transfection of COS-1 cell membranes 52

viii 1.7. COS-1 cell crude membrane preparations 53

1.8. Protein concentration of COS-1 cell crude membrane preparations 54

1.9. Specific nucleotidase activity of NTPDase3 expressed in COS-1 cells 55

1.10. Malachite green procedure for nucleotide assays 55

1.11. SDS-PAGE and Western blotting 56

1.12. Deglycosylation 57

2. Project Specific Methods and Materials 58

2.1. NTPDase3 “linker region” proline residue mutagenesis 58

2.1.1. List of oligonucleotides/mutants made 58

2.1.2. Oxidative cross-linking in the presence and absence of substrate 59

2.2. NTPDase3 transmembrane domain polar residue mutagenesis 59

2.2.1. List of oligonucleotides/mutants made 59

2.2.2. Triton X-100 nucleotidase assays 61

2.2.3. Oxidative cross-linking and alkylation of cysteine 61

2.2.4. NTPDase3 nucleotidase assays in the presence of Tween 20 62

2.3. NTPDase3 alternatively spliced variant (NTPDase3β) 63

2.3.1. IMAGE clone encoding NTPDase3β 63

2.3.2. List of oligonucleotides/mutants made 63

2.4. NTPDase3 single nucleotide polymorphisms (SNPs) 64

2.4.1. Source of SNPs 64

2.4.2. List of oligonucleotides/mutants made to mimic known SNPs 64

2.4.3. Treatment of wild-type and mutant NTPDase3 with Concanavalin A 65

ix Chapter III: Results 66

1. Mutagenesis of NTPDase3 “linker region” proline residues 66

1.1. Rationale for selection and analysis of NTPDase3 “linker region” proline residues 66

1.2. Characterization of the NTPDase3 proline to alanine mutants 69

1.3. Rationale for selection of transmembrane “sensor” mutations and experimental

approach 76

1.4. Characterization of the TM “sensor” mutations (V42C and G489C NTPDase3) 77

1.5. Characterization of proline to alanine mutants in the V42C or G489C NTPDase3

background 81

2. Mutagenesis of NTPDase3 transmembrane domain polar residues 87

2.1. Rationale for selection and analysis of TM polar residue mutants 87

2.2. Characterization of the NTPDase3 polar residue to alanine mutants 90

2.3. Characterization of the polar residues singly mutated to cysteine 96

2.4. Paired cysteine mutagenesis and generation of a model incorporating the inter-

and intra-molecular hydrogen bonding pattern of the conserved polar residues

in the transmembrane helices of NTPDase3 100

2.5. Tween 20 stimulation of NTPDase3 107

3. NTPDase3 alternatively spliced variant 117

3.1. Rationale and experimental approach 117

3.2. NTPDase3β alternatively spliced isoform 118

3.3. Expression and enzymatic activities of NTPDase3 isoforms 122

3.4. Co-transfection of NTPDase3a and NTPDase3β 123

4. NTPDase3 single nucleotide polymorphisms (SNPs) 127

x 4.1. Rationale and experimental approach 129

4.2. Characterization of NTPDase3 SNPs 129

Chapter IV: Discussion 135

1. Regulation of the enzymatic activity of membrane-bound, cell surface NTPDases 135

2. Structural and functional roles of specific structural elements in NTPDase3 137

2.1. The conserved proline residues are important components of the linkage between

the active site and transmembrane domain movements 137

2.2. The conserved polar residues stabilize transmembrane domains and promote

oligomerization 139

2.3. NTPDase3β: A possible modulator of nucleotidase activity and purinergic receptor

signaling 144

2.4. Biochemical characterization of NTPDase3 single nucleotide polymorphisms 145

3. Summary and Future Directions 145

3.1. Determine if a polyproline II helix exists in the linker region of NTPDase3 and

identify the specific amino acids involved in its formation/stability 146

3.2. To further define the structural motifs that mediate oligomerization via the

transmembrane domain of the NTPDases 147

3.3. Further biochemical characterization of the single nucleotide polymorphisms of

NTPDase3 148

Chapter V: References 149

xi Appendix: Publications and Manuscripts under Review 175

xii List of Figures and Tables

Figures: Page

1. The structural elements of the most common nucleotides 3

2. De novo synthesis of purine nucleotides 5

3. Synthesis of AMP and GMP from IMP 6

4. De novo synthesis of pyrimidine nucleotides 7

5. Timeline of highlights of ATP signaling 9

6. General methods of nucleotide release 14

7. Purinergic signaling pathways 16

8. Major extracellular nucleotide hydrolyzing 22

9. Intracellular nucleotidase control of protein folding and processing 29

10. The three structural regions of the NTPDases 37

11. Multiple sequence alignment for portions of NTPDase sequences between the N-

and C-terminal transmembrane helices and the extracellular domain containing

the enzyme active site 67

12. Alanine substitution at P53, P472, P475, and P485 caused a decrease in

expression level in mutant NTPDase3 enzymes 70

13. P53A and P481A NTPDase3 mutants exhibit a decrease in nucleotidase

activities as compared to the C10S NTPDase3 background mutant 72

14. Alanine substitution of NTPDase3 proline residues are properly processed and

trafficked to the cell membrane 73

15. Double alanine substitution of the proline residues cause a decrease in

xiii expression level in mutant NTPDase3 enzymes 74

16. P53A/P481A and P484A/P485A NTPDase3 mutants exhibit a decrease in

nucleotidase activities as compared to the C10S NTPDase3 background mutant 75

17. Cysteine substitution at V42 and G489 in NTPDase3 exhibits similar nucleotidase

activities as the C10S/C501S/C509S NTPDase3 background mutant 78

18. ATP binding at the extracellular domain decreased the cross-linking efficiency

of V42C and G489C NTPDase3 mutants 79

19. Oxidative cross-linking decreases ATP hydrolysis 80

20. Ca-ATPase activity of the V42C proline to alanine mutants 82

21. Ca-ATPase activity of G489C proline to alanine mutants 83

22. P53A and P481A abolish the ATP-induced decrease in cross-linking efficiency 84

23. P481A abolishes the ATP-induced decrease in cross-linking efficiency 86

24. Rationale for NTPDase3 mutations generated and analyzed in this study 88

25. Alanine substitution of the polar residues cause a decrease in expression level,

but similar nucleotidase activities in most mutant NTPDase3 enzymes 91

26. Alanine substitution of NTPDase3 polar residues are properly processed and

trafficked to the cell membrane 94

27. Alanine substitution at Q44 eliminates the detergent inhibition of Triton X-100

on NTPDase3 enzymatic activities 95

28. Characterization of NTPDase3 polar residues to single cysteine mutants 97

29. Specific activities of NTPDase3 double cysteine mutants 101

30. Oxidative cross-linking of NTPDase3 polar residue hydrogen bonding pair

(double cysteine) mutants 103

xiv 31. Model of the transmembrane domain helical interactions in the NTPDase3 dimer 105

32. NTPDase3 S39C/Q44C, T41C/Q44C, and T495C/Q44C mutant enzymatic activities 108

33. Analysis for higher order oligomeric structure by oxidative cross-linking of

S39C/Q44C, T41C/Q44C, and T495C/Q44C NTPDase3 double mutants 109

34. The stimulatory effect of Tween 20 is specific for NTPDase3 and conserved among

species 110

35. Tween 20 promotes oligomerization mediated by the TMDs of NTPDase3 112

36. Detergents Tween 20 and Triton X-100 do not change the enzymatic activity of

sNTPDase3 115

37. Effect 0.1% Tween 20 or Triton X-100 on CuPhen cross-linking efficiency of the

TMDs of a mutant NTPDase3 116

38. cDNA and translated amino acid sequences of NTPDase3β, the truncated splice

variant of NTPDase3 119

39. A diagram of the exon map of the human NTPDase3 gene representing the

splicing events that occur 121

40. Western blot analysis of co-transfection experiments using the KLH1 or KLH11

antibody 125

41. Percent activity of co-transfections normalized per molecule active protein 128

42. Enzymatic activity of NTPDase3 SNPs 130

43. ConA stimulatory effect on NTPDase3 SNPs 131

44. Triton X-100 detergent effect on NTPDase3 SNPs 132

45. Tween 20 stimulatory effect on NTPDase3 SNPs 133

xv Tables: Page

1. Members of the E-NTPDase family 25

2. Enzymatic characterization of NTPDase1-3 and 8 32

3. Results of co-transfection of different NTPDase3 proteins in COS cells 124

xvi List of Abbreviations

3-(N-morpholino)propanesulfonic acid (MOPS)

4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (HEPES)

5-phosphoribosyl α-diphosphate (PRPP)

5’-aminoimidazole ribonucleotide (AIR) acetylcholine (ACh) adenosine (Ado) adenosine 5’-diphosphate adenosine 5’-monophosphate (AMP) adenosine 5’-triphosphate (ATP) alkaline phosphatases (AP) apyrase conserved regions (ACRs) arylazidoaminopropionyl-ATP (ANAPP3)

ATP-binding cassette (ABC) proteins

cyclic (cAMP)

cyclic monophosphate (cGMP)

concanavalin A (Con A)

copper phenanthroline (CuPhen)

5’-monophosphate (CMP; cytidylate)

cytidine 5’-triphosphate (CTP)

deoxyribonucleic acid (DNA)

disuccinimidyl suberate (DSS)

dithiothreitol (DTT)

xvii ecto-nucleoside triphosphate diphosphohydrolases (eNTPDases) ecto-nucleotide phosphodiesterase/pyrophosphatases (E-NPPs) endoglycosidase H (EndoH) endothelium-derived relaxing factor (EDRF) ethylenediaminetetraacetic acid (EDTA) formylglycinamide ribonucleotide (FGAR) formylglycinamidine ribonucleotide (FGAM) gamma-amino (GABA) glycinamide ribonucleotide (GAR) guanosine 5’-diphosphate (GDP) guanosine 5’-monophosphate (GMP) guanosine 5’-triphosphate (GTP) (Hyp) inorganic phosphate (Pi) (Ino) inosine 5’-diphosphate (IDP) inosine 5’-monophosphate (IMP; inosinate) inosine 5’-triphosphate (ITP)

International Union of Pharmacology (IUPHAR) isotonic wash buffer (IWB) maleimide polyethylene glycol-5000 (MalPEG) multi-drug resistance (MDR) glycoprotein

N-ethylmaleimide (NEM)

xviii N-succinylo-5-aminoimidazole-4-carboxamide ribonucleotide (SAICAR) nicotinamide dinucleotide (NAD+; NADH) non-adrenergic, non-cholinergic (NANC) noradrenaline (NA) or norepinephrine (NE) orotidine 5’-monophosphate (orotidylate; OMP) p-chloromercuriphenylsulfonate (pCMPS) phosphatidylinositol (PI) phospholipase C (PLC) plasma cell differentiation antigen-1 (PC-1) polymerase chain reaction (PCR) purine nucleoside phosphohydrolase (PNP) real time-polymerase chain reaction (RT-PCR) ribonucleic acid (RNA) phosphate pyrophosphokinase (PRPP synthetase) single nucleotide polymorphisms (SNPs) sodium azide (NaN3) sodium dodecyl sulfate (SDS) tissue homogenization buffer (THB) transmembrane (TM) transmembrane domain (TMD) 5’-diphosphate (UDP) uridine 5’-monophosphate (UMP; uridylate) uridine 5’-triphosphate (UTP)

xix 5’-monophosphate (XMP ; xanylate)

xx Chapter I: Introduction

1. Evolutionary Appearance of Nucleotides

Nucleotides are the basic structural units that form ribonucleic acid (RNA) and deoxyribonucleic acid (DNA), the compounds that contain the genetic instructions used in the development and functioning of all known living organisms. Nucleotides also play important roles in cellular by serving as sources of chemical energy, participate in cell signaling, and are utilized as cofactors of enzymatic reactions. In the early stages of the evolution of life, it is thought that thermal processes mediated phosphorylation of nucleosides and the formation of pyrophosphate bonds (1). These resulting compounds became the starting materials for nucleotide synthesis in terrestrial and aquatic life (2). Since adenosine 5’- triphosphate (ATP) has a propensity to bind the metal ion , which promotes dephosphorylation and the generation of energy, it is thought to be particularly suitable for early development. Inosine 5’-triphosphate (ITP) and guanosine 5’-triphosphate (GTP) also show a metal ion-promoted dephosphorylation, but their reaction rates are lower. The were also available on the early earth. It is suggested they were incorporated into living systems in a passive way, possibly directed by the (3). Therefore, purines such as ATP may have played a crucial and active role in early evolution. The formation of ATP from adenosine 5’- monophosphate (AMP), or other nucleosides, seems to have paved the way for the synthesis of oligonucleotides, then RNA, and eventually DNA, a molecule more reliable for storing genetic information (4).

1 1.1. Nucleotide structure

As shown in Figure 1, a nucleotide is composed of a nitrogenous base, a five carbon

sugar (either ribose (RNA) or deoxyribose (DNA)), and one to three phosphate groups (at either

the 2’, 3’, or 5’-carbon atom of the sugar moiety, the 5’-carbon site occurring most commonly).

Together, the nitrogenous base and the sugar constitute a nucleoside. Nucleosides can contain

two different types of nitrogenous bases, purines (adenine and ) or pyrimidines (,

, and thymine).

1.2. Nucleotide synthesis

Nucleotides can be synthesized in vitro and in vivo. In vivo nucleotides are synthesized

via the de novo pathway or the salvage pathway. De novo synthesis of nucleotides begins with

their metabolic precursors: amino acids, ribose 5’-phosphate, CO2, and NH3. Salvage pathways

recycle free bases and nucleosides released from nucleic acid breakdown. Both pathways are important in cellular metabolism.

1.2.1. De novo purine nucleotide synthesis

The de novo pathway for purine nucleotide synthesis seems to be identical in nearly all living organisms. The two parent purine nucleotides are AMP and guanosine 5’-monophosphate

(GMP), containing the purine bases adenine and guanine. Eleven enzymatic reactions lead to the first intermediate with a complete purine ring,

2

Figure 1. The structural elements of the most common nucleotides. Nucleotides have three characteristic components: 1. a nitrogenous base, 2. a pentose sugar, and 3. a phosphate. The molecule without the phosphate group is called a nucleoside. The nitrogenous bases are derivatives of two parent compound classes, pyrimidines (cytosine, uracil, and thymine) and purines (adenine and guanine). The base of the nucleotide is joined covalently (at N-1 of pyrimidines and N-9 of purines) in an N-β-glycosyl bond to the 1’ carbon of the pentose, and the phosphate is most commonly esterified to the 5’-carbon atom.

3 inosine 5’-monophosphate (IMP; inosinate) (Figure 2). From IMP, AMP and GMP are synthesized in separate two step reactions (Figure 3).

1.2.2. De novo pyrimidine nucleotide synthesis

The common pyrimidine nucleotides are cytidine 5’-monophosphate (CMP; cytidylate) and uridine 5’-monophosphate (UMP; uridylate), which contain the pyrimidines cytosine and uracil. During de novo pyrimidine nucleotide synthesis (Figure 4), the six-member pyrimidine ring is made first and then attached to ribose 5’-phosphate. Pyrimidine nucleotide synthesis begins with the formation of N-carbamoylaspartate from aspartate and carbamoyl phosphate and yields orotate through subsequent reaction steps. Orotate reacts with 5-phosphoribosyl α- diphosphate (PRPP) to yield orotidine 5’-monophosphate (orotidylate; OMP) which is decarboxylated to form UMP. From UMP, other pyrimidine nucleotides are derived through kinase reactions to yield uridine 5’-triphosphate (UTP). Cytidine 5’-triphosphate (CTP) is derived from the conversion of UTP by cytidylate synthetase. Upon the accumulation of excess concentrations of CTP, pyrimidine nucleotide synthesis is also regulated by a feedback inhibition, where CTP inhibits the enzyme aspartate transcarbamoylase, the first reaction in the process.

4

5 5-Phosphoribosyl-1-pyrophosphate (PRPP) N -CAIR Glutamine 1 Glutamate 7

PPi 5-Amino-4-carboxyamino- 5-Phospho-β-D-ribosylamine Imidazole ribonucleotide 2 ATP Aspartate ATP ADP + Pi 8 ADP + Pi Glycinamide ribonucleotide (GAR) N-Succinylo-5-aminoimidazole-4- 10 N -Formyl H4 carboxamide ribonucleotide (SAICAR) 3 folate

Formylglycinamide ribonucleotide (FGAR) 9 Fumarate Glutamine 5-aminoimidazole-4-carboxamide 4 Glutamate ATP ribonucleotide (AICAR) ADP + Pi 10 10 N -Formyl H4 Formylglycinamidine ribonucleotide (FGAM) folate ATP N-Formylaminoimidazole-4- 5 ADP + Pi carboxamide ribonucleotide (FAICAR) H20 5-Aminoimidazole ribonucleotide (AIR) 11 H2O HCO3 ATP 6 Inosine-5’-monophosphate (IMP); ADP + Pi 5 5 Inosinate N -Carboxyaminoimidazole ribonucleotide (N -CAIR)

Figure 2. De novo synthesis of purine nucleotides. Formation of 5-phosphoribosylamine, the

product of the first reaction, is the first committed step in purine synthesis. The enzymes are: 1.

glutamine-PRPP amidotransferase, 2. GAR synthetase, 3. GAR transformylase, 4. FGAR

amidotransferase, 5. FGAM cyclase (AIR synthetase), 6. N5-CAIR synthetase, 7. N5-CAIR mutase, 8. SAICAR synthetase, 9. SAICAR lyase, 10. AICAR transformylase, and 11. IMP synthase.

5 Fumarate Adenosine-5’- GDP + Pi GTP Adenylosuccinate monophosphate (AMP); 2 Adenylate Aspartate 1

Inosine-5’-monophosphate (IMP); Inosinate

H2O NAD+ NADH + H+ 3 Gln Glu ATP AMP + PPi Xanthine-5’- Guanosine-5’- monophosphate (XMP); monophosphate (GMP); Xanylate 4 Adenylate H2O

Figure 3. Synthesis of AMP and GMP from IMP. The reaction enzymes are: 1. adenylosuccinate synthetase, 2. adenylosuccinate lyase, 3. IMP dehydrogenase, and 4. XMP- glutamine amidotransferase.

6 Aspartate Carbamoyl aspartate Phosphate transcarbamoylase Pi N-Carbamoylaspartate dihydroorotase

H2O L-Dihydroorotate + dihydroorotate NAD

dehydrogenase NADH + H+ Orotate orotate PRPP phosphoribosyl- transferase PPi Orotidylate orotidylate decarboxylase CO2 Uridylate (UMP) kinases 2 ATP 2 ADP Uridine-5’-triphosphate Gln cytidylate Glu synthetase ATP ADP + Pi Cytidine-5’-triphosphate (CTP)

Figure 4. De novo synthesis of pyrimidine nucleotides. The pyrimidine is constructed from carbamoyl phosphate and aspartate. Then the ribose 5’-phosphate is added to the completed pyrimidine ring by orotate phosphoribosyltransferase.

7 2. The Concept of Purinergic Signaling

In 1929 (Figure 5), ATP was identified in muscle extracts independently by Karl

Lohmann at the Kaiser Wilhelm Institute for Biology in Berlin and by Cyrus Hartwell Fiske and

Yellagaprada Subbarow at Harvard University (5, 6). This led individuals to believe ATP was primarily involved in muscle contraction. In the years following the discovery of ATP it became clear that the three chemically and functionally different parts of the molecule were synthesized by independent metabolic routes and that ATP was generated during the breakdown of glucose to lactic acid (anaerobic glycolysis), within the citric acid cycle, and predominantly during aerobic oxidation (oxidative phosphorylation) (7). The recognition that ATP was utilized for energy requiring functions allowed Lippman to introduce his general hypothesis for energy transfer in living cells (8). As a result, ATP was soon recognized as the universal source of readily available chemical energy for all living cells.

It was the essential intracellular roles of ATP that led to considerable resistance to the concept that ATP could exert any extracellular physiological action. The main argument was that cells would not release a molecule so fundamental to life under normal physiological conditions. This idea was reinforced by the size and charge of ATP, indicating it could not cross the plasma membrane by simple diffusion (9, 10). However, the biological effects of extracellular nucleotides were discovered early. In 1929, Drury and Szent-Gyorgyi at the

University of Cambridge published the first paper demonstrating the extracellular actions of

AMP and adenosine on the heart and coronary blood vessels (11), causing negative chronotropic

effects and dilation of coronary vessels. Several studies followed, confirming the role

8

2010 2008 sol NTPDase2 crystal structure 2005 First 3-D NTPDase3 model proposed 2000 1998 NTPDase3 identified 1996 NTPDase1 identified 1993 , 1994 P2X receptor 1990 1990 5’-nucleotidase identified 1985 P2X and P2Y receptors differentiated 1985 AP and NPP1 identified 1980

1974 ATP in cholinergic synaptic vessicles 1970 1970 ATP recognized as a neurotransmitter

1960 1957 Term “ecto-ATPase” coined

1950 1948 Chemical synthesis of ATP 1945 Extracellular hydrolysis of ATP 1940

1934 Biological activity of ATP 1930 1929 ATP discovered

Figure 5. Timeline of highlights of ATP signaling. ATP was discovered in 1929 and its biological activity recognized in 1934. In 1957, the term “ecto-ATPase” was coined. In 1970,

ATP was recognized as a neurotransmitter and in the mid 1980s and 90s the purinergic receptors and the ecto-nucleotidases were discovered. In 1998, NTPDase3, the enzyme used throughout this dissertation, was cloned, sequenced, expressed, and characterized by our laboratory (12).

9 of purine nucleotides in the cardiovascular system (13). In 1934, J.H. Gillespie from Queen’s

University in Belfast extended these studies, providing evidence that extracellular ATP affects a

variety of tissues including heart, vasculature, and smooth muscle (14). Non-vascular effects of

purine nucleotides were later established in the gut, bladder, vagina, and other visceral organs

and vascular beds (15, 16).

Initially, adenosine (17) or AMP (18) were the purines thought to mediate the

extracellular actions of nucleotides, which are both degradation products of ATP. Studies soon established the length of the phosphate chain conferred differences in activity and ATP often had stronger effects than the other purinergic compounds (19, 20). In 1948, ATP was chemically synthesized by Alexander Todd, and scientists no longer isolated ATP from muscle extracts for use in their experiments (21). More detailed reports followed on the hydrolysis of ATP and ADP by intact yeast cells (22), human erythrocytes (23), ascites tumor cells (24), and nucleated avian erythrocytes (25). This led to the term “ecto-ATPase”, which was coined in 1957 by W.A.

Engelhardt from the Academy of Sciences in Moscow, during his findings of surface-located

ATPase activity on avian blood cells (25) as well as he introduced the terms “ecto-enzyme” and

“ecto-apyrase” (25, 26).

For many years the concept of neurotransmission was dominated by the dogma that each neuron releases only a single transmitter, known as “Dale’s Principle” (27). This arose from a widely adopted misinterpretation of his suggestion that the same neurotransmitter was stored in and released from all terminals of a single neuron. However, Dale’s Principle did not specifically exclude the possibility that more than one transmitter may be associated with the same neuron. In 1953 and 1954, Holton observed that ATP could be released from stimulated sensory neurons, which raised the possibility that ATP may be a neurotransmitter at sensory

10 nerve endings (28, 29). Soon after, it was recognized that adenosine and ATP were co-stored

with catecholamines in adrenal chromaffin cells (30, 31). Subsequently, ATP was shown to be

co-released with adrenaline from chromaffin cells (32, 33). In 1970, Geoffrey Burnstock, then working at the University of Melborne, proposed the existence of a “non-adrenergic, non- cholinergic” (NANC) regulation of smooth muscle contraction in the gut and bladder, in which

ATP was identified as the principal transmitter (34). Soon after, Burnstock and colleagues

provided the first evidence for sympathetic co-transmission involving ATP together with

noradrenaline (NA). In that study, they demonstrated that stimulation of periarterial sympathetic

nerves leads to release of tritium from guinea-pig taenia coli pre-incubated in 3H-adenosine

(which is taken up and largely converted to 3H-ATP), and that the release of both tritium and NA

was blocked by guanethidine (35). This led Burnstock and his colleagues to propose the use of

the term “purinergic” (36), and demonstrate that ATP fully fulfills the criteria for a

neurotransmitter as well as a co-transmitter.

Since the mid 1980’s, scientists have identified and functionally characterized many

purinergic (and pyrimidinergic) receptors (37) and the ecto-nucleotidases that regulate these

receptors by hydrolysis of their extracellular nucleotide agonists. We now know that not only

ATP, but also adenosine 5’-diphosphate (ADP), UTP, uridine 5’-diphosphate (UDP),

dinucleoside polyphosphates, UDP glucose, and even nicotinamide adenine dinucleotide (NAD+) activate nucleotide receptors and act as signaling molecules between cells. These signaling molecules are rapidly degraded to terminate their action on purinergic receptors by ecto- nucleotidases. The first ecto-nucleotidases cloned were an alkaline phosphatase (38), nucleotide phosphodiesterase/pyrophosphatase 1 (NPP1) (39), followed by ecto-5’-nucleotidase (40). In

1996, the first nucleoside triphosphate diphosphohydrolase (NTPDase), human CD39/NTPDase1

11 was identified (41), and in 1998, human NTPDase3, the enzyme studied in this dissertation, was

identified by our laboratory (12). Now it is widely recognized that purinergic signaling is an

evolutionarily ancient system involved in many neuronal and non-neuronal mechanisms, in both

short-term and long-term (trophic) events (42).

3. ATP: The Neurotransmitter

Intracellular ATP is primarily utilized to drive energy-requiring processes such as active

transport, cell motility, and biosynthesis, whereas extracellular ATP functions as a signaling

molecule at purinergic receptors. There is a multi-step cascade of ATP “release-signaling-

inactivation” that includes: 1. transient release of ATP and ADP; 2. triggering of signaling events

through a series of ligand-gated ionotropic (P2X) and metabotropic (P2Y) receptors; 3.

nucleotide breakdown by membrane-bound and soluble nucleotidases; 4. interaction of the

resulting adenosine with its own nucleoside-selective (P1) receptors; and 5. extracellular

adenosine re-uptake and metabolism (43).

ATP is a ubiquitous molecule synthesized and released by almost every mammalian cell.

Various excitatory/secretory tissues such as nerve terminals and chromaffin cells (44, 45), pancreatic acinar cells (46), and circulating platelets (47) store ATP and ADP, together with other neurotransmitters in specialized granules (called synaptic vesicles, chromaffin granules, or platelet dense granules) at concentrations up to 1 M. Specifically, platelet dense granules contain very high concentrations of both ATP (up to 1 M) and ADP (up to 0.5 M) (47). ATP is also co- stored in the synaptic vesicles with acetylcholine (ACh) at concentrations in the millimolar range. Cholinergic synaptic vesicles can also store (at lower concentrations) GTP, ADP, and

12 AMP (chromaffin granules also UTP (48)). There is always more ACh in the isolated vesicle fractions than ATP, with the molecular ratio of ACh to ATP in cholinergic vesicles varying

between 5 and 10, depending on the tissue source (49). It is also well established that

noradrenaline (NA) and ATP exist as co-transmitters in sympathetic nerves in a molar ratio

estimated to be 7:1 to 12:1, NA:ATP (50-54). However, the molar proportions of NA:ATP vary in different tissues, and species, during development and aging, and in different pathophysiological conditions (55).

It was previously assumed the main source of ATP acting on purinergic receptors originated from damaged or dying cells. It is now accepted that ATP release from healthy cells is a physiological or pathophysiological response to hypoxia, inflammation, mechanical stress, and some agonists (56). At basal level, many cells release nanomolar concentrations of ATP, and it is likely different mechanisms underlie constitutive versus stress-stimulated nucleotide release (67-69). ATP is also released from the following non-excitatory tissues: astrocytes and glial cells (44); bone cells, joints, and keratinocytes (44, 61); cardiomyocytes (62); epithelial (57) and endothelial cells (58); erythrocytes (63); fibroblasts (59); hepatocytes (60);

monocytes/macrophages (65); neutrophils (64); and other hematopoietic cells (66).

There are three general mechanisms by which intracellular ATP can be released (Figure

6) from cells. The first is cell lysis, which may occur due to cell death or tissue damage.

Although this is not a physiological release of ATP, it contributes to pathophysiological

mechanisms. Second is vesicular release, resulting from neuronal and non-neuronal cells taking

up and packaging ATP. The vesicular ATP is then released upon stimulation (70) via both Ca-

dependent and Ca-independent mechanisms (71, 72). The third major mechanism for

13

ATP Extracellular ATP ATP ATP ATP Pool ATP (nanomolar) ATP

ATP ATP ATP NE ACh ATP Intracellular Cell ATP ATP Pool Vesicular -3 ATP ATP-binding (~10 M) ATP Damage ATP ATP ATP ATP ATP Release Cassette (ABC) ATP ATP ATP ATP proteins

Figure 6. General methods of nucleotide release. There are three main mechanisms for release of intracellular ATP: 1. cell lysis/damage; 2. vesicular release; and 3. ATP binding cassette (ABC) proteins.

14 intracellular ATP release is by activation of ATP-binding cassette (ABC) proteins. ABC proteins constitute a family of transport ATPases found in virtually all cells including bacteria, plants, and animals (73). One such ABC protein is the P-glycoprotein (multi-drug resistance

(MDR) glycoprotein), which is responsible for microorganisms acquiring resistance to a variety of normally toxic drugs via extrusion of the drugs from the microorganism (74).

Once released from the cell, extracellular adenosine nucleotides are rapidly hydrolyzed to adenosine by ecto-nucleotidases to terminate P2 receptor signaling and to limit desensitization and internalization of the P2 purinergic receptors. When adenosine levels generated by the action of ecto-nucleotidases in the extracellular space are high, adenosine is further metabolized to inosine by adenosine deaminase, or transported into cells by means of transporters and phosphorylated to AMP by adenosine kinase. Additional intracellular phosphorylation events convert AMP to ADP and finally, back to ATP.

4. Purinergic Receptors

The concept of extracellular nucleotides acting as agonists at specific receptors was first

proposed by Dr. Geoffrey Burnstock in 1976 (75). In 1978, Burnstock suggested a subdivision

of purinergic receptors into P1 (adenosine) and P2 (ATP/ADP) receptors based on several

criteria (76) (Figure 7). At the P1 receptors, an agonist potency order of adenosine > AMP >

ADP > ATP was proposed. Methylxanthines such as and were proposed to

be selective antagonists and activation of the P1 receptor was suggested to lead to changes in

intracellular cAMP levels. At the P2 receptors, an agonist potency order of ATP > ADP > AMP

15

Ecto- Nucleotidase Nucleotidase 5’-nucleotidase ATP ADP AMP Ado

Extracellular

COOH NH2 P2X P2Y 1-7 1,2,4,6,11,12,13,14 P11,2A,2B,3 Intracellular

Figure 7. Purinergic signaling pathways. The purinergic receptors are divided into P1

(adenosine) and P2 (ATP/ADP) receptors. The P2 receptors, directly modulated by ecto- nucleotidase hydrolysis of nucleotides, are divided into the non-selective P2X ligand-gated ion channels and the P2Y G-protein coupled receptors. ATP is the only agonist at the P2X receptors, while either ATP or ADP can act as agonists at some P2Y receptors. Membrane-bound and soluble nucleotidases quickly terminate agonist response by hydrolyzing NTPs and NDPs to

NMPs and the ecto-5’-nucleotidase specifically hydrolyses AMP to adenosine.

16 > adenosine was proposed. Methylxanthines were not antagonists at P2 receptors and activation

of the P2 receptor was suggested not to lead to changes in intracellular cAMP levels, but in some

cases to promote prostaglandin synthesis. In 1985, Burnstock proposed a pharmacological basis

for distinguishing subtypes of P2 receptors (77), designated P2X and P2Y. The P2X receptors have an agonist potency order of α,β-methyleneATP > β,γ-methyleneATP > ATP = 2-

methylthioATP; antagonism by arylazidoaminopropionyl-ATP (ANAPP3), and selective

desensitization following administration of α,β-methyleneATP. These contraction mediating

P2X receptors were localized in the vas deferens and urinary bladder of guinea-pig, rat, frog, as

well as in rat ventricle, and also in the smooth muscle of the rat femoral artery and rabbit central

ear artery. The P2Y receptors were characterized as having an agonist potency order of 2-

methylthioATP >> ATP > α,β-methyleneATP = β,γ-methyleneATP; weak antagonism by

ANAPP3 and desensitization following administration of α,β-methyleneATP. These P2Y

receptors were present in the guinea-pig taenia coli and the longitudinal muscle layer of the

rabbit portal vein, where they mediate relaxation, and also caused relaxation in the rat vascular endothelial cells, femoral artery, and pig aorta.

