DEFINING THE CATALYTIC AND KINETIC MECHANISM AND NATURAL FUNCTION OF THE HIGHLY CONSERVED ACYL-AMP HYDROLASE, HINT1

A DISSERTATION SUBMITTED TO THE FACULTY OF THE GRADUATE SCHOOL OF THE UNIVERSITY OF MINNESOTA BY

SANAA BARDAWEEL

IN PARTIAL FULFILMENT OF THE REQUIREMENTS

FOR THE DEGREE OF DOCTORE OF PHILOSOPHY

DR. CARSTON R. WAGNER, ADVISER

DECEMBER 2010

© Sanaa Bardaweel, 2010

ACKNOWLEDGEMENTS

I would like to gratefully and sincerely thank my thesis adviser, Dr. Carston R.

Wagner, for his guidance, understanding and patience. His mentorship was paramount in providing a well rounded experience consistent with my long-term career goals. He encouraged me to not only grow as an experimentalist but also as an instructor and an independent researcher. In every sense, none of this work would have been possible without him.

I gratefully acknowledge Dr. Michael Sadowsky for his advice, supervision, and

crucial contributions, which made him a backbone of this research and so to this thesis. I

am also deeply thankful to Dr. Patrick Hanna, for the valuable advice and discussions that

inspired me through out the work. Many thanks must also go to my thesis committee

members, Dr. Chengguo Xing and Dr. Howard Towle, for their valuable input and

suggestions that improved my dissertation.

I would like to make a special reference to Dr. Yusuf Abul-Hajj for his unflinching encouragement and support in various ways. His support and guidance during the very first steps in my PhD have had a remarkable influence on my entire progress in the program.

I am grateful to former members of Dr. Wagner’s laboratory, Dr. Brahma Ghosh,

Dr. Brandie Brummer, Dr. Yan Jia, Dr. Qing Li, Dr. Brian White and Dr. Xin Zhou for

their help and great friendship. I also thank current members of the lab, Dr. Adrian

Fegan, Dr. Matthew Cuellar, Dr. Sidath Kumarapperuma, Amit Ganger and Ryan Holton

for the numerous and productive discussions that helped me improve my knowledge

during my graduate study.

i

I owe my loving thanks to my mother who raised me with a love of science and supported me in all my life. Her endless faith inspired me to become the person who I am today. A special thank-you goes to my wonderful sons, Hashem, Yamen and Othman who made my graduate life a lot less stressful. I hope that I may serve as an example for them as my mother has for me.

My final, and most heartfelt, acknowledgment must go to my husband Rizq. His support, encouragement and unwavering love have turned my journey through graduate school into a pleasure. For never once complaining through all difficult times and challenges, for taking all responsibility for the care of our family, for sharing all the good and bad times with endless support, for all that, this dissertation is dedicated to him.

ii

ABSTRACT

Histidine triad binding (Hints) are members of the histidine triad (HIT) superfamily of nucleotidyl transferases and hydrolyases. It has been recently demonstrated that Hints are efficient phosphoramidases and therefore activators of potent antiviral and anticancer pronucleotides. In spite of their high evolutionary conservation among all kingdoms of life, and the several regulatory functions in which

Hints have been implicated, a clear connection between their observed function and their catalytic efficiency has not been elucidated. To gain a comprehensive understanding of the essential role of these ubiquitous enzymes, our laboratory has devoted a considerable effort toward the delineation of the principles governing Hints catalysis and cellular function. Such understanding will provide an unprecedented ability to assess the role of these highly conserved, but functionally unknown enzymes.

Since Hints are found in both prokaryotes and eukaryotes, we have attempted to understand their function, mechanism, and structural determinants in prokaryotes, under the assumption that their role may be at least partially conserved among members of the tree of life. Recently, we have demonstrated by E. coli disruption studies that the bacterial Hint enzyme is necessary for growth under high salt conditions, and when alanine is a carbon and nitrogen source. Through a combination of phenotypic screening and complementation experiments with wild-type and ecHinT knock-out E. coli strains, we have shown that catalytically-active ecHinT is required for growth on D-alanine. In addition, using Hint-inhibitors and active-site mutants, we have demonstrated that expression of catalytically-active ecHinT is essential for the activity of the enzyme D- amino acid dehydrogenase (DadA) (equivalent to D-amino acid oxidase in eukaryotes), a

iii necessary component of the D-amino acids metabolic pathway. These results are considered as the first report in literature that shows a successful connection between a discovered Hint-related phenotype and the catalytic activity of Hint.

Previously, we have demonstrated that lysyl-AMP generated by LysRS is a substrate for both human and E. coli Hints. In addition, we have shown that the ability of Hint to hydrolyze lysyl-AMP depends on its enzymatic activity. Here, we demonstrate that the molecular determinants governing this regulation appear to reside in the C- terminus region of Hint. Interestingly, the ecHinT-DadA interaction appears also to be governed by both ecHinT-activity and the C-terminus loop.

We have also expanded our scope to look at possible toxicity of D-alanine in E. coli strains lacking dadA or hinT . Our results demonstrate that E. coli mutants lacking dadA or hinT are highly susceptible to D-alanine toxicity and that the catalytic activity of

Hint is an essential requirement to protect E. coli from the observed toxicity of D-alanine.

Based on careful analysis of the combined results from the ecHinT-LysRS and ecHinT-

DadA potential interactions, and comprehensive understanding of the D-alanine metabolic pathway in bacteria, we proposed a possible regulatory mechanism of Hint,

LysRS and DadA on global protein translational processes to prevent D-amino acids toxicity in E. coli .

iv

Table of Contents

Acknowledgements ...... i

Abstract ...... iii

Table of Contents ...... v

List of Tables ...... xi

List of Figures ...... xiii

List of Schemes ...... xvi

List of Abbreviations ...... xvii

Chapter One: Introduction: Histidine Triad Nucleotide Binding Proteins ...... 1

I. Discovery and Classification of HIT Superfamily...... 2

II. Phosphoramidase Activity of Hints ...... 6

III. Structural Properties of Hints ...... 12

IV. Structural Studies of Hints ...... 13

V. Role of Hint Homodimerization ...... 14

VI. Possible Biological Functions of Hints ...... 18

VII. Prokaryotic Hint ...... 20

VIII. The Interaction of Hints with Aminoacyl tRNA Synthetases ...... 23

IX. Current Research ...... 27

Chapter Two: Probing the Impact of the ecHinT C-Terminal Domain on Structure

and Catalysis...... 28

v

I. Introduction ...... 29

II. Experimental Procedures ...... 32

A. Crystallization and X-ray Data Collection ...... 32

B. Structure Determination and Refinement ...... 32

C. Site-directed Mutagenesis ...... 35

D. Expression and Purification of Recombinant Proteins ...... 35

E. Circular Dichroism Spectroscopy ...... 37

F. Size-exclusion Chromatography ...... 37

G. Steady-state Kinetics ...... 38

H. Pre-steady-state Kinetics ...... 38

I. pH-dependence of Steady-state Kinetics ...... 39

J. Lysyl-AMP-dependent Adenylation of ecHinT by LysU ...... 39

K. PDB Accession Numbers ...... 39

III. Results and Discussion ...... 40

A. Structural Properties of ecHinT ...... 40

B. Ligand Binding ...... 48

C. Design of C-terminus Mutants and Protein Purification...... 54

D. Secondary Structure Analysis by Circular Dichroism (CD) Spectroscopy ...... 54

E. Thermal Stability ...... 55

F. Size-exclusion Chromatography ...... 60

G. Effect on Catalysis ...... 60

H. Lysyl-AMP-dependent Adenylation of ecHinT by LysU ...... 71

vi

IV. Summary and Concluding Remarks ...... 74

Chapter Three: HinT, a Histidine Triad Nucleotide Binding Protein, is Essential for

Alanine Metabolism in Escherichia coli ...... 76

I. Introduction ...... 77

II. Experimental Procedures ...... 80

A. Bacterial Strains, Media, and Growth Conditions ...... 80

B. PCR Verification of Mutants ...... 80

C. Verification of hinT Deletion Mutant by Loss of Activity ...... 80

D. Phenotype Analysis Using Biolog™ GN2-MicroPlates ...... 83

E. Carbon Source Utilization Assays ...... 83

F. Induction of D-amino Acid Dehydrogenase Activity and Enzyme Assays ...... 84

G. Preparation of Membrane Fractions...... 84

H. Reverse Transcription-PCR ...... 85

I. Site-directed Mutagenesis ...... 85

J. Phenotype Complementation Studies ...... 86

K. General Synthetic Procedures and Materials ...... 86

L. Synthesis of 2’,3’-Isopropylidine-5’-O-(4-Chlorophenoxy)Carbonyl Guanosine

(Guanosine -5’-Carbonate Acetonide) ...... 87

M. Synthesis of 2’,3’-Isopropylidine-5’-O-[(3-Indolyl)-1-Ethyl]Carbamoyl

Guanosine (Guanosine -5’-Carbamate Acetonide) ...... 87

vii

N. Synthesis of 5’-O-[(3-Indolyl)-1-Ethyl]Carbamoyl Guanosine (Guanosine-5’-

Tryptamine Carbamate, TpGc) ...... 88

III. Results and Discussion ...... 89

A. Operon Structure of Associated with hinT ...... 89

B. Phenotype of ecHinT ...... 89

C. D-Amino Acid Dehydrogenase Transcription and Activity Testing ...... 93

D. Phenotype Rescue by ecHinT ...... 98

E. Design, Synthesis and Characterization of ecHinT Inhibitor ...... 100

Chapter Four: Synthesis and Evaluation of Potential Inhibitors of Human and

Escherichia coli Histidine Triad Nucleotide Binding Proteins ...... 111

I. Introduction ...... 112

II. Experimental Procedures ...... 113

A. General Procedure for synthesis of 2’,3’-Isopropylidine-5’-(indole-3- propionyl)

nucleosides: (Adenosine Acetonide-5’-Indole-3-Propionate (1) ...... 113

B. General Procedure for Synthesis of 5’-(indole-3- propionyl) nucleosides

(Adenosine 5’-Indole-3-Propionate(2)...... 114

C. General Procedure for Synthesis of 5’-O-(4-chlorophenoxy) carbonyl 3’-azido-3’-

deoxy thymidine (AZT-5’-carbonate (3) ...... 115

D. General Procedure for Synthesis of 5’-O-[(3-indolyl)-1-ethyl]carbamoyl 3’-azido-

3’-deoxy thymidine (AZT-5’-carbamate (4) ...... 115

viii

E. General Procedure for Synthesis of 5’-O-[(3-indolyl)-1-ethyl]carbamoyl 3’-

amino-3’-deoxy thymidine (AZT-3’-NH2-5’-carbamate(5) ...... 116

F. General Procedure for Synthesis of 2’,3’-Isopropylidine-5’-O-(4-

chlorophenoxy)carbonyl nucleosides (Adenosine-5’-carbonate acetonide (6) ....116

G. General Procedure for Synthesis of 2’,3’-Isopropylidine-5’-O-[(3-indolyl)-1-

ethyl]carbamoyl nucleosides (Adenosine 5’-tryptamine-carbamate acetonide (7)117

H. General Procedure for Synthesis of 5’-O-[(3-indolyl)-1-ethyl]carbamoyl

nucleosides (Adenosine 5’-tryptamine-carbamate (8) ...... 118

I. Phosphoramidase Assay...... 118

J. Phenotype Testing ...... 119

III. Results and Discussion ...... 119

A. Inhibitors’ Synthesis and Evaluation ...... 119

B. Phenotype Testing ...... 130

IV. Summary and Concluding Remarks ...... 130

Chapter Five: Possible Physiological Role of The Histidine Triad Nucleotide

Binding Protein in Escherichia coli : Regulation of D-alanine Detoxification ...... 133

I. Introduction ...... 134

II. Experimental Procedures ...... 138

A. Bacterial Strains, Media, and Growth Conditions ...... 138

B. Growth Inhibitory Effects of D-amino Acids on Escherichia coli ...... 138

C. Growth Kinetics ...... 139

ix

D. D-alanine Toxicity in Presence of ecHinT Inhibitors ...... 139

E. Mammalian Cell Lines and Cell Culturing ...... 140

F. Cell Viability Assay ...... 140

G. D-alanine Cytotoxicity in MCF-7, MDA-MB-231 and MIA PaCa-2 Cell Lines 141

H. D-alanine Cytotoxicity in HPB-MLT and Raji Cell Lines ...... 141

III. Results ...... 142

A. D-amino Acids Toxicity in E. coli ...... 142

B. Effect of ecHinT Inhibition on E. coli Sensitivity to D-alanine ...... 148

C. D-alanine Toxicity in Mammalian Cell Lines ...... 152

IV. Discussion ...... 152

BIBLIOGRAPHY ...... 162

x

List of Tables

Chapter One

Table 1. Hydrolysis rates of substrates by human Hint and E. coli Hint ...... 9

Table 2. Sequence alignment for ecHinT, human and rabbit Hint ...... 22

Chapter Two

Table 1. Data collection and refinement statistics for echinT and its H101A mutant GMP

complexes ...... 33

Table 2. Mutagenic forward primers for the C-terminus mutants ...... 36

Table 3. Steady-state and pre-steady-state kinetic parameters for ∆114-119 and ∆117-

119 C-terminus deletion mutants ...... 65

Table 4. Steady-state kinetic parameters for L114A, H116A, K117A, and L119A C-

terminus alanine mutants ...... 66

Chapter Three

Table1. Bacterial strains used in this study ...... 81

Table 2. Primers used for sequence verification PCR reaction ...... 82

Chapter Four

Table 1. Inhibition constants determined in HEPES buffer (pH7.2) at 25ºC ...... 123

xi

Chapter Five

Table 1. Summary of IC 50 values ...... 147

Table 2. Summary of the ecHinT-LysRS and ecHinT-DadA studies ...... 160

xii

List of Figures

Chapter One

Figure 1. Reactions catalyzed by HIT superfamily ...... 3

Figure 2. P-N bond hydrolysis of adenosine phosphoramidates by Hint ...... 8

Figure 3. X-Ray Crystallographic structure of hHint1 active-site bound AMP (PDB:

1KPF) ...... 10

Figure 4. a) X-ray crystallographic structure of hHint1 with bound AMPCP (PDB:

1AV5) ...... 15

Figure 4. b) Structure of the hHint1 C-terminus loop ...... 16

Figure 5. Rabbit Hint1 guanosine monophosphate binding pocket ...... 17

Figure 6. Comparison of the C-terminal domain for ecHinT (PDB code 3N1S)

(green/cyan) and hHint (PDB code 1av5) ...... 24

Figure 7. Catalytic processes involving lysyl-tRNA synthetase and Hints ...... 25

Figure 8. Proposed catalytic mechanism for lysyl-adenylate hydrolysis by Hints ...... 26

Chapter Two

Figure 1. Difference electron density map for ecHinT ...... 42

Figure 2. ecHinT structure superimposed on the structure of hHint1 ...... 44

Figure 3. ecHinT C-terminus structure ...... 46

Figure 4. Environment of GMP binding ...... 49

Figure 5. a) Environment around GMP in domain A showing the disposition of the

HxHxHxx motif of wild-type ecHinT compared with that of the H101A mutant...... 52

xiii

Figure 5. b) View of binding pocket in ecHinT showing GMP phosphate interactions ....53

Figure 6. Secondary structure analysis ...... 56

Figure 7. Thermal stability studies...... 58

Figure 8. Size-exclusion chromatograms from the Superdex G 200 column ...... 61

Figure 9. pH-dependence rate profile for wild-type ecHinT, ∆114-119 and ∆117-119 C-

terminus deletion mutants ...... 67

Figure 10. Stopped flow trace of ecHinT exhibited a biphasic profile, a burst phase

followed by a linear phase ...... 70

Figure 11. Adenylation of wild-type ecHinT and ∆114-119 mutant by ec LysU ...... 72

Figure 12. Time-dependence of ecHinT and ∆114-119 mutant adenylation by ec LysU .73

Chapter Three

Figure 1. Sequence verification PCR ...... 90

Figure 2. Phosphoramidase assay to detect ecHinT activity in E. coli cell-free lysates ....91

Figure 3. Summary of metabolic fingerprints ...... 92

Figure 4. Bacterial growth curves of wild-type E. coli BW25113 and ∆hinT mutant in M9

medium in the presence of either 20 mM glucose or D,L-alanine...... 94

Figure 5. Alanine transport and metabolism in E. coli ...... 95

Figure 6. RT-PCR was performed with equivalent amounts of mRNA obtained from

wild-type E. coli BW25113, ∆hinT, and ∆ycfL mutants using dadA primers ...... 97

Figure 7. ecHinT structural and activity requirement for phenotype rescue ...... 99

Figure 8. Inhibition of phenotype rescue by ecHinT inhibitor ...... 102

xiv

Figure 9. Superposition of the eight independent monomers observed in the structure of the ecHinT-GMP complex ...... 107

Figure 10. Phenotype rescue by the alanine scan mutants ...... 108

Chapter Four

Figure 1. Structures of a) Tryptamine 5’-adenosine phosphoramidate and its b) ester and

c) carbamate analogues ...... 122

Figure 2. All compounds exhibited a non-competitive inhibition profile ...... 126

Figure 3. ecHinT inhibition resulted in impaired phenotype ...... 132

Chapter Five

Figure 1. D-amino acids IC 50 determination in E. coli strains: wild-type BW25113, ∆hinT and ∆dadA ...... 143

Figure 2. Sensitivity of E. coli strains, wild-type BW25113, ∆hinT and ∆dadA , to L- alanine treatment ...... 146

Figure 3. Growth kinetic curves of three E. coli strains: wild-type BW25113, ∆hinT and

∆dadA ...... 149

Figure 4. D-alanine and D-lysine Synergism ...... 150

Figure 5. Effect of ecHinT inhibition on E. coli sensitivity to D-alanine ...... 151