Subsequent identification of P2X and P2Y receptor subtypes led Abbracchio and

Burnstock to propose two major P2 receptor families: the non-selective P2X family of ligand- gated ion channel receptors and a P2Y family of G protein-coupled receptors (78). Currently there are seven P2X and eight P2Y receptors recognized in the human genome. Most of these receptors are activated by nucleotide concentrations within the range of 10-6 to 10-4 M (66) and

are regulated by the ecto-nucleotidases via hydrolysis of extracellular nucleotides. The

generation of adenosine and its action on the P1 receptors is an important, indirect affect of

extracellular nucleotide hydrolysis.

17 4.1. P1 (Adenosine) Receptors

P1 receptors are seven transmembrane G-protein coupled receptors. Adenosine is their

preferred, endogenous agonist. To date, four P1 receptor subtypes have been cloned and

characterized, namely: A1, a ubiquitously expressed receptor that couples to Gi/o, decreasing

cAMP, and causing inhibitory actions in most tissues; A2A, an abundant receptor in basal

ganglia, vasculature, and platelets that couples to Gs to increase cAMP levels; A2B, which interacts with Gs to increase cAMP levels; and A3, which interacts with Gi/o to decrease cAMP

levels (79-82).

4.2. P2X Receptors

P2X receptors are non-selective, ligand-gated ion channels that open in response to the

binding of extracellular ATP, but do not respond to any other naturally occurring nucleotides.

There are seven genes in humans (P2X1-7) that encode P2X subunits. Each subunit shows a

topology similar to the NTPDases, in which there are: two transmembrane helices with an

intracellular amino- and carboxyl-terminus; a large extracellular loop with 10 conserved cysteine

residues forming 5 disulfide bonds; and an ATP-binding site, which may involve regions of the

extracellular loop adjacent to the TM helices. All P2X receptors are thought to form functional

trimers (83-86).

The P2X receptor family shows many pharmacological and operational differences (87).

The kinetics of activation, inactivation, and deactivation vary considerably among P2X receptor

subtypes. The pharmacology of the recombinant P2X receptor subtypes expressed in oocytes or

18 other cell types is often different from the pharmacology of P2X receptor-mediated responses in

naturally occurring tissues. This is partly due to the existence of hetero- and homo-trimers of

various P2X receptor subtypes that form the ion pores. Alternatively spliced variants of P2X

receptor subtypes may also add to the diversity of P2X receptor properties. In addition, the

presence of ecto-nucleotidases in tissues that rapidly hydrolyze ATP is usually not a factor when

examining recombinant receptors, but is in vivo.

4.3. P2Y Receptors

Metabotropic P2Y receptors (P2Y1, P2Y2, P2Y4, P2Y6, P2Y11, , P2Y13, and P2Y14) are seven transmembrane G-protein coupled receptors. Most P2Y receptors couple to a single G protein (P2Y1,2,4,6, all couple to Gq/11; P2Y11 couples to both Gq/11 and Gs; P2Y12,13 couple to Gi;

and P2Y14 couples to Gi/o) (67). The P2Y receptor subtypes differ in agonist activation, where:

P2Y1,6,12 are activated principally by nucleoside diphosphates; P2Y2,4 are activated mainly by

nucleoside triphosphates; some P2Y receptor subtypes are activated by both purine and

pyrimidine nucleotides (P2Y2,4,6), and others are activated by purine nucleotides alone

(P2Y1,11,12) (67). In response to nucleotide activation, recombinant P2Y receptors either activate

phospholipase C (PLC) and release intracellular , or modulate adenylyl cyclase and alter

cAMP levels (67). P2Y receptor subtypes can also form heteromeric complexes and most

recently, adenosine (P1) receptors were shown to form a heteromeric complex with P2Y1 receptors (67, 88). There is little evidence that P2Y5,9,10 sequences are nucleotide receptors or

that they affect intracellular signaling cascades. Consequently, these proteins were dropped from

19 International Union of Pharmacology (IUPHAR) P2Y receptor nomenclature and termed

“orphan” receptors.

5. Functional Significance of Purinergic Receptor Signaling

The majority of the acute signaling events of purinergic receptors revolve around ATP as

a neurotransmitter, co-released with norepinephrine (NE), acetylcholine (ACh), and other

neurotransmitters and neuropeptides. Acute purinergic signaling also plays a role in exocrine

and endocrine secretion by regulation of ion transport in epithelial cells to stimulate Cl- transport and maintain moist airway surfaces (89); vascular tone, mediated not only by P2X, but also P2Y receptor subtypes that cause the release of endothelium-derived relaxing factor (EDRF) and potent vasodilatation (90-92); and insulin secretion from pancreatic β-cells (93). Other acute signaling roles for the purinergic receptors include: ionotropic, chronotropic, and arrhythmogenic effects in the myocardium (62), immune responses and control of leukocyte traffic between the blood and tissues (66); as well as activation and aggregation of platelets at sites of vascular injury (47).

There is increasing experimental data demonstrating the long-term (trophic) roles purines and pyrimidines have in embryonic development (94), cell proliferation and growth (95), induction of apoptosis and anticancer activity (66, 96), atherosclerotic plaque formation (97), as

well as wound healing and bone formation and resorption (61).

20 6. Ecto-nucleotidases

Extracellular nucleotides are rapidly hydrolyzed to adenosine to terminate nucleotide

signaling (Figure 8) at P2 receptors. There are several enzyme families involved in the metabolism and control of extracellular nucleotide concentrations, including: alkaline

phosphatases (AP); the ecto-5’-nucleotidase; the ecto-nucleotide

phosphodiesterase/pyrophosphatases (E-NPPs); and finally, the ecto-nucleoside triphosphate

diphosphohydrolases (eNTPDases), which are considered the key players in extracellular

nucleotide hydrolysis.

6.1. Alkaline phosphatase (AP)

Alkaline phosphatases are ubiquitous enzymes present in many organisms from bacteria

to humans. As their name implies, APs have a pH optimum for hydrolysis in the alkaline range,

from pH 8 to 11 (98). With few exceptions, these enzymes are homodimeric and each catalytic

site contains three metal ions (two and one magnesium) necessary for enzymatic activity

(98). Three isoenzymes are tissue-specific (intestinal AP, placental AP, and germ cell AP) with

90-98% homology. The fourth isoenzyme, tissue-nonspecific alkaline phosphatase (TNAP) is only ~50% identical in sequence to the other three isoenzymes, and is abundantly expressed in bone, liver, kidney, and at lower levels in other tissues (99). Mammalian APs display broad substrate specificity towards different phosphomonoesters and other phosphorylated compounds, including various fatty acid chains, inorganic polyphosphates, glucose-phosphates, β-

21 1

345 ATP ADP AMP Ado Ino Hyp

2 2

Figure 8. Major extracellular nucleotide hydrolyzing enzymes. There are five major

families of enzymes that hydrolyze extracellular nucleotides and nucleosides: 1. E-NPP: ATP Æ

AMP + PPi; 2. E-NTPDase: ATP Æ ADP + Pi and ADP Æ AMP + Pi; 3. Ecto-5’-nucleotidase:

AMP Æ adenosine (Ado) + Pi; 4. Adenosine deaminase: Adenosine Æ Inosine (Ino); and 5.

Purine nucleoside phosphohydrolase (PNP): Inosine Æ hypoxanthine (Hyp). In addition,

alkaline phosphatase (AP) is able to catalyze the hydrolysis of all three phosphates of ATP, resulting in adenosine production.

22 glycerophosphate, and bis(p-nitrophenyl) phosphate, catalyzing release of inorganic phosphate from these varied substrates (98).

6.2. Ecto-5’-nucleotidase/CD73

Currently, there are seven human 5’-nucleotidases isolated and characterized. Five of the enzymes are located in the cytosol, one in the mitochondrial matrix, and one attached to the outer

plasma membrane. This latter enzyme, known as ecto-5’-nucleotidase, is expressed to a variable extent in different tissues, with abundant expression in the colon, kidney, brain, liver, heart, and lung (100). Cell surface-associated ecto-5’-nucleotidase, also known as CD73, efficiently hydrolyses 5’-AMP with Km of 3-50 µM, and shows no activity towards either nucleoside 2’-

and 3’-monophosphates or nucleoside di- and triphosphates (100). The mammalian ecto-5’-

nucleotidase consists of two glycoprotein subunits that are tethered by non-covalent bonds (101), bind zinc and other divalent metal ions at the N-terminal domain (102), and is anchored to the plasma membrane at the C-terminus by phosphatidylinositol (PI) (100, 103). This ecto-enzyme belongs to a large superfamily of dinuclear metallophosphoesterases acting on diverse substrates such as Ser/Thr phosphoproteins, various nucleotides, and sphingomyelin (102).

6.3. Ecto-nucleotide phosphodiesterase/pyrophosphatase (E-NPP)

The E-NPP family of enzymes consists of seven structurally related ecto-enzymes (NPP1

– NPP7) that possess broad substrate specificity. They are capable of hydrolyzing

pyrophosphate and phosphodiester bonds in di-nucleotides, nucleic acids, nucleotide sugars, as

23 well as choline phosphate esters and lysophospholipids (104). Only the first three members of

this family, NPP1 (previously called plasma cell differentiation antigen-1, PC-1), NPP2

(autotoxin, aka phophodiesterase 1α), and NPP3 (gp130RB13-6, aka B10 or phosphodiesterase 1β),

are capable of hydrolyzing various nucleotides, and therefore relevant in the context of

purinergic signaling (105).

6.4. Nucleoside triphosphate diphosphohydrolases (NTPDases)

In 1999, at the “Second International Workshop on Ecto-ATPases and Related

Ectonucleotidases” held in Belgium, the participants established a unifying nomenclature for the

NTPDases, and other extracellular nucleotide hydrolyzing enzymes. The protein nomenclature

for nucleoside triphosphate diphosphohydrolase is “NTPDaseX”, where X is 1-8 (106) (Table 1).

The gene names for the NTPDases are “ENTPDX”, where X is 1-8, with the gene numbers corresponding to the protein numbers (Table 1).

There are eight members of the human NTPDase family of enzymes that hydrolyze extracellular and intra-luminal nucleotides in a divalent cation (Ca+2 and Mg+2) dependent

manner. There are six, membrane-bound enzymes (NTPDases 1-4, 7 and 8), whose carboxyl-

and amino-terminal ends are each anchored to the membrane (107). Two enzymes lack a

carboxyl-terminal transmembrane (TM) helix (NTPDase5 and NTPDase6), and can be secreted

as soluble enzymes after cleavage of their respective N-terminal signal sequences (108, 109).

NTPDase 1-3 and 8 reside on the cell surface, thereby modulating purinergic receptors and

nucleotide signaling.

24

Protein Name Previously used protein names Comments (Gene Name)

NTPDase1 CD39, ecto-ATP diphosphohydrolase, Cell-surface membrane protein, hydrolyzes ATP and (ENTPD1) ectoapyrase ADP almost equally well, maintains hemostasis

NTPDase2 CD39L1, ecto-ATPase Cell-surface membrane protein, hydrolyzes ATP much (ENTPD2) faster than ADP (10-50 X as fast)

NTPDase3 CD39L3, HB6 Cell-surface membrane protein, hydrolyzes ATP (ENTPD3) slightly faster than ADP (3-5 X as fast)

NTPDase4 Golgi UDPase, LALP70 Found in Golgi, and in lysosomal membranes, may (ENTPD4) function in glycosylation of proteins

NTPDase5 CD39L4, PCPH proto-oncogene Found in internal membranes and can be secreted as (ENTPD5) a soluble enzyme after cleavage of signal sequence

NTPDase6 CD39L2 Found in internal membranes and can be secreted as (ENTPD6) a soluble enzyme after cleavage of signal sequence

NTPDase7 LALP1 Very similar to NTPDase4 (ENTPD7)

NTPDase8 ATPDase Cell-surface membrane protein, hydrolyzes ATP (ENTPD8) (chicken liver, oviduct and stomach, slightly faster than ADP (2-3 X as fast), less sensitive and mouse liver) to detergent inactivation

Table 1. Members of the E-NTPDase Family. There are eight members of the human

NTPDase family of enzymes, NTPDase1-8. The protein nomenclature is "NTPDaseX", for

"ecto-nucleoside triphosphate diphosphohydrolase, where X is 1-8. The gene names for these enzymes are "ENTPDX", where X is 1-8, with the gene numbers corresponding to the protein numbers.

25 7. Molecular Cloning, Sequencing, Expression, and Characterization of the NTPDases

7.1. The membrane-bound, cell surface NTPDases (NTPDase1-3 and 8)

NTPDase1-3 and 8 are all membrane-bound, cell surface enzymes that hydrolyze extracellular nucleotides. NTPDase1/CD39 was originally identified as the major surface marker of EBV-transformed B lymphoblastoid cells, and later was found primarily on activated immune cells (110). In 1996, NTPDase1/CD39 became the first human gene reported to encode an

NTPDase (41). By transfection of COS-7 cells with NTPDase1 cDNA, Wang et al demonstrated the 510-amino acid, 70 kDa protein hydrolyzed ATP and ADP equally well in the presence of divalent cations (Ca+2 and Mg+2). In addition, its enzymatic activity was not inhibited by P- or

V-type ATPase inhibitors (41).

Hicks-Berger and Kirley expressed and characterized NTPDase2/CD39L1 in mammalian

COS-1 cells, demonstrating NTPDase2 is a ~66 kDa glycoprotein that hydrolyzes NTPs at a

ratio of 20:1 as compared to NDPs (111). The activity of NTPDase2 depends on its

glycosylation state (treatment with PNGase-F decreases enzyme activity) and also rapidly

decreases in activity with time during incubation at 37oC. To the contrary, the activity of

NTPDase2 is stimulated by lectin concanavalin A (Con A) and the chemical cross-linking agent disuccinimidyl suberate (DSS). Based on Northern blot hybridization, NTPDase2 is expressed in the heart, brain, placenta, skeletal muscle, kidney, and pancreas (112).

Smith and Kirley isolated the NTPDase3 (aka HB6 and CD39L3) clone from a human brain cDNA library and expressed the protein in mammalian COS-1 cells (12). NTPDase3 encodes a 529-amino acid glycoprotein, with a molecular weight of ~80 kDa. It has seven

26 putative N-glycosylation sites, one cAMP- or cGMP-dependent protein kinase phosphorylation

site, four protein kinase C (PKC) phosphorylation sites, and two casein kinase phosphorylation

sites. The nucleotide hydrolysis ratio of the expressed protein is approximately 3:1

(ATPase:ADPase activity). Based on Northern blot hybridization, NTPDase3 is highly

expressed in the brain, pancreas, spleen, and prostate, with moderate to low expression in other

tissues (112). Similar to other NTPDases, NTPDase3 is sensitive to inhibition by millimolar

concentrations of sodium azide (NaN3).

NTPDase8 is the newest member of the NTPDase family that was cloned from human

liver RNA by real time-polymerase chain reaction (RT-PCR) and expressed in HEK293 cells

(113). It is a 495-amino acid protein that is sensitive to detergent inhibition by Triton X-100 and

NP-40. Consistent with many other membrane-bound NTPDases, NTPDase3 preferentially

hydrolyzes ATP over ADP, and has higher enzyme activity in the presence of Mg+2 as compared

to Ca+2.

7.2. The intraluminal NTPDases (NTPDase4-7)

NTPDase4 and 7 are intracellular, membrane-bound proteins found in the Golgi and

lysosomal membranes. NTPDase4 was cloned from a human brain cDNA library and expressed

in mammalian COS-7 cells (114). It encodes a 610-amino acid protein with two transmembrane

domains (TMDs). The enzyme preferentially hydrolyzes uridine 5’-diphosphate (UDP) as a

substrate. It is stimulated by divalent cations in the following order: Ca+2 >> Mg+2 > Mn+2. Due to its localization, it is suggested NTPDase4 may play a role in glycosylation of proteins and

27 lipids by hydrolyzing UDP, an inhibitor of galactosyltransferase, decreasing its accumulation in

the Golgi apparatus (Figure 9).

NTPDase7/LALP1 was cloned from human brain and expressed in HEK293 cells (115).

It encodes a 604-amino acid protein with a molecular weight of approximately 70 kDa.

NTPDase7 preferentially hydrolyzes NTPs (UTP, GTP, and CTP) and may play a role in the

salvage of nucleotides from the lysosomal/autophagic vacuole lumen.

NTPDase5 and 6 are found in intracellular membrane organelles, but can also be secreted

as soluble enzymes upon cleavage of their respective N-terminal signal sequences (108, 109).

Therefore, these enzymes can hydrolyze both extracellular and intraluminal nucleotides.

NTPDase5 was cloned from a human brain cDNA library as the first reported human, secreted

NTPDase. It was also expressed in bacteria and refolded from inclusion bodies. The soluble form of the enzyme is 410-amino acids long with a molecular weight of 45.5 kDa. It prefers the hydrolysis of NDPs (specifically GDP) over NTPs, and Mg+2 over Ca+2 as the divalent cation

supporting nucleotide hydrolysis. NTPDase6 was expressed in mammalian COS-1 cells where

90% of the enzyme activity was found in the cell media, indicating the presence of a soluble

protein. NTPDase6 is released as a 50 kDa soluble protein, resulting from cleavage of its N-

terminal sequence after Ala-78. It has four cysteine residues that are conserved among the other

NTPDases, which presumably form two disulfide bonds (109). In addition, NTPDase6 has two

potential N-glycosylation sites (glycosylation not necessary for activity) and prefers the

hydrolysis of NDPs (specifically GDP and IDP) over NTPs.

28

Cytoplasm

ER/Golgi Lumen

Folded protein NDPase (NTPDase 4, 5 or 6) UDP UMP UMP ( - Antiporter ) Glucose Transferase + UDP-glucose UDP-glucose Folding protein

Folding protein

Chaperones ATP AMP AMP Misfolded protein ADP Protein Synthesis ATP

Figure 9. Intracellular nucleotidase control of protein folding and processing? Intraluminal

nucleotides are hydrolyzed by NTPDase4, 5, and 6. These nucleotidase, which preferentially

hydrolyze UDP and GDP, are located in the lumen of the ER and Golgi. After protein synthesis

on ribosomes, each protein must be folded correctly. Chaperone proteins, which require ATP, are expressed in the lumen of the ER and Golgi to assist in this process. During the folding

process, many proteins must be glycosylated in the Golgi to be properly processed. NTPDase4,

5, and 6 hydrolyze UDP to UMP. Accumulation of UDP present in the lumen of the ER and

Golgi causes a feedback mechanism that inhibits glucose transferase, the enzyme necessary for

adding glucose molecules to proteins. Thus, intracellular NTPDases regulate the amount of

intraluminal UDP present, likely modulating glycosylation of proteins.

29 8. The Biochemical and Structural Characterization of the NTPDases

8.1. Enzymology of the NTPDases

The NTPDases are the key players in extracellular nucleotide hydrolysis to control the

level of these molecules at the cell surface. ATP, ADP, UTP, and UDP are the biologically most

relevant substrates of the NTPDases, which initiate a number of cellular responses via the

activation of ionotropic P2X and metabotropic P2Y receptors. Therefore, the NTPDases regulate

many physiological processes by altering and controlling the concentration of extracellular

nucleotide agonist at P2 (and P1) receptors.

The NTPDases exhibit E-type ATPase activity, meaning: their nucleotidase activity is

+2 dependent upon millimolar Ca or Mg+2; they are insensitive to specific inhibitors of P-type, F-

type, and V-type ATPases (vanadate, ouabain, and oligomycin, respectively); and sensitive to

millimolar NaN3 inhibition of enzyme activity (116). Typically, the NTPDases have a low

specificity for the hydrolysis of different nucleoside tri- and di-phosphates, but nucleoside

monophosphates and non-nucleoside phosphates are not substrates. The NTPDases also

hydrolyze nucleotides over a broad range of pH, and do not form phosphorylated intermediates during catalysis. In addition, the NTPDases are glycoproteins whose enzyme activity is stimulated by lectins. They are also relatively insensitive to inactivation by proteases, but most are inactivated by many biological detergents, being especially sensitive to Triton X-100 inactivation. Currently, there is a lack of potent and specific inhibitors for both the family of enzymes and for the individual NTPDase isoenzymes.

30 Four out of the eight members of the NTPDase family (NTPDase1-3 and 8) are relevant

to the control of P2 receptor signaling. These enzymes are localized at the surface of the plasma

membrane and hydrolyze nucleotides in the range of concentrations that activate P2 receptors.

Considering each of the membrane-bound, cell surface NTPDases (NTPDase1-3 and 8) were

partially characterized, expressed in different laboratories, and under different conditions,

comparison of the biochemical properties of these enzymes was complicated until Kukulski et al

expressed them all together (117) (Table 2). Even though NTPDase5 (108) and NTPDase6 (109)

can be present at the surface of the plasma membrane and secreted as soluble enzymes following

a proteolytic cleavage, their high Km values and relatively low specific activities make it unlikely

that these enzymes regulate P2 receptor signaling (in comparison to purified, full-length rabbit

NTPDase2 (118), which has a specific activity of approximately 400,000 µmol/mg/hr).

Therefore NTPDase5 and NTPDase6 were not included in the comparative analysis shown in

Table 2.

8.2. Inhibitors of the NTPDases

Modulation of nucleotide signaling has mainly focused on identifying inhibitors

(antagonists) for the purinergic receptors. Over the last decade, the NTPDase family of enzymes

has emerged as a potential new pharmacological target. It is evident that inhibition or

modulation of the NTPDases may prolong the effect of nucleotides at their respective receptors.

Currently, nucleotide analogues (particularly P2 receptor agonists/antagonists) and molecules related to the -Evans blue dye families are being used to inhibit the NTPDases. Gendron

31

max V UDP

m M) (U/mg prot.) (U/mg M) ND ND K µ 67 67 ± 3 ± 0.17 0.01 135 ± 10 ±0.79 0.04 ( 1780 ± 40 ± 1.37 0.08

max V

UTP

m M) (U/mg prot.) (U/mg M) K µ ( 58 ±6 0.57 ±0.03 0.57 ±6 58 47 ± 4 1.05 ± ±47 1.05 0.03 4 393 ± 30 ±3.9 0.2 124 ± 10 ±1.17 0.06

max V ADP

m M) (U/mg prot.) (U/mg M) K ND ND ND µ ( 31 31 ± 1 0.18 ± 0.01 22 22 ± 1 0.75 ± 0.01 265± 20 ±0.61 0.03 max V ATP m M) (U/mg prot.) (U/mg M) µ K ( 75 ± 0.79 10 ± 0.03 70 ± 2 2.3 ± 0.03 17 ±17 ± 0.94 1 0.02 46 ±46 ± 0.74 5 0.04 7 – 10 4.5 –4.5 11 4.5 –4.5 8.5 4.5 –4.5 8.5 pH range Enzyme activity > 50% activity ~2:1 ADP ATP: Ratio ~3-4:1 ~1-2:1 ATPase Apyrase Apyrase Apyrase 10 – 40:1 +2 +2 +2 +2 = Mg = Mg > Mg > Mg +2 +2 +2 +2 Ca Ca Ca Ca Divalent Cation Preference NTPDase hNTPDase1 hNTPDase2 hNTPDase3 rNTPDase8

32 Table 2. Enzymatic characterization of membrane-bound, cell surface human NTPDase1-3 and rat 8 (from reference (119)). Reactions were carried out for 10 minutes in the presence of

5 mM CaCl2 and 80 mM Tris buffer pH 7.4. In all assays, less than 10% of the substrate was hydrolyzed. Km and Vmax values were estimated with nucleotide concentrations ranging from

0.02 to 1.5 mM (119). The results are expressed as the mean ± SEM of three separate experiments, each performed in triplicate.

33 et al described some inhibitory nucleotide analogue compounds (120). Brunschweiger et al

have synthesized a nucleotide analogue derived from uridine-5’-carboxamide that is the first

potent and selective human NTPDase2 inhibitor (121). However, FPL 67156 (6-N,N-diethyl-

β,γ-dibromomethylene-D-ATP) was the first ATP analogue described as a novel, selective

inhibitor of the NTPDases with some 50-fold selectivity over the purinergic receptors (122).

More recently, ARL 67156 (the only commercially available inhibitor of the NTPDases; aka FPL

67156) was shown to be a weak competitive inhibitor of human NTPDase1 (Ki = 11 ± 3 µM) and

NTPDase3 (Ki = 18 ± 4 µM) (123). However, human NTPDase2 and NTPDase8 were less

efficiently inhibited by ARL 67156. At high concentrations of substrate, ARL 67156 was not

able to efficiently inhibit any of the NTPDases, and high concentrations of ARL 67156 were

shown to inhibit the P2Y2, P2Y4, P2Y12, and P2X receptors. Overall, inhibitory nucleotide analogues face the seemingly insurmountable obstacle of the high likelihood of affecting other nucleotide handling enzymes and purinergic receptors as well as their intended NTPDase target.

Non-nucleotide NTPDase inhibitors also exists. Our laboratory was the first to discover that several P2 receptor antagonists such as suramin, Cibacron Blue, and Coomassie Blue G-250 inhibited a purified NTPDase (124). Three structure-activity requirements were found for this class of inhibitors: 1. the position of the sulfonate groups on the naphthalene core is important

(the 2-sulfonate group was found to be required for inhibitory activity); 2. the amino-hydroxy- naphthalene disulfonate structures of these antagonists are involved in the binding to the

NTPDases; and 3. the inhibitory potencies and affinities of these compounds acting as antagonists at P2 receptors increase with size (125).

Polyoxometalates hold some promise as small molecule NTPDase inhibitors (64), but are not yet sufficiently potent or specific to be truly useful. A recent publication (126) gives rise to

34 hope that specific and fairly potent small molecule inhibitors for NTPDases might be possible.

One inhibitor (compound #21) had a sub-micromolar IC50 for rat NTPDase1 and another very

intriguing compound (#20) seems to be specific for NTPDase3 over NTPDase1 and NTPDase2,

at least for the rat enzymes (126). However, this compound does not show the same specificity

for human NTPDase3 (personal communication from the senior author of that published study,

Dr. C.E. Muller, and confirmed by unpublished experiments performed in our laboratory).

Lastly, Munkonda et al are the first to report inhibitory monoclonal antibodies specific

for human NTPDase3 that inhibit the enzyme activity by 60 – 70% (127). In this study,

Munkonda et al demonstrates the importance of the native conformation of NTPDase3 for

epitope recognition as denaturation of the protein with SDS or reducing agents, indication

abolished binding of the monoclonal antibodies. Additional data demonstrated the SDS-resistant

parts of the inhibitory epitope were located in two fragments of the C-terminal lobe (i.e. Leu220-

Cys347 and Cys347-Pro485) and site-directed mutagenesis revealed the importance of Ser297 and the fifth disulfide bond (Cys399-Cys422) for antibody binding. As yet unpublished epitope localization data has further pinpointed the amino acids critical for antibody binding. This could serve as a springboard for development of antibody inhibitors of other NTPDases, if like

NTPDase3, each NTPDase has a specific epitope for antibody binding, and consequent inhibition.

Regardless, the potential therapeutic outcomes of specific NTPDase inhibition are substantial, as delineated by Gendron et al (120, 128). It is evident the ideal NTPDase inhibitor should not be a P2 receptor agonist/antagonist and should not be dephosphorylated by the enzyme. Unfortunately, the use of non-specific purinergic receptor and NTPDase inhibitors has led to the misinterpretation of some purinergic signaling data; specifically, which purinergic

35 receptors are involved in the many and varied physiological responses to exogenously added nucleotides.

8.3. Site-directed mutagenesis and computational comparative protein modeling of the

membrane-bound, cell-surface NTPDases

Site-directed mutagenesis and computational comparative protein modeling are useful tools to characterize and elucidate key structural features of the NTPDases. The general topography of the membrane-bound, cell surface NTPDases resemble a variety of channel forming proteins such as the P2X receptor or members of the epithelial Na+ channel (129). In addition, cross-linking experiments have indicated NTPDase1-3 form homo-oligomeric complexes starting with dimers. Rat and human NTPDase1 were found to form up to tetrameric oligomers (130, 131), whereas chicken NTPDase2 (132) and human NTPDase3 (133) preferentially form dimers. However, no clear picture has emerged regarding how the N- and C- terminal TM helices interact within and between monomers to form the native oligomeric complexes of the NTPDases.

As shown in Figure 10, membrane-bound, cell surface NTPDases can be divided into three major structural regions: a large extracellular domain, a “linker region”, and a transmembrane domain (TMD) (Figure 10). The extracellular region contains the enzyme active site, five apyrase conserved regions (ACRs), and five conserved disulfide bonds (107).

36 472 476 53 481 P52 484 P Extracellular 485

Cell Membrane

Intracellular NH COOH 2

Figure 10. The three structural regions of the NTPDases. Using NTPDase3 as an example, the three structural regions of an NTPDases monomer are shown. The extracellular domain is a group of multi-colored alpha helices and beta sheets. The “linker” region is depicted as red (C- terminal) and blue (N-terminal) lines labeled with the conserved proline residues of NTPDase3.

The TM helices are depicted as yellow cylinders

37 Through extensive site-directed mutagenesis of human NTPDase3, many of the structural and functional elements of the extracellular domain of the NTPDases are defined (134). The “linker region” is a stretch of amino acids in the primary sequence of NTPDase3 that connects the TMD

(the N- and C-terminal TM helices) and extracellular domain (location of the enzyme active site), and contains several conserved proline residues (N-terminal: P52 and P53; C-terminal:

P472, P476, P481, P484, P485). Interestingly, proline residues confer unique structural rigidity needed for some conformational changes in other proteins, and may facilitate the cross-talk between the extracellular and TM domains of NTPDase3. Lastly, the TMD is thought to mediate substrate specificity (135), sensitivity to membrane perturbation (136, 137), and oligomerization of the membrane-bound, cell surface NTPDases.

8.3.1. The structural features of the extracellular domain of the NTPDases

The extracellular domain contains five ACRs that share nearly identical sequence and spatial conservation with related members of the NTPDase family (139). Based on the conservation of the amino acids within the ACRs of the E-type ATPases (specifically two

“DXG” phosphate binding motifs) with the phosphate binding motifs of the actin/heat shock protein/sugar kinase superfamily of proteins, ACR1 and ACR4 were identified as phosphate binding domains in the extracellular domain, suggesting their role in nucleotide binding and subsequent hydrolysis (139). The importance of these residues in ACR1 and ACR4 was established by mutation of the aspartic acids and , which resulted in the inactivation of

NTPDase3 without a change in the protein expression level (140). There are also two highly conserved tryptophan residues (W187 and W459), which are the only two tryptophan residues

38 conserved in the extracellular domain of nearly all the NTPDases (141). Alanine substitution at

W187 (located in ACR3) resulted in inactivation of NTPDase3 and adversely affected the

cellular localization of the enzyme. Data supported the possibility of altered tertiary structure,

causing NTPDase3 to be improperly folded and retained within the cell. To the contrary, alanine

substitution at W459 (located in ACR5) resulted in an increased ATPase:ADPase ratio (an enhanced NTPase activity, but diminished NDPase activity). The W459A mutant was coupled with a previously described mutant, D219E (located in DXG2/ACR55), which together nearly converted NTPDase3 to an NTPase. In addition, the double mutant (D219E/W459A) was substantially less sensitive to inhibition by NaN3, a more potent inhibitor of ecto-apyrases than ecto-ATPases. This data indicated that a major change in nucleotide tri- and di-phosphate specificity of an NTPDase is possible with only a few changes in the amino acid sequence.

In actin, the conserved glycine residues are very important for its structure and function.

Kirley et al mutated the most highly conserved glycine residues (excluding the two DXG

glycines previously mutated in ACR1 and ACR4) to alanine in the extracellular domain of

NTPDase3 to determine their structural and functional significance (142). Alanine substitution

at G98 (in ACR1a) resulted in an approximately 60% decrease in enzyme activity. Alanine

substitution at G263 (in ACR4a) resulted in an inactive protein with severe folding/processing

problems. Alanine substitution at G462 (in ACR5) resulted in an approximately 85% decrease in

enzymatic activity with some enzyme folding and processing problems. This data led to the

identification of two additional ACR regions, ACR1a and ACR4a, in the extracellular domain of

NTPDase3.