Figure 6. D-alanine toxicity in mammalian cell lines ...... 153

Figure 7. The structure of peptidoglycan ...... 156

Figure 8. D,L-alanine metabolism and possible ecHinT-LysRS regulation ...... 161

xv

List of Schemes

Chapter Two

Scheme 1. Proposed kinetic mechanism for both phosphoramidate and acyl-AMP hydrolysis by Hints ...... 69

Chapter Three

Scheme 1. General synthetic scheme of TpGc ...... 101

Chapter Four

Scheme 1. General synthetic scheme of compounds 1 and 2 ...... 121

Scheme 2. General synthetic scheme of compounds 3 and 4 ...... 124

Scheme 3. General synthetic scheme of compound 5...... 125

xvi

List of Abbreviations

AARS Aminoacyl tRNA synthetase

ADP Adenosine diphosphate

AIPA Adenosine indole propionic acid

Ala Alanine

AMP-lysine AMP-N-ε-(N-α -acetyl lysine methyl ester) 5'-phosphoramidate

AMP Adenosine 5'-monophosphate

AMPCP Adenosine 5'-( α,β-methylene) diphosphate

AMP-NH2 Adenosine 5'-monophosphpramidate

AP 3A Diadenosine p1,p3-triphosphate

AP 4A Diadenosine p1,p4-tetraphosphate

Arg Arginine

Asp Aspartic acid

ATP Adenosine 5'-triphosphate

AZT 3'-Azido-3'-deoxythymidine

CD Circular dichroism

Cdk7 Cyclin dependent kinase 7 cDNA Complementary DNA

CNS Central nervous system

DadA D-amino acid dehydrogenase

DCPIP 2,6-dichlorophenol-indolphenol

DMBA 7,12-dimethylbenzanthracene

DMEM Dulbecco's modified eagle medium

xvii

DNA 2'-Deoxyribonucleic acid

DTT Dithiothreitol

E. coli Escherichia coli ec LysU E. coli lysyl tRNA synthetase ec DHFR E. coli dihydrofolate reductase ec Hin T E. coli histidine triad nucleotide-binding protein

EDC Ethyl dimethylaminopropyl carbodiimide

EDTA Ethylenediaminetetracetic acid

ESI-HRMS electrospray ionization high-resolution mass spectrometry

FAD Flavin adenine dinucleotide

Fhit Fragile histidine triad

GalT Galactose-1-phosphate uridylyltransferase

Glu Glutamic acid

Gly Glycine

GMP Guanosine 5'-monophosphate

GMP-Lysine GMP-N-ε-(N-α-acetyl lysine methyl ester ) 5'-phosphoramidate

GTP Guanosine 5'-triphosphate

h Hour

1H NMR Proton nuclear magnetic resonance

HEPES 4-(2-Hydroxyethyl)piperazine-1-ethanesulfonic acid

HI-FBS Heat-inactivated fetal bovine serum

HINT Histidine triad nucleotide binding protein

HIT Histidine Triad

xviii hHint1 Human histidine triad nucleotide-binding protein1

His Histidine

HPLC High performance liquid chromatography

IPTG Isopropyl -β-D-thiogalactopyranoside

IC 50 Inhibitory concentration

Kan Kanamycine

kDa Kilodalton

Ki Inhibition constant

LB Luria bertani medium

LC Liquid chromatography

ESI-MS/MS Electrospray ionization tandem mass spectrometry

Lys Lysine

LysRS Lysyl-tRNA synthetase

M Molar

MALDI-TOF Matrix assisted laser desorption ionization-time of flight

Min Minute

MITF Microphthalmia-associated transcription factor

mg milligram

ml milliliter

mM millimolar

MOPS 3-(N-morpholino) propanesulfonic acid

MRE Mean residue ellipticies

mRNA Messenger RNA

xix

MTX Methotrexate

Mw Molecular weight

NADPH β-Nicotinamide adenine dinucleotide 2'-phosphate-reduced form

nm Nanometer

NMP Nucleotide monophosphate

NMR Nuclear magnetic resonance

OD Optical density

ORF Open reading frame

PAGE Polyacrylamide gel electrophoresis

PCR Polymerase chain reaction

PBS Phosphate-buffered saline

PDB

PEG Polyethylene glycol

31 pNMR Phosphorus nuclear magnetic resonance

Pi Inorganic phosphate

PM Phenotype microarray

PMSF Phenylmethanesulfonyl fluoride

P-N bond Phosphorus -Nitrogen bond

Ppase Yeast inorganic pyrophosphatase

Ppi Inorganic pyrophosphate

PVDF Polyvinylidene fluoride qRT-PCR Quantitative real time polymerase chain reaction

RNA Ribonucleic acid

xx

RNase H Ribonuclease H rpm Revolutions per minute

RPMI Roswell park memorial institute medium

SD Standard deviation

SDS Sodium dodecyl sulfate

SDS-PAGE Sodium dodecyl sulfate polyacrylamide gel electrophoresis

SEC Size exclusion chromatography

SLS Static light scattering

TB Terrific broth

TFIIH Transcription factor II H

TLC Thin layer chromatography

TMP Trimethylphosphate

TpAd Tryptamine adenosine phosphoramidate

TpGc Tryptamine guanosine carbamate

Tris Tris(hydroxymethyl)aminomethane tRNA Transfer RNA

Trp Tryptophan

UDP Uridyl diphosphate

USF2 Upstream simulatory factor 2

UV Ultraviolet

Val Valine vol Volume

WT Wild type

xxi

X-gal 5-Bromo-4-chloro-3-indolyl-β-D-galactopyranoside

µM micromolar

µl microliter

xxii

Chapter One – Introduction

Histidine Triad Nucleotide Binding Proteins

1

I. Discovery and Classification of HIT Superfamily

HIT proteins were initially recognized and classified based on their virtual

sequence similarity regardless of their functions. Generally, the HIT superfamily can be

divided into five branches 1; GalT, Aprataxin 2, DcpS/DCS-1, Fhit and Hint. 1; 3; 4

Although members of each branch carry out a distinct enzymatic activity (Fig. 1), the absolutely conserved motif, His-X-His-X-His-XX (X is a hydrophobic residue), can be found in all members of the superfamily. 2; 5

The histidine triad nucleotide binding proteins (Hint) are the most ancient branch and are considered to be the ancestor of the HIT superfamily. Search results from the

Basic Local Alignment Search Tool (BLAST) program analyses indicate that Hints are found in all kingdoms of life. 6 While prokaryotes typically carry only one Hint gene,

eukaryotes generally contain two to three different Hint genes. 6 For example, three human Hint genes have been identified (hHint1, hHint2 and hHint3), while Drosophila have two genes resulting from (Chou, T. F. and Wagner, C. R. unpublished communication ). Although mammalian Hints share high sequence identity

(93%), eukaryotic Hints, generally, show larger variations within their core amino acid sequences. Remarkably, several predicted proteins from bacteria and archaea have been identified that have greater identity to mammalian Hints than do eukaryotic counterparts. 7

Hints are homodimeric proteins that have been shown to function as purine

phosphoramidases. 8; 9; 10; 11 Brenner and coworkers were the first to describe the

phosphoramidase activity of Hint1, by demonstrating that rabbit Hint1 and yeast Hnt

2

Figure 1. Reactions catalyzed by the HIT superfamily.

O O GalT N N Gal-1-P O O N O O O N O Glc-O P O P O O Gal-O P O P O O O- O - O - O - Glc-1-P OH OH OH OH

NH 2 NH 2 N N N N

O O N N Aprataxin O N N DNA-O P O P O O HO P O O O - O - O -

OH OH DNA-PO 3 OH OH

Me O Me O N N NH NH N N O O N NH2 O N NH 2 DcpS HO P O O3P O P O P O O O O- O - O - OH OH OH OH PPi NH2 NH2 N N N N

O O N N Fhit O N N AMP-O P O P O O HO P O O O - O - O - OH OH ADP OH OH

NH 2 NH2 N N N N

O N N O N N H Hints Alkyl Group N P O O HO P O O O- O- OH OH OH OH Alkyl Amine

NH 2 NH2 N N N N O O N N Hints O N N Alkyl Group O P O O HO P O O - O O - OH OH Alkyl acid OH OH

3 convert adenosine 5’-monophosphoramidate (AMP-NH 2) to AMP and ammonia and adenosine lysine phosphoramidate to AMP and lysine (Fig. 1). 9

The evolutionarily conserved fragile histidine triad (Fhit) protein is composed of

147 amino acids and represents the second branch of the histidine triad (HIT) protein superfamily. 2 The Fhit branch is found only in animals and fungi. 2 Like Hint1, Fhit is a homodimer with two identical purine mononucleotide binding sites. 12 Fhit, like all

members of the HIT superfamily, possesses nucleotide-binding and nucleotide-

hydrolyzing activity. Fhit catalyzes the hydrolysis of Ap nA (n = 3 or 4), producing AMP as one of the two mononucleotide products (Fig. 1). 12

The human Fhit gene is disrupted in many tumors resulting in impaired or missing expression of the encoded Fhit protein. 13 Loss of Fhit expression was reported to represent an early event during neoplasia and was correlated with tumor progression and detrimental prognosis. 14; 15 Consistent with its tumor-suppressor function,

expression of Fhit in multiple Fhit-negative tumor cells resulted in inhibition of tumor

growth in Fhit knock-out mice. 16 The molecular basis of the tumor-suppression activity of Fhit appears to be mediated by a tyrosine phosphorylation-dependent modulation of the Akt-survival pathway. 17; 18 Site-directed mutagenesis of the nucleophilic histidine 96 of the HIT motif to asparagine generates a catalytically inactive mutant Fhit protein.

However, the enzymatically dead protein still shows tumor-suppressive activity, suggesting that the hydrolysis of diadenosine polyphosphates is not involved in the tumor-suppressive function of Fhit. 19; 20

Galactose-1-phosphate uridylyltransferase (GalT) is the second enzyme in the

Leloir pathway of galactose utilization. 21; 22 The GalT branch consists of nucleoside

4 monophosphate transferases, including galactose-1-phosphate uridylyltransferase, diadenosine tetraphosphate phosphorylase, and adenylyl sulfate-phosphate adenylytransferase. 2

Though GalT and Hint are both dimers, the GalT monomer is more than twice the size of the Hint dimer. Except for the slightly altered HIT motif, members of the GalT branch share little overall sequence similarity with Hint or Fhit branch members.

Nevertheless, the three-dimensional structure of the monomer GalT is similar to that of the Hint or Fhit homodimer. 2; 7 GalT also has the same folding structure in the HIT region of the protein and mode of nucleotide binding as the latter two proteins. 2; 7

The fourth branch is Aprataxin, a zinc finger containing member of the HIT

superfamily that is mutated in ataxia-oculomotor apraxia. 1; 2; 23 Human Aprataxin is

expressed in two splice forms 23 ; the minor splice form that encodes for a Hint domain and an apparent C-terminal zinc finger, and the major splice form that encodes for the N- terminus domain. 23; 24 The Hint domain of Aprataxin, which shows 31% amino acid identity with rabbit and human Hint, is believed to contain a dimerization interface as well as a phosphoramidase catalytic domain. 2 In addition to the phosphoramidase activity, members of this branch were shown to have Ap 4A hydrolase activity and

DNA/RNA binding properties. 1 Recently, the physiological substrates for Aprataxin

have been proposed to be abortive adenylated DNA ligation intermediates (Fig. 1). 25

The last branch of the HIT superfamily is the scavenger mRNA decapping enzyme, DcpS/DCS-1, a 7-methyl-GpppG hydrolase (Fig. 1). 3; 4 Although DcpS shares no obvious with other HIT proteins, except for the conserved HIT

5 motif, its 80 kDa size determined by gel filtration suggests that it too has the potential to form a homodimer. 26

Human DcpS was cloned and purified as an enzyme that hydrolyzes compounds

such as 7meGpppG and small capped oligoribonucleotides and has been proposed to

function in the hydrolysis of short oligomers that remain after 3’ to 5’-exonucleolytic

degradation of mRNA (Fig. 1). 3 However, the specificity for methylated guanosine cap demonstrates that DcpS is distinct from the previously reported Fhit, which can efficiently hydrolyze both methylated and unmethylated structures. 27; 28

The central histidine of the HIT motif in DcpS (His-277) is involved in its pyrophosphatase activity as it catalyzes the formation of a covalent nucleotidyl phosphohistidyl intermediate (Fig. 1). 3 Consistent with the significance of the central

histidine in the pyrophosphatase activity, substitution mutagenesis of the central histidine

in DcpS (H277N) demonstrates that DcpS also contains a functional HIT motif. 3

II. Phosphoramidase Activity of Hints

Rabbit Hint was initially purified, as an abundant cytosolic protein, based on its

nucleotide binding properties using adenosine agarose affinity columns. 29 Hint

phosphoramidase activity was first reported by Brenner and coworkers when they

screened a series of compounds as Hint substrates and found that Hint hydrolyzes the

natural product adenosine-5’-monophosphoramidate (AMPNH 2) in an active-site dependent manner at second order rates exceeding 1×10 6 M-1 s-1. 9 Wagner and coworkers later demonstrated by 31 P NMR that human Hint1 and a previously uncharacterized E. coli ORF, ycfF (renamed ecHinT), were purine nucleoside phosphoramidases. 8

6

A sensitive and continuous fluorescence assay to measure the phosphoramidase activity of Hint was developed by Wagner and coworkers. 11 Using this assay, the general outline of human and E. coli Hint phosphoramidase substrate specificities was determined. 11 A series of substrates linking the naturally fluorogenic indole derivatives to nucleoside 5’-monophosphates was synthesized and their steady-state kinetic parameters of hydrolysis by Hint were evaluated (Fig. 2). The substrate specificity studies revealed a preference for purines over pyrimidines and unhindered over sterically arylated amines. Consistent with the observed hydrogen bonding between the 2’-OH group of adenosine monophosphate and the active-site residue (Fig. 3), maintenance of an electrophilic or hydrogen bonding group at the ribose 2’-position appears to be an essential requirement for a Hint substrate (Fig. 3). 11 In addition, Acyl nucleoside

monophosphate was found to be an excellent substrate with kcat /K m values ranging from

10 6 to 10 7 M-1s-1 (Table 1). 11

Despite strong evidence that Hints are phosphoramidases, the identity of the natural phosphoramidate substrates has been a mystery. With the exception of enzyme active-site lysine-adenylates that are intermediates for enzymes, such as DNA and RNA ligases 30 , and the observation that adenylyl sulphate transferase, in the presence of high

31 concentrations of ammonia, could generate AMP-NH 2 , common naturally occurring macromolecular or small molecule purine based or pyrimidine based phosphoramidates have not been reported. Recently, both bacterial and human Hint1 have been shown to hydrolyze acyl-AMP generated by lysyl tRNA synthetase, which suggests that acyl-AMP generated by aminoacyl tRNA synthetases may be a natural substrate of Hints. 32

7

Figure 2. P-N bond hydrolysis of adenosine phosphoramidates by Hint.

NH2 H NH2 N N N N N H O N N N H O N N N P O O HO P O O OH HINT NH OH + 2 OH OH OH OH

8

Table 1. Hydrolysis rates of the most efficiently hydrolyzed Hint1.

NH2 NH O 2 N N N N N NH O O N N O N N O N N H NH2 HN O P O H N P O O P O O O- N O HN O- HN O- OH OH OH OH OH OH GMP-Tryptamine AIPA AMP-Tryptamine

-1 -1 -1 3 kc at (s ) Km(µ M) kcat /K m (M s )x 10

hHint1 ecHinT hHint1 ecHinT hHint1 ecHinT AMP-Tryptamine 2.3±0.1 4.5±0.1 0.13±0.02 5.2±0.2 17000±2000 870±50 GMP-Tryptamine 2.3±0.1 4.0±0.4 0.21±0.02 6.0±1 11000±1000 700±200 AIPA 1.98±0.02 4.0±0.1 0.04±0.002 4.0±0.4 53000±500 1000±200

9

Figure 3. X-ray crystallographic structure of hHint1 active-site bound AMP (PDB:

1KPF).

10

In vitro studies indicate that human Hint1 can bind various , including

33; 34 AMP, ADP, and the diadenosine polyphosphates Ap 3A and Ap 4A. Rabbit Hint1 also binds several purine nucleosides and nucleoside 5’-phosphates. 29 In vitro, the human and rabbit Hint1 proteins do not hydrolyse dinucleoside polyphosphates (Ap nA) or ATP 5; 7 , but they do hydrolyse ADP. 33; 35

Although the HIT superfamily is comprised of at least five distinct subfamilies,

extensive investigations of their catalytic and kinetic mechanisms of action have only

been carried out on GalT and Fhit. In both cases, catalysis proceeds with the formation

of a histidine-NMP intermediate and inversion of the phosphorous configuration,

followed by transfer to either water (Fhit) or galactose (GalT). For GalT, uridinylation of

the nucleophilic histidine by UDP-glucose and subsequent deuridylylation by galactose-

1- phosphate were found to be rapid with rates of 281 s -1 (40ºC) and 166 s -1 (40ºC),

36 respectively. Since neither one of these steps was found to coincide with the k cat value

(62 s -1 at 40ºC), it was proposed that either the product release or a conformation change may be the rate limiting. 36 Most importantly, structural analysis of the GalT active-site

reveals that two of the conserved histidines and a close cysteine formed a coordination

complex with an atom of Fe. 37

While conversion of Ap 3A to ADP and AMP by Fhit proceeds through a double

displacement mechanism, less is known about the specific catalytic steps. 38 When the putative nucleophilic histidine was replaced by glycine, free histidine was found to rescue the enzymatic activity, in addition, AMP-imidazole and AMP N-methylimidazole were found to be substrates. 39 Results of steady-state kinetic and mutagenesis studies have implied that one member of the catalytic triad, His-98, is probably responsible for the

11 donation of a proton to the leaving group, as well as the substrate binding. 40 This is in marked contrast to GalT, in which the corresponding histidines are coordinated to Fe.