The NTPDases are also heavily glycosylated in their extracellular domain. To determine

the role of glycosylation at the most conserved glycosylation site of the NTPDases (located

39 between ACR1 and ACR1a), N81 and T83 were mutated in NTPDase3 (143). Mutation of either

the asparagine residue (N81D or N81E) where the sugars are attached, or the threonine residue

(T83A) that is part of the consensus sequence for glycosylation, resulted in a decreased

nucleotidase activity (more severely affecting Mg-ATPase than Ca-ATPase activity). In addition, the inhibitory affect of treatment with a deglycosylation enzyme (PNGase-F) as well as

stimulation of activity by concanavalin A (ConA) was attenuated. This data suggests that

glycosylation is important for the activity and may maintain the oligomeric structure of

NTPDase3 and other membrane-bound, cell surface NTPDases, since glycosylation is not

important for the activity of the soluble, monomeric NTPDases.

It is known that the membrane-bound, cell surface NTPDases form dimers and higher

order oligomers (144). Lysine 79 is conserved in all cell surface NTPDases (NTPDase1-3 and

8), but not in the soluble (NTPDase5 and 6) and intracellular (NTPDase4 and 7) enzymes. In human NTPDase3, all substitutions at this site resulted in a loss of 70 – 80% of activity (145).

Only K79R, which maintains the positive charge at this residue, remained able to be stimulated/stabilized by Con A. Interestingly, the residual activity of all the mutants at the K79 site became resistant to Triton X-100 detergent inhibition. Since Triton X-100 is known to dissociate NTPDase oligomers into monomers, this data suggests K79 may be involved in the dimer interface of the membrane-bound, cell surface NTPDases, partially mediated by the extracellular domain.

Recently, Ivanenkov et al determined the ten conserved cysteine residues in the

extracellular domain of all membrane-bound, cell surface NTPDases form five disulfide bonds.

When these disulfide bonds are disrupted in NTPDase3, the protein demonstrates five distinct

biochemical phenotypes (107). This data strongly suggests disulfide bonds between C92-C116,

40 C261-C308, C289-C334, C347-C353, and C399-C422. Due to the conservation of the 10 cysteines, it is likely the disulfide bonds exist in all cell surface NTPDases. The third (C289-

C334) and fifth (C399-C422) disulfide bonds are also present in the soluble NTPDases and are

critical for biosynthetic processing and function. In addition, computational comparative protein

modeling of the NTPDase3 sequence revealed a high degree of structural fold similarity with a

bacterial exopolyphosphatase (PDB 1T6C). This established a 3-D model of the extracellular

portion of NTPDase3, which is consistent with the assignment of several of the disulfide bonds.

A recent crystal structure of the extracellular region of rat NTPDase2 (146) confirmed almost all the site-directed mutagenesis conclusions using NTPDase3, and forms the basis for the much more accurate 3-D models for the extracellular domain of NTPDase3 and other NTPDases currently in use.

8.3.2. The structural features of the transmembrane domain of the NTPDases

.

Many ecto-proteins are anchored to the cell membrane by a single transmembrane domain (TMD) or a lipid link. One of the most intriguing features of the membrane-bound, cell surface NTPDases is the presence and function of its two TM helices, one at both the N- and C- terminal of the enzymes. Most studies regarding the role of the TMD have focused on oligomerization. The predominant, active form of these enzymes appear to be a dimer, although higher-order oligomers are also observed, and the soluble NTPDases (which lack a TMD) are monomeric (131, 133, 144). However, no structural data exists and very limited site-directed mutagenesis studies were performed to perturb and analyze helix interactions in the TMD of the

NTPDases.

41 Wang et al deleted the amino terminal, carboxyl terminal, or both TM helices of

NTPDase1/CD39 and compared the activity of each truncated protein with native NTPDase1

(131). Their goal was to determine whether the TM helices are required for enzymatic activity, and whether disruption of the interactions between the helices is the mechanism for detergent inhibition effects. Truncation of both TM helices resulted in a catalytically active, secreted, soluble protein with an enzyme activity ~90% lower than native NTPDase1. Deletion of either

TM helix resulted in a catalytically active enzyme that was properly processed and trafficked to the cell surface, but both mutant proteins had reduced enzymatic activities similar to the double truncation mutant. Wang et al further proposed that not only the presence and membrane insertion of the TM helices, but also the interactions between them are required for optimal activity (131). Glutaraldehyde cross-linking and sucrose density gradient sedimentation in the presence of detergents showed that NTPDase1 formed oligomers. In addition, detergents that reduced the enzyme activity of NTPDase1 dissociated the oligomeric structure (i.e. Triton X-

100), and that the soluble extracellular domain lacking both TM helices existed only as monomers. One detergent, digitonin, had only a small effect on enzyme activity and preserved the oligomers. Taken together, this data suggested that both TM helices are necessary for full enzymatic activity and they may mediate oligomer formation via TM helix interactions.

Murphy et al have conducted one of the few studies to date that use site-directed mutagenesis to determine the function of a specific set of amino acids in the TMD of NTPDase3

(147). Cysteine to serine substitution was used to test the functional and structural significance of the three non-extracellular cysteine residues (C10, located on the short N-terminal cytoplasmic tail; and C501 and C509, located in the C-terminal TM helix of NTPDase3). None of these cysteines were found to be essential for enzymatic activity. However, cysteine 10 was

42 found to be responsible for dimer formation occurring via oxidation during membrane preparation, as well as for dimer cross-linking resulting from exogenously added sulfhydryl- specific cross-linking agents. Cysteine 501 was found to be the site of chemical modification by

a sulfhydryl-specific reagent, p-chloromercuriphenylsulfonic acid (pCMPS), leading to inhibition

of enzyme activity. The latter result was further confirmed by Chiang et al using NTPDase2, in

which pCMPS also decreases its enzymatic activity via cysteine modification in the TMD (148).

In 2004, Grinthal and Guidotti used a cysteine substitution/copper phenanthroline

disulfide bond-induced cross-linking strategy to probe for possible intra- and intermolecular interactions in the TMD of NTPDase1 (149). They demonstrated that the TM helices of rat

NTPDase1 have strong intra- and inter-molecular interactions between the N- and C-terminal

TM helices, and that these interactions are stronger near the extracellular side of the membrane.

In addition, they showed that the TM helices undergo dynamic motions that are necessary for

maximal enzymatic activity, and that these TM interactions and motions are regulated by

nucleotide binding at the extracellular active site in NTPDase1 (149). Lastly, Chiang et al (148)

using human NTPDase2 and our laboratory using human NTPDase3 (138) also noted a similar

reduction in enzymatic activity of these NTPDases upon oxidative cross-linking of introduced

cysteine residues in their respective TM helices. These data suggest the TM helices of the

membrane-bound, cell surface NTPDases must move or re-orient relative to one another during

nucleotide hydrolysis to facilitate optimal enzymatic activity.

43 9. Functional Significance of the NTPDases

The physiological and pathophysiological functions of the NTPDases are currently under

investigation. Many putative functions have been suggested. Data has implicated a role for the

NTPDases in secretion (150), cell adhesion (151), cancer and malignant transformation (152,

153), pain perception, purinergic receptor system modulation, adenosine recycling, and platelet

aggregation (for a review, see (116)).

Strong evidence was presented that NTPDase1, found in vascular endothelial cells, is one of the three important endothelial thromboregulatory systems that inhibit platelet aggregation in response to agonists, along with nitric oxide (NO) and the eicosanoids (154, 155). This proposed function is supported by the generation of NTPDase1 null mice having defects in blood clotting and related processes (156, 157).

NTPDase2 is expressed in the germinal zones of the brain and suggested to be important for the development of certain areas of the brain (158). It is also expressed in the glial cells of the peripheral nervous system (159). In addition, NTPDase2 is expressed in smooth muscle, where contraction is under the control of ATP via purinergic signaling. More recently,

NTPDase2 was shown to play a critical role in the development of the eye, since over-expression of NTPDase2 in the frog caused ectopic eye-like structures to form, and occasionally even complete duplication of the eye (160).

NTPDase3 is expressed at a relatively high level in brain (112). Immunolocalization using three different anti-peptide antibodies demonstrated NTPDase3 localization in rat brain is highly restricted to axons and associated with synapse-like structures, a pattern that strongly suggests NTPDase3 acts as a regulator of synaptic function (161). The high density of

44 NTPDase3 positive terminals on noradrenergic neurons suggests that NTPDase3 may play an

important role in regulation of noradrenergic autonomic function. Moreover, very high densities

of NTPDase3 were detected in the mediobasal hypothalamus, with the overall mesencephalic

pattern of staining associated closely with hormone responsive brain regions. This pattern of expression, in addition to the observed co-localization with hypocretin-1/orexin-A, suggests that

NTPDase3, by regulating the extracellular turnover of ATP, may modulate feeding, sleep-wake, and other behaviors through diverse homeostatic systems. Appelbaum et al (162), using zebra fish, confirmed our laboratory’s localization of NTPDase3 in the brain to the hypothalamus.

They suggested that NTPDase3 in this location functions in conjunction with P2X8 receptors. In

addition, they also found expression of NTPDase3 in primary sensory neurons in the spinal cord,

where it is likely involved in ATP-mediated pain transmission in conjunction with P2X2 and/or

P2X3 receptors.

Most likely, NTPDase3 has other physiological functions outside the nervous system.

Interestingly, "virtual Northern" data (available via

https://escience.invitrogen.com/ipath/search.do) suggests that pancreatic beta cells, responsible

for insulin production, may have the highest expression level of NTPDase3 of any tissue/cell type. Consistent with this, our laboratory has confirmed the presence of NTPDase3 protein in these cells by Western blot and immunoprecipitation (unpublished results). We have also demonstrated that NTPDase3 endogenously expressed in human pancreatic beta cells can be inhibited in situ by a monoclonal antibody specific for human NTPDase3 (127). Since insulin secretion is modulated by purinergic signaling, this suggests that regulation of NTPDase3 may be of clinical significance for the development and treatment of diabetes.

45 Lastly, soluble NTPDase6 was reported to be co-released with ATP in some mammalian sympathetic nerves (163, 164). Therefore, soluble NTPDase6 may be involved in terminating the effects of the co-released ATP neurotransmitter.

10. Dissertation Scope and Objectives

Site-directed mutagenesis and computational comparative protein modeling have served as useful tools to formulate testable hypotheses, generate models for the enzyme active site, and to elucidate the structural elements that modulate the enzymatic properties of the NTPDase family of enzymes. The recently published crystal structure of the extracellular domain of rat

NTPDase2 (146) confirmed and extended much of the site-directed mutagenesis work done by our laboratory using human NTPDase3 (134). Now it seems the major structural and functional features of the extracellular active site of the NTPDases are fairly well defined. However, this crystal structure does not tell the complete story of the structural elements that regulate the enzymatic mechanism of the NTPDases. The expressed and refolded extracellular domain of rat

NTPDase2 that was crystallized is monomeric. However, the membrane-bound, cell surface

NTPDase enzymes are oligomeric, and the oligomeric state effects enzyme activity. Our laboratory has determined that the membrane-bound, cell surface NTPDases can be divided into three main domains: 1. the extracellular domain, consisting of two lobes between which is the location of the enzymatic active site; 2. the “linker region”, a stretch of amino acids in the primary sequence of NTPDase3, between the TM and extracellular domain; and 3. the TMD, containing the N- and C-terminal TM helices and short cytoplasmic tails (Figure 10).

Questions regarding the structural element(s) that facilitate the cross-talk between the

46 extracellular and transmembrane domains during nucleotide hydrolysis as well as the orientation

and/or interactions of the TM helices are still unanswered.

The objectives of this dissertation are: 1. to better understand the structural elements that mediate the intra-protein signal transduction mediated by substrate binding at the extracellular domain, causing movements and/or re-orientations of the TM helices during nucleotide hydrolysis, and 2. to determine how the N- and C-terminal TM helices interact in human

NTPDase3 to form the native, dimeric enzyme. The rationale for these studies and the specific objectives are described below:

A) The “linker region” of NTPDase3 joins the N- and C-terminal lobes of the extracellular domain to the N- and C-terminal helices in the TMD, respectively. Several conserved proline residues (N-terminal: P52 and P53; C-terminal: P472, P476, P481, P484, and P485) are located in the linker region of NTPDase3. Due to the unique structure of proline and its involvement in domain rearrangements in other proteins, we investigated the role of the conserved proline residues in mediating the linkage of substrate binding in the extracellular domain to movements of the TM helices during nucleotide hydrolysis in human NTPDase3.

B) Polar residues are known to drive the association of TM helices and have effects on protein processing and function. NTPDase3 contains several conserved polar residues (N-terminal: S33,

S39, T41, and Q44; C-terminal: T490, T495, and C501) in its TMD. Therefore, we investigated the role(s) of these conserved polar residues in the TMD of NTPDase3, and strove to understand how the N- and C-terminal TM helices interact within and/or between monomers, in order to

47 enhance the knowledge of NTPDase function and its modulation by oligomerization and

membrane perturbation.

C) The existence of alternatively spliced variants is a common phenomenon among enzymes,

including some NTPDases. The cDNA encoding a truncated variant of NTPDase3, lacking the

C-terminal TM helix, was found in a human lung library by polymerase chain reaction (PCR).

We expressed and characterized the enzymatic properties of this alternatively spliced variant to

test the hypothesis that truncation would result in an enzyme with diminished enzymatic

properties as compared to full-length NTPDase3.

D) Single nucleotide polymorphisms (SNPs) can alter the primary structure and cause a distinct

biochemical phenotype, which may be physiologically important. Human NTPDase3 has five

known SNPs, namely: I24V, R264Q, A496V, L505F, and E440D. Three of these SNPs (I24V,

A496V, and L505F) are located in the TMD, and R264Q is located in the enzymatic active site.

We expressed these mutant proteins to test our hypothesis that one or more of these SNPs would

substantially modify the enzymology, and therefore possibly the function, of human NTPDase3.

48 Chapter II: Materials and Methods

1. General Materials and Methods

1.1. Purification of oligonucleotides for site-directed mutagenesis

350 µL TE Buffer was added to the commercial custom synthetic oligonucleotides in an eppendorf tube, followed by vortexing and incubating in a 37°C water bath for 15 minutes. The samples were vortexed again, and then sonicated in a water bath sonicator for 5 minutes. After centrifugation of the samples for 3 minutes at maximum speed, the supernatants were transferred into a clean eppendorf tube. 35 µL of 3 M NaOAc pH 7.0 was added to each, the samples vortexed after adding 900 µL of 100% and mixing, and the samples were incubated at -

20°C for 75 minutes. Afterwards, the samples were centrifuged for 5 minutes at maximum speed in a microcentrifuge at room temperature (22oC). The supernatants were carefully removed by

pipetting, and the pellets were gently washed with 1 mL 70% ethanol, followed by centrifugation

for 3 minutes. The supernatants were discarded, the pellets were centrifuged again for 3 minutes,

the rest of the ethanol was removed, and the pellets were air dried for 30 minutes. The

oligonucleotide pellets were then dissolved in 200 µL 1X TE buffer. To calculate the

oligonucleotide concentrations, 20 µL of each stock solution was diluted with 730 µL 1X TE buffer and the OD260 was measured using a Beckman DU-800 spectrophotometer (1 OD260 = 35

µg/µL oligonucleotide). The oligonucleotide solutions were stored at -20°C.

49 1.2. Site-directed mutagenesis of NTPDase3

Mutagenesis of NTPDase3 in the mammalian pcDNA3 vector was performed using a

QuikChange Site-Directed Mutagenesis kit (Stratagene). 10 ng of template DNA (encoding

NTPDase3) was added to 5 µL 10X buffer, 125 ng sense oligonucleotide, 125 ng anti-sense

oligonucleotide, and 1 µL dNTP mix and brought to a total volume of 50 µL with MilliQ H2O in a PCR tube. The solution was gently mixed and 1 µL PfuTurbo DNA polymerase added. For a control sample, template DNA was omitted. The samples were placed in a Techne TC-3000

Thermocycler for 16 cycles of PCR amplification: 30 seconds at 95°C, 1 minute at 55°C, and 10 minutes at 68°C. The samples were cooled to room temperature (22oC) and the parental DNA

digested with 1 µL DpnI at 37°C for 1 hour.

1.3. Bacterial transformation and inoculation of PCR products

A 15 mL Falcon polystyrene round-bottom tube (Fisher) was pre-chilled on ice for 5

minutes. 50 µL XL1-Blue Supercompetent Cells (Stratagene) and 1 µL PCR product was added

to the tube, swirled gently, and incubated on ice for 30 minutes. The bacteria were heat shocked

in a 42°C water bath for 30 – 45 seconds, incubated on ice for 2 minutes, and 500 µL of LB

media (preheated to 42°C) was added and the tube incubated in a 37°C air shaker for 1 hour.

100 µL of the sample was spread onto an LB-carbenicillin plate, incubated at room temperature

for 10 minutes, and then incubated overnight at 37°C upside down. The following day, several

bacterial colonies were picked from the plate and added to 5 mL LB containing 5 uL (100 ng/uL)

50 carbenicillin in a 15 mL Falcon polystyrene round-bottom tube. The colonies were grown overnight in a 37°C air shaker for plasmid cDNA isolation the following day.

1.4. Isolation of DNA from XL1-Blue Supercompetent Cells

Isolation of plasmid cDNA from the bacterial XL1-Blue Supercompetent Cells

(Stratagene) was performed using a QIAprep Spin Miniprep kit (Qiagen). From the 5 mL overnight sample of LB and bacteria, 1.5 mL was placed into an eppendorf tube, microcentrifuged in a Marathon micro A (Fisher) at maximum speed for 3 minutes, and the supernatant discarded. This was repeated twice, using the same liquid culture and the same tube.

250 µL of Buffer P1 was added to re-suspend the cells. 250 µL Buffer P2 was added to lyse the cells and the eppendorf tube was gently inverted 4-6 times to mix. 350 µL Buffer N3 was added to neutralize the reaction and the eppendorf tube was immediately, but gently, inverted 4-6 times.

The samples were centrifuged for 10 minutes at maximum speed. A QIAprep spin column was placed in a 2 mL collection tube (both supplied in the kit) and the supernatant from the eppendorf tube was applied to the QIAprep spin column by pipetting. The QIAprep spin column was centrifuged for 1 min and the flow through discarded. The QIAprep spin column was washed with 750 µL Buffer PE and centrifuged for 1 minute. The flow through was discarded and the spin column was centrifuged for 1 minute to remove any residual wash buffer. The

QIAprep spin column was placed in a clean 1.5 mL eppendorf tube and 50 µL Buffer PE was added to the spin column, incubated at room temperature (22°C) for 1 minute, and centrifuged for 1 minute to collect the plasmid cDNA. To calculate the cDNA concentration, 5 µL of the cDNA preparation was added to 195 µL of MilliQ H2O, and the absorbance at 260 nm was

51 measured using a Beckman DU-800 spectrophotometer (1 OD260 = 50 ng/µL). The plasmid

cDNA was stored at -20°C until used.

1.5. Restriction enzyme digestion of DNA

1 µL isolated cDNA, 1 µL KpnI (New England BioLabs), 1 µL XbaI (New England

BioLabs), 1 µL 10X multi-core buffer (Fisher), and 6 µL MilliQ H20 were mixed in an

eppendorf tube. The samples were centrifuged in a microcentrifuge (National Labnet Company) for 1 minute and incubated in a 37°C water bath for 1 hour. 5 µL of DNA loading buffer was added and the restriction enzyme digested samples were loaded onto a 0.8% agarose gel. (Make

a 0.8% agarose gel by adding 0.4 grams of electrophoresis grade agarose (Bethesda Research

Laboratories) to 1 mL 50X TAE buffer (60.5 grams Tris Base, 14.3 mL glacial acetic acid, and

25 mL 0.5 M EDTA, pH 8 in 250 mL MilliQ H2O) and 49 mL MilliQ H2O. The solution was

heated for 1.5 minutes, swirled, and heated for an additional 45 seconds in a microwave. The

solution was cooled for 5 minutes, at which time 10 µL of ethidium bromide (5 mg/mL) was added. After mixing, the solution was poured into the gel cast and allowed to cool and solidify.

250 mL of 1X TAE Buffer was added to the gel chamber. The samples were electrophoresed at room temperature (22°C) for 1 hour at 80V.

1.6. Transient transfection of COS-1 cells

COS-1 cells were grown in Dulbecco’s modified Eagle’s medium (DMEM) with 10% bovine serum and 2% mixture of antibiotics and antimycotics (Invitrogen), known as complete media. 4 µg plasmid cDNA encoding NTPDase3 (or empty pcDNA3 vector as a control) was

52 diluted into 750 µL of 37oC serum-free, antibiotic-free DMEM. The solution was mixed after the addition of 20 µL Plus Reagent (Invitrogen) and incubated at room temperature (22oC) for 15 minutes. 30 µL of Lipofectamine (Invitrogen) was added to another 1.5 mL eppendorf tube containing 750 µL of serum-free, antibiotic-free DMEM. The diluted Lipofectamine Reagent,

Plus Reagent, and plasmid cDNA were combined into one eppendorf tube and incubated at 22oC for 15 minutes. At this point, plates of COS-1 cells for transfection were removed from the incubator, the media removed by aspiration, and 5 mL of 37oC serum-free, antibiotic-free

DMEM added. This washing was repeated a second time. The third time the 37oC serum free, antibiotic free DMEM was added, the DNA complexes (DNA, DMEM, Lipofectamine, and Plus

Reagent mixture) were also added to the cells. The plates were swirled and allowed to incubate at 37oC for five (5) hours. 13 mL of complete media was then added and the plates were placed back in the incubator. The next morning, the media was removed by aspiration and 10 mL of

37oC complete media was added to the plates. The cell culture plates were returned to the incubator and the COS-1 cells were harvested the next day.

1.7. COS-1 cell crude membrane preparation

Media was removed from the culture dishes by aspiration, and each culture dish was rinsed twice with 10 mL isotonic wash buffer (IWB; 20 mM MOPS, pH 7.4, 140 mM NaCl, and

5 mM KCl, prepared the day of use). 1 mL tissue homogenization buffer (THB; 250 mM sucrose, 30 mM MOPS, 2 mM EDTA, adjusted to pH 7.4 with NaOH), prepared fresh the day of use, was added to each cell culture plate. The cells were removed from the culture plate with a scraper and the cell suspension transferred to a glass-Teflon homogenizer on ice. The culture

53 plates were rinsed with 1 mL THB, the remaining cells scraped off and added to the cell

suspension in the homogenizer. Each sample was homogenized with 25 passes of the Teflon

pestle (on ice). The homogenized extract was centrifuged in a 50 Ti rotor for 1 hour at 40,000

RPM at 4oC in a Beckman centrifuge. The supernatant was discarded and the COS cell pellet re- suspended in 150 µL THB. The cell suspension was transferred to a homogenizer on ice. The centrifuge tube was rinsed with an additional 150 µL THB, which was added to the homogenizer. The COS cell membranes were homogenized with 25 passes of the Teflon pestle and transferred to a 1.5 mL eppendorf tube on ice. The COS cell membranes were aliquoted and stored at -80oC until used.

1.8. Protein concentration of COS-1 crude cell membrane preparations

Protein concentrations were determined using the Bio-Rad protein assay reagent with the

modifications of Stoscheck (165). The crude COS-1 cell membrane preparations were diluted

1:5 with THB. 20 µL of COS-1 cell membranes (or water for blanks, or bovine serum albumin

for standards) was added to a 12 x 75 mm glass test tube (Fisher). 20 µL of 5 M NaOH was

added to the test tube and vortexed. 960 µL of diluted (1:5) commercial BioRad protein assay

dye reagent was added to each tube and vortexed. The samples were incubated for at least 5 (but

not more than 30) minutes at room temperature (22oC) and the absorbance at 595 nm versus

water was measured using a SPECTRONIC 20 GENESIS spectrophotometer. Protein

concentrations were determined using the “LOWRY” program included in “A collection of

radioligand binding analysis programs” by G.A. McPherson, using a non-linear curve fit.

54 1.9. Specific nucleotidase activity of NTPDase3 expressed in COS-1 cells

For most experiments, nucleotidase activities were determined by measuring the amount

o of inorganic phosphate (Pi) released from nucleotide substrates (Sigma) at 37 C using modifications of the Amidol technique of Fiske and Subbarow (166) as previously described

(140). Either 5 mM MgCl2 or 5 mM CaCl2 (both in 20 mM MOPS pH 7.1 buffer) was used for cation-dependent nucleotidase activities. The reactions were initiated by the addition of nucleotide to a final concentration of 2.5 mM in the 0.3 mL assay solutions. Hydrolysis was allowed to proceed for 30 minutes or 1 hour, depending on the substrate used. The activities were corrected for pcDNA3/COS-1 cell membrane background (membranes from COS-1 cells transfected with an empty vector), as well as differences in expression levels as determined by quantitative Western blotting of each sample.

1.10. Malachite green phosphate determination procedure for nucleotidase assays

Malachite green solution (167) was made fresh on the day of use by combining, with stirring, 10 mL color reagent (60 mL concentrated sulfuric acid to 300 mL MilliQ H2O, cool to

22oC, add 0.44 grams Malachite green), 2.5 mL ammonium molybdate solution (dissolve 7.5

grams ammonium molybdate in 100 mL MilliQ H2O), and 220 µL 10% Tween 20. The solution was incubated at room temperature (22oC) with stirring for 30 minutes. 0.4 µg crude COS-1 cell

membrane protein was typically assayed. The calculated volume of MilliQ H2O (620 µL minus the volume of sample) and 160 µL of 5X buffer (resulting in 5 mM MgCl2 or 5 mM CaCl2 in 20 mM MOPS, pH 7.1 buffer after dilution) was added to a test tube (Fisher). The calculated

55 volume of sample was added, gently mixed, and pre-incubated at 37oC for 3 minutes. Substrate

was added to a final concentration of 0.25 mM to start the reaction. Hydrolysis was typically

allowed to proceed for 8 minutes (Ca-ATPase), 16 minutes (Ca-ADPase), 16 minutes (Mg-

ATPase), or 32 minutes (Mg-ADPase) at 37oC. The reaction was quenched by adding 200 µL

Malachite green solution and incubated at 22oC for 10 minutes. The absorbance at 630 nm was

measured to determine the amount of inorganic phosphate (Pi) released.

1.11. SDS-PAGE and Western blotting

Pre-cast 10-well or 15-well 4-15% polyacrylamide gradient mini-gels (Bio-Rad) were used to resolve aliquots of crude COS membrane proteins (0.5 – 2 µg, depending on the sample and the experimental purpose), usually after boiling for 5 minutes in SDS sample buffer containing 30 mM dithiothreitol (DTT). Following SDS-PAGE electrophoresis, the proteins were electro-transferred to Immun-Blot PVDF membrane (Bio-Rad) for 3 hours at 33 V in cold

10 mM CAPS/NaOH, pH 11. After transfer, the PVDF membrane was incubated for 1 hour in blocking solution (5% nonfat dry milk in Tris-buffered saline, TBS) at room temperature (22oC), and then incubated overnight at room temperature in blocking solution containing 0.02% sodium azide and a 1:5000 dilution of rabbit polyclonal primary antisera against human NTPDase3

(either KLH1 antibody, generated against the cytoplasmic C-terminal peptide (amino acid residues 515-529) of human NTPDase3 (133) or KLH11 antibody, generated against an amino acid sequence in the extracellular domain (amino acid residues 311-327) of human NTPDase3

(107)). After washing the blot in a TBS containing 0.05% Tween 20 (TTBS), a goat anti-rabbit

HRP-conjugated secondary antibody (Pierce) was applied at a 1:4000 dilution for 1 hour at room

56 temperature followed by washing in TTBS and applying the Pierce SuperSignal West Dura

Extended Duration Substrate to the PVDF membrane for 5 minutes to detect immunoreactivity by chemiluminescence. Chemiluminescence was recorded and quantified using a FluorChem IS-

8800 system (Alpha Innotech).

1.12. Deglycosylation

Partial deglycosylation was performed using Endo H (New England Biolabs) according to the manufacturer’s instructions. Briefly, COS-1 cell crude membrane protein was solubilized with 0.2% Triton X-100 containing 1 mM EDTA for 30 minutes at 22oC. The solubilized samples were then centrifuged in a Beckman air-driven centrifuge for 20 minutes at 100,000g.

The supernatant was combined with 10X Glycoprotein Denaturing Buffer (to yield final concentrations of 0.5% SDS and 1% β-mercaptoethanol) and boiled for 10 minutes. After cooling to room temperature (22oC), each sample was combined with the appropriate reaction buffer (NEBuffer G5), aliquoted, and incubated with, or without (controls), Endo H at 37oC for 1 hour or 3 hours. After incubation, the samples were combined with SDS sample buffer containing 100 mM DTT, run on an SDS-PAGE gel, and analyzed by Western blot as described above.

57 2. Project Specific Materials and Methods

2.1. NTPDase3 linker region proline residue mutagenesis

2.1.1. List of oligonucleotides/mutants made

V42C, 5’-GGTACTTGTGAGTATCACTTGCATCCAGATCCACAAGCAAG-3’;

G489C, 5’-CCACCTGTCTTTGTGTGCACCCTCGCTTTC-3’;

P52A, 5’-CACAAGCAAGAGGTCCTCGCTCCAGGACTGAAGTATG-3’;

P53A, 5’-CAAGAGGTCCTCCCTGCAGGACTGAAGTATGC-3’;

P472A, 5’-CTGACCAACCAGATCGCAGCTGAAAGCCCTC-3’;

P476A, 5’-GATCCCAGCTGAAAGCGCTCTGATCCGTCTGC-3’;

P481A, 5’-CCTCTGATCCGTCTGGCCATAGAACCACCTG-3’;

P484A, 5’-CGTCTGCCCATAGAAGCACCTGTCTTTGTGGG-3’;

P485A, 5’-CTGCCCATAGAACCAGCTGTCTTTGTGGGCAC-3’

P52A/P53A, 5’-GCAAGAGGTCCTCGCTGCAGGACTGAAGTATG-3’

P481A/P484A, 5’-CTCTGATCCGTCTGGCCATAGAAGCACC-3’

P484A/P485A, 5’-CTGCCCATAGAAGCAGCTGTCTTTGTGGGC-3’

The altered codons used to introduce the mutations are bolded and underlined (the

complementary anti-sense oligonucleotides also needed for mutagenesis are not shown). The

mutants were made in the wild-type like “free sulfhydryl-less” C10S/C501S/C509S

NTPDase3 background to eliminate all other free sulfhydryls (147). The resulting cDNA

58 constructs were sequenced by the University of Cincinnati DNA Core Facility to verify the

presence of the desired mutation and the absence of any unwanted changes.

2.1.2. Oxidative cross-linking in the presence and absence of substrate

Copper phenanthroline (CuPhen) was the oxidative cross-linker used and was prepared

fresh each day, just before use, by combining cupric sulfate and 1,10-phenanthroline (Sigma)

at a 1:3 ratio in 20% ethanol (168). Cysteine-substituted protein from COS-1 cell crude

membranes was diluted in 50 mM Tris-HCl, pH 7.5 to a final concentration of 0.1 mg/mL

and oxidized with a final concentration of 5 mM CuPhen in the absence or presence of a final

concentration of 5 mM ATP (Sigma) at 37oC for 5 minutes. The reaction was stopped by

adding an equal volume of non-reducing SDS loading buffer containing 20 mM N-

ethylmaleimide (NEM) and 20 mM EDTA. The samples were then loaded on a gel for SDS-

PAGE and Western blotting as described above.

2.2. NTPDase3 transmembrane domain polar residue mutagenesis

2.2.1. List of oligonucleotides/mutants made

S33A, 5’-GGTGGTCTTGCTTGTGGCTATTGTGGTACTTGTG-3’;

S39A, 5’-GAGTATTGTGGTACTTGTGGCTATCACTGTCATCCAGATCC-3’;

T41A, 5’- CTTGTGAGTATCGCTGTCATCCAGATCCACAAGC-3’;

Q44A, 5’-GTGAGTATCACTGTCATCGCGATCCACAAGCAAGAGG-3’;

T490A, 5’-CCTGTCTTTGTGGGCGCCCTCGCTTTCTTC-3’;

59 T495A, 5’-CCTCGCTTTCTTCACAGCAGCGGCAGCCTTG-3’;

C501A, 5’-CACAGCGGCAGCCTTGCTGCTCTGGCATTTCTTGCATAC-3’;

S33C, 5’-CTTGGTGGTCTTGCTTGTGTGTATTGTGGTACTTGTGAG-3’;

S39C, 5’-GTGAGTATTGTGGTACTTGTGTGTATCACTGTCATCCAGATC-3’;

T41C, 5’-GTGGTACTTGTGAGTATCTGTGTCATCCAGATCCACAAG-3’;

Q44C, 5’-CTTGTGAGTATCACTGTCATCTGTATCCACAAGCAAGAGGTCCTC-3’;

T490C, 5’-CACCTGTCTTTGTGGGCTGCCTCGCTTTCTTCACAGC-3’;

T495C, 5’-CACCCTCGCTTTCTTCTGCGCGGCAGCCTTGCTGTC-3’;

Q44C/S39C, 5’-GTGAGTATTGTGGTACTTGTGTGTATCACTGTCATCTGTATC-3’;

Q44C/T41C, 5’-GTGGTACTTGTGAGTATCTGTGTCATCTGTATCCACAAG-3’

The altered codons used to introduce the mutations are underlined and bolded and the

complementary anti-sense oligonucleotides also required for mutagenesis are not shown.

The polar residue to alanine substitution mutants were made in the “wild-type like” C10S

NTPDase3 background and the single as well as double cysteine mutants were made in the

“free sulfhydryl-less” C10S/C501S/C509S NTPDase3 background (which has enzymatic

properties similar to WT C10S (147)), except for the S33C/C501 mutant which was made in

the C10S/C509S NTPDase3 background (also having enzymatic properties similar to WT

C10S (147)). DNA Analysis, LLC sequenced the resulting cDNA constructs to verify the

presence of the desired mutation and the absence of any unwanted changes.

60 2.2.2. Triton X-100 nucleotidase assays

In some experiments, COS-1 cell crude membrane protein (0.05 mg/mL) was

solubilized in 1% Triton X-100 and 25 mM MOPS buffer, pH 7.1 at 22oC for 10 minutes

with occasional vortexing. In most experiments, the effects of Triton X-100 on nucleotidase

activities were measured by dilution of COS membranes into assay solutions containing

0.1% Triton X-100, using either 5 mM CaCl2 or 5 mM MgCl2 (both in 20 mM MOPS, pH

7.1 buffer) and nucleotide at a final concentration of 0.25 mM. The malachite green

phosphate assay described above (167) was used, due to the interference of Triton X-100

with the modified Amidol (Fiske and Subbarow) assay described above.

2.2.3. Oxidative cross-linking and alkylation of cysteine

Copper phenanthroline (CuPhen) was used as the oxidative cross-linker to form

disulfides from pairs of properly positioned free cysteine sulfhydryls. The reagent was

prepared just before use by combining cupric sulfate and 1,10-phenanthroline (Sigma) at a

1:3 ratio in 20% ethanol as described above (168). Cysteine-substituted COS-1 cell crude

membrane protein (0.1 mg/ml) was cross-linked with a final concentration of 0.5 mM

CuPhen in 50 mM HEPES, pH 7.5 at 37oC for 20 minutes. For samples directly analyzed by

SDS-PAGE, the reactions were stopped by adding an equal volume of non-reducing SDS

loading buffer containing 20 mM EDTA. For samples alkylated with maleimide

polyethylene glycol-5000 (MalPEG), the cross-linking reaction was stopped with 0.5 volume

of a solution to yield final concentrations of 20 mM EDTA and 1% SDS in 20 mM HEPES,

61 pH 7.5. The samples were incubated at room temperature (22oC) for 10 minutes with

occasional vortexing, followed by the addition of MalPEG to a final concentration of 5 mM.

The samples were alkylated at room temperature for 20 minutes with occasional vortexing,

and the reaction was stopped by adding an equal volume of non-reducing SDS-PAGE

loading buffer. The samples were then heated at 60oC for 10 minutes and loaded onto a gel

for SDS-PAGE followed by Western blot analysis as described above.

2.2.4. NTPDase3 nucleotidase assays in the presence of Tween 20

COS-1 cell crude membrane protein (0.25 µg/ml) was pre-incubated for 5 minutes at

o 37 C in 0.1% Tween 20 and 5 mM MgCl2 or 5 mM CaCl2 (both in 20 mM MOPS, pH 7.1

buffer). After addition of nucleotide to a final concentration of 0.25 mM to start the reaction,

hydrolysis was typically allowed to proceed for 5 minutes (Ca-ATPase), 10 minutes (Ca-

ADPase), 10 minutes (Mg-ATPase), or 20 minutes (Mg-ADPase) at 37oC. Nucleotidase

activities were measured using a malachite green inorganic phosphate assay as described

above. The absorbance at 630 nm was measured to determine the amount of inorganic

phosphate (Pi) released.

62 2.3. NTPDase3 alternatively spliced variant (NTPDase3β)

2.3.1. IMAGE clone encoding NTPDase3β

A bacterial stock of the NTPDase3β IMAGE clone (an alternatively spliced human

NTPDase3, IMAGE clone ID 5174918; GenBank Accession BC029869) in pCMV-SPORT6

vector was obtained from Invitrogen. Plasmid DNA was purified (Qiagen Miniprep) and

sequenced (University of Cincinnati DNA Core Facility), and digested with NotI and KpnI

(Fisher Scientific). A 1% low-melting point agarose gel was used to purify the gene-

containing plasmid insert after endonuclease digestion. The resulting fragment was ligated

using T4 DNA ligase (Invitrogen) into mammalian expression vector pcDNA3, which was

also digested with the same endonucleases. The resulting plasmid was transformed into

XL1-Blue Supercompetent bacterial cells (Stratagene) and plated on carbenicillin containing

LB-Agar plates.

2.3.2. List of oligonucleotides/mutants made

A511stop, sense primer: 5’-GCATACCTGTGTTCATTGAACCAGAAGAAAGAGGC-3’

A511stop, antisense primer: 5’-GCCTCTTTCTTCTGGTTCATGAACACAGGTATGC-3’

The A511Stop mutant was made in the wild type-like C10S NTPDase3 background.

The DNA Core facility at the University of Cincinnati produced the synthetic oligonucleotide

primers needed to make the NTPDase3 mutant lacking the 19 amino acids comprising the C-

63 terminal cytoplasmic tail, and sequenced the resulting cDNA construct to verify the presence

of the desired mutation and the absence of any unwanted changes.

2.4. NTPDase3 single nucleotide polymorphisms (SNPs)

2.4.1. Source of SNPs

The SNPs were identified based on data obtained from the HapMap database

(www.hapmap.org). This information is catalogued at the National Center for Biological

Technology Information website (http://www.ncbi.nlm.nih.gov) and describes the different

nucleotide variants, where they occur in the DNA, and how they are distributed among

people within populations and among various human populations in different parts of the

world.

2.4.2. List of oligonucleotides/mutants made to mimic known SNPs

I24V, 5’-CCGAACTCCAACCGTCATTGCCTTGGTG -3’;

A496V, 5’-CGCTTTCT TCACAGTGGCAGCCTTGCTG-3’;

L505F, 5’-GCTGTGTCTGGCATTTTTTGCATACCTGTGTTC-3’;

E440D, 5’-CGGTTACAAATTCACAGATGAGACTTGGCCCCAAATAC-3’;

R264Q, 5’-CTTCCAGTGCTATGGCCAGAATGA GGCTGAGAAG-3’

The altered codons used to introduce the mutations mimicking the SNPs are bolded and

underlined. The complementary anti-sense oligonucleotides necessary for the mutagenesis

64 are not shown. A QuikChange Site Directed Mutagenesis kit (Stratagene) was used to

mutate wild-type NTPDase3 in the mammalian pcDNA3 vector. The mutagenized cDNA

was transformed into Stratagene XL1 Blue Supercompetent Cells and grown overnight in 5

mL of LB broth. The mutagenized plasmid DNA was purified from the overnight cultures

using the Qiagen Miniprep system. The plasmid DNA constructs were sequenced by the

University of Cincinnati DNA Core Facility to verify the presence of the desired mutation

and the absence of any unwanted changes.

2.4.3. Treatment of wild-type and mutant NTPDase3 with concanavalin A

Concanavalin A (Con A) lectin (Sigma) was prepared at a concentration of 5 mg/mL in

20 mM MOPS buffer containing 100 mM NaCl, 1 mM MnCl2, and 1 mM CaCl2, pH 7.4.

Wild-type and mutant crude membrane preparations (2 µg) were incubated in a total volume

of 300 µL containing 5 µL of 5.0 mg/mL Con A at 37oC for 15 minutes. Afterwards,

nucleotidase assays were performed in the presence of 5 mM MgCl2 and 2.5 mM substrate

(ATP) as described above.

65 Chapter III: Results

1. Mutagenesis of NTPDase3 “Linker Region” Proline Residues

1.1. Rationale for selection and analysis of NTPDase3 “linker region” proline residues

The “linker region” is one of three structural domains in NTPDase3. It consists of a

stretch of amino acids between the TMD (location of the N- and C-terminal TM helices) and the

extracellular domain (location of the enzyme active site). It also contains several conserved

proline residues (N-terminal: P52 and P53; C-terminal: P472, P476, P481, P484, P485 – see

Figure 11). Previous studies have demonstrated that interactions between the extracellular and

TM domains occur in NTPDases (135, 169, 170); however, the specific structural features that

couple these interactions are unknown.

Proline is a structurally unique amino acid with a distinctive cyclic structure. As a result

of its side-chain binding to the backbone amide position, proline acts only as a hydrogen bond

acceptor since the amide proton is replaced by a -CH2 group, eliminating its ability to be a

hydrogen bond donor. The bulkiness of the N-CH2 group places restrictions on the conformation

of the residue preceding proline (171), thereby destabilizing α-helix secondary structures (172,

173). The elimination of the hydrogen bond donor, together with the bulkiness of the side-chain,

produces the well-known “helix-breaker” (and β-sheet breaker) secondary structure effect of

proline residues. Lastly, the ring structure of proline locks the available backbone φ dihedral angle at approximately 650, resulting in a conformational rigidity unique to

66 ___

TLAF LMVL 481 489 489 G S S G 481

VFLGLLII VFLGLLII AFLGILAF AFLGILAF VFMGVLAF VFMGVLAF VFMGVLAF VFMGVLAF VFVGTLAF VFVGTLAF VFV VFVGTLAF VFVGTLAF VFV VFVGTLAF VFVGTLAF VFVGTLAF VFVGTLAF VFMGVLAF VFMGVLAF VFMGVLAF VFMGVLAF AFLGILAF AFLGILAF VFLGLLII VFLGLLII SAFRCTVAF SAFRCTVAF QVFTGVMFF SAFRCTVAF SAFRCTVAF QVFTGVMFF QVFTGVMFF P P PP PP PP PP PP PP PP PP PP PP PP PP PP P P PP LK MR MK MD IQ IQ LPHSTYI IE IE LPHSTYI LSHSTYVFLMVL LSHSTYVFLMVL IE IE LSHSTYVFLMVL IE IE IQ IQ MD LK MR MK P P P P P P P P P P P P P P P P P P P P P P LSP LSP LST LST LIRL QIQL LIHV LIRL LIQL LIHL LIRL LIRL GLRKGTDFSSWVVLLLL GLRKGTDFSSWVVLLLL AQWRAESYGVMVAKVVF LIRL LIRL GLRKGTDFSSWVVLLLL AQWRAESYGVMVAKVVF LIRL LIRL LIHL LIQL LIRL LIRL QIQL LIHV P P P P P P P P. P P. P P P P. P P. P P P P P P P P P P AES AEN AES AES AGS AGS AES AEQ ___ C-term.TM______C-term.TM_ AES ADP LRDIQQEAFRASHTHWRGVSF ADA AES AEQ AEQ AES AEQ ADP LRDIQQEAFRASHTHWRGVSF ADA AES AES AGS AGS AES AES AEN AES P P P P P P P P P P P P P P P P P P P P P P P P P P P P 472 476481 484/5 472 476481 484/5

ACR5______ACR5______TNQI TNQI TNMI ___ TNMI ___

Horse NTPDase3 NTPDase3 Horse TNQI Opossum NTPDase3 TNMI TNMI NTPDase3 Platypus Dog NTPDase3 NTPDase3 Dog TNQI Rat NTPDase3 NTPDase3 Rat TNQI Mouse NTPDase3 NTPDase3 Mouse TNQI

Human NTPDase3 Human NTPDase1 Rat Human NTPDase4/7 NTPDase4/7 TRFL Human Human NTPDase2 NTPDase2 Human TNLI NTPDase8 Human TGMI Chimpanzee NTPDase3 TNQI NTPDase3 Chimpanzee NTPDase1 Human TNMI Macaque NTPDase3 TNQI Human NTPDase3 Human NTPDase1 Rat Macaque NTPDase3 NTPDase3 TNQI Macaque Human NTPDase2 NTPDase2 Human TNLI Human NTPDase4/7 TRFL Chimpanzee NTPDase3 TNQI NTPDase3 Chimpanzee NTPDase1 Human TNMI NTPDase8 Human TGMI Mouse NTPDase3 TNQI NTPDase3 Mouse Rat NTPDase3 NTPDase3 Rat TNQI Dog NTPDase3 NTPDase3 Dog TNQI Horse NTPDase3 NTPDase3 Horse TNQI Opossum NTPDase3 NTPDase3 TNMI Opossum Platypus NTPDase3 TNMI NTPDase3 Platypus

C-terminal Conserved Conserved C-terminal

GLKYGIVFDAG GLKYGIVFDAG GLKYGIVLDAG GLKYGIVLDAG GLKYGVVLDAG GLKYGVVLDAG GLKYGIVLDAG GLKYGIVLDAG GLKYGIV-DAG GLKYGIV-DAG GLKYGIV-DAG GLKYGIV-DAG GLKYGIVLDAG GLKYGIVLDAG ALKYGIVLDAG NVNYGIVVDCG P P P P P TDIKFGIVFDAG ENVKYGIVLDAG ENVKYGIVLDAG PP PP PP PP P PP P P PP ___ACR1______52 53 IQIHKQEVL VGLTQNKAL 34 A V N-term.TM___ SIT ALL N-terminal Conserved Prolines Prolines Conserved N-terminal Opossum NTPDase3 AITLIQINQKEVLA NTPDase3 TITLVQINQRETLS Platypus Rat NTPDase3 TLTLIQIHHPQVLS Dog NTPDase3 AITLIQFHHKEVLL NTPDase3 Horse TITVIQIHQKEVL Mouse NTPDase3 TLTLIQIRHPQVL Macaque NTPDase3 SITVIQIHKQEVL Chimpanzee NTPDase3 SITVIQIHKQEVL NTPDase3 Chimpanzee NTPDase4/7 LARVTDIEATDTNN Human Human NTPDase8 LLLVE..ATSVLL ___ 42 Human NTPDase3 Human NTPDase1 Rat Human NTPDase1 NTPDase1 Human ALLAVGLTQNKAL Human NTPDase2 LLCVP..TRDVRE

67 Figure 11. Multiple sequence alignment for portions of NTPDase sequences between the N- and C-terminal TM helices and the extracellular domain containing the enzyme active site.

Conserved proline residues found in human NTPDase3 are numbered and bolded. The locations of the cysteine substitutions useful in rat NTPDase1/CD39 (A34C and S481C, (149)), are bolded and underlined, and guided the choice of locations of cysteine substitutions used in this work as cross-linking “sensors” (V42C and G489C, which are underlined and bolded in the human

NTPase3 sequence (top line of each alignment)). To give a perspective as to the location of these conserved proline residues in the overall structure of NTPDase3, the locations of the ends of the N-terminal TM helix, ACR1, ACR5, and the C-terminal TM helix are indicated by lines and italic labeling under the alignments.

68 proline (174, 175). Thus, proline residues confer unique structural characteristics needed for some conformational changes in proteins.

Multiple sequence alignments of the N- and C-terminal linker regions that connect the

TM helices with the extracellular active site lobes of several known NTPDase3 sequences (rat

NTPDase1 and human NTPDase 1, 2, 4/7, and 8) revealed complete conservation of P472, a

high degree of conservation of P53, P476 and P481 (completely conserved in the NTPDase3

enzymes), and a lesser degree of conservation of P52, P484, and P485 (Figure 11). This led to

the hypothesis that these proline residues may provide the structural rigidity in the linker region

of NTPDase3 that is necessary to couple the movements of the active site to the transmembrane

domain during nucleotide hydrolysis. To investigate the role of the conserved proline residues in

coupling the two domain movements, proline to alanine substitutions were coupled with specific

TM helix residue to cysteine “sensor” substitutions used to monitor TMD cross-linking

efficiency in the presence and absence of ATP.

1.2. Characterization of the NTPDase3 proline to alanine mutants

Single proline to alanine mutants were made in the “wild-type like” C10S NTPDase3,

background (denoted as C10S) to eliminate the artificial dimerization via oxidation of cysteine

10 (147), and expressed the mutants in COS-1 cells. The mutants were characterized for

enzymatic activity, expression level, and proper glycosylation and delivery to the cell surface in their native oligomeric state. As seen in Figure 12, alanine substitution caused a significant decrease in the expression level of P53, P472, P476, and P485, as compared to

69 1.0

0.8 * * * * 0.6

0.4

0.2 Relative Expression Level

0.0 C10S P52A P53A P472A P476A P481A P484A P485A

Figure 12. Alanine substitution at P53, P472, P475, and P485 causes a decrease in

expression level in mutant NTPDase3 enzymes. One microgram (µg) of COS-1 cell crude membrane protein for each mutant and 0.25 – 1.0 µg of the C10S NTPDase3 as a standard curve was loaded onto an SDS-PAGE gel and detected and quantified by chemiluminescence. The data is presented as the mean relative expression level ± standard deviation from three separate experiments. The asterisks represent statistical significance (p ≤ 0.05), comparing the C10S background control mutant to the proline to alanine mutants.

70 the background C10S NTPDase3. However, once corrected for expression level, all the mutants

exhibit enzymatic activities similar to C10S except: P53A, which has a significant decrease in

both Ca- and Mg-dependent nucleotidase activities; P481A, which has a significant decrease in

both Ca- and Mg-dependent nucleotidase activities; and mutants P484A and P485A, which

demonstrate a small, but significant decrease in Ca-ATPase activity (Figure 13).

During studies which established the location of 5 extracellular disulfide bonds in human

NTPDase3, endoglycosidase H (EndoH) deglycosylation analysis was shown to be a simple and reliable assay for proper protein processing and delivery of human NTPDase3 to the cell membrane (107). All of the proline to alanine mutants were properly processed and trafficked to the cell-surface as assessed by endoglycosidase H (EndoH) (Figure 14).

Often a structural motif in proteins necessary for protein function consists of multiple amino acids. Therefore, double proline to alanine substitutions (P52A/P53A, P53A/P481A,

P481A/P484A, and P484A/P485A) were made in the C10S background to determine if replacing two prolines on the same linker arm or one on each linker arm would cause a greater decrease in enzymatic activity of NTPDase3. As shown in Figure 15, each of the double proline to alanine mutants experienced a decrease in expression level as compared to C10S. However, once corrected for expression level, all the double mutants had specific activities similar to C10S except: P53A/P481A, which demonstrated a significant decrease in Ca-dependent nucleotidase activities and Mg-ATPase; and the P484A/P485A mutant, which demonstrated a significant decrease in Ca-ADPase and Mg-dependent nucleotidase activities (Figure 16).

Taken together, this data demonstrates a clear role for P53 and P481 in facilitating the full enzymatic activity of NTPDase3. The effect of decreased enzymatic activity was also evident in the P484A/P485A double mutant, eliminating two vicinal proline residues near the C-

71

CaATPase 350 CaADPase MgATPase MgADPase 300 * * 250 *

mol/mg/hr) 200 µ

150 *

100 * * 50 * * Specific Activity ( * * 0 C10S P52A P53A P472A P476A P481A P484A P485A NTPDase3

Figure 13. P53A and P481A NTPDase3 mutants exhibit a decrease in nucleotidase activities as compared to the C10S NTPDase3 background mutant. Nucleotidase activities were assayed as described in the methods. Values represent the mean ± standard deviation from three separate experiments. Asterisks represent statistical significance (p ≤ 0.05), comparing the

C10S background control mutant to the proline to alanine mutants.

72 - + - + - + - + - + - + - + - + EndoH

Native glycosylation

C10S P52A P53A P472A P476A P481A P484A P485A

Figure 14. Alanine substitution of NTPDase3 proline residues are properly processed and presumably trafficked to the cell membrane. C10S and NTPDase 3 proline to alanine substituted mutant proteins were treated with (+) and without (-) Endo H for 1 hour at 37oC.

NTPDase3 has seven putative glycosylation sites, one being an immature high-mannose glycan sensitive to Endo H, resulting in a small shift in the electrophoretic mobility of C10S following

Endo H treatment (+ C10S lane). Protein bands with electrophoretic mobility similar to that of

C10S are marked as “native glycosylation” since they exhibit similar shifts after Endo H digestion, and this band represents the enzyme that is delivered to the cell surface. None of the mutants demonstrated any non-glycosylated “core protein” at about 59 kDa.

73 1.0

0.8 * * * 0.6 *

0.4

Relative Expression Level Expression Relative 0.2

0.0 C10S P52A/P53A P53A/P481A P481A/P484A P484A/P485A

Figure 15. Double alanine substitution of the proline residues causes a decrease in

expression level in mutant NTPDase3 enzymes. One microgram (µg) of COS-1 cell crude membrane protein for each mutant and 0.25 – 1.0 µg of the C10S NTPDase3 as a standard curve was loaded onto an SDS-PAGE gel and detected and quantified by chemiluminescence. The data is presented as the mean relative expression level ± standard deviation from three separate experiments. The asterisks represent statistical significance (p ≤ 0.05), comparing the C10S background control mutant to the proline to alanine mutants. The standard deviation for the

P484A/P485A mutant is too small for representation on this graph.

74

CaATPase CaADPase MgATPase 300 MgADPase

250

200 *

mol/mg/hr) µ

150

* 100 *

Specific Activity ( 50 * * *

0 C10S P52A/P53A P53A/P481A P481A/P484A P484A/P485A

NTPDase3

Figure 16. P53A/P481A and P484A/P485A NTPDase3 mutants exhibit a decrease in

nucleotidase activities as compared to the C10S NTPDase3 background mutant.

Nucleotidase activities were assayed as described in the methods. Values represent the mean ±

standard deviation from three separate experiments. Asterisks represent statistical significance

(p ≤ 0.05), comparing the C10S background control mutant to the double alanine substituted mutants.

75 terminal TM helix. More importantly, the results suggest that P53 on the N-terminal linker and

P481 on the C-terminal linker are functionally important, and may play a role in coupling the movements of the extracellular domain (site of nucleotide binding and hydrolysis) to the TM helices that facilitate full enzymatic activity of NTPDase3.

1.3. Rationale for selection of transmembrane “sensor” mutations and experimental

approach

To test the hypothesis that some of the conserved proline residues couple the movements of the extracellular domain to the TM helices, cysteine residues (V42C and G489C) were singly introduced in the C10S/C501S/C509S WT-like NTPDase3 (147), which is devoid of any free sulfhydryls to allow site-specific, oxidative, intermolecular cross-linking of the TM helices of human NTPDase3. These amino acids correspond to those of rat CD39/NTPDase1 that were shown to be excellent “sensors” of ATP binding effects in the extracellular domain on TM helix cross-linking (149). Thus, V42 (corresponding to A34 in rat NTPDase1) and G489

(corresponding to S481 in rat NTPDase1) in human NTPDase3 were singly mutated to cysteine

(Figure 11). These mutants were used as the “background” and “sensors” for subsequent proline to alanine mutations that were characterized for nucleotidase activity and changes in copper phenanthroline (CuPhen) cross-linking efficiency in the presence and absence of ATP.

76 1.4. Characterization of the transmembrane “sensor” mutations (V42C and G489C

NTPDase3)

The nucleotidase activities of the V42C and G489C mutants were similar to the

C10S/C501S/C509S NTPDase3 control, with a slight, but significant increase in the Ca-ATPase of V42C and a small, but significant decrease in the Ca- and Mg-dependent nucleotidase activities for G489C (Figure 17). The cross-linking efficiency of V42C and G489C in the presence and absence of 5 mM ATP was then assessed, to determine if these mutants could be used as TM specific, cross-linking “sensors” for the subsequent proline residue mutations. In the presence of 5 mM CuPhen, both V42C (Figure 18A) and G489C (Figure 18B) mutants were cross-linked to form dimers (it should be noted that the KLH1 antibody reacts much stronger with dimeric than monomeric NTPDase3, and thus only monomer bands were used in the calculation of the cross-linking efficiency). In the presence of 5 mM ATP, CuPhen-induced cross-linking efficiency of both V42C and G489C mutants decreased (the percent monomer remaining increased). Oxidative cross-linking of V42C and G489C also caused a decrease in the enzymatic activity of NTPDase3 (Figure 19), which is consistent with analogous experiments previously published using rat NTPDase1 (149). These data demonstrate the importance of the

TM helices of NTPDase3 to move relative to one another during nucleotide hydrolysis to achieve

full enzymatic activity, since CuPhen-induced disulfide restrictions caused a decrease in

enzymatic activity. Also, the data confirms the use of V42C and G489C NTPDase3 mutants as

appropriate “sensors” for active-site dependent changes of the TM helices of NTPDase3. This

provides the necessary “background” and “sensor” mutants to test the hypothesis that the cross-

talk between the active site and the TM helices may depend on the presence of some of the

77 CaATPase CaADPase MgATPase 100 * MgADPase

80 *

mol/mg/hr)

µ 60

* 40

20 * * Specific Activity (

0 C10S/C501S/ V42C G489C C509S NTPDase3

Figure 17. Cysteine substitution mutants V42C and G489C in NTPDase3 exhibits similar nucleotidase activities as the C10S/C501S/C509S NTPDase3 “background”. Nucleotidase activities were assayed as described in the methods. Values represent the mean ± standard deviation from three separate experiments. Asterisks represent statistical significance (p ≤ 0.05), comparing the C10S/C501S/C509S background control mutant to the V42C or G489C mutant.

78 100

80 * 60 D 40

M 20 Percent Monomer (M) 0 - + + - + + CuPhen

- - + - - + ATP 100

80 *

D 60

40

20 M Percent Monomer (M) Monomer Percent 0 - + + - + + CuPhen - - + - - + ATP

Figure 18. ATP binding at the extracellular domain decreases the cross-linking efficiency of V42C and G489C NTPDase3 mutants. Cross-linking was performed and the amount of monomer quantified as described in the methods. Top Panel. V42C NTPDase3 mutant.

Bottom Panel. G489C NTPDase3 mutant. The monomer (M) and dimer (D) bands are indicated. Values represent the mean percent monomer remaining ± standard deviation from three separate experiments. Asterisks represent statistical significance (p ≤ 0.05), comparing the amount of monomer remaining between the CuPhen alone and the CuPhen + ATP treated samples.

79

Control CuPhen

1.0

0.8

0.6 Activity

0.4 * Abs 660 nm * 0.2

0.0 C10S/ V42C G489C C501S/C509S NTPDase3

Figure 19. Oxidative cross-linking decreases ATP hydrolysis. Oxidative cross-linking was

conducted as described in the methods with CuPhen at a final concentration of 5 mM. The

nucleotidase reaction was carried out in 5 mM CaCl2 and ATP to a final concentration of 2.5 mM

as described in the methods. The values represent the average Abs 660 nm (Ca-ATPase activity)

± standard deviation from three separate experiments. Asterisks denote statistical significance (p

≤ 0.05), comparing the enzyme activity of the sample in the presence and absence of CuPhen.

80 conserved proline residues located in the “linker region” of NTPDase3.

1.5. Characterization of proline to alanine mutants in the V42C or G489C NTPDase3

background

To determine if the proline residues play a role in “coupling” ATP binding at the enzyme active site to movements of the TM helices, site-directed mutagenesis was used to singly substitute the conserved proline residues in the “linker regions” to alanine in both the V42C and

G489C NTPDase3 backgrounds. Proline to alanine substitutions at positions 53 and 481 resulted in significant decreases in Ca-ATPase activity as compared to the V42C background control

(Figure 20). In addition, proline to alanine substitutions at positions 484 and 485 resulted in smaller, but significant decreases in Ca-ATPase activity as compared to the V42C background control (Figure 20). Proline to alanine substitutions at positions 53, 481, and, to a lesser extent,

484, resulted in significant decreases in Ca-ATPase activity as compared to the G489C background control (Figure 21). Alanine substitution at P472 resulted in a significant increase in

Ca-ATPase activity as compared to G489C, while the P485A mutant demonstrated a small, but significant increase in Ca-ATPase activity as compared to G489C (Figure 21).

Each mutant was also oxidized with 5 mM CuPhen in the presence and absence of 5 mM

ATP to assess CuPhen cross-linking efficiency (measured by quantifying the monomer band remaining relative to the untreated control). In the V42C background, the ATP-induced decrease in CuPhen cross-linking efficiency was abolished by proline to alanine substitution at positions

53 and 481 (Figure 22). Proline to alanine substitutions at position 481 abolished the ATP-

81 450

375

* 300 * *

mol/mg/hr) µ

225 *

150

75

Ca-ATPase Activity ( 0 V42C P52A P53A P472A P476A P481A P484A P485A

V42C NTPDase3 Mutants

Figure 20. Ca-ATPase activity of the V42C proline to alanine mutants. Values represent the mean ± standard deviation from three separate experiments. Asterisks represent statistical significance (p ≤ 0.05), comparing the V42C background control mutant to the proline to alanine mutants.

82

375 ) *

300 *

mol/mg/hr µ ( 225 * *

150 *

75

Ca-ATPase Activity 0 G489C P52A P53A P472A P476A P481A P484A P485A

G489C NTPDase3 Mutants

Figure 21. Ca-ATPase activity of G489C proline to alanine mutants. Values represent the

mean ± standard deviation from three separate experiments. Asterisks represent statistical

significance (p ≤ 0.05), comparing the G489C background control mutant to the proline to alanine mutants.

83

CuPhen CuPhen + ATP 70 * * 60 *

50 * * 40 *

30

20

(M) Monomer Percent 10

0 V42C P52A P53A P472A P476A P481A P484A P485A

V42C NTPDase3

Figure 22. P53A and P481A abolish the ATP-induced decrease in cross-linking efficiency.

Cross-linking was performed and the amount of monomer quantified as described in the methods. Values represent the mean percent of monomer remaining ± standard deviation from three separate experiments. Asterisks represent statistical significance (comparing each mutant in the presence and absence of ATP, p ≤ 0.05).

84 induced decrease in CuPhen cross-linking efficiency in the G489C background (Figure 23).

Interestingly, alanine substitution at position 485 in the G489C background resulted in an inability of the mutant to cross-link at G489C by CuPhen, both in the presence or absence of

ATP (Figure 23). Taken together, these data strongly suggest specific proline residues, namely

P53 and P481, facilitate the cross-talk between binding and hydrolysis of substrate at the enzyme active site to movements or reorientations of the TM helices that are necessary to achieve full enzymatic activity of NTPDase3.

85

CuPhen

CuPhen + ATP

140

120

100

80 * * *

60 * * * 40

Percent Monomer (M) Monomer Percent 20

0 G489C P52A P53A P472A P476A P481A P484A P485A

G489C NTPDase3

Figure 23. P481A abolishes the ATP-induced decrease in cross-linking efficiency. Cross- linking was performed and the amount of monomer quantified as described in the methods.

Values represent the mean percent of monomer remaining ± standard deviation from three separate experiments. Asterisks represent statistical significance (comparing each mutant in the presence and absence of ATP, p ≤ 0.05).

86 2. NTPDase3 Transmembrane Domain Polar Residue Mutagenesis

2.1. Rationale for selection and analysis of TM polar residue mutants

Generally, TM α-helices consist of 20-30 hydrophobic amino acids that have a central region rich in aliphatic residues and phenylalanines, and short border regions enriched in tryptophan and tyrosine (176). Strongly polar residues such as arginine, asparagine, , , glutamine, histidine, and lysine are under-represented in TM helices (177,

178). Mutations of these residues often cause adverse effects on enzyme function, suggesting conserved structural and functional roles (179-181). In addition, the non-covalent association of native TM helices is reported to be mediated by several mechanisms, including: side chain packing interactions mediated by a heptad motif of leucine (182); a G-X-X-X-G motif, first described in human glycophorin A (183); and the formation of hydrogen bonds via TMD polar residues (184).