The mechanism of Hint catalysis has been proposed to proceed through a two-step mechanism that is analogous to Fhit catalysis. In the first step, the active-site nucleophilic His-112 is thought to form a covalent Hint-AMP intermediate followed by release of nucleoside monophosphate after water hydrolysis of the intermediate. The proposed mechanism was supported by site-directed mutagenesis studies, in which activity was abolished by mutations of the nucleophilic histidine. Zhou et al have provided the first elucidated kinetic study of hHint1 by pre-steady-state and steady-state kinetic analyses, pH-vs-rate analysis and viscometric studies (Zhou, X. and Wagner, C.

R. unpublished data), thus setting a foundation for a comprehensive understanding of the rules governing Hint1 catalysis in particular and the HIT superfamily in general.

III. Structural Properties of Hints

Crystallographic data have been reported for six human Hint1 and five rabbit Hint proteins as either the wild-type apo or AMP- or GMP-bound complexes. 5; 6; 10; 33 Two persistent errors have been reported in the Hint protein scientific literature. When first isolated, the sequence of bovine Hint was annotated as a PKC inhibitor-1. 41; 42 However,

subsequent studies have demonstrated that, while Hint binds to PKC, it is a modulator of

PKC function and not a PKC inhibitor. 5; 29; 43 It was also reported that “bovine PKCI-1”

dried onto nitrocellulose filters binds Zn +2 . 42 This observation was supported when the

active-site of Hint1 was examined and shown to contain four histidines, which were

proposed to bind to Zn +2 . 44 In 1996, a crystal structure of the “zinc form” of Hint was

published that had no zinc electron density and no change in structure from the non-zinc

12 form of the protein. 5 Brenner and coworkers reported that metal binding at the

nucleotide binding site in Hint would be competitive with binding to the ribose ring of

the nucleoside or nucleotide. 7

IV. Structural Studies of Hints

X-ray crystallography studies have demonstrated that hHint1 exists as a homodimeric protein with α + β overall folding topology (Fig. 4a). Each monomer contains a five-stranded antiparallel sheet and two helices. The two monomers are brought together in the dimer to form a 10-stranded antiparallel sheet, such that the central helix from one monomer packs against the central helix from the other monomer.5

Residues in the nucleotide binding site are not involved in the dimerization interface and are positioned 25Å apart on the opposite ends of each monomer (Fig. 4a). The carboxy terminal amino acids of both monomers wrap around one another allowing the carboxylate group of Gly-126 to form a salt bridge with Arg-119 from the other monomer. In addition, the C-terminus of each monomer resides in close proximity to the catalytic residues from the opposite monomer and does form a range of contacts with the nucleotide binding site (Fig. 4b).

In the crystal structures of the complexes, Hint-GMP, Hint-AMP and Hint-8Br-

AMP, two identical nucleotide-binding sites were found in the homodimeric structure of

Hint. 7 A set of conserved hydrophobic residues composes the binding site for the purine base, while conserved nonpolar and polar residues form the binding site for the ribose.

Conserved polar residues, including His-110 and His-112 from the HIT motif, make up the binding site for the α-phosphate (Fig. 5). 7 Sequences, from 17 HIT proteins, within

the nucleotide binding region have been compared and only six residues were found to be

13 identical in every HIT protein (Phe-19, His-51, Leu-53, His-110, His-112, His-114). Five of the six absolutely conserved residues make direct contact with the nucleotide. 7

V. Role of Hint Homodimerization

To study the importance of homodimerization on the catalytic activity of Hint1,

Wagner and coworkers designed the monomer version of Hint1 by destabilizing the

dimerization interface. Replacement of Val97 of hHint1 with Asp, Glu, or Arg resulted

in monomeric mutants of Hint1, which were characterized by a combination of size-

exclusion chromatography, static light scattering, and chemically induced dimerization

studies. Significant perturbations of the active-site residues were not detected by

molecular dynamics simulations of the monomeric hHint1. The combined kinetic and

structural results demonstrate that, for monomeric hHint1, the efficiency (k cat /K m) of

acylated-AMP hydrolysis, but not maximal catalytic turnover (k cat ), is dependent on homodimerization. 45

14

Figure 4. a) X-ray crystallographic structure of hHint1 with bound AMPCP (PDB:

1AV5). hHint1 exists as a homodimeric protein with α + β overall folding topology.

Residues in the nucleotide binding site are not involved in the dimerization interface and

are positioned 25Å apart on the opposite ends of each monomer.

15 b) Structure of the hHint1 C-terminus loop. The C-terminus of each monomer resides in close proximity to the catalytic residues from the opposite monomer. A tryptophan residue (Trp-123) in the C-terminus of one monomer (blue) is in close proximity to the active-site of the other monomer (yellow).

16

Figure 5. Rabbit Hint1 guanosine monophosphate binding pocket: the signature histidine triad residues are mainly responsible for stabilizing the binding to the α-phosphate,

whereas the base moiety of the nucleoside is sandwiched between two phenylalanines

and an isoleucine.

17

VI. Possible Biological Functions of Hints

During the last decade, evidence has begun to accumulate that Hints are involved

in a wide array of biological processes. Two-hybrid screening experiments revealed that

Hint1 interacts with Cdk7, the catalytic subunit of the cyclin dependent kinase activation

complex Cdk7-cyclin H-MAT1. 46 Analogous interactions have been found between

HNT and Kin28, the yeast orthologs of Hint and Cdk7. 9 Nevertheless, Hint1 mouse knock-out studies indicate that Hint1 is not required for Cdk7 function. 47 hHint1 has also been shown to directly interact with human Pontin and Reptin in the TCF-β-catenin transcription complex. 48 Further investigations with the knock-out mice have shown that

at 2~3 years of age, both heterozygous and homozygous mice were been found to have an

increased susceptibility to the induction of ovarian and mammary tumors by the

carcinogen 7, 12-dimethylbenzanthracene (DMBA) and to the production of spontaneous

tumors. 49 Up regulation of Hint1 and the significantly reduced in vivo tumorigenicity of

5-aza-dC-treated human non-small cell lung cancer cell line NCI-H522 demonstrated that hHint1 might be a tumor-suppressor. 50 The recent observation that hHint1 is involved in the modulation of apoptosis, independent of its enzymatic activity, suggests a possible mechanism for its tumor-suppressor activity. 51 Consistent with its potential tumor- suppressive function, over-expressed Hint1 was found to inhibit cell growth, as well as activator protein-1 activity in the human colon cancer cell line SW480. 52 In addition, it has been suggested that the tumor-suppressing ability of Hint1 is related to a potential role in enhancing DNA repair processes associated with the histone variant, H2AX and the gene, ataxia telangiectasia mutated (ATM). 53

18

Because Hint is widely expressed in the brain, including the frontal cortex, Hint1 has been suggested to have a physiological function in the CNS. 54 Studies with knock-

out mice have revealed a decreased ability for spontaneous locomotion and

supersensitivity to amphetamine, suggesting that Hint1 is involved in the regulation of

postsynaptic dopamine transmission. 55 Recent studies have provided evidence that control of the activity of morphine on the function of the µ-opiod receptor is associated with zinc-dependent binding of Hint1 to the N-terminal cysteine rich domain of specific

GTPase activating proteins of receptor-activated GazGTP subunits (RGSZ’s). 56 Further analysis has revealed that the Hint1-RGSZ complex functions as an adaptor that facilitates control of the downstream activity of PKC γ by regulating its association with the µ-opiod in a zinc-dependent manner; thus preventing desensitization of the µ-opiod receptor. 56

Like Hint1, human Hint2 is a purine nucleotide phosphoramidase. 57 hHint2 has

61% sequence similarity to hHint1 and is thought to be a homodimer. 57 hHint2 is found

exclusively in mitochondria and has been shown to be a mitochondrial apoptotic

sensitizer, which is down-regulated in hepatocellular carcinomas. 57 It also appears to be

involved in the regulation of calcium-independent steroidogenesis. 58 hHint2 silencing in

H295R cells resulted in a marked reduction of the steroidogenic response. The duration

of the mitochondrial calcium signal induced by angiotensin II was also reduced upon

hHint2 silencing, but not affected after its overexpression, suggesting that under basal

conditions, hHint2 is optimally expressed to regulate steroidogenesis. 58

Little is known about Hint3. Wagner and coworkers have discovered that hHint3,

which is only 28% sequence identical to hHint1, is likely to be a distinct branch of the

19

HIT superfamily, since it is a multimeric oligomer and strongly prefers acyl-adenylates over purine nucleotide phosphoramidates as substrates. 59 Although of unknown physiological consequence, a natural monomeric polymorph has been identified. 59 In contrast to hHint1, when ectotopically expressed, hHint3 appears in both the cytoplasm and nucleus. 59

VII. Prokaryotic Hint

E. coli Hint (ecHinT) was first discovered by Wagner and coworkers when an

open reading frame, designated ycfF , at the 16-min position (161090–1161467) on the E.

coli genetic map was found to have 47% amino acid sequence identity to rabbit and

human Hint1. 8 Moreover, after cloning, purifying and characterizing the activity of the

purified protein in vitro , it was shown that the hinT gene is indeed a nucleoside

monophosphoramidase. Based on these finding, ycfF was renamed as hinT . 8 Further characterization demonstrated that bacterial Hint is homodimeric and capable of hydrolyzing adenosine and guanosine 5’-phosphoramidate monoesters significantly faster than mammalian Hint. 8 Analysis of the lysates from a constructed hinT knock-out strain of E. coli demonstrated that all of the cellular nucleoside phosphoramidase activity is due to ecHinT and that it is the only purine phosphoramidase expressed by E. coli . 8

An observed association of ecHinT with growth under high salt conditions was

the first suggestion of a possible physiological role of Hint in bacteria. 8 Interestingly,

the reported phenotype appeared not to be connected to the phosphoramidase activity of

ecHinT, since the expression of either wild-type ecHinT or the catalytically-impaired

mutant, H101A, appeared to rescue the knock-out strain from the cation-dependent

phenotype. 8

20

Despite their ubiquity and high sequence similarity to eukaryote Hint1 (i.e, ecHinT is 48% indentical to hHint1), only a few studies of prokaryote Hints have been reported. ecHinT is alleged to form stable protein-protein interactions with six species: a putative oxidoreductase and formate dehydrogenase (b1501), the heat shock protein 70

(Hsp70), the β-subunit of DNA polymerase III ( dnaN ), a membrane-bound lytic murein transglycosylase D ( dniR ), ET-Tu elongation factor ( tufA ), and a putative synthetase

(yjhH ). 60 In addition, Mycoplasma HinT has been shown to interact with two

membrane proteins (P60 and P80). 61; 62 However, the physiological and biochemical

importance of these interactions remains unresolved.

Sequence analysis of both prokaryote and eukaryote Hints revealed that a

significant amount of sequence diversity between species resides in the C-terminus.

Although ecHinT and hHint1 share nearly 50% sequence similarity, the C-termini are not

sequence similar nor of the same length. Alignment of the human (126 residues), rabbit

(126 residues) and E. coli Hint (119 residues) sequences (Table 2) reveals that there is a

12-residue deletion in the N-terminus of ecHinT compared to the mammalian sequence

(residues 5-16), a single residue insertion at residue 79 of ecHinT, and a four-residue extension of the C-terminus of ecHinT compared to mammalian sequences (Table 2). To probe the impact of the C-terminus of Hints, Wagner and coworkers constructed two chimeric proteins in which the C-terminal loops were switched between hHint1 and ecHinT. The Human/ec chimera, which contains the C-terminus of ecHinT, exhibited

nearly identical specificity constants (k cat /K m) to those found for ecHinT, whereas the specificity constants of the ecHinT/Hs chimera, which contains the C-terminus of hHint1, were found to approximate those for hHint1. The conclusion was that substrate

21

Table 2. Sequence alignment for ecHinT, human and rabbit Hint.

1 9 19 29 39 ec _HINT MAEE------TIFS KIIRREIPSD IVYQDDLVTA FRDISPQAPT HILIIPNILI 1 11 21 31 41 51 human_HINT madeiakaqv arpggdtifg kiirkeipak iifeddrcla fhdispqapt hflvipkkhi rabbit_HINT madeiakaqv arpggdtifg kiirkeipak iifeddqcla fhdispqapt hflvipkkhi

49 59 69 79 89 99 ec_HINT PTVNDVSAEH EQALGRMITV AAKIAEQEGI AEDGYRLIMN TNRHGGQEVY HIHMHLLGGR 61 71 81 91 100 110 human_HINT sqisvaeddd esllghlmiv gkkcaadlgl –nkgyrmvvn egsdggqsvy hvhlhvlggr rabbit_HINT sqisaaedad esllghlmiv gkkcaadlgl –kkgyrmvvn egsdggqsvy hvhlhvlggr

109 119 ec _HINT PLGPMLAHKGL 120 126 human_HINT qmhwppg---- rabbit_HINT qmnwppg----

22 specificity could be transferable by C-terminal loop exchange between hHint1 and ecHinT. 63

Recently, the first comprehensive high-resolution crystal structures of the full

length N-terminal and C-terminal residues of wild-type and the H101A mutant of ecHinT

were reported. 64 As described in Chapter Two, the structure of ecHinT shows the general α and β type fold observed for other Hint proteins 5; 6; 10; 33 and is characterized by five antiparallel β-strands and a central helix (residues 56-76). The homodimer is

formed by the interactions of the central helix and the joining of β-strands into a 10- stranded antiparallel sheet. The C-terminal residues of both monomers fold across one another, but unlike the human Hint1 structures, the terminal residues do not form a salt bridge as observed between domains A (Gly126) and domain B (Arg119) of the human

Hint1 structure (Fig. 6). 33

VIII. The Interaction of Hints with Aminoacyl tRNA Synthetases

Recently, Wagner reported the first evidence of a connection between Hint1 proteins and LysRS. 32 A series of catalytic radiolabeling, mutagenesis, and kinetic experiments was conducted with purified LysRSs and Hints from human and E. coli. The results of these studies have confirmed that lysyl-AMP generated by LysRS is a natural substrate for ecHinT and hHint1 (Fig. 7). 32 The proposed mechanism for the lysyl-AMP

hydrolysis by Hint proceeds through the formation of an adenylated-Hint intermediate

followed by water hydrolysis of the intermediate (Fig. 8). 32 Site-directed mutagenesis studies of the active-site histidine triad abolished Hint labeling when either His-101 of ecHinT or His-112 of hHint1 was replaced by either alanine or glycine. 32

23

Figure 6. Comparison of the C-terminal domain for ecHinT (PDB code 3N1S)

(green/cyan) and hHint (PDB code 1av5) (yellow/red) highlighting the hydrogen bond between G126 (red) and R119 (yellow) in hHint1.

24

Figure 7. Catalytic processes involving lysyl-tRNA synthetase and Hints.

25

Figure 8. Proposed catalytic mechanism for lysyl-adenylate hydrolysis by Hints.

26

Consistent with pyrophosphate being an inhibitor for aminoacyl tRNA synthetase, incubations in the presence of pyrophosphatase resulted in enhanced formation of Hint-

AMP. 32 However, the rationale for the interaction of Hints and LysRS is not apparent, since the over-expression of Hints appears not to inhibit protein translation or to have any other deleterious effects on bacteria.

IX. Current Research

The research described in the current thesis elucidates the contributions to the characterization of ecHinT from the aspect of function, mechanism, and structural determinants. The current work provides the first strong evidence of a biological function of the highly conserved Hint in E. coli . In Chapter 2, the impact of the C-terminus as a

structural determinant of ecHinT activity and function is described. In Chapter 3, our

discovery of a bacterial phenotype for ecHinT that is dependent on the enzymatic

catalytic activity is described. Structural and activity requirements of ecHinT for the

discovered phenotype are elaborated in this chapter. Chapter 4 describes the synthesis

and evaluation of novel Hint inhibitors. The prepared inhibitors are considered as the first

cell permeable Hint inhibitors reported in the literature. Finally, in Chapter 5, the

importance of ecHinT on D-amino acids detoxification in E. coli is described.

27

Chapter Two

Probing the Impact of the ecHinT C-Terminal Domain on Structure and Catalysis

28

I. Introduction

Histidine triad nucleotide binding protein (Hint) belongs to a ubiquitous

superfamily consisting primarily of nucleoside phosphoramidates and acyl-AMP

hydrolases, dinucleotide hydrolases and nucleotidylyl transferases. The histidine triad

(HIT) superfamily has a characteristic C-terminal active-site motif, HXHXHXX, where

X is a hydrophobic residue. 63 Based on their enzymatic function, sequence composition, and structural similarity, HIT proteins have been classified into five branches; fragile HIT

(Fhit), Hint, galactose-1-phosphate uridyl transferase (GalT) 2, Aprataxin 1, and

DcpS/DCS-1. 3; 4

Hint is considered as the ancestor of the HIT protein superfamily and is highly conserved from bacteria to humans. While prokaryote genomes, including a wide array of both Gram-negative and Gram-positive bacteria, typically encode one Hint gene, eukaryotes generally express multiple forms of Hint. Although E. coli gene disruption studies have suggested that the bacterial enzyme is necessary for growth under high-salt conditions 8, the cellular function of Hint and the rationale for its evolutionary conservation in bacteria have remained a mystery.

E. coli Hint (ecHinT) is alleged to form stable potential protein-protein interactions with six species: a putative oxidoreductase and formate dehydrogenase

(b1501), the heat shock protein 70 (Hsp70), the β-subunit of DNA polymerase III ( dnaN ), a membrane-bound lytic murein transglycosylase D ( dniR ), ET-Tu elongation factor

(tufA ), and a putative synthetase ( yjhH ). 60 In addition, Mycoplasma HinT has been shown to interact with two membrane proteins (P60 and P80). 601; 62 However, the

physiological and biochemical importance of these interactions has remained unresolved.