Multiple sequence alignments of the N- and C-terminal TM helices of all known

NTPDase3 sequences to date revealed complete conservation of polar residues serine 33 (S33) and glutamine 44 (Q44), high conservation of threonine 41 (T41) and cysteine 501 (C501), and a much more limited degree of conservation for serine 39 (S39), threonine 490 (T490), and threonine 495 (T495) (Figure 24A). Helical wheel predictions of the TM helices revealed that 3 pairs of polar residues, namely Q44 and T490, S39 and T495, and S33 and C501 are at approximately the same depth in the membrane, and located on the same face of the TM alpha helices (Figure 24B), suggesting that each pair has the potential for inter- or intra-molecular hydrogen bonding (Figure 24C). Therefore, the polar residues were singly mutated to alanine

87

N-terminal TM Helix C-terminal TM Helix A 33 39 41 44 490 495 501 Human 24 IIALVVLLVSIVVLVSITVIQI 45 486 VFVGTLAFFTVAALLCLAFLAYLC 509 Rat IVTLVVLLVSIVVLVTLTLIQI VFMGVLAFFTAIALLCLAFLFYLC Monkey IIALVVLLVSIVVLVSITVIQI VFVGTLAFFTAAALLCLAFLAYLC Mouse IVTLVVLLVSIVVLVTLTLIQI VFMGVLAFFTAIALLCLAFLLYLC Cat IIALVVLFLSVVVLVTITLIQF AFVGTLAFFTAVALLCLAFLVYLC Chicken VVAQAFLLLSVILIIAIAVIQI LFAGLLIFLTVLALLCLFFLVYLC Chimpanzee IIALVVLLVSIVVLVSITVIQI LFAGLLIFLTVLALLCLFFLVYLC Cow IIALVVLLVSIVVLVSITVIQI Dog IIALVVLLLSVVVLVAITLIQF AFLGILAFFTAVALLCLAFLVYLY Platypus IVALVLLLVSVVVLVTITLVQI VFLGLLIISSAVALLCLVFLVYLC Fish FVRFLCTAASIAALVAVVVIQ– VFAGLVFLFSALTIIMVVFVFIIL Frog IFSLLFLIISVSVIFAIAYVQI VFTGLLTLFTVICALCLIVLLILF Horse IVALVFLLLSVVVLVTITVIQI AFRCTVAFFSVVTLLCLAFLVYLC Oppossum VVALVVFLVSVAVLMAITLIQI VFTGVMFFFAGVALLCLLFLVYLC Pufferfish AGVMLLLLASIAALVAVAVIQ– VFAGLVFLFSALTIV------Shrew IITLVVLLVSVVVLVTITIIQM AFWGILIFFMVMAMMCLAFLVYFY Zebrafish VLAAAFMLASIAVIISIAVVQT LFTGLLLLFSVLTIITLTYLVIAL

T490 Q44 B L37 S33 A497 C501 C L30 I40 F504 F494 T41 F493 Potential TM polar V29 L505 residue hydrogen L500 A498 I34 N-TM V36 C-TM bond pairings: L27 Y507 Helix Helix I43 L491 Q44 – T490 V38 V496 V32 L502 S39 – T495 L31 A503 S39 T495 V28 A506 S33 – C501 V42 V35 A492 L499

88 Figure 24. Rationale for NTPDase3 mutations generated and analyzed in this study. Panel

A. Multiple sequence alignment of the N- and C-terminal TM helices in NTPDase3. The polar residues present in human NTPDase3 which were mutated are bolded and underlined. The sequence in GenBank for the cow NTPDase3 terminates prior to the C-terminal TM helix, and is therefore not known and not shown in the figure. Panel B. Helical wheel analysis of human

NTPDase3 TM helices. The polar TM amino acids are represented by filled black circles. The helices are depicted as viewed from the extracellular side of the cell membrane, with the size of the circles decreasing, representing the amino acids increased distance from the cell surface.

Panel C. The potential TM polar residue hydrogen bond pairings. Based on positions on the helical wheel diagram Q44, S39, and S33 on the N-terminal TM helix, are predicted to be on the same face and at the same depth as T490, T495, and C501, respectively (Panel B). Note that intra-molecular, inter-molecular (or both in the case of Q44 which can form 2 hydrogen bonds) interactions are possible for these polar residues present in the NTPDase3 dimer, which includes

2 N-terminal and 2 C-terminal TM helices.

89 and analyzed for the effect of eliminating each polar residue in the TM helices of NTPDase3.

Then the polar residues were mutated to cysteine, both singly and in pairs, in a “free sulfhydryl-

less” NTPDase3 background (C10S/C501S/C509S NTPDase3) to probe possible hydrogen

bonding pairs involving these residues.

2.2. Characterization of the NTPDase3 polar residue to alanine mutants

To explore the structural and/or functional significance of the conserved polar residues in

the N-terminal (S33, S39, T41, Q44) and C-terminal (T490, T495, and C501) TM helices of

NTPDase3, a site-directed mutagenesis strategy was utilized to singly substitute an alanine

residue (a non-polar amino acid) for each of the polar residues, and the mutants were expressed

in COS-1 cells. The “wild-type like” C10S NTPDase3, denoted as WT (C10S), was used as the

background for these mutations to eliminate the artificial dimerization via oxidation of cysteine

10 which is observed when homogenizing COS membranes expressing wt human NTPDase3

(147). The mutants were characterized for enzymatic activity, expression level, and proper

glycosylation and delivery to the cell surface in their native oligomeric state. As seen in Figure

25A and B, alanine substitution caused a significant and substantial decrease in the expression

level of S33, S39, Q44, T495, and C501 as compared to WT (C10S). However, after correcting for expression level, most of the mutants have specific enzymatic activities similar to the WT

(C10S) enzyme (Figure 25C). Exceptions include: S39A, which, although expressed relatively poorly (26% the level of WT (C10S)), exhibits significant increases in Ca- and Mg-dependent nucleotidase activities; T490A, which exhibits small, but significant decreases in Ca-ATPase specific activity and Mg-dependent nucleotidase activities; T495A, which displays relatively

90

1.0 WT (C10S)

A S33A T41A Q44A 0.25 0.5 1.0 B 0.8

* * 0.6

* * 0.4 WT (C10S) * S39A T490A T495A C501A 0.25 0.5 1.0 0.2

Protein Expression Relative

0.0 WT S33A S39A T41A Q44A T490A T495A C501A C10S NTPDase3

CaATPase CaADPase 600 MgATPase MgADPase C 500 * 400 *

300 *

mol/mg/hr) µ * 200 * * 100 * * * * 75 * * * * 50 *

25

Specific( Activity * 0 WT S33A S39A T41A Q44A T490A T495A C501A C10S NTPDase3

91 Figure 25. Alanine substitution of the polar residues causes a decrease in expression level, but similar nucleotidase activities in most mutant NTPDase3 enzymes. Panel A. Western blot analysis of NTPDase3 mutants detected by the KLH1 C-terminal antibody. One microgram

(µg) of COS-1 cell crude membrane protein for each mutant and 0.25 – 1.0 µg of the WT (C10S)

enzyme as a standard curve was loaded onto an SDS-PAGE gel, blotted, and detected and

quantified by chemiluminescence. Panel B. Relative protein expression as compared to WT

(C10S) NTPDase3. The bar graph represents the relative expression level for each mutant calculated from the WT (C10S) standard curve. Values represent the mean ± standard deviation from three separate experiments. Panel C. Activity of NTPDase3 polar residue to alanine substituted mutants, normalized for expression levels as compared to WT (C10S). Nucleotidase activities were assayed using the same amount of COS-1 cell crude membrane protein per assay.

The activities were determined by measuring the amount of inorganic phosphate (Pi) released from nucleotide substrates (ATP or ADP) in the presence of 5 mM CaCl2 or MgCl2 buffer.

Values represent the mean ± standard deviation from three separate experiments. The asterisks represent statistical significance p ≤ 0.05.

92 small, but significant increases in Ca-dependent nucleotidase activities and Mg-ADPase specific

activity, but a significant decrease in Mg-ATPase specific activity; and C501A, which shows

significant increases in Ca-ADPase and Mg-ADPase specific activities. However, the largest

change is seen for the Q44A mutant, which exhibits significant decreases in both Ca- and Mg-

nucleotidase activities.

Previous studies have shown mutation of a polar residue to a non polar residue may result in improper protein processing and/or abnormal biochemical properties of an enzyme (179). Due

to the significant decreases in expression level and changes in nucleotide hydrolysis as compared

to WT (C10S), possible changes in processing and trafficking of the mutants to the cell-surface

were examined by Endo H (107). As seen in Figure 26, all of the mutants migrated at

approximately 70 kDa with a small electrophoretic shift in mobility upon treatment with Endo H.

This data suggests that all the mutants were properly processed and presumably trafficked to the

cell-surface.

The monomeric effects of alanine substitution on the enzyme were monitored by treating

the mutant proteins with Triton X-100. This detergent is known to cause a decrease in enzymatic

activity by dissociation of NTPDase3 oligomers into monomers (147). As seen in Figure 27,

Q44A, unlike the WT (C10S) and other polar residue to alanine mutants, is insensitive to Ca-

dependent Triton X-100 detergent inhibition and the Mg-dependent Triton X-100 inhibition is

greatly attenuated. Thus, alanine substitution at Q44 may result in compromised TM helix

interactions between two different monomers, presumably accounting for the substantial

decrease in Ca- and Mg-dependent nucleotidase activities (as compared to the WT (C10S) in the

absence of detergent), and explaining the attenuation of further decrease in enzymatic activities

after Triton X-100 treatment (Figure 27). Therefore, the highly conserved and strongly polar

93 - - + - - + - - + - - + - - + - - + - - + - - + Endo H 97-

66- Native glycosylation Core protein

45- WT S33A S39A T41A Q44A T490A T495A C501A (C10S)

Figure 26. Alanine substitution of NTPDase3 polar residues are properly processed and trafficked to the cell membrane. WT (C10S) and NTPDase 3 polar reside to alanine substituted mutant proteins were treated with Endo H for 1 hour at 37°C. There are two negative

(-) control samples: the first represents total COS-1 cell membrane protein and the second one is the Triton X-100 extract control (heated at 37°C for 1 hour; neither is treated with Endo H).

NTPDase3 has seven putative glycosylation sites, one being an immature high-mannose glycan sensitive to Endo H, resulting in a shift in the electrophoretic mobility of WT (C10S). Protein bands with electrophoretic mobility similar to that of WT (C10S) are marked as “native glycosylation” since they exhibit similar shifts after Endo H digestion, and this band represents the enzyme that is delivered to the cell surface. A small amount of non-glycosylated “core protein” is seen in the T490A, T495A, and C501A mutants.

94

160 CaATPase CaADPase 140 MgATPase MgADPase 120

100

80

60

40

20

Percent No Triton Control Activity Control Triton No Percent 0 WT S33A S39A T41A Q44A T490A T495A C501A C10S NTPDase3

Figure 27. Alanine substitution at Q44 eliminates or attenuates the detergent inhibition of

Triton X-100 on NTPDase3 enzymatic activities. Triton X-100 is known to cause dissociation of wt NTPDase3 oligomers to monomers, inhibiting enzyme activity. Therefore, effects of the alanine substitutions of the polar residues on the monomeric enzyme were assessed by measuring activities after Triton X-100 solubilization. The nucleotidase activities were performed by

incubating equal amounts of COS-1 cell crude membrane protein in 0.1% Triton X-100 as

described in the methods. The activities are expressed as the percent control activity measured in

the absence of Triton X-100. Values represent the mean ± standard deviation from three separate

experiments. The dashed horizontal line indicates no change of activity in the presence of Triton

X-100.

95 Q44 residue may play a dominant role in driving TM helix associations to form the native, dimeric NTPDase3.

2.3. Characterization of the polar residues singly mutated to cysteine

The introduction of a free sulfhydryl by cysteine substitution may cause potential problems for protein folding and processing (due to inappropriate disulfide bond formation), so each polar residue was first singly substituted to cysteine in the “free sulfhydryl-less”

C10S/C501S/C509S NTPDase3 background (ensuring that the introduced cysteine is the only free sulfhydryl available for chemical reactivity). All these mutants were enzymatically active

(Figure 28A). Once corrected for expression level, only T41C exhibited specific enzymatic activities similar to the sulfhydryl-less background, unlike: S33C, which exhibited a statistically significant increase in Ca- and Mg-dependent specific activities; S39C, which exhibited a statistically significant increase in Ca- and Mg-dependent specific activities; Q44C, which exhibited an ~60% decrease in Ca- and Mg-dependent specific activities; and T490C, T495C, and C501, which all exhibited small, but statistically significant decreases in Ca- and Mg- dependent specific activities (Figure 28A). Interestingly, similar to Q44A, the Q44C mutant exhibited many of the same biochemical properties, including its attenuation of Triton X-100 detergent inhibition, unlike the other mutants (Figure 28B).

96

160 AB600 CaATPase CaATPase

) * CaADPase CaADPase MgATPase 140 MgATPase 500 MgADPase MgADPase 120

400 * * mol/mg/hr 100 µ ( 300 80 * * * * 60 200 * * * * 40 100 * * * * * * 20

Specific Activity * * * * * * * 0 0

C10S/ S33C S39C T41C Q44C T490C T495C C501 No Triton Control Activity Percent C10S/ S33C S39C T41C Q44C T490C T495C C501 C501S/C509S C501S/C509S NTPDase3 NTPDase3 C D + - + - + - + - + - + - + - DTT BME DTT NR

D D

M M

S33C S39C T41C Q44C T490C T495C C501 Q44C

97 Figure 28. Characterization of NTPDase3 polar residues to single cysteine mutants. Panel

A. Activity of NTPDase3 mutants normalized to the expression level of the C10S/C501S/C509S

NTPDase3 background enzyme. Nucleotidase activities were determined as described in the

methods. Values represent the mean ± standard deviation from three separate experiments. The

asterisks represent p ≤ 0.05. Panel B. Effects of cysteine substitution on the monomeric enzyme

activities. Equal amounts of COS-1 cell crude membrane protein were assayed in the presence

of 0.1% Triton X-100 as described in the methods. The activities are expressed as the percent

control activity measured in the absence of Triton X-100. Values represent the mean ± standard

deviation from three separate experiments. The dashed horizontal line indicates no change of

activity in the presence of Triton X-100. Panel C. Western blot analysis for the presence of

spontaneous, intermolecular disulfide bond formation. Equal amounts of COS-1 cell crude

membrane protein were treated with or without 30 mM DTT and boiled for 5 minutes as

described in the methods. The monomer (M) and dimer (D) bands are labeled. Bands migrating

above the dimer are higher order oligomers. Panel D. Analysis of the Q44C inter-molecular

disulfide linked dimer which is resistant to reduction by 30 mM DTT (Panel C). Equal amounts

of Q44C COS-1 cell crude membrane protein was either treated with no reductant (NR), 200 mM

DTT, or 2.86 M BME, and boiled for 5 minutes prior to electrophoresis, as described in the

methods.

98 In order to better understand the contributions the TMD make to the oligomeric structure

of NTPDase3, it is important to elucidate if, and how, the conserved polar residues interact between the N- and C-terminal TM helices by the cysteine substitution/disulfide bond formation strategy. This should allow development of a working model for TMD interactions. To explore the orientation and spatial proximity via the introduced cysteine residues for spontaneous, inter- molecular disulfide bond formation, the mutants were analyzed by SDS-PAGE with and without reduction by 30 mM DTT. As seen in Figure 28C, most of the cysteine mutants form some spontaneously linked disulfide dimers and higher order oligomers, however the extent of this spontaneous cross-linking is small when quantitatively comparing the two monomer bands of the treated versus untreated samples for each mutant (it should be noted that the KLH1 antibody reacts much more strongly with oligomeric forms of NTPDase3 than with monomeric NTPDase3

on Western blots). Furthermore, each of the spontaneously cross-linked cysteine mutants was

sensitive to DTT reduction, except Q44C. To ensure the spontaneous cross-linking observed in

Q44C was indeed disulfide-mediated, the sample was treated with 200 mM DTT or 2.86 M

(20%) β-mercaptoethanol (BME) at 100oC for 5 minutes prior to SDS-PAGE. As seen in Figure

28D, the spontaneously cross-linked Q44C dimer was still somewhat resistant to reduction by

this very high concentration of DTT, but was fully reduced by the very high concentration of

BME. The spontaneous cross-linking of the Q44C mutant demonstrates proper orientation and

close proximity needed for effective hydrogen bonding, and suggests that Q44 in the N-terminal

TM helix of one monomer interacts with Q44 in the N-terminal helix of the other monomer in

dimeric NTPDase3.

99 2.4. Paired cysteine mutagenesis and generation of a model incorporating the inter- and

intra- molecular hydrogen bonding pattern of the conserved polar residues in the

transmembrane helices of NTPDase3

The helical wheel analysis predicted that Q44 and T490, S39 and T495, and S33 and

C501 are on the same helical face and at the same depth in the membrane of their respective TM helices (Figure 24B). Thus, in a dimeric structure, these residues have the potential for inter- or intra-molecular hydrogen bonding with themselves or each other. To test this hypothesis, each of these putative polar residue hydrogen bonding pairs were replaced with cysteine residues and introduced into the “free sulfhydryl-less” C10S/C501S/C509S NTPDase3 background (the

C10S/C509S background for the S33C/C501 mutant). These mutants have only two free cysteines available for sulfhydryl specific cross-linking via copper phenanthroline (CuPhen) oxidation. Each double mutant cysteine pair was enzymatically active (Figure 29). To assess the proximity and orientation of each cysteine residue to one another in the TMD of NTPDase3, and to determine if each pair might preferentially interact via inter- or intra-molecular TM helix hydrogen bonding, each double cysteine mutant was alkylated with maleimide polyethylene glycol-5000 (MalPEG), oxidized with CuPhen, or treated with the combination of CuPhen followed by MalPEG. MalPEG reacts covalently to alkylate free sulfhydryls and will increases the molecular weight of NTPDase3 approximately 5 kDa for each free cysteine alkylated by

MalPEG. Each double cysteine mutant was first alkylated with MalPEG before oxidative cross- linking with CuPhen to allow differentiation of cysteine(s) present as free sulfhydryls as opposed to those being involved in spontaneous, intra-molecular cross-linking.

100

220 CaATPase A 200 CaADPase 180 MgATPase MgADPase 160

mol/mg/hr) 140 µ 120 100 80 60 40

Enzyme Activity( 20 0 C10S/C501S/C509S S39C/T495C Q44C/T490C NTPDase3

250 ) CaATPase B CaADPase MgATPase 200 MgADPase

mol/mg/hr

µ 150 (

100

50

Enzyme Activity 0 C10S/C509S S33C/C501 NTPDase3

Figure 29. Specific activities of NTPDase3 putative hydrogen bonding pair (double

cysteine) mutants. Panel A. Nucleotidase activities of NTPDase3 mutants normalized to the expression level of C10S/501S/C509S NTPDase3 background enzyme. Activity measurements were done as described in the methods and expressed as the mean ± standard deviation from three separate experiments. Panel B. Nucleotidase activities of NTPDase3 S33C/C501 mutant

normalized to the expression level of the C10S/C509S NTPDase3 background enzyme.

Nucleotidase activities were determined as described in the methods and expressed as the mean ±

standard deviation from three separate experiments.

101 As seen in Figure 30A, neither the S33C nor C501 single cysteine mutants exhibit inter- molecular cross-linking (dimer formation) after oxidation with CuPhen. Furthermore, the

S33C/C501 paired cysteine mutant was insensitive to alkylation by MalPEG or oxidation with

CuPhen as the mutant protein migrated as a monomer on the SDS-PAGE gel under all conditions. This suggests that the cysteines present at positions 33 and 501 (in the S33C/C501 mutant) are spontaneously and quantitatively intra-molecularly cross-linked, and therefore unavailable for exogenous sulfhydryl chemical reactions.

In Figure 30B, both the S39C and T495C single cysteine mutants were shown to inter- molecularly cross-link to a relatively small extent, with S39C cross-linking more efficiently than

T495C. When the S39C/T495C double mutant was oxidized with CuPhen, dimers were formed also indicating inter-molecular cross-linking. Oxidation of the S39C/T495C double cysteine mutant followed by alkylation with MalPEG, reiterates the dimer formation and shows an electrophoretic shift for the monomer, indicating the presence of free sulfhydryls even after oxidative cross-linking. This data suggests that S39 and T495 may favor inter-molecular hydrogen bonding either within or between dimeric NTPDase3.

As seen in Figure 30C, the Q44C single cysteine mutant inter-molecularly and quantitatively cross-links with CuPhen (i.e., no monomer remains). The T490C single cysteine mutant was also shown to inter-molecularly cross-link, but with lesser efficiency. When the

Q44C/T490C double cysteine mutant was alkylated with MalPEG alone, 2 bands were seen with shifted mobilities, indicating the presence of free sulfhydryls in both the monomer and the spontaneously cross-linked dimer. After CuPhen cross-linking, at least 4 bands were seen. The lowest band indicates intra-molecular cross-linking, as evidenced by its electrophoretic shift to a

102

AB- + - + + - + CuPhen - + - + + - + CuPhen - - + - + - - MalPEG - - + - + - - MalPEG

-200 kD -200kD

D D

M M S33C S33C/C501 C501 S39C S39C/T495C T495C - + - + + - + CuPhen C - - + - + - - MalPEG

-200 kDa

D

M

Q44C Q44C/T490C T490C

Figure 30. Oxidative cross-linking of NTPDase3 polar residue hydrogen bonding pair

(double cysteine) mutants. Cysteine-substituted protein (0.1 mg/ml) from COS-1 cell crude

membranes was cross-linked by CuPhen and treated with MalPEG as described in the methods.

Panel A. Lack of incorporation of MalPEG and lack of dimer formation indicates quantitative, intra-molecular cross-linking of the S33C/C501 mutant. Panel B. Inter-molecular interactions of the S39C/T495C mutant. Panel C. Inter- and intra-molecular interactions of Q44C/T490C

mutant. The monomer (M) and dimer (D) bands are indicated. Bands migrating above the dimer

are higher order oligomers.

103 position slightly above the monomer, and the lack of any shift seen after alkylation with

MalPEG. The most dense band, second from the bottom, represents the spontaneously and

CuPhen-induced, inter-molecularly cross-linked Q44C dimer, which contains two free sulfhydryls that are alkylated when treated with MalPEG as indicated by electrophoretic mobility shift higher than the CuPhen-induced Q44C dimer. The third band from the bottom represents a dimer which is apparently inter-molecularly cross-linked between both the N-terminal

(Q44C/Q44C) and C-terminal (T490C/T490C) TM helices of two monomers, since alkylation with MalPEG does not change the electrophoretic mobility of the band treated with CuPhen alone. The faint fourth band from the bottom may represent higher order oligomers.

Integrating all the cross-linking data leads to the putative hydrogen bonding pattern of the conserved polar residues as illustrated in Figure 31. The polar residues mutated, the labeling of the monomer 1 helices (denoted as N1 (N-terminal domain) and C1 (C-terminal domain)) and the solid lines representing the inter- and intra-molecular interactions of monomer 1, are bolded.

Monomer 2, containing TM helices denoted N2 and C2, and the dotted lines representing the intra-molecular interactions of monomer 2 are labeled without bolding. The filled black circles represent the polar residues with the circle size decreasing as each amino acid increases in distance through the cell membrane towards the cytosol, away from the extracellular surface (see

Figure 24 legend for a complete description of the representations in Panel A). This model places S39, T495, and, to a lesser extent, T41 on the “outside” of the interacting faces of the TM helices forming dimeric NTPDase3 (i.e., they are not directly involved in making TM polar dimeric contacts). To test this model, which suggests that S39, T495, and possibly T41, are not involved in hydrogen bonding within a dimer, but may instead be involved in higher order oligomer interactions involving multiple dimers, we used site-directed mutagenesis to make

104 A B

N2 S39 C2 Q44 N2 T41 T495 S33 Q44 T490 T490 T490 C2 C501 C501 C1 Q44 T490 Q44 S33 T495 T41 N1 S33C S33C C501 C501

S39 C1 N1

Figure 31. Model of the transmembrane domain helical interactions in the NTPDase3

dimer. Panel A. Helical wheel model depicting the putative inter- and intra-molecular

hydrogen bonding pattern of the conserved polar residues in the TMD of NTPDase3. The polar

residues mutated, the labeling of the helices, and the solid lines representing the inter- and intra-

molecular interactions of monomer 1, denoted as N1 (N-terminal domain) and C1 (C-terminal

domain), are bolded. Monomer 2, containing TM helices denoted N2 and C2, and the dotted lines representing the inter- and intra-molecular interactions are labeled without bolding. The filled black circles represent the polar residues with the circle size decreasing as each amino acid increases in distance into the membrane from the extracellular surface (see Figure 24 legend for a complete description of the representation in Panel A). Data pairing the Q44C mutation with

S39C, T41C, or T495C mutations suggests that S39 and T495 may contribute to interactions involved in forming higher order oligomers (e.g. tetramers formed by interactions between dimers), since they are available for cross-linking dimers into higher order oligomers. T41 may

105 provide specificity of helical association by taking part in the Polar-X-X-Polar (T41-V42-I43-

Q44) NTPDase3 motif, but is not available for sulfhydryl specific cross-linking of dimers into

higher order oligomers, as shown in the model by its relative inaccessibility from the surface of

the dimer helical bundle. Panel B. A 3-D cartoon model depicting the putative inter- and intra-

molecular hydrogen bonding pattern of the conserved polar residues in the TMD of NTPDase3.

Monomer 1, denoted as N1 (N-terminal domain) and C1 (C-terminal domain), is represented as

solid cylinders with polar residues important for dimer interactions underlined and solid lines

representing the helical interactions. Monomer 2, denoted as N2 and C2, is represented as dotted cylinders with dotted lines representing the helical interactions. Q44 is the “lynch-pin” of interactions responsible for dimer formation, forming presumed hydrogen bonds between the

Q44 residues present in each monomer. In addition, Q44 interacts with T490, each strongly polar glutamine being capable of forming two separate hydrogen bonds. S33 and C501 form a strong intra-molecular association within each monomer.

106 Q44C/S39C, Q44C/T41C, and Q44C/T495C double mutants. Our hypothesis was that upon

oxidative cross-linking by CuPhen, these “outside-facing” polar residues, coupled with the

spontaneous, inter-molecularly cross-linked Q44C mutant (revealing the presumed native dimer),

could form higher order oligomers upon oxidation by CuPhen. Each of the double mutants was

enzymatically active (Figure 32). As seen in Figure 33, the Q44C/S39C (and to a lesser extent,

the Q44C/T495C mutant) forms a small amount of higher order oligomers (above 200 kDa), and

these bands do not electrophoretically shift upon the addition of MalPEG, suggesting the absence

of free sulfhydryls. No higher order oligomers were evident after oxidative cross-linking of the

Q44C/T41C mutant. This data suggests that while Q44 is a main player in driving the

association of the TM helices of dimeric NTPDase3, S39, and possibly T495, may contribute to

higher order oligomer formation. T41 does not seem to play a role in promoting higher order

oligomerization, but instead is most likely teamed with Q44 to form a “Polar-X-X-Polar” motif mediating TMD association (185).

2.5. Tween 20 Stimulation of NTPDase3

Generally, the enzymatic activity of the NTPDases is inhibited by most detergents used to solubilize membrane-bound proteins (116). However, during the course of this study, we discovered that incubation of NTPDase3 with 0.1% Tween 20 increases the Mg-dependent nucleotidase activities of the human enzyme by more than 2.5 fold (Figure 34A). In contrast,

Tween 20 was shown to inhibit the enzymatic activity of human NTPDase1 and 2 by about 50 –

70% (Figure 34A), suggesting the stimulation of activity by Tween 20 is selective for

107

140 CaATPase CaADPase 120 MgATPase MgADPase 100

mol/mg/hr) µ 80

60

40

20 Enzyme Activity (

0 C10S/ Q44C S39C/Q44C T41C/Q44CT495C/Q44C C501S/C509S

NTPDase3

Figure 32. NTPDase3 S39C/Q44C, T41C/Q44C, and T495C/Q44C mutant enzymatic activities. Nucleotidase activities were determined as descried in the methods and corrected to the expression level of the C10S/C501S/C509S NTPDase3 background enzyme. Values are expressed as mean ± standard deviation from three separate experiments.

108

--++--++ - - + + - - + + CuPhen - + - + - + - + - + - + - + - + MalPEG

-200 kDa

D

M

Q44C Q44C/S39C Q44C/T41C Q44C/T495C

Figure 33. Analysis for higher order oligomeric structure by oxidative cross-linking of

S39C/Q44C, T41C/Q44C, and T495C/Q44C NTPDase3 double mutants. The mutants were treated with CuPhen and MalPEG as described in the methods. Since, unlike the parent Q44C, a small amount of higher order oligomers are observed in the S39C/Q44C and T495C/Q44C mutants (indicated by asterisks), S39 and T495 may participate in hydrogen bond interactions involved in forming these oligomers. The monomer (M) and dimer (D) bands are labeled and band(s) above 200 kDa represent a tetramer.

109

A 300

250

200

150

100

50

Percent No Tween Control Activity 0 NTPDase1 NTPDase2 NTPDase3

NTPDase Isoenzyme

B 450 400 350

300 250 200

150 100

Percent No Tween Activity No Tween Percent 50

0 Human Mouse Rat NTPDase3

110 Figure 34. The stimulatory effect of Tween 20 is specific for NTPDase3 and conserved

among species. Panel A. Mg+2-ATPase assay of human, cell surface NTPDase isoenzymes in

the presence of 0.1% Tween 20. Equal amounts of COS-1 cell crude protein were pre-incubated

with 0.1% Tween 20 for 5 minutes at 37oC before starting hydrolysis of 0.25 mM ATP for 10

minutes in 5 mM MgCl2 buffer, pH 7.5. The activities are expressed as the percent control

activity measured in the absence of Tween 20. Values represent the mean ± standard deviation

from three separate experiments. The dashed horizontal line indicates no change of activity in

the presence of Tween 20. Panel B. Mg+2-ATPase assay of NTPDase3 from 3 species in the

presence of 0.1% Tween 20. Human, rat, or mouse NTPDase3 was pre-incubated for 5 minutes

o in 0.1% Tween 20 at 37 C before hydrolysis of 0.25 mM ATP for 10 minutes in 5 mM MgCl2 buffer, pH 7.5. The activities are expressed as the percent control activity measured in the absence of Tween 20. Values represent the mean ± standard deviation from three separate experiments. The dashed horizontal line indicates no change of activity in the presence of

Tween 20.

111 + + + + Glutaraldehyde - + - - Digitonin - - + - Tween 20 -- - + Triton X-100

Dimer

Monomer

Figure 35. Tween 20 promotes oligomerization mediated by the TMDs of NTPDase3.

Panel A. Effect of Tween 20 on glutaraldehyde (lysine specific) cross-linking efficiency via the non-transmembrane regions of NTPDase3. WT (C10S) NTPDase3 was diluted to 0.1 mg/ml in

20 mM MOPS, 5 mM MgCl2 buffer, pH = 7.5. WT (C10S) was solubilized with and without

0.1% detergent (digitonin, Tween 20, or Triton X-100) for 10 minutes at 22oC prior to incubation with 5 mM glutaraldehyde at 22oC for 10 minutes. Note less monomer remaining, and more oligomers formed, in the presence of Tween 20, in sharp contrast to the large attenuation of cross-linking observed with Triton X-100.

112 NTPDase3. To confirm this observation, Figure 34B demonstrates that incubation of human,

mouse, and rat NTPDase3 with 0.1% Tween 20 results in a greater than 2.5 fold increase in Mg-

ATPase activity, indicating the stimulatory effect is not species-specific. Therefore, we hypothesized the detergent Tween 20 stimulates the enzyme activity of NTPDase3 by enhancing the oligomeric stability of the enzyme, mediated by the TM helices. Glutaraldehyde (a lysine specific cross-linker) has been frequently used to assess the oligomerization status of NTPDases

(143). NTPDase3 (WT-like C10S) was therefore glutaraldehyde cross-linked in the absence of

any detergent, as well as after treatment with 0.1% of the detergents digitonin, Tween 20, or

Triton X-100, and analyzed on a reducing SDS-PAGE gel. As seen in Figure 35, NTPDase3

treated with digitonin (which maintains activity and native oligomeric structure (116)) resulted in efficient cross-linking into dimers and higher order oligomers. NTPDase3 treated with Tween

20 resulted in more dimer and higher order oligomer formation following glutaraldehyde cross- linking as compared to the NTPDase3 treated with no detergent or treated with digitonin. In contrast, NTPDase3 treated with Triton X-100, a detergent known to disrupt the oligomeric state of NTPDase3, dramatically decreased the cross-linking efficiency of the normally dimeric

NTPDase3. This data suggests Tween 20, unlike Triton X-100 (and most other detergents), enhances NTPDase3 oligomer formation, resulting in increased nucleotidase activity.

Previous studies have shown that detergents elicit their inhibitory effects on the

membrane-bound NTPDases, but do not inhibit the soluble, extracellular domains of these same

enzymes (131). To test the hypothesis that Tween 20 stimulation of NTPDase3 activity may be

mediated by the TM helices, a soluble NTPDase3 (sNTPDase3) construct (lacking the TM

helices and short cytoplasmic N- and C-terminal tails) was treated with Tween 20 (and Triton X-

100) to assess their effect on activity. No change in the absorbance at 630 nm in the Ca-ATPase

113 Malachite green assay was observed for soluble NTPDase3 (Figure 36), indicating the Tween 20

stimulatory effect is indeed mediated by the TM helices of NTPDase3.

The S39C/T495C mutant was used to further test the hypothesis that Tween 20 increases

the enzymatic activity of NTPDase3 by promoting oligomerization of the enzyme at the TMD

level. The S39C/T495C mutant protein (in the free sulfhydryl-less NTPDase3 background) was treated with or without 0.1% Tween 20 or Triton X-100, followed by oxidative cross-linking with 0.5 mM CuPhen for 20 minutes at 37oC. As seen in Figure 37, treatment of the

S39C/T495C mutant with 0.1% Tween 20 followed by CuPhen oxidation resulted in the monomeric enzyme efficiently intra-molecularly cross-linked, and more efficiently inter- molecularly cross-linked to oligomeric forms. To the contrary, 0.1% Triton X-100 treatment of

S39C/T495C followed by CuPhen oxidation resulted in complete attenuation of dimer and higher order oligomer formation, but intra-molecular (within a monomer) cross-linking was still evident as a decrease in the electrophoretic mobility of the monomer. This data suggests the NTPDase3- selective stimulatory effect of Tween 20 might be explained by this detergent promoting oligomerization via the TM helices of NTPDase3, which contain a set of unique, conserved polar residues not present in NTPDase1 or NTPDase2.

114

0.5

0.4

0.3

0.2

Activity

0.1

0.0 NosNTPDase3 Detergent Tween 20 Triton X-100 Control

Figure 36. Detergents Tween 20 and Triton X-100 do not change the enzymatic activity of sNTPDase3. The Ca-ATPase activity of a soluble construct of human NTPDase3 was conducted as described in the methods using the Malachite green assay. The data represented the mean activity ± standard deviation from three separate experiments.

115

Tween 20 Triton X-100 0 0 0.1 0 0 0.1 [Detergent (vol %)] - + + - + + CuPhen

Dimer

Monomer

Figure 37. Effect of 0.1% Tween 20 or Triton X-100 on CuPhen cross-linking efficiency of

the TMDs of a mutant NTPDase3. S39C/T495C NTPDase3 was diluted to 0.1 mg/ml in 50

mM Hepes, pH = 7.5 and solubilized with and without 0.1% Tween 20 or Triton X-100 for 10

minutes at 22oC prior to oxidative cross-linking with or without 0.5 mM CuPhen at 37oC for 20 minutes. Note the stark contrast in results obtained using Tween 20 versus Triton X-100.

116 3. NTPDase3 Alternatively Spliced Variant

Note: This work is the culmination of a project begun and performed to a large extent by

Dr. Patrick Crawford. My role in this work, which resulted in authorship for me on the resulting paper, was to design and perform the necessary experiments requested by the reviewers and to re-write the manuscript accordingly for acceptance for publication.

3.1. Rationale and experimental approach

Alternative splicing occurs as a normal phenomenon in eukaryotes, where it greatly increases the diversity of proteins that are encoded by the genome. In humans, over 80% of gene products are alternatively spliced, including some NTPDases, namely NTPDase1, 2, and 4 (186,

187). The truncated splice variant of NTPDase1 was obtained from a human placental cDNA library, and is not yet fully characterized (188). Three splice variants of NTPDase2 have been reported: NTPDase2α, the full-length, active form of NTPDase2, which has ATPase activity and is properly processed and trafficked to the cell-surface (189); NTPDase2β, the first

NTPDase2 isoform reported (which lacks 23 amino acids in the extracellular domain) (189), and

NTPDase2γ, which is a C-terminally truncated isoform (190). These two variants of NTPDase2

(2β and 2γ) are due to differential splicing of exon 8 that results in expression of inactive proteins lacking a portion of the extracellular domain near ACR5. These splice variants are not fully biosynthetically processed and are retained in the endoplasmic reticulum. Lastly, an active, alternatively spliced form of the intracellular NTPDase4 has been described that differs in only 8 amino acids, but has broader nucleotide substrate specificity (188). Our laboratory discovered

117 and characterized (191) an alternatively spliced form of human NTPDase3, which is described

below.

3.2. NTPDase3β alternatively spliced isoform

NTPDase3β is a C-terminally truncated, alternatively spliced variant of human

NTPDase3, which was discovered by searching the IMAGE (192) clone libraries (IMAGE ID

5174918; GenBank accession BC029869) and subsequently isolated from a human lung cDNA library. The complete 1608 bp cDNA encoding NTPDase3β was sequenced and found to be a novel, isoform of NTPDase3 with a shorter 5`-untranslated region resulting in a truncated protein, maintaining the 10 extracellular cysteines forming 5 disulfide bonds (107), but lacking

apyrase conserved region 5 (ACR5), a potential N-glycosylation site (Asn 454), the C-terminal

TM helix, and the short cytoplasmic C-terminal peptide. The coding region of NTPDase3β is

identical to the full-length NTPDase3α except for two notable exceptions (Figure 38): a single A

to G point mutation at the beginning of the N-terminal TMD (a known single nucleotide

polymorphism in NTPDase3 according to the HapMap database at http://www.hapmap.org/),

which changes isoleucine 24 to valine in the N-terminal TM helix; and a premature truncation of

NTPDase3β immediately after glutamate 452, which is a valine in NTPDase3α. This truncation is generated using an alternative exon in the genomic DNA of NTPDase3 (Figure 39). Thus, both isoforms of NTPDase3 have 11 exons, the first 10 of which are the same. The alternative exons, exons 11 and 12, are the terminal exons of NTPDase3β and NTPDase3α, respectively.

The difference in the 5’-untranslated regions results from NTPDase3β having only nucleotides

59 through 68 of exon 1, presumably due to the use of an alternative promoter.

118 M F T V L T R Q P C E Q A 1 CTCCGCACAGCTAGGAGAAAAGATGTTCACTGTGCTGACCCGCCAACCATGTGAGCAAGC 60

G L K A L Y R T P T V I A L V V L L V S 61 AGGCCTCAAGGCCCTCTACCGAACTCCAACCGTCATTGCCTTGGTGGTCTTGCTTGTGAG 120

I V V L V S I T V I Q I H K Q E V L P P 121 TATTGTGGTACTTGTGAGTATCACTGTCATCCAGATCCACAAGCAAGAGGTCCTCCCTCC 180 ACR1 G L K Y G I V L D A G S S R T T V Y V Y 181 AGGACTGAAGTATGGTATTGTGCTGGATGCCGGGTCTTCAAGAACCACAGTCTACGTGTA 240

Q W P A E K E N N T G V V S Q T F K C S 241 TCAATGGCCAGCAGAAAAAGAGAATAATACCGGAGTGGTCAGTCAAACCTTCAAATGTAG 300 ACR1a V K G S G I S S Y G N N P Q D V P R A F 301 TGTGAAAGGCTCTGGAATCTCCAGCTATGGAAATAACCCCCAAGATGTCCCCAGAGCCTT 360

E E C M Q K V K G Q V P S H L H G S T P 361 TGAGGAGTGTATGCAAAAAGTCAAGGGGCAGGTTCCATCCCACCTCCACGGATCCACCCC 420 ACR2 I H L G A T A G M R L L R L Q N E T A A 421 CATTCACCTGGGAGCCACGGCTGGGATGCGCTTGCTGAGGTTGCAAAATGAAACAGCAGC 480

N E V L E S I Q S Y F K S Q P F D F R G 481 TAATGAAGTCCTTGAAAGCATCCAAAGCTACTTCAAGTCCCAGCCCTTTGACTTTAGGGG 540 ACR3 A Q I I S G Q E E G V Y G W I T A N Y L 541 TGCTCAAATCATTTCTGGGCAAGAAGAAGGGGTATATGGATGGATTACAGCCAACTATTT 600

M G N F L E K N L W H M W V H P H G V E 601 AATGGGAAATTTCCTGGAGAAGAACCTGTGGCACATGTGGGTGCACCCGCATGGAGTGGA 660 ACR4 T T G A L D L G G A S T Q I S F V A G E 661 AACCACGGGTGCCCTGGACTTAGGTGGTGCCTCCACCCAAATATCCTTCGTGGCAGGAGA 720

K M D L N T S D I M Q V S L Y G Y V Y T 721 GAAGATGGATCTGAACACCAGCGACATCATGCAGGTGTCCCTGTATGGCTACGTATACAC 780 ACR4a L Y T H S F Q C Y G R N E A E K K F L A 781 GCTCTACACACACAGCTTCCAGTGCTATGGCCGGAATGAGGCTGAGAAGAAGTTTCTGGC 840

M L L Q N S P T K N H L T N P C Y P R D 841 AATGCTCCTGCAGAATTCTCCTACCAAAAACCATCTCACCAATCCCTGTTACCCTCGGGA 900

Y S I S F T M G H V F D S L C T V D Q R 901 TTATAGCATCAGCTTCACCATGGGCCATGTATTTGATAGCCTGTGCACTGTGGACCAGAG 960

P E S Y N P N D V I T F E G T G D P S L 961 GCCAGAAAGTTATAACCCCAATGATGTCATCACTTTTGAAGGAACTGGGGACCCATCTCT 1020

C K E K V A S I F D F K A C H D Q E T C 1021 GTGTAAGGAGAAGGTGGCTTCCATATTTGACTTCAAAGCTTGCCATGATCAAGAAACCTG 1080

S F D G V Y Q P K I K G P F V A F A G F 1081 TTCTTTTGATGGGGTTTATCAGCCAAAGATTAAAGGGCCATTTGTGGCTTTTGCAGGATT 1140

119 Y Y T A S A L N L S G S F S L D T F N S 1141 CTACTACACAGCCAGTGCTTTAAATCTTTCAGGTAGCTTTTCCCTGGACACCTTCAACTC 1200

S T W N F C S Q N W S Q L P L L L P K F 1201 CAGCACCTGGAATTTCTGCTCACAGAATTGGAGTCAGCTCCCACTGCTGCTCCCCAAATT 1260

D E V Y A R S Y C F S A N Y I Y H L F V 1261 TGATGAGGTATATGCCCGCTCTTACTGCTTCTCAGCCAACTACATCTACCACTTGTTTGT 1320

N G Y K F T E E T W P Q I H F E K E E * 1321 GAACGGTTACAAATTCACAGAGGAGACTTGGCCCCAAATACACTTTGAAAAAGAAGAATA 1380

1381 GCTGATATGAAGCCTGAAGGATGTCTGGAGTCAGCCTGTTCAATACCAGCTTCAGCAATA 1440 1441 ATTAGATTAAACCAATCAATATCCTGCATTAGGCCTCTGACTATCTACTTTTACATAAAT 1500 1501 AAAATTGAGTCATCTACAAAAAAATAAGGATGTCTGGAGACATAAGGACTGTTGGTAGCA 1560 1561 TGTACTCTCTAAGCTTTGCTTCTAATAAATCTAATTACCAGTAAAGTA 1608

Figure 38. cDNA and translated amino acid sequences of NTPDase3β, the truncated splice variant of NTPDase3. The amino acids that differ between NTPDase3β and full-length

NTPDase3α are indicated in bold. The N-terminal transmembrane domain is underlined and the apyrase conserved regions (ACRs) are labeled and shaded.

120 Start 1 68 69 121 122 TM1 250 251ACR1 369 NTPDase3α 12 3 4 NTPDase3β 1 10 11 6364 192 193 311

P-2 P-1 ACR2 ACR3 370 521 522 682 683ACR4 907 908 1181 5 6 7 8

312 463 464 624 625 849 850 KLH111123 ACR1a ACR4a A511Stop Stop ACR5 P-α 1182 1293 1294 1432 1433 2793 91011 12

1124 1179 1180 1374 1375 1608 TM2 KLH1 P-β St op

Figure 39. A diagram of the exon map of the human NTPDase3 gene representing the splicing events that occur. The boxes represent the 12 exons, which are numbered (exon numbers inside ovals), and the introns are depicted as peaks and valleys connecting the exons above and below the boxes. The cDNA sequence is indicated by the dark line with NTPDase3a above and NTPDase3b below the exons. The seven ACRs are marked by crosshatched boxes within the exons. The transmembrane helices (TM1 and TM2) are represented by boxes filled with vertical lines. The two antibodies (KLH1 and KLH11) used are indicated by shaded boxes within the exons.

121 3.3. Expression and enzymatic activities of NTPDase3 isoforms

Ca- and Mg-dependent nucleotidase assays were performed on the crude membrane preparations of COS-1 cells transiently transfected with NTPDase3α or NTPDase3β. ATP and

ADP nucleotide hydrolysis by NTPDase3α was, as expected, detected in the membrane preparations (12, 107, 140). However, there was no measurable ATPase or ADPase activity detected for NTPDase3β in the membrane preparations. COS cell membrane fractions expressing NTPDase3β protein were further tested for activity using alternative nucleotide substrates (AMP, GDP, UDP, UTP, IDP, CDP, and CTP), but no detectable hydrolysis was observed using any of the nucleotide substrates (191). The cell culture media was assayed for nucleotidase activity and NTPDase3 protein, to determine if the truncated form of NTPDase3 was being released as a soluble enzyme after cleavage of the single TM helix, similar to what has been observed for NTPDase5 and 6. Neither nucleotidase activity nor NTPDase3 protein was detected in any of the soluble fractions (191). Western-blot analysis performed utilizing an antibody (KLH11) against an internal peptide sequence (191), demonstrated that the COS cell crude membrane preparations contained NTPDase3β protein. There was a clear difference in the molecular masses of the two isoforms of NTPDase3 as predicted from their respective amino acid sequences (191). The non-glycosylated theoretical molecular mass of the full-length

NTPDase3α isoform is 59,190 Da, compared to 50,748 Da for the non-glycosylated

NTPDase3β isoform.

122 3.4. Co-Transfection of NTPDase3α and NTPDase3β

Membrane-bound NTPDases, NTPDase3 being one, are expressed on the cell-surface

mostly as homodimers. To assess the effect of expressing NTPDase3α and

NTPDase3β together, COS-1 cells were transfected with NTPDase3α, NTPDase3β, the

A511Stop-NTPDase3 mutant, or a combination of the three DNAs, as shown in Table 3. The co-transfected COS cell membranes were characterized based on the expression level of

NTPDase3 protein (Figure 40) and nucleotidase activities (Table 3 and Figure 40). The

A511Stop-NTPDase3 mutant is a fully active, truncated NTPDase3α protein lacking only the cytoplasmic C-terminal sequence (i.e. amino acids 511-529). This NTPDase3 mutant is not recognized by the KLH1 C-terminal antibody, but still reacts with the internal KLH11 antibody.

These properties allowed the A511Stop mutant to be used as a control in the co-transfection experiments.

Co-transfection with NTPDase3β caused a significant decrease in Ca- and Mg-dependent

nucleotidase specific activities that were substantially greater than can be explained by

NTPDase3β simply being an inactive isoform and having no effect on NTPDase3α expression and activity (Table 3). These activities are approximately 40% of the nucleotidase activities

measured for COS cells transfected with NTPDase3α alone (Table 3, compare the “8,0,0” row

with the “4,4,0” row). Smaller, but similar decreases in activity are also seen when NTPDase3α

is co-transfected with NTPDase3β and A511Stop (the “4,2,2” row). More importantly, when the

123 Expression levels relative to NTPDase3α transfected alone Specific activity (µmol/mg/hr, µg cDNA used to transfect COS cells (from Figure 4) uncorrected for expression levels)

Using Using KLH11 KLH1 detected detected active, upper Ca- Mg- NTPDase3α NTPDase3β A511Stop bands bands Ca-ATP ADP Mg-ATP ADP

8 0 0 100 100 203±31 38±3 119±29 32±6

4 0 4 60±14* 82±1** 173±20+ 34±3+ 98±24# 28±4#

4 4 0 30±6* 34±5** 73±7+ 15±2+ 41±9# 14±1#

4 2 2 31±4 49±6 93±4 19±1 56±6 17±1

Table 3. Results of co-transfection of different NTPDase3 proteins in COS cells. This table summarizes the expression level and enzymatic properties of NTPDase3α when co-transfected with either or both the A511Stop-mutant or NTPDase3β. The relative protein expression level and enzyme activity measurements of NTPDase3α in the presence of A511Stop are statistically different than it is in the presence of the same amount of DNA encoding NTPDase3β. The p- values given for this table were calculated by comparing row “4,0,4” with “4,4,0” and each have an n=3. (*p-value <=0.04, **p-value <=0.003, +p-value<=0.004, #p-value<=0.03)

124

KLH1 KLH11

1 2 3 4 5 6 7 8 9 10

NTPDase3α + NTPDase3α A511Stop NTPDase3β

µg NTPDase3α 8 4 4 4 0 8 4 4 4 0 µg NTPDase3β 0 0 4 2 8 0 0 4 2 8 µg A511Stop 0 4 0 2 0 0 4 0 2 0

Figure 40. Western blot analysis of co-transfection experiments using the KLH1 or KLH11 antibody. COS-1 cells were transfected with a total of 8 µg of NTPDase3 cDNA. Primary antibodies used were KLH1 for lanes 1-5 and KLH11 for lanes 6-10. KLH1 is reactive against the C-terminal cytoplasmic amino acid sequence and only detects NPTDase3α. KLH11 is reactive against a portion of the extracellular domain and recognizes NTPDase3α and A511Stop

(active, upper bands) as well as NPTDase3β (inactive, lower bands).

125 A511Stop-NTPDase3 mutant was co-transfected with full-length NTPDase3α, relatively small decreases in specific activity and expression levels were noticed. This was substantially and statistically significantly different in every case than those observed when co-expressing

NTPDase3α with NTPDase3β (compare the “4,0,4” and “4,4,0” rows in Table 3). The comparison of these two rows is the most important aspect of Table 3 and demonstrates a clear difference between a “control”, non-interfering, co-transfected DNA (A511Stop in the “4,0,4” row) from a co-transfected DNA that results in inhibition of expression of a full-length, wild- type enzyme (the NTPDase3β isoform, in the "4,4,0" row). Thus, co-transfection of NTPDase3β causes a specific effect of decreased expression level of NTPDase3α that is not seen in the presence of the active, normally processed, artificially truncated A511Stop-NTPDase3 mutant.

The A511Stop-NTPDase3 mutant only decreases NTPDase3α expression level by about 40%

(60 ± 14 %) (see Table 3 and lanes 1 and 2 and 6 and 7 of Figure 40), while NTPDase3β more dramatically decreases the expression level of NTPDase3α by about 70% (30 ± 6 %), in the presence and absence of the A511Stop-NTPDase3. Therefore, the active and properly processed

A511Stop was used as a control for non-specific effects that might be caused by any other DNA present during co-transfection, affecting the amount of DNA taken up by the COS cells, or the ability of the COS cells to express the transiently transfected NTPDase3α DNA.

The KLH11 antibody is able to detect all NTPDase3 proteins used in this set of experiments (Figure 40). On the basis of size, by Western blot analysis, NTPDase3α and

NTPDase3β, as well as the A511Stop-NTPDase3 and NTPDase3β are distinguished from one another, but NTPDase3α and the A511Stop-NTPDase3 mutant cannot be distinguished from one another (Figure 40). Thus, KLH11 can be used to distinguish between the active forms of

NTPDase3 (NTPDase3α and A511Stop – the upper bands in lanes 6-9 in Figure 40) and the

126 inactive NTPDase3β (the lower bands in lanes 8-10 in Figure 40). Therefore, the data plotted in

Figure 41 is calculated from the nucleotidase activities reported in Table 3, after correcting for the amount of active and properly processed NTPDase3 by dividing the specific activities in

Table 3 by the relative densities of the upper, active band detected by the KLH11 antibody (as depicted in lanes 6-9 of Figure 40 and quantified in Table 3). Thus, as seen in Figure 41, the activity per molecule of fully processed NTPDase3 (NTPDase3α plus A511Stop) is not changed by the presence of either 2 or 4 µg of NTPDase3β DΝΑ (or by the presence of the A511Stop

DNA), suggesting that the co-expression of NTPDase3β does not inactivate NTPDase3α (or

A511Stop NTPDase3α) through a direct protein-protein interaction, as might occur via hetero- dimerization-induced inactivation of NTPDase3α by NTPDase3β contained in the same oligomer. Thus, both NTPDase3α and the active A511Stop mutant exhibit the same nucleotidase activity per molecule expressed, as detected by the upper band reactive with the

KLH11 antibody, resulting in no significant differences between any of the nucleotidase values,

after correction for expression level (see Figure 41, where none of the values are statistically

different from the “8,0,0” row controls). This suggests that the presence of NTPDase3β protein

interferes with the biosynthetic processing of NTPDase3α protein, decreasing the amount of

NTPDase3α that is properly processed and delivered to the cell-surface in an active form, but

does not alter the specific activity of the smaller amount of NTPDase3α that is properly

processed.

127

Figure 41. Percent activity of co-transfections normalized per molecule active protein.

Specific activity was calculated from the amount of inorganic phosphate released during the reaction as µmol Pi released per mg active protein per hour (see Table 3) and normalized to the expression level of active forms of the protein, NTPDase3α and A511Stop, quantified using the upper bands observed in the KLH11 Western blot (see Figure 40, lanes 6-9, and Table 3). This normalized specific activity was used to generate the percentage normalized specific activities, defining the NTPDase3α, KLH11 normalized, specific activity as 100%. The data is expressed as the mean percent NTPDase3a specific activity ± standard deviation for multiple assays performed on three independent experiments.

128 4. NTPDase3 Single Nucleotide Polymorphisms (SNPs)

4.1. Rationale and experimental approach

Single nucleotide polymorphisms (SNPs) are single nucleotide base changes that can

alter the primary structure of a particular protein. Often, these SNPs lead to phenotypic changes

that result from alterations of the structure and/or function of a protein. Human NTPDase3 has

five non-synonymous, coding region SNPs (I24V, R264Q, A496V, L505F, and E440D), which

we constructed and expressed in COS-1 cells to characterize these mutant proteins. The mutant proteins were characterized based on their enzymatic activities, expression level, Triton X-100 detergent sensitivity, and concanavalin A (ConA) and Tween 20 stimulatory effects.

4.2. Characterization of NTPDase3 SNPs

Using site-directed mutagenesis, the five SNPs were made in the wt NTPDase3 background. All the mutants exhibited similar Ca- and Mg-dependent nucleotidase activities, except: I24V, which demonstrated a small, but significant increase in nucleotidase activities;

L505F, which demonstrated a small, but significant increase in nucleotidase activities; and

R264Q, which demonstrated a substantial and significant decrease in nucleotidase activities

(Figure 42).

The Mg-ATPase activity of each mutant was stimulated approximately 2.5 fold by ConA

(Figure 43), inhibited approximately 90% by Triton X-100 (Figure 44) and stimulated more than

2-fold by Tween 20 (Figure 45), similar to the wt NTPDase3 enzyme. Taken together, these

129 CaATPase CaADPase MgATPase 200 MgADPase

150 mol/mg/hr) µ * 100

50 * * Specific Activity ( * 0 WT I24V R264Q E440D A496V L505F NTPDase3

Figure 42. Enzymatic activity of NTPDase3 SNPs. The nucleotidase activities were conducted as described in the methods and corrected for the expression level of WT NTPDase3.

The expression levels for the mutants, relative to expressed wt NTPDase3: I24V, 0.97; R264Q,

1.23; E440D, 1.22; A496V, 1.14; and L505F, 0.86. The data is presented as the mean activity ± standard deviation from three separate experiments. The asterisks represent statistical significance (p ≤ 0.05), comparing the wt NTPDase3 with the SNPs.

130 300

250

200

150

100

50 % no Con A Control

0 wt I24V R264Q E440D A496V L505F NTPDase3

Figure 43. ConA stimulatory effect on NTPDase3 SNPs. The COS cell membrane proteins were treated with ConA and the Mg-ATPase activity determined as described in the methods.

The data is presented as the mean Mg-ATPase activity ± standard deviation from three separate experiments.

131 20 18 16 14 12 10 8 6 4 2

% no Triton X-100 Control 0 wt I24V R264Q E440D A496V L505F NTPDase3

Figure 44. Triton X-100 Detergent effect on NTPDase3 SNPs. The nucleotidase activities were performed by incubating equal amounts of COS-1 cell crude membrane protein in 0.1%

Triton X-100 as described in the methods. The activities are expressed as the percent control activity measured in the absence of Triton X-100. Values represent the mean ± standard deviation from three separate experiments.

132 300

250

200

150

100

50 % no Tween 20 Control 0 wt I24V R264Q E440D A496V L505F NTPDase3

Figure 45. Tween 20 Stimulatory effect of NTPDase3 SNPs. The Mg-ATPase activity was conducted as described in the methods using the Malachite green assay. The data represented the mean ± standard deviation from three separate experiments.

133 data suggest that, with the exception of R264Q that is located near the active site, these SNPs may not be functionally important. The substantial decreases in nucleotidase activities of R264Q could cause alterations in purinergic receptor signaling via less efficient hydrolysis of ATP and

ADP agonists. These human NTPDase3 SNPs were not pursued any further at this time, since those SNPs that are relevant to this thesis (due to their location in the TMD of NTPDase3: I24V,

A496V, and L505F) were not different from the wild-type human NTPDase3, based on this preliminary biochemical characterization.

134 Chapter IV: Discussion

1. Regulation of the Enzymatic Activity of the Membrane-bound, Cell Surface

NTPDases

The most direct and obvious way to regulate the activity of an enzyme is via its active site. However, the recent analysis of the crystal structure of the extracellular domain of rat

NTPDase2 (146), which contains the active site, does not tell the entire story of the enzymatic mechanisms regulating the NTPDases. This is because nucleotide hydrolysis of the membrane- bound, cell surface NTPDases appears to be regulated by a complex system of coordinated movements between (dependent on the oligomeric stability and structure) and within (dependent upon interactions between the enzyme active site and the transmembrane domains) monomers.

To better understand the structural features regulating hydrolysis (and therefore function), it is necessary to elucidate what the major structural elements of the membrane-bound, cell surface

NTPDases are, and how they work together for optimal enzyme function.

Membrane-bound NTPDase3 expressed on the cell surface has three major structural domains: a large extracellular domain containing the active site, a “linker region”, and a transmembrane domain consisting of 2 TM helices per monomer (134). The extracellular domain contains five disulfide bonds that are conserved among all cell surface NTPDases, which when disrupted in NTPDase3 yield five distinct biochemical phenotypes (107). There are also five apyrase conserved regions (ACRs) in the extracellular region, which are common to all members of the NTPDase family (139), and share sequence homology with the actin/heat shock protein/sugar kinase superfamily of proteins. ACR1 and ACR4 were identified as phosphate

135 binding domains in the extracellular active site by homology with similar regions in the actin superfamily of proteins (193). Also by homology with the actin superfamily of proteins, it has been hypothesized that the extracellular lobes of the NTPDases, upon functional state changes

(i.e. nucleotide binding), undergo an interdomain hinge (butterfly-like) motion of the two lobes during nucleotide hydrolysis.

The “linker region” is a stretch of amino acids in the primary sequence of NTPDase3, between the TMD (location of the N- and C-terminal TM helices) and the extracellular domain

(location of the enzyme active site). The linker region contains a set of conserved proline residues (N-terminal: P52 and P53; C-terminal: P472, P476, P481, P484, and P485). Grinthal and Guidotti, using rat NTPDase1, demonstrated an intra-protein signal transduction phenomenon involving ATP binding in the extracellular domain, which resulted in movement of the transmembrane (TM) helices during nucleotide hydrolysis (149). Thus, NTPDase activities and substrate specificities are modulated not only by mutation of amino acids in the active site

(134, 194), post-translational modifications such as glycosylation (143, 190) and N-acetylation

(195) in the extracellular domain, but also by membrane perturbations of the TM helical interactions (137, 169) and modification of TMD cysteine residues by p- chloromercuriphenylsulfonate (pCMPS) (147, 148).

The transmembrane domain (TMD) of NTPDase3 contains two TM helices, one near the

N-terminus and one near the C-terminus, each of which contains a set of conserved polar residues (N-terminal: S33, S39, T41, and Q44; C-terminal: T490, T495, and C501). Most studies regarding the role of the TMD of the membrane-bound, cell-surface NTPDases have focused on oligomerization, since the predominant, active form of these enzymes appears to be a dimer, although higher-order oligomers are also observed, and the soluble NTPDases (which

136 lack TMDs) are monomeric (131, 133, 144). However, no structural data exists, and no site- directed mutagenesis studies have been performed, to perturb and analyze helix interactions in the TMD of NTPDase3.

The objectives of this dissertation were to: 1. elucidate the structural elements that mediate the intra-protein signal transduction between substrate binding at the extracellular domain and movements and/or re-arrangements of the TM helices during nucleotide hydrolysis; and 2. determine how the N- and C-terminal TM helices interact within and between monomers to form native, dimeric NTPDase3. The experimental evidence presented in this dissertation provides novel insights regarding the enzymatic mechanism of NTPDase3 via : 1. the conserved proline residues linking the active site to TMD movements; 2. the conserved polar residues stabilizing the TMD and promoting oligomerization; 3. the selective stimulation of the enzymatic activity of NTPDase3 by the detergent Tween 20; and 4. the possible modulation of purinergic receptor signaling via expression of the alternatively spliced variant of human NTPDase3,

NTPDase3β, or the R264Q NTPDase3 SNP.

2. Structural and Functional Roles of Specific Structural Elements in NTPDase3

2.1. The conserved proline residues are important components of the linkage between

the active site and transmembrane domain movements

Similar to the results obtained using rat CD39/NTPDase1 (149), mutations in both TM helices of human NTPDase3 near the extracellular surface (V42C and G489C in NTPDase3) cross-linked efficiently in the presence of CuPhen. In the presence of ATP, the CuPhen-induced

137 cross-linking efficiency of the V42C and G489C NTPDase3 mutants was attenuated, resulting in

an increase in the amount of monomer remaining after treatment. Disulfide cross-linking of

V42C and G489C via CuPhen also inhibited the nucleotidase activity of the mutants. This data suggests the importance of the TM helix rearrangements and/or movements during nucleotide hydrolysis that facilitate full enzymatic activity and may be generalized to all cell surface

NTPDases.

When proline to alanine mutations were added to the cross-linking “sensor” background

NTPDase3 enzymes (V42C and G489C), alanine substitution at positions P53 and P481 in the

V42C sensor background caused a substantial decrease in activity, while a significant, but smaller decrease in activity was noted for P484 and P485. Alanine substitution at positions P53 and P481 in the G489C sensor background resulted in a substantial decrease in activity, while the

P484A mutant exhibited a smaller decrease in activity. This data suggests P53, P481, and P484, which are completely conserved in all sequenced NTPDase3 enzymes, but not in all other

NTPDases, are needed for optimal enzymatic activity of NTPDase3.