29

In the last decade, evidence has begun to accumulate that Hints are involved in a wide array of biological processes. Mouse Hint1 gene knock-out studies have demonstrated that Hint1 acts as a tumor suppressor. 49; 55 The recent observation that human Hint (hHint1) is involved in the modulation of apoptosis, independent of its enzymatic activity, suggested a possible mechanism for its tumor-suppressor activity. 51

Consistent with its potential tumor-suppressing function, over-expressed Hint1 was found to inhibit cell growth and activator protein-1 activity in the human colon cancer cell line

SW480. 52 Hint1 has also been implicated as a modulator of central nervous system sensitivity to amphetamine. 65 Hint2, which is highly homologous to Hint1, is found in

mitochondria and appears to function as a regulator of apoptosis. 57 Recently, protein- protein interaction studies have found that the transcription factor, TFIIH, complexes of

MITF or USF2, and complexes of lysyl-tRNA synthetase (LysRS) are associated with human Hint1 (16, 17). 34; 66

Crystallographic data for six human Hint1 and five rabbit Hints as the wild-type

(wt) apo and inhibitor-bound complexes have been reported. 5; 6; 10; 33 More recently, a number of Hint-like protein structures from pathogenic organisms have been deposited in the Protein Data Bank (PDB) by several Structural Genomics Consortia. These data reveal that the typical length of the Hint is about 120 residues and show that the protein functions as a homodimer, which is characterized by an α + β fold that contains a five-

stranded antiparallel sheet and two helices. The monomers come together to form a

homodimer with a 10-stranded antiparallel sheet that makes extensive contacts between a

helix and the carboxy-terminal amino acids of one protomer and the corresponding

residues in the other protomer. 5 The histidine signature sequence of the superfamily is

30

HXHXHXX near the nucleoside phosphate binding site. There is a fourth histidine involved in the binding pocket that is not in the contiguous sequence. Structural data reported for mammalian species reveal an incomplete trace of the full-length monomers, as electron density was not interpretable for the first N-terminal 13 residues in the human or rabbit Hint structures.

Sequence analysis of both prokaryote and eukaryote Hints revealed that the greatest sequence diversity between species resides in the C-terminus. Although ecHinT and hHint1 share a nearly 50% sequence similarity, the C-termini are neither similar in sequence nor of the same length. Alignment of the human (126 residues), rabbit (126 residues) and E. coli Hint (119 residues) sequences reveals that there is a 12-residue deletion in the N-terminus of ecHinT compared to the mammalian sequence (residues 5-

16), a single-residue insertion at residue 79 of ecHinT, and a four-residue extension of the

C-terminus of ecHinT compared to mammalian sequences.

In a previous work, we have shown that the C-terminal motif of Hint1 proteins appears to be responsible for mediating phosphoramidate and lysyl-AMP substrate specificity. 11 In our effort to understand the mechanism of ecHinT catalysis, we report the first structural characterization of a bacterial Hint protein. Our data reveal the presence of four unique homodimers in the asymmetric unit of the monoclinic crystal lattice of ecHinT with an observation of conformational flexibility in terminal loop regions. The reported crystal complexes with GMP are the first structural data reported for a bacterial Hint protein. In addition, through deletion mutagenesis, steady-state and pre-steady-state kinetics, and LysRS-generated lysyl-AMP hydrolysis studies, we have probed the importance of the C-terminus of ecHinT in catalysis.

31

II. Experimental Procedures

A. Crystallization and X-ray Data Collection (Data collection and analysis were

carried out by Dr. Vivian Cody).

Recombinant wild-type ecHinT and its H101A mutant were cloned, isolated, and

purified as described previously. 8 Crystals were grown using microbatch under paraffin oil at 20 oC from enzyme incubated with GMP prior to crystallization. Protein droplets

contained 40% polyethylene glycol (PEG) 20K, 0.1 M Mg acetate, and 0.1 M Na acetate

(pH 5.0) and the sample buffer contained 20 mM Tris (pH 7.0), 1 mM EDTA, 10%

glycerol, and 2 mM GMP. The protein concentration was 7.8 mg/ml for wild-type

ecHinT and 6.7 mg/ml for the H101A mutant protein. Crystals grew over several days’

time and were cryopreserved in PEG 400 at 16%. Data were collected on beamline 11-1

at the Stanford Synchrotron Radiation Laboratory (SSRL) using the remote access

facility. 67; 68; 69 Crystals of both complexes were monoclinic, belonged to space group

P2 1 and diffracted to 1.3 Ǻ and 1.8 Ǻ resolutions for the wild-type and H101A mutant

complexes, respectively. Data were processed with HKL2000 in Denzo and scaled with

SCALEPACK for the wild-type structure and with Mosflm and SCALA for the H101A

mutant. 70 The diffraction statistics for these structures are shown in Table 1.

B. Structure Determination and Refinement

These structures were solved by molecular replacement using the coordinates of human Hint (PDB code 1av5) 5 in the program Molrep. 71 Inspection of the resulting difference electron density map was performed with the program COOT running on a

Mac G5 workstation and revealed density for the complex. 72 The sequence changes from the search model were made during the map-fitting process. To monitor

32

Table 1. Data collection and refinement statistics for ecHinT and its H101A mutant

GMP complexes.

Data collection ec hinT wt GMP H101A ec hinT GMP PDB number 3N1S 3N1T Space group P2 1 P2 1 Cell dimensions ( Ǻ) 75.18 65.30 99.12 49.33 64.66 74.75 β = 109.75 β = 109.04 Beamline SSRL 11-1 SSRL 11-1 Resolution ( Ǻ) 1.30 (1.45) 1.70 (1.76) Wavelength ( Ǻ) 0.975 0.975 Rmerge 0.165 0.080 a,b Rsym (%) 0.115 0.108 Completeness (%) a 95.2 (77.7) 86.4(25.8) Observed reflections 222,096 140,655 Unique reflections 149,986 41,405 I/ σ(I) 5.3 9.5 Multiplicity a 3.4 (2.5) 3.4 (2.1)

Refinement and model quality

Resolution range ( Ǻ) 37.96 – 1.45 35.33-1.72 No. of reflections 149,986 39,308 R-factor c 17.4 19.4 d Rfree -factor 20.3 25.2 Total protein atoms 9071 3973

Total water atoms 1068 317 Average B-factor ( Ǻ2) 20.4 24.8 Luzzati error 0.163 0.212 Rms deviation from ideal Bond lengths ( Ǻ) 0.01 0.02 Bond angles ( o) 1.52 2.03 Ramachandran plot Residues in most favored regions (%) 97.3 97.8 Residues in additional allowed regions (%) 2.6 2.0 Residues in generously allowed regions (%) 0.1 0.2 Residues in disallowed regions (%) 0.0 0.0

a The values in parentheses refer to data in the highest resolution shell. b Rsym = ΣhΣi|I h,i - | / ΣhΣi|I h,i |, where is the mean intensity of a set of equivalent reflections. c R-factor = Σ|F obs – Fcalc | / ΣFobs , where F obs and F calc are observed and calculated structure factor amplitudes. d Rfree -factor was calculated for R-factor for a random 5% subset of all reflections.

33 the refinement, a random subset of all reflections was set aside for the calculation of R free

(5%). The final cycles of refinement were carried out using the program Refmac5 from the CCP4 suite of programs. 73 The Ramachandran conformational parameters from the

last cycle of refinement, as generated by PROCHECK 74 , showed that more than 95% of

the residues have the most favored conformation and only 10 residues are located in the

disallowed regions of the protein. These residues are near the terminal residues that had

poor electron density.

Analysis of the data showed that the wild-type ecHinT crystallized with four

homodimers in the asymmetric unit of the monoclinic lattice, while the H101A mutant

complex crystallized with two homodimers in the asymmetric unit. Non-crystallographic

symmetry was not used during the refinement, and each monomer was refined

independently. There are 119 residues in each ecHinT monomer, compared with 126

residues in the human protein. The final cycle of refinement for ecHinT revealed

residues 1-117 in chain A, residues 2-119 in chain B, residues 1-116 in chain E, residues

2-119 in chain F, residues 2-116 in chain I, residues 1-119 in chain J, and residues 2-116

in chains M and N. There were eight copies of GMP. Twenty-eight copies of ethylene

diol from the buffer, as well as 1113 water molecules, were located in the structure. This

is the first report on the complete trace of the residues in a Hint protein, as previous

structures reported no interpretable electron density for the first 14 N-terminal residues

and for the last 5 C-terminal residues. In the ecHinT H101A mutant protein, the final

refinement revealed residues 5-112 for chain A, residues 5-116 for chain B, residues 5-

114 for chain E, and residues 4-114 for chain F. Four copies of GMP and 317 water

molecules were observed in the electron density.

34

C. Site-directed Mutagenesis

C-terminus deletion and alanine scan variants were generated from E. coli hinT -

pSGA02 expression vector harboring the gene encoding wild-type ecHinT using the

Quick-Change mutagenesis kit (Stratagene) in accordance with the manufacturer’s

protocol. Mutagenic primers are listed in Table 2. Encoding sequences of the C-

terminus mutants were confirmed by DNA automated sequencing (Biomedical Genomic

Center).

D. Expression and Purification of Recombinant Proteins

E. coli hinT -disrupted strain BB2 was used for the expression of ecHinT proteins

to avoid possible wild-type ecHinT contaminants, as previously described. 8 BB2 cells

containing the respective plasmids were grown at 37ºC in terrific broth (TB) medium

containing 100 µg/ml ampicillin and 50 µg/ml chloramphenicol to an OD 600 of 0.6.

Expression was induced by adding isopropyl-β-D-thiogalactopyranoside to a final concentration of 0.5 mM, and the culture was incubated for a further 10 h. Cells were harvested by centrifugation for 15 min at 6000 g, and the pellet was resuspended in buffer

A [20 mM Tris (pH 7.0), 1 mM EDTA, and 1 mM DTT]. Cell lysate was obtained by

15-s sonication (nine times) in lysis buffer [1 mg/ml lysozyme, 20 mM Tris (pH 7.00), 1 mM EDTA, 1 mM DTT, and protease inhibitor]. Lysates were centrifuged at 25,000 g at

4ºC for 30 min. ecHinT proteins were purified by an AMP-agarose affinity column

(Sigma) followed by a PD-10 desalting column (GE Healthcare). The purified proteins were exchanged with buffer A and concentrated with an Amicon stirred cell with a YM-

10 membrane (Millipore).

35

Table 2. Mutagenic primers for the C-terminus mutants.

Protein Primer

117-119 GGACCAATGCTGGCGCATTAAGCGCGCTCGAGGGTACCC

114-119 GGCCGTCCGCTGGGACCAATGTAAGCGCGCTCGAGGGTA

l114A GACCAATGGCGGCGCATAAAGGTCTGTAA

H116A TGCTGGCGGCTAAAGGTC

K117A TGGCGCATGCAGGTCTGTAA

L119A ATAAAGGTGCGTAAGCGCGTC

36

Homogeneity was analyzed by SDS-PAGE and gel filtration chromatography. Protein concentrations were determined with the Bradford protein assay (Bio-Rad). The E. coli

BS68 strain harboring the pBAS39 (derived from pET3a) plasmid for expressing ec LysU as a C-terminal His6-tag fusion protein was a gift from Dr. Paul Schimmel (The Scripps

Research Institute). 75 LysU protein was purified by Ni +2 agarose binding according to a previously published procedure. 76

E. Circular Dichroism Spectroscopy

Circular dichroism (CD) experiments were performed on a Jasco J710 spectropolarimeter equipped with a temperature-controlled water bath. Proteins at a concentration of 0.1 mg/ml in buffer A were analyzed in a quartz cuvette with a path length of 1 mm under N 2. The spectra of the wild-type and the mutants were determined in the far-UV region (190–260 nm). The spectra were accumulated and averaged over nine scans, and the buffer background was subtracted from the protein spectra. The measured ellipticities were converted into mean residue ellipticities, MRE, using Jasco software. To determine the thermal denaturation of proteins, CD unfolding measurements were obtained at 222 nm with a temperature gradient of 1°C/min between

10°C and 80°C.

F. Size-exclusion Chromatography

Size-exclusion chromatography was conducted to determine apparent molecular weights (Mw) using a Superdex G 200 column (GE Healthcare) at a flow rate of 0.5 ml/min. Protein samples (25 µM) were prepared in buffer A with 5% (vol/vol) glycerol.

Proteins were eluted with P500 buffer [0.5 M NaCl, 50 mM potassium phosphate, and 1 mM EDTA, (pH 7.0)] and the retention time of proteins was monitored by absorbance at

37

280 nm using an in-line UV detector (Beckman Gold 168). The molecular standards used were blue dextran (200 kDa), albumin (66 kDa), DHFR2-1DDG (36 kDa), carbonic anhydrase (29 kDa), cytochrome c (12.4 kDa), and aprotinin (6.5 kDa) (Sigma).

G. Steady-state Kinetics

Turnover rates were measured by following the hydrolysis of either tryptamine

5’-adenosine phosphoramidate (TpAd) or adenosine 5’-indole-3-propionic adenylate

(AIPA) as fluorogenic substrates, by the ecHinT proteins, as previously described. 11

Excitation wavelength was set at 280 nm; fluorescence emission was measured at 360 nm, and all the kinetic assays were performed in duplicate at 25ºC. The rate of hydrolysis of the substrates was determined by measuring the increase in fluorescence intensity upon the addition of the enzyme over the course of the reaction. The Michaelis-

-1 Menten constants, k cat (s ) and K m (µM) were determined by a nonlinear regression analysis of the initial velocity versus concentration using the JMP IN 7 software (SAS

Institute, Inc., Cary, NC, USA).

H. Pre-steady-state Kinetics

Pre-steady-state kinetic experiments were performed on a fluorescence two- syringe stopped-flow apparatus (model SX.18MV, Applied Photophysics). The reaction rates of the ecHinT adenylation were monitored at 25°C with either AIPA or TpAd in

HEPES buffer [20 mM, and 1 mM MgCl 2, (pH 7.2)]. The sample was excited at 280 nm

and the fluorescence emission at 340 nm was monitored with a cutoff filter of 320 nm.

The time-course curves were fitted, using the JMP IN 7 software with the equation: P(t) =

-kobst A0 – A(e ) + k hydro •t, where A0 is the fluorescence intensity at time zero, A is the

amplitude, k hydro is the hydrolysis rate of the intermediate, k obs is the pseudo-first-order

38 rate constant for the adenylation step, and t is time. The results represent the average of five experiments.

I. pH-dependence of Steady-state Kinetics

The pH-dependence of the steady-state turnover rates was determined over a pH range of 5.8-9.0 using phosphate buffer (0.1M, pka=7.2). Hydrolysis rates were determined at 25ºC and k cat values were plotted versus the pH.

J. Lysyl-AMP-dependent Adenylation of ecHinT by LysU

E. coli LysU (2 µM) was incubated with [ α-32 P]ATP (0.2 µM, 800 Ci/mmol, MP

Biomedicals) in buffer B [10 µl, 25 mM Tris-HCl (pH 7.8), 100 mM NaCl, 2 mM MgCl 2,

1 mM dithiothreitol, 113 µM Lysine, 0.02 unit/µl, inorganic pyrophosphate, and protease

inhibitor tablet (Roche Applied Science)] at 25°C for 1 min, then ecHinT proteins (5 µM)

were added and incubated for 10 min. The reaction was terminated by the addition of

SDS sample buffer (4x, 5 µl; Invitrogen). The reaction mixture was boiled for 3 min, and

the proteins were separated by SDS-PAGE and electroblotted onto a polyvinylidene

difluoride membrane. Labeled proteins were visualized by subjecting dried

polyvinylidene difluoride membranes to autoradiography with a storage phosphor screen

for 12 h, followed by scanning with a Storm 840 PhosphorImager.

K. PDB ACCESSION NUMBERS

Coordinates for wild-type ecHinT and H101A ecHinT have been deposited in the

PDB with accession codes 3N1S and 3N1T respectively.

39

III. Results and Discussion

A. Structural Properties of ecHinT

This is the first report on the crystal structures of full-length Hint from E. coli (a

119 residue homodimeric protein) and its H101A mutant, both in complex with GMP.

There are four unique homodimers in the asymmetric unit of the monoclinic crystal lattice for the wild-type ecHinT-GMP complex and two unique homodimers in the asymmetric unit for the H101A mutant complex. In all cases, GMP is bound in both monomeric domains, unlike the example of the hHint1 complex, which revealed that only one ligand was bound to the homodimer. 5; 33

The quality of the electron density maps for the ecHinT GMP complex was such that nearly all the N-terminal and C-terminal residues could be traced for the four independent homodimeric pairs of ecHinT in the asymmetric unit of the crystal lattice

(Fig. 1). As structural data reported for human or rabbit Hints revealed no interpretable electron density for the first 14 N-terminal residues, our structural analysis of ecHinT is the first report to show the complete trace of the N-terminal domain of a Hint protein.

Analysis of electron density for both the N-terminal and the C-terminal residues of ecHinT-GMP complex indicates that the loops in some monomers can adopt more than one conformation. The largest variations are observed in the N-terminal and C-terminal regions and for the loop regions opposite the guanine ring of GMP. The observation of conformational flexibility in the terminal loop regions could explain the presence of multiple homodimers in the asymmetric unit of these structures.

The structure of ecHinT shows the general α+β− type folds observed for other Hint proteins 5; 6; 10; 33 and is characterized by five antiparallel β− strands and a central helix

40

(residues 56-76). The homodimer is formed by the interactions of the central helix and the joining of β− strands into a 10 −stranded antiparallel sheet. β− sheet interactions across

the homodimeric interface are made by residues 83-89 of both domains (Fig. 2a), while

on the opposite surface, the helices are nearly perpendicular to each other, crossing Gly63

(Fig. 2b), similar to that of the hHint1 structure (Gly75). 5; 33 The central helix of ecHinT is longer than that of the human Hint1 and is displaced relative to it (Fig. 2b).

The C-terminal residues of both monomers fold across one another; however, unlike the human Hint1 structure, the terminal residues do not form a salt bridge, as observed between domains A (Gly126) and domain B (Arg119) of the hHint1 structure

(Fig. 3). 33 Rather, the C-terminal residues in ecHinT extend beyond those of the human

Hint1 sequence (Fig. 3). Additionally, the N-terminus of ecHinT reveals an extended loop region with little tertiary structure among the eight monomers in the asymmetric unit of the wild-type structure. Also shown (Fig. 3b) are the residues of ecHinT removed in the ∆114-119 mutant, which reduces the surface interface of the homodimer and make it more like that of the hHint1 structure.

41

Figure 1. Difference electron density map for a) ecHinT showing the GMP binding

pocket. 2Fo-Fc (1 σ) (blue); Fo-Fc (3 σ) (green).

42 b) ecHinT H101A mutant with GMP.

43

Figure 2. ecHinT structure superimposed on the structure of hHint1 a) View of ecHinT

domains A (green) and B (cyan) superimposed on the structure of hHint (PDB code 1av5)

(yellow/red) illustrating the beta sheet structure and the interleaving of the C-terminal

residues of each domain.

44 b) View of the opposite side of the A-B homodimer showing the helical dimer interface crossing at Gly63 (space fill) and the binding of GMP compared with hHint. The B domains were used to fit these structures. Drawn with PyMol .

45

Figure 3. ecHinT C-terminus structure a) Comparison of C-terminal domain for ecHinT

(green/cyan) and hHint (PDB code 1av5) (yellow/red) highlighting the hydrogen bond

between G126 (red) and R119 (yellow) in hHint1.

46 b) Surface representation of residues removed for the ecHinT ∆114-119 mutant. Drawn

with PyMol.

47

B. Ligand Binding

Hint protein superfamily is characterized by the conserved histidine triad (His-X-

His-X-His-XX) where X is a hydrophobic residue. Structural data reveal the proximity of a fourth His that can be considered part of the histidine triad catalytic site (Fig. 4a). These data show that the substrate GMP or AMP binds in a cleft near the

His triad binding site, with the phosphate oxygen atoms involved in hydrogen −bonding

contacts with the side chain functional groups of His101, His103, and Asn88, and the

backbone functional groups of Glu96 and Val97, in addition to solvent. Interactions of

the GMP indicate that the nucleoside ring occupies an open cleft region that is in contact

with the C-terminus of other domains which pack against this region. Furthermore, the

C-terminal residues of ecHinT extend farther than the C-terminal Gly126 of human

Hint1, which forms a salt bridge with Arg119 of the other domain (Fig. 3). This

extended structure in ecHinT permits a closer involvement of the C-terminus with the

nucleoside binding site.

The structure of hHint1 (PDB code 1av5) was shown to bind only one ligand in

the homodimer. When domain B is used to superimpose the two structures, which both

have a bound ligand, there is a good fit of the two B domains (Fig. 4b); however, there is

a displacement in the positions of the secondary structural features of domain A,

compared to the fit of domain B (Fig. 4a). Part of the shift may be caused by a collapse

of the binding pocket, as the superposition of the structure (PDB code 1kpe), with a

ligand bound in each domain, shows a similar pattern of shifts in domain A that is

intermediate between that of ecHinT and that of hHint1 (PDB code 1av5).

48

Figure 4. Environment of GMP binding a) in domain A to signature His (39, 99, 101,

103 in ecHinT) (green) and hHint (yellow).

49 b) Similar view with ecHinT domain B (cyan) and hHint with Ap2 in domain B (red).

Drawn with PyMol.

50

Comparison of the ecHinT H101A mutant-GMP complex with the wild-type structure reveals that the effect of changing His to Ala at this position permits a small shift in the phosphate group of GMP such that the contact distance from the phosphate oxygen to His101, as observed in the wild-type ecHinT structure (3.03/3.33 Ǻ for domains A and B, respectively), is shorter for the mutant model (2.35/2.91 Ǻ). The structure of the transition-state analogue ADV from human Hint1 (PDB code 1kpe) 5 shows the formation of a covalent link of the transition state analogue with His110

(2.52 Ǻ). The equivalent contact in the ecHinT-GMP complex is 3.43 Ǻ. These data

support the diminished catalytic activity observed for the ecHinT H101A mutant, as the

active-site nucleophile is no longer available (Fig. 5).

51

Figure 5. a) Environment around GMP in domain A showing the disposition of the

HxHxHxx motif of wild-type ecHinT (green) compared with that of the H101A mutant

(violet).

52 b) View of binding pocket in ecHinT showing GMP phosphate interactions. Drawn with

PyMol.

53

C. Design of C-terminus Mutants and Protein Purification

The significance of the C-terminus of Hint1 proteins in mediating their phosphoramidate and acyl-AMP hydrolyase activity was studied previously. 63 Studies with chimeric ecHinT, in which the E. coli C-terminus was replaced with the hHint1 C-

terminus, and chimeric hHint1, in which the hHint1 C-terminus was replaced with the

ecHinT C-terminus, revealed that the substrate specificity of each enzyme could be

transferred to the other by a simple C-terminal switching. 63 To gain further insights into

the critical role of the C-terminal loop in governing the catalysis and substrate specificity

of Hint1 proteins, we designed four alanine mutants of the C-terminus and two sequential

deletion mutants, that shortened the length of the C-terminus by three ( ∆117-119) and six

(∆114-119) amino acids. Plasmids encoding the DNA sequences for the mutants were generated, expressed in an E. coli strain which lacks the ability to endogenously express wild-type ecHinT, and purified using affinity chromatography. The purification of ecHinT wild-type using an AMP-agarose column was reported previously. 8 The same

protocol was followed to purify the mutants with minor modifications.

Compared to wild-type and ∆117-119 mutant, the purified yield of ∆114-119 mutant, using the AMP-agarose column, was significantly reduced, consistent with our earlier finding that the C-terminal loop plays an important role in ligand binding. 63

D. Secondary Structure Analysis by Circular Dichroism (CD) Spectroscopy

To investigate the robustness of the secondary structure of ecHinT to the sequential deletion of 3 and 6 residues from the C-terminal loop, far-UV CD spectra of the wild-type and the deletion mutants were determined. Relative to the wild-type, the

CD spectra of the recombinant C-terminus deletion proteins show that the helical

54 structure of the mutants has likely been increased. However, since The CD spectra of all three proteins shared a minimum at 222 nm at 25ºC, it is likely that shortening the C- terminal loop by 3 or 6 amino acid residues did not lead to major changes in the secondary structure (Fig. 6a). The secondary structures of the four alanine mutants were also studied. Similar to wild-type, all four mutants shared a minimum point at 222 nm at

25ºC (Fig. 6b), suggesting that the overall secondary structure was not affected by any of the alanine mutations in the C-terminus loop. Interestingly, although the H116A mutant shared the same minimum point at 222 nm, analysis of its CD spectrum revealed the presence of more β-sheets in the secondary structure.

E. Thermal Stability

To determine the effect of C-teminus deletion on the thermal stability of ecHinT,

the temperature dependence of the CD spectra was measured. The denaturation curves

showed midpoints at 55 ± 0.4ºC and 49 ± 0.90ºC for ∆117-119 and ∆114-119 ecHinT, respectively. Wild-type ecHinT has a Tm value of 61 ± 0.5ºC under the same experimental conditions. Based on these results, it is apparent that the C-terminal loop has a significant impact on ecHinT stability (Fig. 7a). Nevertheless, given that the C- terminus of hHint1 is three amino acids shorter than the ecHinT C-terminus, it is likely that protein stability is also influenced by other structural motifs. Thermal stability studies of the four alanine mutants did not reveal any significant changes in Tm values of the alanine mutants relative to wild-type ecHinT (Fig. 7b).

55

Figure 6. Secondary structure analysis of a) ∆114-119 and ∆117-119 C-terminus deletion

mutants relative to wild-type ecHinT.

56

b) L114A, H116A, K117A, and L119A C-terminus alanine mutants relative to wild-type ecHinT.

57

Figure 7. Thermal stability studies of a) ∆114-119 and ∆117-119 C-terminus deletion mutants relative to wild-type ecHinT.

58 b) L114A, H116A, K117A, and L119A C-terminus alanine mutants relative to wild-type ecHinT.

59

F. Size-exclusion Chromatography

To determine the dimerization state of the proteins under native conditions, samples of wild-type ecHinT, ∆114-119, ∆117-119 and the four alanine mutants were analyzed on a Superdex G 200 column with in-line UV detector. Wild-type ecHinT eluted with a retention time of 36.5 min corresponding to the theoretical mass (26 kDa) for the dimer (Fig. 8). In good agreement with wild-type ecHinT, only homodimeric protein was observed for all mutants with retention times of 36 min ( ∆114-119), 36.5 min

(∆117-119) (Fig. 8a), and 36 ± 0.5 min for all the four alanine mutants (Fig. 8b).

G. Effect on Catalysis

Previously, it has been demonstrated that Hint1 catalysis proceeds by the

formation of an active-site His-NMP intermediate, followed by intermediate hydrolysis.

32 Evidence in support of this hypothesis for ecHinT has arisen from several observations. First, the ecHinT active-site residue, His101, has been shown to be essential for catalytic activity. 8 Second, His101-dependent intermediate formation was directly

observed during the ecHinT and hHint1 hydrolysis of LysRS-generated lysyl-AMP. 32

Recently, we have shown that the C-terminus of Hint1 is a significant contributing factor in determining the phosphoramidate substrate specificity. 63 Given the proximity of the

C-terminus to the active-site of ecHinT and our observation that it can adopt multiple conformations ( vide supra ), we chose to investigate the role of the C-terminus in catalysis

by comparing the catalytic efficiencies of the deletion mutants, ∆117-119 and ∆114-119.

In addition, from previous studies of engineered monomeric hHint1 45 and C-terminal chimeras of hHint1 and ecHinT 63 , we suspected that the C-terminus may play a role in the hydrolysis of lysyl-AMP generated by E. coli LysRS.

60

Figure 8. Size-exclusion chromatograms from the Superdex G 200 column equipped

with UV (absorbance of 280 nm): a) ∆114-119 (pink) and ∆117-119 (blue) C-terminus deletion mutants compared to wild-type ecHinT (green).

61 b) L114A, H116A, K117A, and L119A C-terminus alanine mutants compared to wild- type ecHinT.

30 30 Det 168-280nm Det 168-280nm Det 168-280nm Det 168-280nm Det 168-280nm Det 168-280nm WildType-1 114-1 116-1 117-1 119-1

20 20

10 10 m A U m A U

0 0

-10 -10 5.0 7.5 10.0 12.5 15.0 17.5 20.0 22.5 25.0 27.5 30.0 32.5 35.0 37.5 40.0 Minutes

62 c) An overlay of all the C-terminus deletion and the C-terminus alanine mutants with wild-type ecHinT.

30 30 Det 168-280nm Det 168-280nm Det 168-280nm Det 168-280nm Det 168-280nm Det 168-280nm Det 168-280nm 117 114 114-1 116-1 117-1 119-1 WildType-1

25 25

20 20

15 15

10 10 m A U m A U

5 5

0 0

-5 -5

-10 -10 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 Minutes

63

To characterize the effect of C-terminus deletion on ecHinT catalysis, the steady- state and pre-steady-state kinetic parameters were determined with a model substrate phosphoramidate substrate, TpAd 11 , and a model acyl-AMP substrate 45 , AIPA. When

first assessed by steady-state kinetics, ∆117-119 and ∆114-119 mutants exhibited an approximately 9-fold reduction in catalytic efficiency (i.e. k cat /K m values) compared to

wild-type ecHinT for the hydrolysis of both substrates. The basis for this perturbation

differed for each mutant. For ∆114-119, the reduction in k cat /K m was largely due to the 5-

fold reduction in k cat , with only a modest 2-fold increase in K m. In contrast, a larger 5-

fold increase in the K m and only a modest 2-fold decrease in k cat were observed for the

∆117-119 mutant (Table 3). Catalytic efficiencies in the steady-state step, with the two model substrates, for the four alanine mutants were also evaluated (Table 4). All alanine mutants exhibited 3-5 fold reduced catalytic efficiencies relative to the wild-type. Except for the H116A mutant, the reduction in k cat /K m was mainly due to an increase in K m values.

To verify that the observed reduction in the steady-state parameters for the

deletion mutants was not because of a shift in the pH-versus-rate profile relative to wild-

type ecHinT, k cat and K m values were determined for the hydrolysis of TpAd at variable

pH points. Wild-type ecHinT and the two deletion mutants exhibited bell-shaped pH

profiles from pH 5.8 to 9.0, with pKa values of 6.9 and 8.5 and pH optima of 7.8 (Fig. 9).

64

Table 3. Steady-state and pre-steady-state kinetic parameters for ∆114-119 and ∆117-

119 C-terminus deletion mutants compared to wild-type ecHinT, using AIPA or TpAd as substrate, in HEPES buffer (pH 7.2, 20 mM HEPES, 1 mM MgCl 2) at 25°C.

NH2 NH 2 N N N N O N N O H O N N N P O O O- HN O P O O HN - O OH OH OH OH

AIPA TpAd

Rate constant WT 114-119 117-119 WT 114-119 117-119

-1 k2 (s ) 135±5 35±2 39±3 116±6 22±1 33±2 adenyl Km (µM) 7.5±2.0 18±2 16±1 11.0±2 23 ±3 21±4 adenyl 6 -1 -1 k2/K m (×10 s M ) 18.0±1 1.9±0.5 2.4±0.4 10.5±0.2 0.96±0.10 1.6±0.3 -1 kcat (s ) 4.0±0.1 0.79±0.10 1.9±0.2 4.5±0.1 0.88±0.10 2.3±0.8 Km (µM) 4.2±0.4 13±1 15±2 5.2±0.2 9.2±0.8 23.8±2.4 3 -1 -1 kcat /K m (×10 s M ) 952±200 59±4 123±15 870±50 96±18 97±24

65

Table 4. Steady-state kinetic parameters for L114A, H116A, K117A, and L119A C- terminus alanine mutants compared to wild-type ecHinT, using AIPA or TpAd as substrate, in HEPES buffer (pH 7.2, 20 mM HEPES, 1 mM MgCl 2) at 25°C.

TpAd AIPA

-1 -1 Protein Kcat (s ) Km(µM) Kcat /K m Kcat (s ) Km(µM) Kcat /K m (×10 3 s-1M-1) (×10 3 s-1M-1) WT ecHinT 4.5±0.1 5.2±0.2 870 ± 50 4±0.1 4.2±0.4 952 ± 200

∆114-119 0.88±0.1 9.2±0.8 96 ± 18 0.79±0.1 13±1 59 ± 4

∆117-119 2.3±0.8 23.8±2.4 97 ± 24 1.9±0.2 15±2 123 ±15

L114A 2.4±0.6 13.9±2.5 173±11 2.8±0.5 16.6±1.6 169±7

H116A 2.3±0.5 3.2±0.84 719±22 3.0±0.2 4.6±.3 652±20

K117A 3.5±0.3 12.5±1.2 280±13 2.9±0.7 14.7±1.6 197±12

L119A 1.9±0.2 8.3±0.4 229±15 2.2±0.1 9.1±0.8 242±14

66

Figure 9. pH-dependence rate profile for wild-type ecHinT, ∆114-119 and ∆117-119 C- terminus deletion mutants.

67

It had been previously proposed that the kinetic mechanism for both phosphoramidate and acyl-AMP hydrolysis by Hints is likely to proceed through active- site adenylation, followed by the enzyme intermediate hydrolysis (Scheme 1). 32 If the rate of adenylation is faster than intermediate hydrolysis, we hypothesized that one should be able to observe a burst of tryptamine or indolylic acid after treatment with

TpAd or AIPA, respectively. This fast first step (k2) would then be followed by a slower,

possibly rate-limiting, intermediate hydrolysis step (k cat ). Consistent with our hypothesis, transient-state kinetic measurements revealed a burst, followed by a linear rate, corresponding to the observed steady-state turnover (Fig. 10). At pH 7.2, the burst rate

-1 -1 (k 2) for wild-type ecHinT was found to be 116 + 6 s and 135 + 5 s for TpAd and

AIPA, respectively (Table 3). The burst rate values (k 2) for the deletion mutants, ∆117-

119 and ∆114-119, were only modestly reduced by approximately 3.5 to 6-fold, while the

burst Km increased by only 2 to 3-fold. Thus, reduction of the C-terminal length by either 3 or 6 residues resulted in a moderate reduction in the efficiency of the burst step.

Presumably, the burst phase corresponds to either the rate of enzyme adenylation and/or a conformational step. Ongoing studies should clarify the specific nature of the catalytic and kinetic mechanisms of ecHinT and the role of the C-terminus. Nevertheless, neither deletion mutation appears to significantly alter the overall kinetic pathway.

68

Scheme 1. Proposed kinetic mechanism for both phosphoramidate and acyl-AMP hydrolysis by Hints.

k1 k2 kcat E + S E•S AMP-E + P E + AMP k-1 S = TpAd or AIPA P = Trptamine or Indole-3-acetic Acid

69

Figure 10. Stopped flow trace of ecHinT exhibited a biphasic profile, a burst phase followed by a linear phase.

70

H. Lysyl-AMP-dependent Adenylation of ecHinT by LysU

Previously we had demonstrated that lysyl-AMP generated by ec LysU is a native substrate for ecHinT. 32 In situ adenylation of ecHinT can be observed after treatment

with ec LysU and α-32 P-labeled ATP. As a proposed mechanism for the adenylation

reaction, nucleophilic attack at the phosphorous atom of either TpAd or AIPA by active-

site residue, His-101, is believed to be responsible for ecHinT-AMP intermediate

formation. 32 Evidence for this mechanism was provided by the inability of the

catalytically impaired active-site mutant H101A to hydrolyze lysyl-AMP and form the

labeled ecHinT intermediate. 32 To assess the role of the C-terminus on the hydrolysis of

lysyl-AMP generated by LysRS, we examined the ability of the deletion mutants to form

the enzyme-AMP intermediate. As can be seen in Fig. 11, the intensity of wild-type

ecHinT labeling by ec LysU (lane 2) was found to be at least 20-fold higher when

compared to ∆114-119 deletion mutant labeling (lane 4) while labeling of ∆117-119 was not observed. Since the extent of labeling could be observed for the ∆114-119 deletion mutant relative to wild-type ecHinT, the rate of the labeling reaction was determined and compared to the rate observed for wild-type ecHinT (Fig. 12). Although labeling could be observed within 10 s for wild-type ecHinT, the rate of intermediate formation for the

∆114-119 mutant was found to be at least 50-fold slower while again no labeling of the

∆117-119 mutant was observed. Taken together, these results provide insights into the crucial role of the C-terminus on the hydrolysis of lysyl-AMP generated by LysRS.