Assessment of the proline to alanine mutations for possible “uncoupling” effects revealed that mutations at positions P53 and P481 in the V42C background were indeed capable of uncoupling ATP binding at the active site to reorientations or movements of the TM helices, as assessed by abolishment of the ATP-dependent decrease in TM cross-linking efficiency. In the

G489C background, proline to alanine substitution at position P481 also caused this effect.

Surprisingly, alanine substitution at position P485 abolished the CuPhen-induced cross-linking of G489C both in the presence and absence of ATP, although this was not true for cross-linking of V42C after alanine substitution at the same P485 residue. This suggests that the placement of the cysteine residue used as the cross-linking sensor is important for the ability to measure the

138 effects of the proline mutations, and that such proline to alanine mutations may affect the TM

helix closer to the mutation more than the TM helix on the other end of the protein. Thus,

alanine substitution at P485, which is only 4 amino acids away from the G489C mutation used as

a sensor in the C-terminal helix, changes the orientation or freedom of movement of the C-

terminal TM helix of NTPDase3 in a way that abolishes intermolecular, CuPhen-induced cross-

linking at G489C, but not cross-linking at V42C, which is located in the N-terminal TM helix.

Thus, consistent with our hypothesis, some proline to alanine mutants exhibited a decrease in enzymatic activities, an “uncoupling” of the interactions between the active site and

TMD, or both. Importantly, other proline to alanine mutants also located in the linker region had little, if any, effect, demonstrating that the results obtained are not simply non-specific artifacts, in the sense that any proline to alanine substitution in this area of the protein would result in decreases in activity and changes in ATP-induced cross-linking efficiencies as a result of adverse effects on protein folding or processing. Only alanine substitution at positions P53 and P481 resulted in both decreases in nucleotidase activities and abolishment of the decreased cross- linking efficiency of the TM helices in the presence of ATP, suggesting that these two proline residues are important components of the linkage between movements of the extracellular and

TM domains necessary for optimal function of NTPDase3.

2.2. The conserved polar residues stabilize transmembrane domains and promote

oligomerization

The significant decreases in expression levels of S33A, S39A, Q44A, T495A, and C501A mutants were not surprising, since similar polar-to-apolar amino acid substitutions in the TMDs

139 of several unrelated proteins are disease-phenotypic mutations, often accompanied by adverse effects on processing and normal expression of the mutant proteins (196). However, each of the

NTPDase3 polar residues to alanine mutants was properly glycosylated and apparently trafficked to the cell surface, as suggested by Endo H deglycosylation analyses. Somewhat surprising, most of the polar residue to alanine mutations, once corrected for their lower expression levels, did not adversely affect enzymatic activity. Taken together, these data suggest that the conserved polar residues in the TMD of NTPDase3 are important for proper protein folding and normal protein expression, presumably via TM hydrogen bonding.

Interestingly, the Q44A and Q44C mutants exhibited similar enzymatic characteristics, specifically, a decrease in nucleotidase activities, expression level, and the relative insensitivity to Triton X-100 detergent inhibition (Triton X-100, as well as many other detergents, has been shown to inhibit activity of NTPDase3 as well as most other cell surface NTPDases (116)). In fact, Ca-dependent activities actually increased for the Q44 mutants, probably due to unmasking of active sites by making a sub-population of sealed COS membrane vesicles leaky to the nucleotide substrates. The Mg-dependent activities are decreased by Triton, but this inhibitory effect is greatly attenuated as compared to the WT (C10S) sample and the other polar residue mutants. Thus, it seems there may be either a second component of the inhibitory effect of

Triton X-100 that is detected when assaying for Mg-nucleotidase activities, or a stabilizing effect on NTPDase3 mediated by calcium. Nevertheless, the insensitivity of the Q44A mutant to

Triton inhibition suggests that alanine substitution at Q44 may result in compromised TM helix interactions between two different monomers, presumably accounting for the approximately

50%-70% decrease in specific nucleotidase activities (as compared to the WT (C10S), both in the absence of detergent), and explaining the attenuation of further decreases in enzymatic

140 activities after Triton X-100 treatment (Triton X-100 is known to dissociate the oligomeric

structure of several NTPDases, including human NTPDase3). In addition, the spontaneous,

intermolecular disulfide bond formation observed in the Q44C mutant (resistant to even very

high concentrations of DTT, but could be reduced by high concentrations of the more

hydrophobic reduction agent, BME) most likely limits the dynamic motions of the TM helices.

This decreases activity, since both appropriate hydrogen bonding and movement of the TM

helices during nucleotide hydrolysis are necessary to achieve full enzymatic activity.

Taken together, this data is consist with results and conclusions of Sal-Man et al (184,

185), suggesting the existence of a Polar-X-X-Polar (T41-X-X-Q44) TM-TM association motif in the N-terminal helix of NTPDase3. In that study, Sal-Man et al demonstrated the importance of a strong polar residue present in this motif to drive the association of the TM helices in the E.

coli Tar-1 receptor. We conclude that Q44 of one monomer is in close proximity and in the

correct orientation to the Q44 of the other monomer in a dimer, and that the N-terminal TM

helices of dimeric NTPDase3 interact efficiently near the extracellular surface of the cell

membrane. The current data demonstrating the importance of the strongly polar Q44 residue,

which is located at the interface of the N-terminal TM helix with the extracellular portion of

NTPDase3, is also consistent with data published by Grinthal and Guidotti (149) using rat

NTPDase1. In that study, a cysteine substitution and oxidative cross-linking strategy was also

utilized, and identified strong inter- and intra-molecular interactions between the TM helices of

rat NTPDase1 involving residues near the extracellular surface of the membrane, and further

demonstrated that disulfide bond formation between two TM helices decreased enzymatic

activity of NTPDase1.

141 Our laboratory previously noted the attenuation of Triton X-100 inhibition following site-

directed mutagenesis of lysine 79 in the extracellular domain of NTPDase3 (145). Lysine 79 is

conserved in all membrane-bound, cell surface NTPDases, but not in the soluble, monomeric or

intracellular NTPDases. It is also located in a region hypothesized to stabilize oligomerization

(between ACR1 and a glycosylation site, N81) via the extracellular domain. Mutation of lysine

79 to any amino acid besides arginine (which maintains the positive charge) rendered NTPDase3

insensitive to Triton inhibition. Therefore, it is likely that, in addition to the highly conserved,

and strongly polar, Q44 in the TMD, amino acid residues in the extracellular portion of

NTPDase3 also contribute to interactions that enable and stabilize dimerization of NTPDase3.

Our laboratory previously determined that the dominant oligomeric structure of

NTPDase3 is a dimer, with higher order oligomers also possible, including a tetramer, suggested

to be a “dimer of dimers” (147). In that study, it was noted that the interactions between two

monomers in each dimer is stronger than the interactions between the two dimers forming the

tetramer (147). The cysteine substitution/CuPhen cross-linking strategy used in this study

revealed: 1. Q44C appears to form strong intermolecular disulfides with Q44C and, being a

strongly polar residue with the ability to form two hydrogen bonds, Q44C also formed

intramolecular disulfides with T490C; 2. S33C and C501 appear to spontaneously and

quantitatively form intramolecular disulfides with each other; and 3. S39C and T495C

inefficiently form some intermolecular disulfides, resulting in higher order oligomers. This

cross-linking data suggests a putative hydrogen bonding pattern of the conserved polar residues

that places S39 and T495 on the “outside” of the TM helices forming dimeric NTPDase3.

Additional cross-linking data suggested S39 and T495 may be involved in higher order oligomer interactions (e.g., to form the “dimer of dimers” (147)).

142 These cross-linking results suggest that Q44 is the “lynch-pin” residue in the association

of the N-terminal TM helixes to form dimeric NTPDase3. S39 (and possibly T495) may form intermolecular hydrogen bonds to promote interactions that facilitate formation of the tetrameric

and higher order oligomeric forms of NTPDase3. T41, due to its inability to form higher order

oligomers when combined with Q44C, is not likely to be directly involved in higher order

oligomer formation, but instead may work in concert with Q44 to provide the specificity of the

N-terminal helix associations via a “Polar-X-X-Polar” (T41-X-X-Q44) TM-TM association

motif (197).

Lastly, during the course of this study, the detergent Tween 20 was shown to exhibit an

isoenzyme specific, species-conserved, stimulatory effect on the nucleotidase activity of

NTPDase3. This stimulation was not observed when a soluble construct, comprising the

extracellular domain of human NTPDase3, was treated with Tween 20 under identical

conditions. This discovery of the stimulatory effect of Tween 20 on full-length NTPDase3 may

facilitate future structural studies and serve as a useful tool for identifying the presence and

importance of NTPDase3 in various tissues and processes. Considering this stimulatory effect is

isoenzyme-specific (not observed in the closely related human NTPDase1 or NTPDase2), it is

not unlikely that the unique TMD hydrogen bond interactions mediated by the conserved polar

residues of NTPDase3 promote the association of the TM helices in the presence of Tween 20. It

should be noted that the conserved NTPDase3 TMD polar residues examined in this dissertation

are not conserved in other membrane-bound NTPDases, suggesting that the results reported are

specific for NTPDase3, and that other NTPDases have different mechanisms and amino acids facilitating TM helix interactions and TMD oligomeric associations. This most likely contributes to the differential susceptibility to membrane perturbation (e.g., detergent activation

143 and inactivation) (136, 137) and substrate specificities (119, 170) seen in different family

members of the NTPDase enzymes.

2.3. NTPDase3β: A possible modulator of nucleotidase activity and purinergic receptor

signaling

No activity was found in the media or membranes of COS cells transfected with truncated

NTPDase3β. Co-transfection of full-length NTPDase3α with NTPDase3β decreased the

expression level and enzymatic activity of NTPDase3α detected in the COS cell membranes.

Since NTPDases are known to form active homo-oligomers (198), it is possible that the inactive,

truncated NTPDase3β could form hetero-oligomeric structures with the active, full-length

NTPDase3α isoforms, causing a decrease in the enzymatic activity of NTPDase3α, as suggested

by Mateo et al (190). However, the activity per molecule of full length NTPDase3α, as detected

by the NTPDase3α specific KLH1 antibody, did not decrease in co-transfection experiments

including NTPDase3β. In addition, Endo H deglycosylation revealed that NTPDase3β is not

delivered to the cell surface where it could form hetero-oligomers with NTPDase3α. This is in

contrast to what was found for the P2X7 receptor, which has identical membrane topology to the

cell surface NTPDases (199), and for the human serotonin receptor 5-HT3A (200), which both form hetero-oligomers containing both full length and truncated proteins. In contrast,

NTPDase3β appears to be interfering with the processing of the full-length isoform (as judged by the large decrease in expression levels of the NTPDase3α when co-transfected with

NTPDase3β). To ensure this was not a non-specific effect of the presence of another cDNA co- expressed by the same COS cells the active, A511Stop NTPDase3 mutant (missing just the non- critical C-terminal cytoplasmic tail) was co-transfected with NTPDase3α, supporting the

144 differential effect. It also seems unlikely that the effect of the co-transfection of NTPDase3β on

NTPDase3α expression is mediated via transcriptional means. Thus, we speculate that the

physiological role for the enzymatically inactive, truncated isoform of NTPDase3, NTPDase3β,

might be to decrease the ATPase and ADPase activity at the cell surface by decreasing the

synthesis and cell surface delivery of full-length, active NTPDase3α. In vivo, this would lead to

extended activation of P2 purinergic receptors by extracellular nucleotides, and increased P2 signaling in cells expressing the truncated isoform of NTPDase3, NTPDase3β.

2.4. Biochemical characterization of NTPDase3 single nucleotide polymorphisms

Other than significant decreases in the nucleotidase activity of R264Q, the NTPDase3

SNPs exhibit similar biochemical characteristics to wt NTPDase3. However, if R264Q

NTPDase3 is expressed instead of wt NTPDase3, this would result in slower hydrolysis of extracellular nucleotide agonists, leading to extended activation of P2 receptors by extracellular nucleotides, which may lead to a distinct phenotype.

3. Summary and Future Directions

This dissertation provides new insights regarding the structural elements that modulate

the enzymatic activity and mechanism of NTPDase3, as well as other membrane-bound, cell

surface NTPDases. Specifically, the findings extend our knowledge of two key structural

domains of NTPDase3: 1. the linker region, which contains a set of conserved proline residues

that mediate the intra-protein signal transduction between substrate binding in the extracellular

145 domain and movements and/or rearrangements of the TM helices; and 2. the TMD, which contains a set of conserved polar residues that provide NTPDase3 with an isoenzyme specific hydrogen bonding pattern, facilitating TM helix association, and mediating differential susceptibility to membrane perturbation. This body of work broadens our understanding of how the enzymatic activity of the membrane-bound, cell surface NTPDases is regulated by associations between (affecting the stability of the oligomeric structure of the NTPDase) and within (modulating interactions between the extracellular and TM domains) monomers to regulate nucleotide hydrolysis. These findings suggest the following experiments:

3.1. Determine if a polyproline II helix exists in the linker region of NTPDase3 and

identify the specific amino acids involved in its formation/stability

The conserved proline residues in the linker region of NTPDase3 constitute a proline-rich region that mediates interactions between the extracellular and TM domains, suggesting the possible existence of a polyproline II (PPII) helix. Typically, PPII helices are shorter than five amino acids, contain multiple proline resides, and often contain positively charged amino acids.

(201-203). These basic characteristics are consistent with the amino acid sequences of the linker regions of NTPDase3 that contain the functionally important P53 and P481 residues (N-terminal:

QEVLPP53GLK; C-terminal: QIPAESPLIRLP481IEPP). One approach to investigate this hypothesis is to make synthetic peptides of varying lengths and amino acid compositions to confirm the existence of this putative secondary structure, as well as to determine by mutagenesis the amino acids critical for formation of this secondary structure (in addition to P53

146 and P481), which may be involved in forming the hypothesized PPII helix in the N- and C-

terminal linker regions of human NTPDase3.

3.2. To further define the structural motifs that mediate oligomerization via the

transmembrane domain of the NTPDases

It is well established that the membrane-bound, cell surface NTPDases form homo-

oligomers from dimers up to tetramers (144). In this dissertation, data was presented suggesting

a Polar-X-X-Polar (Q44-X-X-T41) TM-TM motif existed in the N-terminal TM helix of

NTPDase3. There are likely other, non-polar, TM helix interactions involved in TMD

oligomerization and stability. One possibility is a putative leucine zipper motif consisting of L30

and L37 in the N-terminal TM helix of human NTPDase3. According to the helical wheel

analysis, L30 and L37 are on the same face of the TM helix as the presumed Polar-X-X-Polar

motif, possibly giving rise to the hydrophobic pocket making the Q44C mutant dimers

particularly resistant to DTT reduction. Considering these leucine residues are more conserved among the different membrane-bound NTPDases, the putative leucine zipper motif may mediate

TM helix interactions that are not specific for individual NTPDases (drive hetero-

oligomerization), while the Polar-X-X-Polar motif provides specificity for native, homo-

oligomeric NTPDase3.

In addition, the C-terminal transmembrane helix of NTPDase3 contains a set of

conserved phenylalanine residues (F487, F493, F494, F504), which may also promote non-polar

TM helix interactions, possibly by forming a “phenylalanine zipper” utilizing F487 and F494

(204-206).

147

3.3. Further biochemical characterization of the single nucleotide polymorphisms of

NTPDase3

Some SNPs are known to be disease phenotypic (207), and considering the lack of genetically modified animal models, as well as the lack of small molecule specific inhibitors of

NTPDase3, expression and further biochemical characterization of SNPs found in NTPDase3 may lead to a better understanding of the physiological function(s) of NTPDase3. Especially interesting is the SNP that is very close to the active site, R264Q. Of course, genetic studies showing a linkage between a disease phenotype and a particular SNP would be instrumental in determining the physiological function(s) of NTPDase3. Future generation and characterization of an NTPDase3 knockout mouse may indicate physiological functions of NTPDase3, and suggest phenotypes likely to occur with the lower activity R264Q NTPDase3 SNP.

148 Chapter V: References

(1) Sawai, H., and Orgel, L. E. (1975) Prebiotic peptide-formation in the solid state. III.

Condensation reactions of glycine in solid state mixtures containing inorganic

polyphosphates. J Mol Evol. 6, 185-97.

(2) Cairns-Smith, A. G. (1985) The first organism. Sci Am 252, 74-82.

(3) Sigel, H. (1992) Have adenosine 5'-triphosphate (ATP4-) and related purine-nucleotides

played a role in early evolution? ATP, its own 'enzyme' in metal ion facilitated

hydrolysis. Inorgan Chim Acta 198-200, 1-11.

(4) Waldrop, M. M. (1989) Did life really start out in an RNA world? Science. 246, 1248-9.

(5) Fiske, C. H., and Subbarow, Y. (1929) Phosphorus compounds of muscle and liver.

Science 70, 381-382.

(6) Lohmann, K. (1929) Uber die Pyrophosphatfraktion im Muskel. Naturwissenschaften 17.

(7) Kalckar, H. M. (1969) Biological Phosphorylations: Development of Concepts., Prentice

Hall, Englewood Cliffs, NJ.

(8) Lippman, F. (1941) Metabolic generation and utilization of phosphate bond energy.

Enzymology 1, 99.

(9) Glynn, I. M. (1968) Membrane and cation transport. Br. Med.

Bull. 24, 165-169.

(10) Chaudry, I. H. (1982) Does ATP cross the cell plasma membrane. Yale J. Biol. Med 55,

1-10.

(11) Drury, A. N. a. S.-G., A. (1929) The physiological activity of adenine compounds with

special reference to their action upon the mammalian heart. Journal of Physiology 68,

213-237.

149 (12) Smith, T. M., and Kirley, T. L. (1998) Cloning, sequencing, and expression of a human

brain ecto-apyrase related to both the ecto-ATPases and CD39 ecto-apyrases. Biochim.

Biophys. Acta 1386, 65-78.

(13) Bennett, D. W., and Drury, A. N. (1931) Further observations relating to the

physiological activity of adenine compounds. J. Physiol. (Lond) 72, 288-320.

(14) Gillespie, J. H. (1934) The biological significance of the linkages in adenosine

triphosphoric acid. J. Physiol 80, 345-359.

(15) Burnstock, G. (1972) Purinergic Nerves. Pharmacol Rev 24, 509-581.

(16) Burnstock, G. (1997) The past, present, and future of purine nucleotides as signalling

molecules. Neuropharmacol. 36, 1127-1139.

(17) Lindler, F., and Rigler, R. (1931) Uber die Beeinflussung der Weite Der

Herzkranzgefasse durch Produkte des Zellkernstoffwechsels. Pflugers Arch 226, 697-

708.

(18) Hoffman, W. S. (1930) The isolation of crystalline adenine nucleotide from blood. J.

Biol. Chem. 63, 675.

(19) Drury, A. N. (1936) The physiological activity of nucleic acid and its derivatives.

Physiol. Rev. 16, 292-325.

(20) Moir, T. W., and Downs, T. D. (1972) Myocardial reactive hyperemia: Comparative

effects of adenosine, ATP, ADP, and AMP. Am. J. Physiol. 222, 1386-1390.

(21) Baddiley, J., Michelson, A. M., and Todd, A. R. (1949) Nucleotides. Part II. A synthesis

of adenosine triphosphate. J. Chem. Soc., 582-586.

(22) Rohstein, A., and Meier, R. (1948) The relationship of he cell surface to metabolism I.

Phosphates in the cell surface of living yeast cells. J. Cell. Comp. Physiol. 32, 77-95.

150 (23) Clarkson, E. M., and Maizels, M. (1952) Distribution of phosphate in human

erythrocytes. J. Physiol. (Lond) 116, 112-128.

(24) Acs, G., Ostrowski, W., and Straub, F. B. (1954) Adenylpyrophosphatase activity on he

surface of ascites tumor cells. Acta Physiol. Hung. 6, 261-263.

(25) Engelhardt, W. A. (1958) Enzymes as structural elements of physiological mechanism.

Proc. Int. Symp. Enzyme Chem. 2, 163-166.

(26) Wenkstern, T. W., and Engelhardt, W. A. (1959) The adenosine polyphosphatase

localized on the surface of nucleated red corpuscles. Folia Haematol. Int. Mag. Kin.

Morphol. Blutforsch. 76, 422-431.

(27) Dale, H. (1935) Pharmacology of nerve endings. Proc R Soc Med 28, 319-332.

(28) Holton, F. A. P., and Holton, P. (1953) The possibility that ATP is a transmitter at

sensory nerve endings. J. Physiol. 119 (Supple.), 50P-51P.

(29) Holton, F. A. P., and Holton, P. (1954) The capillary dilator substances in dry powder of

spinal roolts. J. Physiol. 126, 124-140.

(30) Hillarp, N. A., Hogberg, B., and Nilson, B. (1955) Adenosine triphosphate in the adrenal

medulla of the cow. Nature 176, 1032-1033.

(31) Blaschko, H., Born, G. V. R., D'Lorio, A., and Eade, N. R. (1956) Observations on the

distribution of catecholamines and adenosine triphosphate in the bovine adrenal medulla.

J. Physiol. (Lond) 133, 548-557.

(32) Carlsson, A., Hillarp, N. A., and Hokfelt, B. (1957) The concomitant release of adenosine

triphosphate and catecholamines from the adrenal medulla. J. Biol. Chem. 227, 243-252.

151 (33) Douglas, W. W., and Poisner, A. M. (1966) Evidence that the secreting adrenal

chromaffin cell releases catecholamines directly from ATP-rich granules. J. Physiol.

(Lond) 183, 236-248.

(34) Burnstock, G., Campbell, G., Satchell, D. G., and Smythe, A. (1970) Evidence that

adenosine triphosphate or a related nucleotide is the transmitter substance released by

non-adrenergic inhibitory nerves in the gut. Br. J. Pharmacol. 40, 668-688.

(35) Su, C., Devan, J. A., and Burnstock, G. (1971) [3H]-Adenosine triphosphate: release

during stimulation of enteric nerves. Science 173, 337-339.

(36) Burnstock, G. (1971) Neural nomenclature. Nature 229, 282-283.

(37) Burnstock, G., and Knight, G. E. (2004) Cellular distribution and functions of P2 receptor

subtypes in different systems. Int Rev Cytol 240, 31-304.

(38) Kam, W., Clauser, E. M., Kim, Y. S., Kan, Y. W., and Rutter, W. J. (1985) Cloning,

sequencing, and chromosomal localization of human term placental alkaline phosphatase

cDNA. Proc Natl Acad Sci U S A 82, 8715-8719.

(39) van Driel, I. R., Wilks, A. F., Pietersz, G. A., and Goding, J. W. (1985) Murine plasma

cell membrane antigen PC-1: molecular cloning of cDNA and analysis of expression.

Proc Natl Acad Sci U S A 82, 8619-8623.

(40) Misumi, Y., Ogata, S., Hirose, S., and Ikehara, Y. (1990) Primary structure of rat liver 5'-

nucleotidase deduced from the cDNA presence of the COOH-terminal hydrophobic

domain for possible post-translational modification by glycophospholipid. J. Biol. Chem.

265, 2178-2183.

(41) Wang, T. F., and Guidotti, G. (1996) CD39 is an ecto-(Ca2+, Mg2+)-apyrase. J. Biol.

Chem. 271, 9898-9901.

152 (42) Burnstock, G. (1996) Purinoceptors: Ontogeny and Phylogeny. Drug Development

Research 39, 204-242.

(43) Yegutkin, G. G. (2008) Nucleotide- and nucleoside-converting ectoenzymes: Important

modulators of cascade. Biochim Biophys Acta. 1783, 673-94. Epub

2008 Feb 12.

(44) Burnstock, G. (2007) Physiology and pathophysiology of purinergic neurotransmission.

Physiol. Rev. 87, 659-797.

(45) Zimmermann, H. (1996) Biochemistry, localization and functional roles of ecto-

nucleotidases in the nervous system. Prog. Neurobiol. 49, 589-618.

(46) Sorensen, C. E., and Novak, I. (2001) Visualization of ATP release in pancreatic acini in

response to cholinergic stimulus. Use of fluorescent probes and confocal microscopy. J.

Biol. Chem. 276, 32925-32932.

(47) Marcus, A. J., Broekman, M. J., Drosopoulos, J. H., Islam, N., Pinsky, D. J., Sesti, C.,

and Levi, R. (2003) Heterologous cell-cell interactions: thromboregulation,

cerebroprotection and cardioprotection by CD39 (NTPDase-1). J Thromb Haemost 1,

2497-509.

(48) Zimmermann, H. (1994) Signalling via ATP in the nervous system. Trends Neurosci. 17,

420-6.

(49) Nagy, A., Backer, R. R., Morris, S. J., and Whittaker, V. P. (1976) The preparation and

characterization of synaptic vesicles of high purity. Brain Res 109, 285-309.

(50) von Euler, U. S., and Lishajko, F. (1963) Catecholamines and adenosine triphosphate in

isolated adrenergic nerve granules. Acta Physiol Scand 59, 495-496.

153 (51) Stjarne, L., and Lishajko, F. (1966) Comparison of spontaneous loss of catecholamines

and ATP in vitro from isolated bovine adrenomedullary, vesicular gland, vas deferens,

and splenic nerve granules. J Neurochem 13, 1213-1216.

(52) De Potter, W. P., Smith, A. D., and Schaepdryver, D. (1970) Subcellular fractionation of

splenic nerve: ATP, chromogranin A, and dopamine β-hydroxylase in noradrenergic

vessicles. Tissue Cell 2, 529-546.

(53) Helle, K. B., Lagercrantz, H., and Stjarne, L. (1971) Biochemistry of catecholamine

storage: some similarities between whole sympathetic nerve trunk vesicles and the

membranes of the adrenomedullary vessicles. Acta Physiol Scand 81, 565-567.

(54) Lagercrantz, H., and Stjarne, L. (1974) Evidence that most noradrenaline is stored

without ATP in sympathetic large dense core nerve vesicles. Nature 249, 843-844.

(55) Burnstock, G. (1995) Noradrenaline and ATP: cotransmitters and neuromodulators. J

Physiol Pharmacol 46, 365-84.

(56) Bodin, P., and Burnstock, G. (2001) Purinergic signalling: ATP release. Neurochem Res

26, 959-969.

(57) Schwiebert, E. M., and Zsembery, A. (2003) Extracellular ATP as a signaling molecule

for epithelial cells. Biochem. Biophys. Acta. 1615, 7-32.

(58) Bodin, P., and Burnstock, G. (2001) Evidence that release of adenosine triphosphate from

endothelial cells during increased shear stress is vesicular. J Cardiovasc Pharmacol. 38,

900-8.

(59) Gerasimovskaya, E. V., Ahmad, S., White, C. W., Jones, P. L., Carpenter, T. C., and

Stenmark, K. R. (2002) Extracellular ATP is an autocrine/paracrine regulator of hypoxia-

induced adventitial fibroblast growth. Signaling through extracellular signal-regulated

154 kinase-1/2 and the Egr-1 transcription factor. J Biol Chem. 277, 44638-50. Epub 2002

Sep 18.

(60) Roman, R. M., and Fitz, J. G. (1999) Emerging roles of purinergic signaling in

gastrointestinal epithelial secretion and hepatobiliary function. Gastroenterology. 116,

964-79.

(61) Hoebertz, A., Arnett, T. R., and Burnstock, G. (2003) Regulation of bone resorption and

formation by purines and pyrimidines. Trends Pharmacol Sci. 24, 290-7.

(62) Vassort, G. (2001) Adenosine 5'-triphosphate: a P2-purinergic agonist in the

myocardium. Physiol Rev. 81, 767-806.

(63) Sprague, R. S., Stephenson, A. H., and Ellsworth, M. L. (2007) Red not dead: signaling

in and from erythrocytes. Trends Endocrinol Metab. 18, 350-5. Epub 2007 Oct 23.

(64) Eltzschig, H. K., Eckle, T., Mager, A., Kuper, N., Karcher, C., Weissmuller, T.,

Boengler, K., Schulz, R., Robson, S. C., and Colgan, S. P. (2006) ATP release from

activated neutrophils occurs via connexin 43 and modulates adenosine-dependent

endothelial cell function. Circ Res. 99, 1100-8. Epub 2006 Oct 12.

(65) Wong, C. W., Christen, T., Roth, I., Chadjichristos, C. E., Derouette, J. P., Foglia, B. F.,

Chanson, M., Goodenough, D. A., and Kwak, B. R. (2006) Connexin37 protects against

atherosclerosis by regulating monocyte adhesion. Nat Med. 12, 950-4. Epub 2006 Jul 23.

(66) Bours, M. J. L., Swennen, E. L. R., Di Virgilio, F., Cronstein, B. N., and Dagnelie, P. C.

(2006) Adenosine 5'-triphosphate and adenosine as endogenous signaling molecules in

immunity and inflammation. Pharmacol. Ther. 112, 358-404.

(67) Abbracchio, M. P., Burnstock, G., Boeynaems, J. M., Barnard, E. A., Boyer, J. L.,

Kennedy, C., Knight, G. E., Fumagalli, M., Gachet, C., Jacobson, K. A., and Weisman,

155 G. A. (2006) International Union of Pharmacology LVIII: update on the P2Y G protein-

coupled nucleotide receptors: from molecular mechanisms and pathophysiology to

therapy. Pharmacol Rev. 58, 281-341.

(68) Lazarowski, E. R., Boucher, R. C., and Harden, T. K. (2000) Constitutive release of ATP

and evidence for major contribution of ecto-nucleotide pyrophosphatase and nucleoside

diphosphokinase to extracellular nucleotide concentrations. J Biol Chem. 275, 31061-8.

(69) Okada, S. F., Nicholas, R. A., Kreda, S. M., Lazarowski, E. R., and Boucher, R. C.

(2006) Physiological regulation of ATP release at the apical surface of human airway

epithelia. J Biol Chem. 281, 22992-3002. Epub 2006 Jun 5.

(70) White, T. D. (1978) Release of ATP from a synaptosomal preparation by elevated

extracellular K+ and by veratridine. J Neurochem. 30, 329-36.

(71) White, T. D. (1984) Characteristics of neuronal release of ATP. Prog

Neuropsychopharmacol Biol Psychiatry 8, 487-93.

(72) White, T. D., and Leslie, R. A. (1982) Depolarization-induced release of adenosine 5'-

triphosphate from isolated varicosities derived from the myenteric plexus of the guinea

pig small intestine. J Neurosci. 2, 206-15.

(73) Demolombe, S., and Escande, D. (1996) ATP-binding cassette proteins as targets for

drug discovery. Trends Pharmacol Sci. 17, 273-5.

(74) Senior, A. E., Al-Shawi, M. K., and Urbatsch, I. L. (1995) The catalytic cycle of P-

glycoprotein. FEBS Lett 377, 285-289.

(75) Burnstock, G. (1976) Purinergic receptors. J Theor Biol. 62, 491-503.

156 (76) Burnstock, G. (1978) A basis for distinguishing two types of purinergic receptors., in Cell

Membrane Receptors for Drugs and Hormones: A Multidisciplinary Approach (Straub,

R. W., and Bolis, L., Eds.) pp 107-118, Reven Press, New York.

(77) Burnstock, G., and Kennedy, C. (1985) Is there a basis for distinguishing two types of

P2-purinoceptor? Gen Pharmacol 16, 433-40.

(78) Abbracchio, M. P., and Burnstock, G. (1994) Purinoceptors: Are there families of P2X

and P2Y purinoceptors. Pharmac. Ther. 64, 445-475.

(79) Cobb, B. R., and Clancy, J. P. (2003) Molecular and cell biology of adenosine receptors.

Curr Top Membr 54, 151-181.

(80) Daly, J. W. (1985) Adenosine receptors. Adv Cyclic Nucleotide Protein Phosphorylation

Res 19, 29-46.

(81) Fredholm, B. B., AP, I. J., Jacobson, K. A., Klotz, K. N., and Linden, J. (2001)

International Union of Pharmacology. XXV. Nomenclature and classification of

adenosine receptors. Pharmacol Rev. 53, 527-52.

(82) Ralevic, V., and Burnstock, G. (1998) Receptors for purines and pyrimidines. Pharmacol

Rev. 50, 413-92.

(83) Barrera, N. P., Ormond, S. J., Henderson, R. M., Murrell-Lagnado, R. D., and

Edwardson, J. M. (2005) Atomic force microscopy imaging demonstrates that P2X2

receptors are trimers but that P2X6 receptor subunits do not oligomerize. J Biol Chem.

280, 10759-65. Epub 2005 Jan 18.