71

Figure 11. Adenylation of wild-type ecHinT and ∆114-119 mutant by ec LysU. Lane1, 3:

controls, [ α-32 P] ATP (0.2 µM) was incubated in buffer A (10 µl, 25 mM Tris HCl, pH

7.8, 100 mM NaCl, 2 mM MgCl 2, 1 mM dithiothreitol, 113 µM Lysine, 0.02 unit/µl,

Inorganic Pyrophosphate, protease inhibitor tablet) at 25°C for 1 min followed by the addition of ecHinT proteins and incubated for 10 min. Lane 2,4,5: E. coli LysU (2 µM)

was pre-incubated with [ α-32 P] ATP at 25°C for 1 min before the addition of ecHinT proteins.

1.2 1 1

0.8

0.6

0.4

0.2

Normalized P-32 Intensity 0.049 0 ecHint ∆114-119

72

Figure 12. Time-dependence of ecHinT and ∆114-119 mutant adenylation by ec LysU.

[α-32 P] ATP (0.2 µM) was incubated in buffer A in presence of ec LysU(2 µM) for 1 min followed by the addition of either ecHinT (5 µM) or ∆114-119 (16 µM). Proteins were incubated for: 10, 20, 30, and 40 seconds before the reaction was terminated.

73

IV. Summary and Concluding Remarks

In summary, we report the first comprehensive high-resolution crystal structure of the full length N-terminal and C-terminal residues of wild-type and the H101A mutant of the histidine triad nucleotide binding protein from Escherichia coli (ecHinT) complexes with GMP. These data reveal the presence of four unique homodimers in the asymmetric unit of the monoclinic crystal lattice and two unique homodimers in the asymmetric unit of the H101A mutant protein. The presence of multiple homodimers in the asymmetric unit can be explained by the conformational flexibility observed in the terminal loop regions of these structures, although the role of His-101 on the conformations remains unclear. In an effort to probe the impact of the C-terminus loop on the structural and kinetic characteristics of ecHinT, two C-terminus deletion mutants of ecHinT were constructed. Our results show that sequential deletions of 3 and 6 amino acids from the C-terminus did not result in major perturbations of the secondary structure

o of the protein. However, a significant shift by approximately 10 C in the T m value was observed for the C-terminus deletion mutants, suggesting that the C-terminus is important for maintaining ecHinT stability. Kinetic studies of wild-type ecHinT and the C-terminus deletion mutants revealed that the catalytic efficiency of ecHinT with diffusible model substrates is only modestly influenced by the C-terminal loop. In contrast, the ability of the C-terminus deletion mutants to hydrolyze the native substrate lysyl-AMP generated by LysRS was greatly impaired by the loss of as few as three amino acids. One would expect that if ecHinT could only hydrolyze lysyl-AMP after release from LysRS, little difference in the catalytic efficiency in hydrolysis of the model substrates and the enzyme generated substrates would be observed. This should be especially true for the ∆117-119

74 mutant, which was the most catalytically efficient of the two mutants ( vide supra ).

Nevertheless, unlike wild-type ecHinT and the ∆114-119 mutant, enzyme intermediate formation was not observed. Since it is well-recognized that dissociation of amino acid adenylates from amino acyl-tRNA synthetases is highly disfavored 77 , the ability of ecHinT to hydrolyze LysRS generated lysyl-AMP suggests a mechanism that may rely on a direct transfer process.

Recent results of E. coli protein-protein interaction studies demonstrated that the elongation factor (EF-Tu) may bind not only to lysyl, alanyl, and isoleucyl-tRNA synthetase but also ecHinT. 60 These results, combined with our recent finding that lysyl-AMP generated by LysRS is a physiological substrate for Hints 32 , suggest that

Hints might function as regulators of protein translation processes. Whether this

regulation occurs as a pre-transfer editing or scavenging of inadvertent aminoacyl-

adenylates, it appears to be mediated by the proposed mechanism of direct transfer of

lysyl-AMP. It is also possible that the interaction of Hints with aminoacyl tRNA

synthetases (AARS) facilitate an as yet undetermined non-canonical function of AARS.

Ongoing studies attempting to unravel the physiological function of Hints in E. coli

should shed light on either hypothesis .

75

Chapter Three

HinT, a Histidine Triad Nucleotide Binding Protein, is Essential for

Alanine Metabolism in Escherichia coli

76

I. Introduction

Histidine triad nucleotide binding proteins (Hints) are conserved from bacteria to humans. 6 While prokaryotes typically encode only one Hint gene, eukaryotes generally contain two to three different Hint genes. 2 The physiological role of Hints in mammalian cells is just now being delineated. Two hybrid screening experiments revealed that Hint1 interacts with Cdk7, the catalytic subunit of the cyclin dependent kinase activation complex Cdk7-cyclin H-MAT1. 46 However, Hint1 mouse knock-out

studies indicated that Hint1 is not required for Cdk7 function. 47 The human Hint1

(hHint1) has also been shown to directly interact with human Pontin and Reptin in the

TCF-β-catenin transcription complex 48 , and it has been reported that hHint1 might be a

tumor suppressor. 49; 50 In addition, recent studies have shown that hHint1 is involved

in the regulation of postsynaptic dopamine transmission. 55 Recently, we reported that

Hints hydrolyze lysine-AMP generated by LysRS in eukaryotes, thus suggesting that

Hints have a specific role in regulating LysRS. 32

E. coli Hint (encoded by hinT ) has high sequence similarity to human Hint1

(ecHinT is 48% identical to hHint1 at the amino acid level) and has been shown to form stable potential protein-protein interactions with six proteins: a putative oxidoreductase and formate dehydrogenase (b1501), the heat shock protein 70 (Hsp70), the β-subunit of

DNA polymerase III ( dnaN ), a membrane-bound lytic murein transglycosylase D ( dniR ),

ET-Tu elongation factor ( tufA ), and a putative synthetase ( yjhH ). 60 In addition, the

HinT from Mycoplasma has been shown to interact with two membrane proteins (P60 and P80). 61; 62 However, despite these determined interactions, and the diversity of

functions that Hints have been shown to be involved in, from regulators of tumor

promotion to amphetamine sensitivity, the physiological and biochemical importance of 77 the relationship of Hints and these proteins remain unresolved. Moreover, the foundational reason for their wide-spread conservation across all three kingdoms of life remains enigmatic.

Crystal structure studies of hHint1 and, recently, ecHinT have revealed that both proteins are homodimers containing an active-site with four conserved histidines. 5; 64

While similar, close inspection of the two structures revealed little sequence similarity between their C-termini. In contrast to hHint1, the longer C-terminus of ecHinT was found to adopt eight different conformations in the unit cell. 64 Chimeric domain swap mutants, in which the C-termini of both ecHinT and hHint1 have been switched, have demonstrated the importance of the C-termini on model substrate specificity. 63

Moreover, deletion mutagenesis studies have shown that the loss of just three C-terminal side-chains abolishes the ability of ecHinT to hydrolyze LysRS-generated lysine-AMP while having only a modest effect on the catalytic efficiency of the enzyme with model substrates. 64 Although catalytic insights of Hint activity have been garnered from

these studies, a defined biochemical rationale for the function of Hints in general, and

ecHinT in particular, has remained elusive.

Phenotype characterization is an essential approach for understanding structure-

function relationships among a variety of biological systems. While several advanced

and comprehensive technologies have been developed to sequence and identify functions

of genes and their products, as well as assign them to particular metabolic pathways, the

function of many genes in most organisms that have been sequenced to date remains

unknown. For example, although E. coli is considered to be amongst the most genetically-characterized microorganisms, about 30-40% of its genes have unknown

78 function. 78 In an effort to determine the function of many of these “unknown” genes, a

library of single gene knock-out mutants of all nonessential E. coli K-12 genes has been

generated. 79 The metabolic profiles of many of these knock-out mutants have been

characterized using Phenotype MicroArrays (PM) that allow testing of a large number of

cellular phenotypes in 96-well microplates. 80 Based on the same redox chemistry,

Biolog™ phenotypic screening plates have been developed as a simplified universal reporter of metabolism in a single bacterium.

Given the high sequence similarity between hHint1 and ecHinT, we hypothesized that determining the function of Hint in E. coli may reveal a conserved biochemical and

physiological role for Hints in general. In E. coli K-12, in silico analysis indicates that

hinT is located in an apparent operon consisting of the hinT(ycfF)-ycfL-ycfM-ycfN(thiK)- ycfO(nagZ)-ycfP genes. In the study reported here, we investigated the role of hinT in E. coli by using single gene deletion mutants and metabolic analyses. In addition, we investigated the role of ecHinT catalysis and structure on the observed ecHinT phenotype with a combination of site-directed mutagenesis and chemical biological studies. Our results show that ecHinT catalytic activity and the C-terminal domain are required for E. coli to grow on D-alanine as a sole carbon and energy source by the regulation of D- amino acid dehydrogenase activity. Taken together, our results demonstrate that ecHinT plays an essential role in the regulation of the fate of alanine in cellular compartments and thus link for the first time the catalytic activity of a Hint protein with a physiological function.

79

II. Experimental Procedures

A. Bacterial Strains, Media, and Growth Conditions

The bacterial strains used in this study were obtained from the E. coli Genetic

Stock Center at Yale University and are listed in Table 1. All strains were received as

glycerol stocks on filters, sub-cultured twice on Luria–Bertani (LB) agar medium 81 , and incubated at 37°C for 48 h before testing. Liquid cultures were grown aerobically in LB medium supplemented with kanamycin (50 µg/ml) at 37°C, with shaking, for 18 h.

Strains were stored as 10% glycerol stocks at - 80°C.

B. PCR Verification of Mutants

The mutations in gene knock-out mutants were verified using the polymerase chain reaction (PCR) technique. Primers sequences for each PCR reaction are listed in

Table 2. DNA from freshly isolated colonies served as templates for PCR and reactions were initiated following a ‘‘hot start’’ protocol at 95°C. The first PCR reaction was used to confirm the expected size of hinT operon with a single gene deletion and insertion of a kan resistance gene. The second PCR reaction confirmed the junction points of the kan resistance gene in each knock-out mutant. PCR conditions were as follows: 94°C for 2 min followed by 30 cycles of 94°C for 30 s, 55°C for 30 s, and 72°C for 2 min.

C. Verification of hin T Deletion Mutant by Loss of Activity

The turnover rates of the fluorogenic substrate tryptamine 5’-adenosine phosphoramidate (TpAd) by ecHinT protein in cell-free lysates of wild-type E. coli

BW25113 and the hinT deletion mutant ( ∆hinT ) was determined using a previously

80

Table1. Bacterial strains used in this study

Gene deleted Strain Gene Size (Kb) Gene Product

hinT JW1089 0.360 Purine nucleoside Phosphoramidase ycfL JW1090 0.378 Predicted protein

ycfM JW5157 0.642 Predicted outer membrane lipoprotein ycfN JW1092 0. 825 Thiamine kinase

nagZ JW1093 1.026 Beta-N-glucosaminidase

ycfP JW5158 0.543 Conserved protein

81

Table 2. Primers used for sequence-verification PCR reaction

Mutant Forward primer Reverse primer

hinT CGTCTACGGCACACCGCGTAA GAATATAGTTTCTTCTGCCAC

ycfL TGGCGCATAAAGGTCTGTAA GGCGTAGCGACTCATTTTTGTC

ycfM GGGGGCGCACAAAGTCAGACAA GGGATTATTGCTGCGAATCGGC

ycfN CTGGTCAGGTAAAGGTGCCGTT GACATCCAACATTACTGGACCC

ycfO CCTGGCGGCAGCTATTAATAAA CCGGACTGTTAGAGTCAAAACC

ycfP AAGCGTTCAAAACCCTCGGGTAA CGCCGCCGACAATCACAATC

82 described phosphoramidase assay. 11 The excitation wavelength was 280 nm and

fluorescence emission was measured at 360 nm. All kinetic assays were performed in

duplicate at 25ºC. The substrate hydrolysis rate was determined by measuring

fluorescence increase following addition of 5-10 µl of lysates (8 ng total protein).

D. Phenotype Analysis Using Biolog™ GN2-MicroPlates

Phenotypic analyses of hinT operon mutants containing single gene deletions

were determined using Biolog™ GN2-MicroPlates (Biolog Inc., Hayward, CA). Mutants

were inoculated onto R2A agar plates containing, per liter, 0.5 g casamine acids, 0.5 g of

D-glucose, 0.3 g sodium pyruvate, 0.3 g K 2HPO 4, 0.05 g MgSO 4·7H 2O, 0.5 g proteose

peptone, 0.5 g soluble starch, 0.5 g yeast extract, and 15 g agar and incubated overnight

at 37°C. Cells were swabbed from the surface of the agar plate, suspended in GN/GP-IF

inoculating fluid (Biolog Inc., Hayward, CA) to a final OD 600nm of 0.7, and inoculated

into the GN2-MicroPlate (150 µl per well). The microplates were incubated at 30°C for

24 h, and absorbance values at 540 nm and 630 nm were determined using an EL808

Ultra Microplate reader (Bio-Tek Instruments Inc., Winooski, VT).

E. Carbon Source Utilization Assays

Overnight cultures of wild-type and mutant E. coli strains were grown at 37°C,

with shaking at 200 rpm, in LB medium supplemented with kanamycin (50 µg/ml) as

needed. Cell pellets were obtained by centrifugation at 6000 g at 4°C for 10 min. Pellets

were washed twice and re-suspended in minimal medium containing (per liter): 6.8 g

Na 2HPO 4, 3.0 g KH 2PO 4, 0.5 g NaCl, 1.0 g NH 4Cl, 0.001 g CaC1 2, 0.001 g MgSO 4, and

0.004 g thiamine (pH 7.0). Carbon sources, D,L-alanine, glycerol, or glucose, were

added, after autoclaving the medium, to a final concentration of 20 mM when needed.

83

F. Induction of D-amino Acid Dehydrogenase Activity and Enzyme Assays

Induction of D-amino acid dehydrogenase in wild-type and mutant strains was achieved by adding 100 mM D,L-alanine to minimal medium containing 10% (vol/vol) glycerol as the sole carbon source. Cultures were grown at 37°C for 24 h prior to collection of cell pellets. D-amino acid dehydrogenase activity was measured by following the reduction of 2,6-dichlorophenol-indophenol (DCPIP) at 600 nm as previously described. 82 Reaction mixtures contained 50 mM potassium phosphate

buffer (pH 7.0), 10 mM KCN, 0.5 mM phenazine methosulfate, 10 mM DCPIP, and 100

mM of D,L-alanine. The reduction of absorbance at 600 nm was measured for 5 min

using Varian Cary 50 UV-visible spectrophotometer (Palo Alto. CA, USA). The amount

of DCPIP reduced was determined using a molar extinction coefficient of 21,500. All

assays were produced in triplicate, and rates were determined from the first 2 min of

activity. Activity is expressed as nmoles DCPIP reduced/min/mg protein, and rates

between replicates varied by < 5%.

G. Preparation of Membrane Fractions

Wild-type and mutant strains were grown to late exponential phase (OD 600nm =

0.9-1.1) in minimal medium containing 100 mM D,L-alanine and 10% (vol/vol) glycerol,

harvested by centrifugation at 10,000g, and washed once in 0.85% NaCl. The pellets

were stored at -80°C. Cell pellets were thawed, suspended in MOPS Buffer (10 mM

MOPS (pH 7.5), 20 mM MgCl 2, and 10% glycerol) and passed twice through a French

Pressure Cell at 11,000 psi. Cell lysates were centrifuged for 10 min at 14,500g and the

resulting supernatant was centrifuged for 80 min at 37,000g in a Beckman

84

Ultracentrifuge. The membrane fractions were obtained as previously described. 83

Aliquots of membrane fractions were stored at -80°C and used only once after thawing.

H. Reverse Transcription-PCR

Total RNA was isolated from E. coli strains using Qiagen RNeasy minikit

(Qiagen, Valencia, CA) according to the manufacturer’s protocol. RNA concentration was determined spectrophotometrically at 260 nm and RNA quality was confirmed by electrophoresis of 1 µg of total RNA from each preparation on a 1.2 % denatured agarose

gel. The following primers were used for RT-PCR analysis: LP

5’GTTTCGATAACCGCATTCGT3’ and RP 5’CCGGTATTCAGCCACAGATT3’. The

RT-PCR conditions were adapted from Qiagen OneStep RT-PCR Kit handbook and

were as follows: 50°C for 30 min, 95°C for 15 min followed by 30 cycles of 94°C for 0.5

min, 50°C for 0.5 min, 72°C for 0.5 min followed by a single 10-min cycle at 72°C for

extension.

I. Site-directed Mutagenesis

All hinT mutants were generated from E. coli hinT -pSGA02, the expression

vector harboring the gene encoding wild-type hinT , by using the Quick-Change

mutagenesis kit (Stratagene, Santa Clara, CA) following the manufacturer’s protocol.

Mutagenesis of ecHinT and hHint1 to produce ec /Hs and Hs/ ec chimeric proteins was

described previously. 63

85

J. Phenotype Complementation Studies

The preparation of competent cells was achieved using a Z-Competent E. coli

Transformation Kit and Buffer Set (Zymo Research Corporation, Orange, CA). The

∆hinT mutant was transformed with 20 ng of plasmid DNA encoding either the wild-type

hinT , H101A, H101G , ∆114-119, ∆117-119, L114A, H116A, K117A, L119A, Hs/ ec chimera or ec /Hs chimera. Cultures were grown in LB medium, induced with IPTG (500

µM) when the culture reached OD 600nm = 0.4-0.7, and incubated for an additional 10 h.

Cell pellets were obtained by centrifugation at 6000g for 10 min at 4ºC, washed twice in

M9 minimal medium 81 , and re-suspended in 200 µl of the same medium. A 10-20 µl

aliquot of the cell suspension was inoculated into M9 medium supplemented with 20 mM

D,L-alanine or glucose, with ampicillin (100 µg/ml) and kanamycin (50 µg/ml), to an

initial OD 600nm = 0.06 - 0.08. Cells were incubated at 37ºC and OD 600nm was measured after for 36 - 40 h of growth.

K. General Synthetic Procedures and Materials (Compounds’ synthesis and characterization were carried out by Dr. Brahma Ghosh)

All reagents were purchased from commercial vendors and used without further purification. Anhydrous pyridine was used from previously unopened SureSeal or

AcrosSeal bottles. Analytical thin layer chromatorgraphy (TLC) was performed on 0.25 mm precoated Merck silica gel (SiO 2) 60 F254. Column chromatography was performed on Purasil 60A silica gel, 230–400 mesh (Whatman). 1H and 31 P NMR were recorded on a Varian Mercury-300 or a Varian MR 400 spectrometer. Chemical shifts are reported in ppm relative to residual dimethyl sulfoxide (DMSO-d6) peak. High-resolution mass spectrometry (HRMS) data were obtained on a Biotof II (Bruker) ESI-MS spectrometer.

86

L. Synthesis of 2’,3’-Isopropylidine-5’-O-(4-Chlorophenoxy)Carbonyl Guanosine

(Guanosine -5’-Carbonate Acetonide, 1)

Guanosine-2’,3’- acetonide (0.231 mmol) was dissolved in anhydrous pyridine (7 ml) in a 2-neck round-bottomed flask equipped with an Ar. inlet and a microsyringe. The stirred solution was chilled over a dry ice-acetone bath and 4-chlorophenyl chloroformate

(0.278 mmol, 1.20 eq) was added dropwise under Ar. The cooling bath was removed and the solution was slowly allowed to return to room temperature. Stirring was continued until thin layer chromatographic and ESI-MS analyses showed that all of the starting material had disappeared (after approximately 4 h). Pyridine was evaporated in vacuo by

co-distillation with n-heptane and the crude solid obtained was purified by

chromatography on SiO 2 gel to isolate the guanosine-5’-carbonate as a while solid with a

1 90% yield. The H NMR spectrum was (DMSO-d6): 1.30 (s, 3H, -CH 3), 1.56 (s, 3H, -

CH 3), 4.28- 4.58 (m, 3H, C2’+C3’+ C4’-H), 5.22 (s, 2H, 5’-CH 2), 6.22 (s, 1H, C1’-H),

6.58 (s, 2H, NH 2), 7.22 (d, 2H, -Ph), 7.50 (d, 2H, -Ph), 7.86 (s, 1H, C8-H), 10.74 (s, 1H,

N1-H) ppm.

M. Synthesis of 2’,3’-Isopropylidine-5’-O-[(3-Indolyl)-1-Ethyl]Carbamoyl

Guanosine (Guanosine -5’-Carbamate Acetonide, 2)

A mixture of guanosine-5’-carbonate (0.255 mmol) and tryptamine (0.640 mmol,

2.5 eq) in anhydrous THF (10 ml) was heated to 60ºC (oil bath) under Ar. and stirred at that temperature until TLC showed complete consumption of the starting material (4 h).

The solvent was removed by rotary evaporation and the residue chromatographed on

SiO 2 gel using CH 2Cl 2:MeOH:H 2O (5:2:0.25) to give the title carbamate as white

87 amorphous solid (yield 90%). The 1H NMR spectrum was (DMSO-d6): 1.32 (s, 3H, -

CH 3), 1.58 (s, 3H, -CH 3), 2.80 (t, 2H, -CH 2-), 3.24 (q, 2H, -CH2 NHCO), 4.00- 4.30 (m,

3H, 5’-CH 2+ C4’-H), 5.10 (d, 1H, 5’-C3’-H), 5.24 (d, 1H, C2’-H), 6.00 (s, 1H, C1’-H),

6.58 (s, 2H, NH 2), 6.96 (t, 1H, indole), 7.12 (t, 1H, -indole), 7.18 (s, 1H, indole), 7.36 (d,

1H, indole), 7.40 (t, 1H, -NH), 7.50 (d, 1H, indole), 7.84 (s, 1H, C8-H), 10.72 (s, 1H, N1-

H), 10.82 (s, 1H, NH-indole) ppm.

N. Synthesis of 5’-O-[(3-Indolyl)-1-Ethyl]Carbamoyl Guanosine (Guanosine-5’-

Tryptamine Carbamate, TpGc)

The above 2’, 3’acetonide-protected guanosine carbamate (100 mg) was dissolved in TFA:H 2O (4:1, 5 ml) and stirred at room temperature until TLC and ESI-MS showed no starting material left (60 min). The mixture was stripped to dryness and the residue purified by SiO 2 gel chromatography using CH 2Cl 2:MeOH:H 2O (5:3:0.5) as eluant. The

relevant UV-active fractions were collected and concentrated to afford the final

compound as a white solid (yield 40%). The 1H NMR spectrum was (DMSO-d6): 2.38 (t,

2H, -CH 2-), 3.17 (q, 2H, -CH2NHCO), 4.23- 4.33 (m, 2H, 5’-CH 2), 4.56 (t, 1H, C4’-H)

5.14 (d, 1H, C3’-H), 5.39 (d, 1H, C2’-H), 6.10 (s, 1H, C1’-H), 6.38 (s, 2H, NH 2), 6.96 (t,

1H, indole), 7.15 (t, 1H, -indole), 7.21 (s, 1H, indole), 7.32 (d, 1H, indole), 7.40 (t, 1H, -

NH), 7.48 (d, 1H, indole), 7.80 (s, 1H, C8-H), 10.65 (s, 1H, N1-H), 10.78 (s, 1H, NH- indole) ppm [OHs exchanging with H 2O in NMR solvent] .

88

III. Results and Discussion

A. Operon Structure of Genes Associated with hinT

In E. coli K-12 strains, the hinT gene is located adjacent to ycfL , ycfM , ycfN (thiK ), ycfO (nagZ ), and ycfP . PCR analyses were performed to determine if the deletion/insertion mutations occurred in frame without affecting the operon structure.

Results in Figure 1 show that the operon structure is still intact after deletion of each gene and insertion of a kan resistant gene. Further evidence of hinT deletion was obtained by measuring the phosphoramidase activity in cell-free lysates. Figure 2 shows that the hinT deletion mutant ( ∆hinT ) did not have appreciable phosphoramidase activity that is associated with the expression of hinT in wild-type E. coli BW25113.

B. Phenotype of ecHinT

Despite the fact that Hint is present in E. coli and other prokaryotes, the function of this conserved group of proteins in bacteria has remained largely unknown. To ascertain the role of ecHinT in microbial metabolism, we obtained a series of mutants from the Keio collection containing deletions in the hinT (ycfF ), ycfL , ycfM , ycfN (thiK ), ycfO (nagZ ), and ycfP genes. Biolog™ plates were used for phenotypic screening to estimate the differential metabolic potential between wild-type E. coli BW25113 and the hinT deletion mutant ( ∆hinT ). Results from Biolog™ substrate utilization studies (Fig. 3) indicated that metabolic profiles of wild-type E. coli BW25113 and the ∆hinT mutant

were highly similar, with the exception of the utilization of D- and L-alanine.

89

Figure 1. Sequence verification PCR (representative gel)

MW 1 2 3 4 marker

Lane1: hinT (annealing Temp 38ºC)

Lane2: hinT (annealing Temp 40ºC)

Lane3: hinT (annealing Temp 43ºC)

Lane4: hinT (annealing Temp 43ºC)

Expected fragment size= 1.15 hinT size (0.36Kb) + Kan gen (0.795)

90

Figure 2. Phosphoramidase assay to detect ecHinT activity in E. coli cell-free lysates.

91

Figure 3. Summary of metabolic fingerprints. Dataset of the GN2 plate assay of wild- type E. coli BW25113 and ∆hinT mutant strains was processed as described in experimental procedures. The overlay represents the difference between the ability of the two strains to use 95 carbon sources. This is a representative overlay of three independent experiments.

92

The initial screening results were verified and confirmed by culturing wild-type and mutant strains in minimal media containing D,L-alanine (Fig. 4). These results were highly reproducible, even when a starvation phase was introduced prior to culturing.

To further define the role of genes in the apparent hinT operon in alanine

metabolism, E. coli strains from the Keio collection with verified deletions in ycfL

(JW1090), ycfM (JW5157), ycfN (JW1092), ycfO (JW1093), and ycfP (JW5158) were

evaluated for carbon source utilization. Compared to the wild-type BW25113 strain,

none of the tested mutants showed significant difference in alanine catabolism when

tested using Biolog™ plates and test tube cultures.

C. D-Amino Acid Dehydrogenase Transcription and Activity Testing

Inspection of the metabolic fingerprints of wild-type BW25113 and the ∆hinT mutant revealed that while the ∆hinT mutant failed to grow on both D and L isomers of

alanine, it did grow on serine and glycine. Based on these results, and what is known

about the alanine catabolic pathway in E. coli , we hypothesized that the ∆hinT mutant

was neither deficient in the amino acid transporter system for D,L-alanine nor in alanine

racemase, but most likely in D-amino acid dehydrogenase activity encoded by dadA (Fig.

5). The inducible dadA encodes for the broad specificity bacterial D-amino-acid

dehydrogenase. DadA catalyzes the oxidation of D-amino acids, including alanine, into

their corresponding ketoacid. 84 The bacterial D-amino-acid dehydrogenase is a

heterodimer complex formed by a 45-kDa and a 55-kDa subunit. 85 To avoid overproduction of reactive oxygen species that could damage cellular components, D- amino acid oxidation in bacteria is not coupled to O 2 reduction and the two electrons from the substrate are received by FAD in the small subunit before being transferred to a

93

Figure 4. Bacterial growth curves of wild-type E. coli BW25113 and ∆hinT mutant in

M9 medium in the presence of either 20 mM glucose or D,L-alanine. Measurements were done in duplicate and variants were less than 5% between two independent cultures.

94

Figure 5. Alanine transport and metabolism in E. coli.

95 sulfur-iron center in the large subunit. 85 Bacterial D-amino-acid dehydrogenase appears to have two main functions in bacteria. Initially, it allows growth of bacteria using D- amino acids as the sole carbon, nitrogen, and energy source. 86; 87 Additionally, it prevents the accumulation of D-amino acids in cellular compartments, as some D-amino acid analogues have specific inhibitory effects on bacterial growth. 88; 89 In bacteria, D-

amino acids are used to stabilize the peptidoglycan structure 90 , and have recently been shown to regulate biofilm disorganization 91 and are thought to play a role in quorum sensing.

The inability of the ∆hinT mutant to grow on D,L-alanine can be explained by either failure to induce transcription of the D-amino-acid dehydrogenase or the presence of a defective D-amino-acid dehydrogenase that is incapable of metabolizing D-alanine.

To test these hypotheses, RT-PCR was performed using primers for dadA and an

equivalent amounts of mRNA obtained from wild-type BW25113 or the ∆hinT mutant

grown in the presence of D,L-alanine. Results in Figure 6 show that equivalent amounts

of dadA transcript were produced by either the wild-type strain or the ∆hinT mutant. To

determine if the inability of the ∆hinT mutant to grow on D-alanine was due to a lack of

DadA enzyme activity, membrane fractions were isolated and analyzed for D-amino acid

dehydrogenase activity. Results showed that the ∆hinT mutant possessed 38-fold less D-

amino acid dehydrogenase activity (< 5 nmol/min/mg) relative to the wild-type

BW25113 strain (190 nmol/min/mg).

96

Figure 6. RT-PCR was performed with equivalent amounts of mRNA (100 ng) obtained from wild-type E. coli BW25113, ∆hinT, and ∆ycfL mutants using dadA primers under the described conditions. 1 µg of each PCR reaction was loaded on 1% agarose gel.

97

In this study, we report the first evidence of a regulatory relationship between the evolutionary conserved enzymes; DadA and ecHinT. Since the transcripts encoding for

DadA were produced in the ∆hinT mutant, our results suggested that D-amino acid dehydrogenase activity in E. coli is likely modulated by ecHinT in a post-translational

manner.

D. Phenotype Rescue by ecHinT

Transformation of a plasmid encoding the wild-type hinT into the ∆hinT mutant

(JW1089) allowed rescue of the mutant for growth on D,L-alanine to a level of growth

comparable to wild-type BW25113 grown under the same conditions (Fig. 7). This

indicated that other mutations in strain JW1089 were not involved in the inability of the

∆hinT mutant to grow on D,L-alanine.

A mutational approach was used to fully examine the relationship between ecHinT structure and activity, and DadA activity. Two active-site mutants were prepared and characterized as described previously. 8 ecHinT His-101 residue, which was

previously suggested to be the key catalytic residue for enzymatic activity 8, was

exchanged by site-directed mutagenesis into either alanine (H101A) or glycine (H101G).

Enzymatic activities of these mutants were previously quantified by 31 P-NMR using

adenosine-5´-monophosphoramidate (AMP-NH 2) as a substrate and the catalytic efficiency was found to be 3,0000-fold lower than that of wild-type ecHinT. 8 Plasmids

encoding the catalytically impaired mutants of ecHinT, His-101A or His-101G, failed to

complement the ∆hinT mutant for growth on D,L-alanine. This indicated that the catalytically active ecHinT is needed for DadA activity (Fig. 7).

98

Figure 7. ecHinT structural and activity requirement for phenotype rescue. Each column represents the growth of the ∆hinT mutant that was transformed with a plasmid containing either hinT , a mutant hinT , ec /Hs chimera, Hs/ ec chimera, or hHint1 , except for the first column of wild-type E. coli BW25113. OD 600 values were determined 48 h

after inoculation into M9 medium in presence of 20 mM D,L-alanine.

99

E. Design, Synthesis and Characterization of ecHinT Inhibitor

Tryptamine guanosine carbamate (TpGc) was prepared according to synthetic

Scheme 1. The rationale for designing such an inhibitor was based on the structural activity relationship of Hint substrates that was reported previously 11 , which found that the phosphoramidate substrate tryptamine 5’-guanosine monophosphate (TpGd) (Fig. 8a)

-1 5 - was an excellent substrate with a kcat value of 4.0 s and a k cat /K m value of 7.0 x 10 M

1s-1. 11 It was hypothesized that substitution of the hydrolysable phosphoramidate moiety with a carbamate linker (TpGc) would result in an inhibitor of ecHinT and not a substrate

(Fig. 8b). In addition, the incorporation of guanosine would insure that potential off- target inhibition of Tryptophan tRNA synthetase would be minimal.

Using the previously described spectrophotometric phosphoramidase assay 11 , the

ecHinT Ki value for TpGc was determined. Although designed to be a competitive

inhibitor, TpGc was found to be a non-competitive inhibitor of ecHinT with a Ki value of

42 µM (Fig. 8c). Consistent with our previous finding, that ecHinT catalytic activity is

required for the phenotype, when compared to the growth of ∆hinT E. coli strain on D,L-

alanine, cultures of wild-type BW25113 grown in the presence of D,L-alanine and 100

µM of TpGc exhibited a similarly diminished growth capacity (Fig. 8d).

100

Scheme 1. General synthetic scheme of TpGc.

O O

N N NH Cl O NH N N NH a N N NH HO 2 O O 2 O O

O O O O

1

b

O O N N O NH O NH NH NH c N N N NH2 N NH2 N O N O H O H O O O OH OH TpGc 2

a) p-Cl phenyl chloroformate, pyridine, RT, 4h, 65% b) tryptamine, anhyd THF, 60°C, 4h, 90% c) TFA:H2O (4:1), RT, 60 min, 40%

101

Figure 8. Inhibition of phenotype rescue by ecHinT inhibitor.

a) The structure of tryptamine 5’-guanosine monophosphate (TpGd).

O

N NH O H N N NH N P O O 2 HN O- OH OH

TpGd

102 b) The structure of tryptamine guanosine carbamate (TpGc).

O

N NH O

N N NH N O O 2 HN H

OH OH

TpGc

103 c) TpGc exhibits a non-competitive inhibition profile.

104 d) Wild-type E. coli BW25113 was grown in M9 medium supplemented with either 20 mM glucose (blue bar) or 20 mM D,L-alanine (brown bar) in the presence or absence of

100 µM TpGc. OD 600 values were determined after 48 h.

105

To probe structural elements of ecHinT that are determinant for the discovered

phenotype, non-active-site deletion mutations that reside in the C-terminus region of

ecHinT were prepared and characterized as described previously. 64 Recently, we

investigated the C-terminus region of ecHinT protein and found that this loop

accommodates several conformational forms in the 1.3°A crystal structure (Fig. 9). 64

We designed, and fully characterized, two deletion mutants of ecHinT C-terminus in which three or six amino acids were deleted from the 11 amino acids C-terminal loop.

The two deletion mutants were equivalent in their secondary structure to wild-type ecHinT and preserved the ability to form homodimers. The catalytic efficiencies of the deletion mutants were within one order of magnitude lower than that for the wild-type ecHinT. Despite the partial activity possessed by these mutants, both mutants failed to rescue ∆hinT when growing on D,L-alanine (Fig. 7).

To further evaluate the impact of the C-terminal loop on the observed phenotype and in an effort to determine if the phenotype rescue is dependent on specific residues in the C-terminus of ecHinT, four alanine mutants of the C-terminal loop, L114A, H116A,

K117A, and L119A were designed and characterized as previously described in Chapter

Two. All of the alanine mutants retained partial activities and were structurally equivalent to wild-type ecHinT when studied by CD spectroscopy. The ability each of these mutants to rescue ∆hinT when growing on D,L-alanine was studied as shown in Figure 10. All the alanine mutants were able to rescue the phenotype, suggesting that the length rather than the sequence of the C-terminus is what determines the ability of ecHinT to function as a regulator of DadA activity.

106

Figure 9. Superposition of the eight independent monomers observed in the structure of

the ecHinT-GMP complex.