(84) Khakh, B. S., Burnstock, G., Kennedy, C., King, B. F., North, R. A., Seguela, P., Voigt,

M., and Humphrey, P. P. (2001) International union of pharmacology. XXIV. Current

157 status of the nomenclature and properties of P2X receptors and their subunits. Pharmacol

Rev. 53, 107-18.

(85) Mio, K., Kubo, Y., Ogura, T., Yamamoto, T., and Sato, C. (2005) Visualization of the

trimeric P2X2 receptor with a crown-capped extracellular domain. Biochem Biophys Res

Commun. 337, 998-1005. Epub 2005 Oct 3.

(86) Nicke, A., Baumert, H. G., Rettinger, J., Eichele, A., Lambrecht, G., Mutschler, E., and

Schmalzing, G. (1998) P2X1 and P2X3 receptors form stable trimers: a novel structural

motif of ligand-gated ion channels. Embo J. 17, 3016-28.

(87) Gever, J. R., Cockayne, D. A., Dillon, M. P., Burnstock, G., and Ford, A. P. (2006)

Pharmacology of P2X channels. Pflugers Arch. 452, 513-37. Epub 2006 Apr 29.

(88) Prinster, S. C., Hague, C., and Hall, R. A. (2005) Heterodimerization of g protein-

coupled receptors: specificity and functional significance. Pharmacol Rev. 57, 289-98.

(89) Boucher, R. C., Knowles, M. R., Olivier, K. N., Bennett, W., Mason, S. J., and Stutts, M.

J. (1995) Mechanisms and therapeutic actions of in the lung.,

Norwell, Kluwer.

(90) De Mey, J. G., and Vanhoutte, P. M. (1981) Role of the intima in cholinergic and

purinergic relaxation of isolated canine femoral arteries. J Physiol. 316, 347-55.

(91) Boeynaems, J. M., and Pearson, J. D. (1990) P2 purinoceptors on vascular endothelial

cells: physiological significance and transduction mechanisms. Trends Pharmacol Sci.

11, 34-7.

(92) Natori, Y., Baba, T., Moriguchi, M., Takeshita, I., and Fukui, M. (1992) Effects of

theophylline on the selective increases in intratumoral blood flow induced by intracarotid

158 infusion of adenosine and adenosine triphosphate in C6 glioma-transplanted rat brains.

Surg Neurol. 37, 8-14.

(93) Loubatieres-Mariani, N. M., Hillaire-Buys, D., Chapel, J., Bertrand, G., and Petit, P.

(1997) P2 purinoceptor agonists: new insulin secretagogues potentially useful in the

treatment of non-insulin-dependent diabetes mellitus., Wiley-Liss, New York.

(94) Foresta, C., Rossato, M., and Di Virgilio, F. (1992) Extracellular ATP is a trigger for the

acrosome reaction in human spermatozoa. J Biol Chem. 267, 19443-7.

(95) Burnstock, G. (2006) Pathophysiology and therapeutic potential of purinergic signaling.

Pharmacol Rev. 58, 58-86.

(96) White, N., and Burnstock, G. (2006) P2 receptors and cancer. Trends Pharmacol Sci. 27,

211-7. Epub 2006 Mar 13.

(97) Di Virgilio, F., and Solini, A. (2002) P2 receptors: new potential players in

atherosclerosis. Br J Pharmacol. 135, 831-42.

(98) Millan, J. L. (2006) Alkaline Phosphatases : Structure, substrate specificity and

functional relatedness to other members of a large superfamily of enzymes. Purinergic

Signal. 2, 335-41. Epub 2006 Jun 17.

(99) Millan, J. L. (2006) Mammalian Alkaline Phosphatases, Wiley-VCH Verlag GmbH &

Co, Weinheim.

(100) Hunsucker, S. A., Mitchell, B. S., and Spychala, J. (2005) The 5'-nucleotidase as

regulators of nucleotide adn drug metabolism. Pharmacol. Ther. 107, 1-30.

(101) Gorrell, M. D., Gysbers, V., and McCaughan, G. W. (2001) CD26: a multifunctional

integral membrane and secreted protein of activated lymphocytes. Scand J Immunol. 54,

249-64.

159 (102) Strater, N. (2006) Ecto-5'-nucleotidase: Structure function relationships. Purinergic

Signal. 2, 343-50. Epub 2006 May 16.

(103) Zimmermann, H. (1992) 5'-nucleotidase: molecular structure and functional aspects.

Biochem. J. 285, 345-365.

(104) Stefan, C., Jansen, S., and Bollen, M. (2005) NPP-type ectophosphodiesterases: unity in

diversity. Trends Biochem Sci. 30, 542-50.

(105) Goding, J. W., Grobben, B., and Slegers, H. (2003) Physiological and pathophysiological

functions of the ecto-nucleotide pyrophosphatase/phosphodiesterase family. Biochim

Biophys Acta. 1638, 1-19.

(106) Zimmermann, H., Beaudoin, A. R., Bollen, M., Goding, J. W., Guidotti, G., Kirley, T. L.,

Robson, S. C., and Sano, K. (1999) in Second International Workshop on Ecto-ATPases

and Related (Vanduffel, L., Ed.) pp 1-9, Shaker Publishing BV,

Maastricht, The Netherlands, Diepenbeek, Belgium.

(107) Ivanenkov, V. V., Meller, J., and Kirley, T. L. (2005) Characterization of Disulfide

Bonds in Human Nucleoside Triphosphate Diphosphohydrolase 3 (NTPDase3):

Implications for NTPDase Structural Modeling. Biochemistry 44, 8998-9012.

(108) Murphy-Piedmonte, D. M., Crawford, P. A., and Kirley, T. L. (2005) Bacterial

Expression, Folding, Purification and Characterization of Soluble NTPDase5 (CD39L4)

Ecto-nucleotidase. Biochimica et Biophysica Acta 1747, 251-259.

(109) Ivanenkov, V. V., Murphy-Piedmonte, D. M., and Kirley, T. L. (2003) Bacterial

Expression, Characterization, and Disulfide Bond Determination of Soluble Human

NTPDase6 (CD39L2) Nucleotidase: Implications for Structure and Function.

Biochemistry 42, 11726-11735.

160 (110) Rowe, M., Hildreth, J. E., Rickinson, A. B., and Epstein, M. A. (1982) Monoclonal

antibodies to Epstein-Barr virus-induced, transformation-associated cell surface antigens:

binding patterns and effect upon virus-specific T-cell cytotoxicity. Int J Cancer. 29, 373-

81.

(111) Hicks-Berger, C. A., and Kirley, T. L. (2000) Expression and Characterization of Human

Ecto-ATPase and Chimeras with CD39 Ecto-Apyrase. IUBMB Life 50, 43-50.

(112) Chadwick, B. P., and Frischauf, A. M. (1998) The CD39-like gene family: Identification

of three new human members (CD39L2, CD39L3, and CD39L4), their murine

homologues, and a member of the gene family from Drosophila melanogaster. Genomics

50, 357-367.

(113) Knowles, A. F., and Li, C. (2006) Molecular cloning and characterization of expressed

human ecto-nucleoside triphosphate diphosphohydrolase 8 (E-NTPDase 8) and its

soluble extracellular domain. Biochemistry 45, 7323-7333.

(114) Wang, T.-F., and Guidotti, G. (1998) Golgi Localization and Functional Expression of

Human Uridine Diphosphatase. J. Biol. Chem. 273, 11392-11399.

(115) Shi, J. D., Kukar, T., Wang, C. Y., Li, Q. Z., Cruz, P. E., Davoodi-Semiromi, A., Yang,

P., Gu, Y., Lian, W., Wu, D. H., and She, J. X. (2001) Molecular cloning and

characterization of a novel mammalian endo-apyrase (LALP1). J Biol Chem 276, 17474-

8.

(116) Plesner, L. (1995) Ecto-ATPases: Identities and functions. Int Rev Cytol 158, 141-214.

(117) Kukulski, F., Sevigny, J., and Komoszynski, M. (2004) Comparative hydrolysis of

extracellular adenine nucleotides and adenosine in synaptic membranes from porcine

brain cortex, hippocampus, cerebellum and medulla oblongata. Brain Res 1030, 49-56.

161 (118) Treuheit, M. J., Vaghy, P. L., and Kirley, T. L. (1992) The Mg2+-ATPase of rabbit

skeletal muscle transverse tubules is a 67 kDa glycoprotein. J. Biol. Chem. 267, 11777-

11782.

(119) Kukulski, F., Lévesque, S. A., Lavoie, E. G., Lecka, J., Bigonnesse, F., Knowles, A. F.,

Robson, S. C., Kirley, T. L., and Sévigny, J. (2005) Comparative hydrolysis of P2

receptor agonists by NTPDases 1, 2, 3 and 8. Purinergic Signalling 1, 193-204.

(120) Gendron, F. P., Benrezzak, O., Krugh, B. W., Kong, Q., Weisman, G. A., and Beaudoin,

A. R. (2002) Purine signaling and potential new therapeutic approach: possible outcomes

of NTPDase inhibition. Curr Drug Targets 3, 229-45.

(121) Brunschweiger, A., Iqbal, J., Umbach, F., Scheiff, A. B., Munkonda, M. N., Sevigny, J.,

Knowles, A. F., and Muller, C. E. (2008) Selective nucleoside triphosphate

diphosphohydrolase-2 (NTPDase2) inhibitors: nucleotide mimetics derived from uridine-

5'-carboxamide. J Med Chem. 51, 4518-28. Epub 2008 Jul 17.

(122) Crack, B. E., Pollard, C. E., Beukers, M. W., Roberts, S. M., Hunt, S. F., Ingall, A. H.,

McKechnie, K. C., AP, I. J., and Leff, P. (1995) Pharmacological and biochemical

analysis of FPL 67156, a novel, selective inhibitor of ecto-ATPase. British Journal of

Pharmacology 114, 475-81.

(123) Levesque, S. A., Lavoie, E. G., Lecka, J., Bigonnesse, F., and Sevigny, J. (2007)

Specificity of the ecto-ATPase inhibitor ARL 67156 on human and mouse

ectonucleotidases. Br J Pharmacol. 152, 141-50. Epub 2007 Jul 2.

(124) Stout, J. G., and Kirley, T. L. (1995) Inhibition of purified chicken gizzard ecto-ATPase

by P2 purinoceptor antagonists. Biochem. Molec. Biol. Int. 36, 927-934.

162 (125) Wittenburg, H., Bultmann, R., Pause, B., Ganter, C., Kurz, G., and Starke, K. (1996) P2-

purinoceptor antagonists: II. Blockade of P2-purinoceptor subtypes and ecto-

nucleotidases by compounds related to Evans blue and . Naunyn

Schmiedebergs Arch Pharmacol. 354, 491-7.

(126) Baqi, Y., Weyler, S., Iqbal, J., Zimmermann, H., and Muller, C. E. (2009) Structure-

activity relationships of anthraquinone derivatives derived from bromaminic acid as

inhibitors of ectonucleoside triphosphate diphosphohydrolases (E-NTPDases). Purinergic

Signal. 5, 91-106. Epub 2008 Jun 5.

(127) Munkonda, M. N., Pelletier, J., Ivanenkov, V. V., Fausther, M., Tremblay, A., Kunzli, B.,

Kirley, T. L., and Sevigny, J. (2009) Characterization of a monoclonal antibody as the

first specific inhibitor of human NTP diphosphohydrolase-3 : partial characterization of

the inhibitory epitope and potential applications. Febs J. 276, 479-96.

(128) Gendron, F. P., Halbfinger, E., Fischer, B., Duval, M., D'Orleans-Juste, P., and Beaudoin,

A. R. (2000) Novel inhibitors of nucleoside triphosphate diphosphohydrolases: chemical

synthesis and biochemical and pharmacological characterizations. J Med Chem 43, 2239-

47.

(129) North, R. A. (1996) P2X receptors: a third major class of ligand-gated ion channels. Ciba

Found Symp 198, 91-105.

(130) Schulte am Esch, J., 2nd, Sevigny, J., Kaczmarek, E., Siegel, J. B., Imai, M., Koziak, K.,

Beaudoin, A. R., and Robson, S. C. (1999) Structural elements and limited proteolysis of

CD39 influence ATP diphosphohydrolase activity. Biochemistry 38, 2248-2258.

163 (131) Wang, T. F., Ou, Y., and Guidotti, G. (1998) The transmembrane domains of ectoapyrase

(CD39) affect its enzymatic activity and quaternary structure. J. Biol. Chem. 273, 24814-

24821.

(132) Kirley, T. L., Gerber, L. K., and Smith, T. M. (1999) Expression and characterization of

chicken muscle ecto-ATPase in mammalian COS cells. IUBMB Life 48, 67-72.

(133) Smith, T. M., and Kirley, T. L. (1999) Glycosylation is essential for functional expression

of a human brain ecto-apyrase. Biochemistry 38, 1509-1516.

(134) Kirley, T. L., Crawford, P. A., and Smith, T. M. (2006) The structure of the nucleoside

triphosphate diphosphohydrolases (NTPDases) as revealed by mutagenic and

computational modeling analysis. Purinergic Signalling 2, 379-389.

(135) Grinthal, A., and Guidotti, G. (2000) Substitution of His59 Converts CD39 Apyrase into

an ADPase in a Quaternary Structure Dependent Manner. Biochemistry 39, 9-16.

(136) Chiang, W. C., and Knowles, A. F. (2008) Transmembrane domain interactions affect the

stability of the extracellular domain of the human NTPDase 2. Arch Biochem Biophys

472, 89-99.

(137) Mukasa, T., Lee, Y., and Knowles, A. F. (2005) Either the carboxyl- or the amino-

terminal region of the human ecto-ATPase (E-NTPDase 2) confers detergent and

temperature sensitivity to the chicken ecto-ATP-diphosphohydrolase (E-NTPDase 8).

Biochemistry 44, 11160-11170.

(138) Gaddie, K. J., and Kirley, T. L. (2009) Conserved Polar Residues Stabilize

Transmembrane Domains and Promote Oligomerization in Human Nucleoside

Triphosphate Diphosphohydrolase 3. Biochemistry 48, 9437-9447.

164 (139) Handa, M., and Guidotti, G. (1996) Purification and cloning of a soluble ATP-

diphosphohydrolase (apyrase) from potato tubers (Solanum tuberosum). Biochem

Biophys Res Commun 218, 916-923.

(140) Smith, T. M., and Kirley, T. L. (1999) Site-Directed Mutagenesis of a Human Brain

Ecto-Apyrase: Evidence that the E-type ATPases are related to the Actin/ Heat Shock 70/

Sugar Kinase Superfamily. Biochemistry 38, 321-328.

(141) Smith, T. M., Lewis Carl, S. A., and Kirley, T. L. (1999) Mutagenesis of two conserved

tryptophan residues of the E-type ATPases: Inactivation and conversion of an ecto-

apyrase to an ecto-NTPase. Biochemistry 38, 5849-5857.

(142) Kirley, T. L., Yang, F., and Ivanenkov, V. V. (2001) Site-Directed Mutagenesis of

Human Nucleoside Triphosphate Diphosphohydrolase 3: The Importance of Conserved

Glycine Residues and the Identification of Additional Conserved Protein Motifs in

eNTPDases. Arch Biochem Biophys 395, 94-102.

(143) Murphy, D. M., and Kirley, T. L. (2003) Asparagine 81, An Invariant Glycosylation Site

Near Apyrase Conserved Region 1, is Essential for Full Enzymatic Activity of Ecto

Nucleoside Triphosphate Diphosphohydrolase 3. Archives of Biochemistry and

Biophysics 413, 107-115.

(144) Failer, B. U., Aschrafi, A., Schmalzing, G., and Zimmermann, H. (2003) Determination

of native oligomeric state and substrate specificity of rat NTPDase1 and NTPDase2 after

heterologous expression in Xenopus oocytes. Eur J Biochem 270, 1802-1809.

(145) Basu, S., Murphy-Piedmonte, D. M., and Kirley, T. L. (2004) Conserved lysine 79 is

important for activity of ecto-nucleoside triphosphate diphosphohydrolase 3 (NTPDase3).

Purinergic Signalling 1, 51-58.

165 (146) Zebisch, M., and Sträter, N. (2008) Structural insight into signal conversion and

inactivation by NTPDase2 in purinergic signaling. Proc Natl Acad Sci U S A 105, 6882-

6887.

(147) Murphy, D. M., Ivanenkov, V. V., and Kirley, T. L. (2002) Identification of Cysteine

Residues Responsible for Oxidative Cross-linking and Chemical Inhibition of Human

Nucleoside Triphosphate Diphosphohydrolase 3. Journal of Biological Chemistry 277,

6162-6169.

(148) Chiang, W. C., and Knowles, A. F. (2008) Inhibition of human NTPDase 2 by

modification of an intramembrane cysteine by p-chloromercuriphenylsulfonate and

oxidative cross-linking of the transmembrane domains. Biochemistry 47, 8775-8785.

(149) Grinthal, A., and Guidotti, G. (2004) Dynamic Motions of CD39 Transmembrane

Domains Regulate and Are Regulated by the Enzymatic Active Site. Biochemistry 43,

13849-13858.

(150) Stout, J. G., Strobel, R. S., and Kirley, T. L. (1995) Identification and

immunolocalization of ecto-ATPDase in chicken stomach. Biochem Mol Biol Int 36, 529-

535.

(151) Kansas, G. S., Wood, G. S., and Tedder, T. F. (1991) Expression, distribution, and

biochemistry of human CD39: role in activation-associated homotypic adhesion of

lymphocytes. J. Immunol. 146, 2235-2244.

(152) Knowles, A. F. (1988) Differential expression of ectoMg2+-ATPase and ectoCa2+-

ATPase activities in human hepatoma cells. Archives of Biochemistry and Biophysics

263, 264-271.

166 (153) Karasaki, S., Simard, A., and de Lamirande, G. (1977) Surface morphology and

nucleoside phosphatase activity of rat liver epithelial cells during oncogenic

transformation in vitro. Cancer Research 37, 3516-3525.

(154) Kaczmarek, E., Koziak, K., Sevigny, J., Siegel, J. B., Anrather, J., Beaudoin, A. R., Bach,

F. H., and Robson, S. C. (1996) Identification and characterization of CD39/vascular

ATP diphosphohydrolase. J. Biol. Chem. 271, 33116-22.

(155) Marcus, A. J., Broekman, M. J., Drosopoulos, J. H. F., Islam, N., Alyonycheva, T. N.,

Safier, L. B., Hajjar, K. A., Posnett, D. N., Schoenborn, M. A., Schooley, K. A., Gayle,

R. B., and Maliszewski, C. R. (1997) The endothelial cell ecto-ADPase responsible for

inhibition of platelet function is CD39. J Clin Invest 99, 1351-1360.

(156) Enjyoji, K., J, S. v., Lin, Y., Frenette, P. S., Christie, P. D., Esch, J. S., 2nd, Imai, M.,

Edelberg, J. M., Rayburn, H., Lech, M., Beeler, D. L., Csizmadia, E., Wagner, D. D.,

Robson, S. C., and Rosenberg, R. D. (1999) Targeted disruption of cd39/ATP

diphosphohydrolase results in disordered hemostasis and thromboregulation. Nat Med 5,

1010-1017.

(157) Marcus, A. J., Broekman, M. J., Drosopoulos, J. H., Islam, N., Pinsky, D. J., Sesti, C.,

and Levi, R. (2003) Metabolic control of excessive extracellular nucleotide accumulation

by CD39/ecto-nucleotidase-1: implications for ischemic vascular diseases. J Pharmacol

Exp Ther 305, 9-16.

(158) Braun, N., Sevigny, J., Mishra, S. K., Robson, S. C., Barth, S. W., Gerstberger, R.,

Hammer, K., and Zimmermann, H. (2003) Expression of the ecto-ATPase NTPDase2 in

the germinal zones of the developing and adult rat brain. Eur J Neurosci 17, 1355-1364.

167 (159) Braun, N., Sevigny, J., Robson, S. C., Hammer, K., Hanani, M., and Zimmermann, H.

(2004) Association of the ecto-ATPase NTPDase2 with glial cells of the peripheral

nervous system. Glia 45, 124-32.

(160) Masse, K., Bhamra, S., Eason, R., Dale, N., and Jones, E. A. (2007) Purine-mediated

signalling triggers eye development. Nature. 449, 1058-62.

(161) Belcher, S. M., Zsarnovszky, A., Crawford, P. A., Hemani, H., Spurling, L., and Kirley,

T. L. (2005) Immunolocalization of ecto-Nucleoside Triphosphate Diphosphohydrolase 3

(NTPDase3) in Rat Brain: Implications for modulation of multiple homeostatic systems

including feeding and sleep-wake behaviors. Neuroscience.

(162) Appelbaum, L., Skariah, G., Mourrain, P., and Mignot, E. (2007) Comparative expression

of p2x receptors and ecto-nucleoside triphosphate diphosphohydrolase 3 in hypocretin

and sensory neurons in zebrafish. Brain Res. 1174, 66-75. Epub 2007 Aug 9.

(163) Todorov, L. D., Mihayalova-Todorova, S., Westfall, T. D., Sneddon, P., Kennedy, C.,

Bjur, R. A., and Westfall, D. P. (1997) Neuronal release of soluble nucleotidases and

their role in neurotransmitter inactivation. Nature 387, 76-79.

(164) Mihaylova-Todorova, S., Todorov, L. D., and Westfall, D. P. (2001) Correlation between

the release of the sympathetic neurotransmitter ATP and soluble nucleotidases from the

guinea pig vas deferens. J Pharmacol Exp Ther 296, 64-70.

(165) Stoscheck, C. M. (1987) Protein assay sensitive at nanogram levels. Anal. Biochem. 160,

301-305.

(166) Fiske, C. H., and Subbarow, Y. (1925) The colorometric determination of phosphorous.

Journal of Biological Chemistry 66, 375-400.

168 (167) Baykov, A. A., Evtushenko, O. A., and Avaeva, S. M. (1988) A malachite green

procedure for orthophosphate determination and its use in alkaline phosphatase-based

enzyme immunoassay. Anal Biochem 171, 266-270.

(168) Careaga, C. L., and Falke, J. J. (1992) Thermal motions of surface alpha-helices in the D-

galactose chemosensory receptor. Detection by disulfide trapping. J Mol Biol 226, 1219-

1235.

(169) Papanikolaou, A., Papafotika, A., Murphy, C., Papamarcaki, T., Tsolas, O., Drab, M.,

Kurzchalia, T. V., Kasper, M., and Christoforidis, S. (2005) Cholesterol-dependent lipid

assemblies regulate the activity of the ecto-nucleotidase CD39. J Biol Chem 280, 26406-

26414.

(170) Grinthal, A., and Guidotti, G. (2002) Transmembrane Domains Confer Different

Substrate Specificities and Hydrolysis Mechanisms on CD39,

CD39L1, and Chimeras. Biochemistry 41, 1947-1956.

(171) Hurley, J. H., Mason, D. A., and Matthews, B. W. (1992) Flexible-geometry

conformational energy maps for the amino acid residue preceding a proline. Biopolymers

(Peptide Science) 32, 1443–1446.

(172) MacArthur, M. W., and Thornton, J. M. (1991) Influence of proline residues on protein

conformation. J Mol Biol 218, 397–412.

(173) Nicholson, H., Tronrud, D. E., Becktel, W. J., and Matthews, B. W. (1992) Analysis of

the effectiveness of proline substitutions and glycine replacements in increasing the

stability of phage T4 lysozyme. Biopolymers 32, 1431–1441.

(174) Balasubramanian, R., Lakshminarayanan, A. V., Sabesan, M. N., Tegoni, G.,

Venkatesan, K., and Ramachandran, G. N. (1971) Studies on the conformation of amino

169 acids. VI. Conformation of the proline ring as observed in crystal structures of amino

acids and peptides. Int J Protein Res 3, 25-33.

(175) Morris, A. L., MacArthur, M. W., Hutchinson, E. G., and Thornton, J. M. (1992)

Stereochemical quality of protein structure coordinates. Proteins 12, 345-364.

(176) Landolt-Marticorena, C., Williams, K. A., Deber, C. M., and Reithmeier, R. A. (1993)

Non-random distribution of amino acids in the transmembrane segments of human type I

single span membrane proteins. J Mol Biol, 602-608.

(177) Tourasse, N. J., and Li, W. H. (2000) Selective constraints, amino acid composition, and

the rate of protein evolution. Mol Biol Evol, 656-664.

(178) Eilers, M., Shekar, S. C., Shieh, T., Smith, S. O., and Fleming, P. J. (2000) Internal

packing of helical membrane proteins. Proc Natl Acad Sci U S A, 5796-5801.

(179) Partridge, A. W., Therien, A. G., and Deber, C. M. (2002) Polar mutations in membrane

proteins as a biophysical basis for disease. Biopolymers (Peptide Science) 66, 350-358.

(180) Partridge, A. W., Therien, A. G., and Deber, C. M. (2004) Missense mutations in

transmembrane domains of proteins: Phenotypic propensity of polar residues for human

disease. Proteins 54, 648-656.

(181) Adamian, L., and Liang, J. (2003) Interhelical hydrogen bonds in transmembrane region

are important for function and stability of Ca2+-transporting ATPase. Cell Biochem

Biophys 39, 1-12.

(182) Gurezka, R., Laage, R., Brosig, B., and Langosch, D. (1999) A heptad motif of leucine

residues found in membrane proteins can drive self-assembly of artificial transmembrane

segments. J Biol Chem 274, 9265-70.

170 (183) Lemmon, M. A., Treutlein, H. R., Adams, P. D., Brünger, A. T., and Engelman, D. M.

(1994) A dimerization motif for transmembrane alpha-helices. Nat Struct Biol 1, 157-63.

(184) Sal-Man, N., Gerber, D., and Shai, Y. (2005) The identification of a minimal

dimerization motif QXXS that enables homo- and hetero-association of transmembrane

helices in vivo. J Biol Chem 280, 27449-27457.

(185) Sal-Man, N., Gerber, D., and Shai, Y. (2004) The composition rather than position of

polar residues (QxxS) drives aspartate receptor transmembrane domain dimerization in

vivo. Biochemistry 43, 2309-2313.

(186) Wang, C. J., Vlajkovic, S. M., Housley, G. D., Braun, N., Zimmermann, H., Robson, S.

C., Sevigny, J., Soeller, C., and Thorne, P. R. (2005) C-terminal splicing of NTPDase2

provides distinctive catalytic properties, cellular distribution and enzyme regulation.

Biochem J 385 (Pt 3), 729-736.

(187) Vlajkovic, S. M., Housley, G. D., Greenwood, D., and Thorne, P. R. (1999) Evidence for

alternative splicing of ecto-ATPase associated with termination of purinergic

transmission. Mol Brain Res 73, 85-92.

(188) Matsumoto, M., Sakurai, Y., Kokubo, T., Yagi, H., Makita, K., Matsui, T., Titani, K.,

Fujimura, Y., and Narita, N. (1999) The cDNA cloning of human placental ecto-ATP

diphosphohydrolases I and II. FEBS Lett 453, 335-340.

(189) Chadwick, B. P., and Frischauf, A.-M. (1997) Cloning and mapping of a human and

mouse gene with homology to ecto-ATPase genes. Mammalian Genome 8, 668-672.

(190) Mateo, J., Kreda, S., Henry, C. E., Harden, T. K., and Boyer, J. L. (2003) Requirement of

Cys399 for processing of the human ecto-ATPase (NTPDase2) and its implications for

171 determination of the activities of splice variants of the enzyme. J Biol Chem 278, 39960-

39968.

(191) Crawford PA, G. K., Smith TM, Kirley TL. (2007) Characterization of an alternative

splice variant of human nucleoside triphosphate diphosphohydrolase 3 (NTPDase3): a

possible modulator of nucleotidase activity and purinergic signaling. Arch Biochem

Biophys. 457, 7-15.

(192) Lennon, G., Auffray, C., Polymeropoulos, M., and Soares, M. B. (1996) The I.M.A.G.E.

Consortium: an integrated molecular analysis of genomes and their expression.

Genomics. 33, 151-2.

(193) Bork, P., Sander, C., and Valencia, A. (1992) An ATPase domain common to prokaryotic

cell cycle proteins, sugar kinases, actin, and hsp70 heat shock proteins. Proc Natl Acad

Sci U S A 89, 7290-7294.

(194) Drosopoulos, J. H., Broekman, M. J., Islam, N., Maliszewski, C. R., Gayle, R. B., 3rd,

and Marcus, A. J. (2000) Site-directed mutagenesis of human endothelial cell ecto-

ADPase/soluble CD39: requirement of glutamate 174 and serine 218 for enzyme activity

and inhibition of platelet recruitment. Biochemistry 39, 6936-6943.

(195) Basu, S., and Kirley, T. L. (2005) Identification of a tyrosine residue responsible for N-

acetylimidazole-induced increase of activity of ecto-nucleoside triphosphate

diphosphohydrolase 3. Purinergic Signalling 1, 271-280.

(196) Krawczak M, C. D. (1997) The human gene mutation database. Trends Genet 13, 121-2.

(197) Senes, A., Ubarretxena-Belandia, I., and Engleman, D. M. (2001) The Calpha ---H...O

hydrogen bond: a determinant of stability and specificity in transmembrane helix

interactions. Proc Natl Acad Sci U S A 98, 9056-9061.

172 (198) Stout, J. G., and Kirley, T. L. (1994) Tissue Distribution of Ecto-Mg-ATPase in Adult

and Embryonic Chicken. Biochemistry and Molecular Biology International 32, 745-753.

(199) North, R. A. (2002) Molecular Physiology of P2X Receptors. Physiol Rev 82, 1013-67.

(200) Bruss, M., Barann, M., Hayer-Zillgen, M., Eucker, T., Gothert, M., and Bonisch, H.

(2000) Modified 5-HT3A receptor function by co-expression of alternatively spliced

human 5-HT3A receptor isoforms. Naunyn Schmiedebergs Arch Pharmacol. 362, 392-

401.

(201) Stapley, B. J., and Creamer, T. P. (1999) A survey of left-handed polyproline II helices.

Protein Sci. 8, 587-95.

(202) Creamer, T. P. (1998) Left-handed polyproline II helix formation is (very) locally driven.

Proteins. 33, 218-26.

(203) Ma, K., Kan, L., and Wang, K. (2001) Polyproline II helix is a key structural motif of the

elastic PEVK segment of titin. Biochemistry. 40, 3427-38.

(204) Unterreitmeier, S., Fuchs, A., Schaffler, T., Heym, R. G., Frishman, D., and Langosch, D.

(2007) Phenylalanine promotes interaction of transmembrane domains via GxxxG motifs.

J Mol Biol. 374, 705-18. Epub 2007 Sep 26.

(205) Hunter, C. A., Singh, J., and Thornton, J. M. (1991) Pi-pi interactions: the geometry and

energetics of phenylalanine-phenylalanine interactions in proteins. J Mol Biol. 218, 837-

46.

(206) Seguin, J., Mayer, C., Robert, B., and Arluison, V. (2008) Thermodynamics of the

beta(2) association in light-harvesting complex I of Rhodospirillum rubrum. Implication

of peptide identity in dimer stability. Febs J. 275, 1240-7. Epub 2008 Feb 6.

173 (207) Friedman, D. J., Talbert, M. E., Bowden, D. W., Freedman, B. I., Mukanya, Y., Enjyoji,

K., and Robson, S. C. (2009) Functional ENTPD1 polymorphisms in African Americans

with diabetes and end-stage renal disease. Diabetes. 58, 999-1006. Epub 2008 Dec 18.

174 Appendix: Publications and Manuscripts under Review

1. Gaddie, K. J. and Kirley, T. L. Conserved Polar Residues Stabilize Transmembrane

Domains and Promote Oligomerization in Human Nucleoside Triphosphate

Diphosphohydrolase3. Biochemistry. 48(40): 9437-9447. 2009.

2. Gaddie, K. J., and Kirley, T. L. Proline Residues Link the Active Site to Transmembrane

Domain Movements in Human Nucleoside Triphosphate Diphosphohydrolase 3

(NTPDase3). Submitted to Purinergic Signalling, November, 2009.

2. P. Crawford, K. J. Gaddie, T.M. Smith and T. L. Kirley. Characterization of an alternative

splice variant of human nucleoside triphosphate diphosphohydrolase 3 (NTPDase3): A

possible modulator of nucleotidase activity in purinergic signaling. Archives of

Biochemistry and Biophysics. 475(1): 7-15. 2006.

175