107

Figure 10. phenotype rescue by the alanine scan mutants. Each column represents the

growth of the ∆hinT mutant that was transformed with a plasmid containing wild-type ecHint, L114A, H116A, K116A, or L119A mutant except for the first column of wild- type E. coli BW25113. OD 600 values were determined 48 h after inoculation into M9

medium in presence of 20 mM D,L-alanine.

108

Although both the N and C-terminal loops of ecHinT can adopt more than one

conformation in the ecHinT-GMP complex structure 64 , the overall fold of the structure is similar to hHint1. 64 To investigate the possibility of phenotype rescue by the human version of Hint protein (hHint1), which is 48% homologous to ecHinT, the E. coli ∆hinT mutant was transformed with a plasmid encoding hHint1 and the ability to grow on D,L- alanine was examined. The hHint1 failed to rescue the growth phenotype suggesting that certain structural characteristics, in addition to catalytic activity, govern the regulation of

DadA activity by ecHinT. Based on our finding that the C-terminus loop of ecHinT plays a significant role in the observed relationship between ecHinT and DadA activity, the ability of two chimeric proteins to rescue ∆hinT cells was studied. A hHint1-ecHinT

chimera (Hs/ ec ), in which the C-terminus of hHint1 was replaced with ecHinT C-

terminus, was unable to rescue the D-alanine growth defect exhibited by E. coli ∆hinT cells. However, and in contrast, a ecHinT-hHint1 chimera ( ec /Hs), in which the C- terminus of ecHinT was replaced with the hHint1 C-terminus, had only a slightly reduced ability to complement the phenotype when compared to wild-type ecHinT (Fig. 7).

Taken together, these results suggest that other inherent structural requirements, in addition to the C-terminus, are critical to the ecHinT-mediated DadA activation.

Ongoing studies should shed light on the post-translational regulatory relationship between ecHinT and DadA and whether this regulation influences the localization, dimerization and/or degradation of DadA.

In summary, by using a series of mutagenic and chemical biological studies, ecHinT has been shown to be necessary for the growth of E. coli on D-alanine by regulating the catalytic activity of a key metabolizing enzyme, DadA. The role of ecHinT

109 has been shown to be dependent on the catalytic activity of ecHinT, components of its C- terminus domain and potentially unidentified structural factors. This marks the first example of a direct relationship between the catalytic activity of a Hint protein and a physiological function. Interestingly, the characteristics of ecHinT that appear necessary for the discovered phenotype are also required for the hydrolysis of LysU-generated lysyl-AMP by ecHinT. 32; 64 Whether this observation is a coincidence or evidence of a link between the ecHinT regulation of DadA and its interaction with LysRS remains to be determined.

110

Chapter Four

Synthesis and Evaluation of Potential Inhibitors of Human and Escherichia coli Histidine Triad Nucleotide Binding Proteins

111

I. Introduction

Histidine triad nucleotide binding protein (Hint) is named after the highly

conserved motif related to the sequence His-X-His-X-His-XX in the active-site, where X

is a hydrophobic amino acid. Hint is considered to be the ancestor of the histidine triad

protein superfamily and can be found in both prokaryotes and eukaryotes. 6 While

eukaryotes generally express multiple forms of Hint (Hint1, Hint2 and Hint3), prokaryote

genomes, including Gram-negative and Gram-positive bacteria, typically encode one

Hint gene. 2

Recently, evidence has begun to emerge that Hint1 may function as a tumor suppressor through multiple molecular mechanisms involving triggering apoptosis 51 ,

regulating gene transcription and cell cycle modulation. 34; 66 Hint1 deleted mice studies

have revealed that an increased susceptibility to the carcinogenic DMBA induced ovarian

and mammary tumors, as well as an increased occurrence of spontaneous tumors. 49

Eukaryotic Hint1 has been shown to associate with, and suppress the β-catenin Wnt signaling pathway transcriptional activity by direct interactions with Reptin and Pontin. 48

Hint1 has also been associated with transcription factors such as, TFIIH 46 , MITF 34; 92 ,

and USF2. 66 In addition, the interactions between MITF and USF2 appear to be mediated by LysRS. 34; 66; 93 Recently, we have demonstrated that lysyl-adenylate (lysyl-

AMP) generated by lysyl-tRNA synthetases (LysRS) is also a substrate for Hints. 32

Despite such a wide variety of possible biological functions in eukaryotes, the cellular function and rationale for the evolutionary conservation of Hint in bacteria has remained a mystery. Moreover, none of the previously proposed functions has been successfully linked to the enzymatic activity of Hints. Recently, we have shown that the

112 expression of the catalytically active Hint is essential for the activity of the enzyme D- amino acid dehydrogenase (DadA) in Escherichia coli . 94 Consequently, we have chosen to synthesize novel Hint1 inhibitors that can be used as chemical probes to study the connection between the newly discovered Hint1-associated phenotype in E. coli and its enzymatic activity.

Hint1 has been found to be a nucleoside phosphoramidase and acyl-AMP hydrolase. 6 Probing the Hint1 base recognition site, substrate specificity for both purine and pyrimidine phosphoramidates was determined.11 Structure–activity relationship studies revealed that Hint1 prefers purine over pyrimidine phosphoramidates. In addition, based on kinetic and crystallographic studies, the 2- and 3-hydroxyl groups of the ribose ring were shown to be required for maximal phosphoramidase efficiency. 11

II. Experimental Procedures

(Compounds’ synthesis and characterization were carried out by Dr. Brahma Ghosh).

A. General Procedure for synthesis of 2’,3’-Isopropylidine-5’-(indole-3- propionyl)

nucleosides: (Adenosine Acetonide-5’-Indole-3-Propionate (1)): [M+ Na] 501.0283

A mixture of the appropriate 2’,3’-isopropylidine nucleoside (0.231 mmol) ,

indole-3-propionic acid (0.254 mmol, 1.1 eq), DCC (0.254 mmol, 1.1 eq), and DMAP

(0.254 mmol, 1.1 eq) was suspended in CH 2Cl 2 (10 ml) and a few drops of DMSO were

added to assist in solubility. The mixture was stirred at RT and after 24 h an additional

equivalent of DCC and DMAP was added. After 36 h, the mixture was diluted with

CH 2Cl 2 (50 ml) and washed with H 2O (3 X 10 ml). The organic layer was collected and concentrated. The residue obtained was purified by SiO 2 gel chromatography using

113

1 CH 2Cl 2:MeOH:H 2O (5:2:0.25) as eluant. H- DMSO-d6: 1.32 (s, 3H, -CH3), 1.54 (s, 3H,

-CH3), 2.54 (s, 1H), 2.64 (m, 2H, -CH2-), 2.90 (t, 2H, -CH2-), 4.12- 4.27 (ddd, 2H, 5’-

CH2), 4.35 (qn, 1H), 4.99 (m, 1H), 5.43 (dd, 1H), 6.18 (d, 1H, C1’-H), 6.95 (t, 1H, indole), 7.05 (t, 1H, indole), 7.08 (s, 1H, indole), 7.20 (d, 1H, indole), 7.34 (br s, 2H, -

NH2), 7.47 (d, 1H, indole), 8.16 (s, 1H, C2-H), 8.29 (s, 1H, C8-H), 10.77 (s, 1H, -NH indole) ppm. 13 C- DMSO-d6: 173.42, 156.59, 152.79, 148.73, 140.08, 136.90, 127.12,

121.84, 121.16, 118.41, 117.93, 114.21, 113.35, 111.09, 90.83, 84.94, 84.32, 81.48,

63.76, 34.76, 26.22, 24.40, 26.22, 24.39, 20.66 ppm.

B. General Procedure for Synthesis of 5’-(indole-3- propionyl) nucleosides

(Adenosine 5’-Indole-3-Propionate(2)) : [M+ 1] : 439.0332

The 2’, 3’acetonide-protected nucleoside 5’-ester (100 mg) was dissolved in

TFA:H 2O (4:1, 5 ml) and stirred at RT for 2 h. The mixture was evaporated to dryness and the residue purified by SiO 2 gel chromatography using CH 2Cl 2:MeOH:H 2O (5:3:0.5)

as eluant. The relevant UV-active fractions were collected and concentrated to afford the

final compounds as white solids. 1H- DMSO-d6: 2.72 (m, 2H, -CH2-), 3.00 (t, 2H, -CH2-

), 4.18- 4.42 (ddd, 2H, 5’-CH2), 4.30 (m, 2H), 4.65 (t, 1H), 6.01 (d, 1H, C1’-H), 7.00 (t,

1H, indole), 7.15 (t, 1H, indole), 7.20 (s, 1H, indole), 7.37 (d, 1H, indole), 7.57 (d, 1H,

indole), 8.20 (br s, 2H, -NH2), 8.38 (s, 1H, C2-H), 8.58 (s, 1H, C8-H), 10.82 (s, 1H, -NH

indole) ppm. 13 C- DMSO-d6: 172.48, 159.81, 150.01, 148.92, 140.74, 136.20, 126.84,

122.34, 120.97, 119.15, 118.27, 118.78, 11.95, 111.37, 87.89, 81.66, 73.14, 70.19, 63.82,

34.31, 20.16 ppm.

114

C. General Procedure for Synthesis of 5’-O-(4-chlorophenoxy) carbonyl 3’-azido-3’- deoxy thymidine (AZT-5’-carbonate (3) ): (M+1) 422.0207

AZT (0.231 mmol) was dissolved in anhydrous pyridine (7 ml) in a 2-neck round- bottomed flask equipped with an Ar. inlet and a microsyringe. The stirred solution was chilled using a dry ice-acetone bath and 4-chlorophenyl chloroformate (0.278 mmol, 1.20 eq) was added dropwise to it under Ar. The cooling bath was then removed and the solution slowly allowed to return to RT and stirred for an additional 3h. Pyridine was evaporated in vacuo by co-distillation with n-heptane. The crude solid obtained was

1 purified by chromatography on SiO 2 gel to afford the title carbonate as a white solid. H-

DMSO-d6: 1.76 (s, 3H, -CH3), 4.14 (m, 1H, C’-H), 4.4- 4.6 (m, 3H, 5’-CH2 + C4’-H),

6.18 (t, 1H, C1’-H), 7.24(d, 2H, -Ph), 7.50 (d, 2H, -Ph), 7.56 (s, 1H, C6-H), 11.38 (s, 1H,

N3-H) ppm.

D. General Procedure for Synthesis of 5’-O-[(3-indolyl)-1-ethyl]carbamoyl 3’-azido-

3’-deoxy thymidine (AZT-5’-carbamate (4)): (M-1] 452.1490

A mixture of the foregoing AZT-5’-carbonate (0.255 mmol) and tryptamine

(0.640 mmol, 2.5 eq) in anhydrous THF (10 ml) was heated to 60ºC (oil bath) under Ar.

Stirring was continued at that temperature for 2.5 h at which point TLC showed complete

disappearance of starting material. The solvent was removed by rotary evaporation and

the residue chromatographed on SiO 2 gel using CH 2Cl 2:MeOH:H 2O (5:2:0.25) to give the

desired carbamate as a white solid. 1H- DMSO-d6: 1.78 (s, 3H, -CH3), 2.30 (m, 1H), 2.46

(q, 1H), 2.84 (t, 2H, -CH2-), 3.30 (q, 2H, -CH2NH-), 4.12- 4.26 (ddd, 2H, 5’-CH2), 4.44

(m, 1H), 6.13 (t, 1H, C1’-H), 6.97 (t, 1H, indole), 7.05 (t, 1H, -indole), 7.14 (s, 1H,

115 indole), 7.33 (d, 1H, indole), 7.45 (t, 2H, -NH + C6-H), 7.51 (d, 1H, indole), 10.80 (s,

1H, NH-indole) ppm.

E. General Procedure for Synthesis of 5’-O-[(3-indolyl)-1-ethyl]carbamoyl 3’-amino-

3’-deoxy thymidine (AZT-3’-NH2-5’-carbamate(5)) : [M+1]: 428.2349

To a solution of the above AZT-5’-carbonate in MeOH (5 ml) was carefully added 10% Pd/C (10% by weight of the starting material). The reaction vessel was evacuated, filled with H 2, and stirred at RT. After 16h the reaction was deemed was complete by ESI-MS analysis. The mixture was diluted 15 ml of MeOH and the catalyst filtered off. The filtrate was concentrated under vacuum and the residue chromatographed

1 on a SiO 2 gel column using CH 2Cl 2:MeOH:H 2O (5:3:0.5) to elute the product. H-

DMSO-d6: 1.76 (s, 3H, -CH3), 1.98 (m, 1H), 2.14 (m, 1H), 2.81 (t, 2H, -CH2-), 3.26 (q,

2H, -CH2NH-), 3.36 (m, 1H), 3.69 (m, 1H), 4.08- 4.23 (ddd, 2H, 5’-CH2), 4.44 (m, 1H)

6.13 (t, 1H, C1’-H), 6.95 (t, 1H, indole), 7.04 (t, 1H, -indole), 7.12 (s, 1H, indole), 7.30

(d, 1H, indole), 7.36 (t, 1H, -NHCO), 7.42 (s, 1H, C6-H), 7.48 (d, 1H, indole), 10.79 (s,

1H, NH-indole) ppm. 13 C- DMSO-d6: 163.94, 156.28, 150.61, 136.42, 127.39, 122.81,

121.09, 118.41, 111.75, 111.56, 109.81, 83.67, 79.40, 64.40, 52.10, 41.39, 25.71, 12.36

ppm.

F. General Procedure for Synthesis of 2’,3’-Isopropylidine-5’-O-(4-

chlorophenoxy)carbonyl nucleosides (Adenosine-5’-carbonate acetonide (6)): (M+1)

462.1246

Adenosine-2’,3’- acetonide (0.231 mmol) was dissolved in anhydrous pyridine (7

ml) in a 2-neck round-bottomed flask equipped with an Ar. inlet and a microsyringe. The

116 stirred solution was chilled over a dry ice-acetone bath and 4-chlorophenyl chloroformate

(0.278 mmol, 1.20 eq) was added dropwise to it under Ar. The cooling bath was then removed and the solution slowly allowed to return to RT. Stirring was continued until

TLC showed all of the starting material had disappeared (4 h, monitored by TLC and

ESI-MS). Pyridine was evaporated in vacuo by co-distillation with n-heptane and the crude solid obtained was purified by chromatography on SiO 2 gel to afford the title carbonated as a white solid. 1H- DMSO-d6: 1.30 (s, 3H, -CH3), 1.58 (s, 3H, -CH3), 4.3-

4.6 (m, 3H, 5’-CH2 + C4’-H), 5.16 (d, 1H, C3’), 5.42 (d, 1H, C2’), 6.21 (s, 1H, C1’-H),

7.20 (d, 2H, -Ph), 7.40 (s, 2H, -NH2), 7.50 (d, 2H, -Ph), 8.18 (s, 1H, C2-H), 8.36 (s, 1H,

C8-H) ppm.

G. General Procedure for Synthesis of 2’,3’-Isopropylidine-5’-O-[(3-indolyl)-1- ethyl]carbamoyl nucleosides (Adenosine 5’-tryptamine-carbamate acetonide (7)):

(M+1) 494.2723

A mixture of the foregoing nucleoside-5’-carbonate (0.255 mmol) and tryptamine

(0.640 mmol, 2.5 eq) in anhydrous THF (10 mL) was heated to 60ºC (oil bath) under Ar. and stirred at that temperature until TLC showed complete consumption of the starting material (12 h). The solvent was removed by rotary evaporation and the residue chromatographed on SiO 2 gel using CH 2Cl 2:MeOH:H 2O (5:2:0.25) to give the appropriate carbamate and white amorphous solids. 1H- DMSO-d6: 1.34 (s, 3H, -CH3),

1.58 (s, 3H, -CH3), 2.80 (t, 2H, -CH2-), 3.20 (q, 2H, -CH2 NHCO), 4.0- 4.40 (m, 3H, 5’-

CH2 + C4’-H), 5.00 (d, 1H, C3’), 5.42 (d, 1H, C2’), 6.18 (s, 1H, C1’-H), 6.90- 7.54 (7H,

indole + -NH2), 8.18 (s, 1H, C2-H), 8.35 (s, 1H, C8-H), 10.81 (s, 1H, -NH indole) ppm.

117

H. General Procedure for Synthesis of 5’-O-[(3-indolyl)-1-ethyl]carbamoyl nucleosides (Adenosine 5’-tryptamine-carbamate (8)): (M+1) 454.0683

The 2’, 3’acetonide-protected nucleoside carbamate (100 mg) was dissolved in

TFA:H 2O (4:1, 5 ml) and stirred at RT until TLC and ESI-MS showed no starting material left (30 min). The mixture was stripped to dryness and the residue purified by

SiO 2 gel chromatography using CH 2Cl 2:MeOH:H 2O (5:3:0.5) as eluant. The relevant UV- active fractions were collected and concentrated to afford the final compounds as white solids. 1H- DMSO-d6: 2.80 (t, 2H, -CH2-), 3.24 (q, 2H, -CH2 NHCO) 4.00 (m, 2H, C4’-

H, + -OH) 4.25 (m, 2H, 5’-CH2), 4.65 (m, 1H, -OH), 5.38 (d, 1H, C3’), 5.52 (d, 1H,

C2’), 5.90 (s, 1H, C1’-H), 6.90- 7.58 (7H, indole + -NH2), 8.16 (s, 1H, C2-H), 8.38 (s,

1H, C8-H), 10.80 (s, 1H, -NH indole) ppm.

I. Phosphoramidase Assay

In the absence of inhibitors, turnover rates were measured by following the

hydrolysis of the fluorogenic substrate tryptamine 5’-adenosine phosphoramidate (TpAd)

by ecHinT and hHint1 proteins as previously described. 11 Excitation wavelength was set

at 280 nm, fluorescence emission was measured at 360 nm and all the kinetic assays were

performed in triplicate at 25ºC. The rate of hydrolysis of the substrate was determined by

measuring the increase of fluorescence intensity upon the addition of the protein over two

-1 minutes. The Michaelis-Menten constants, k cat (s ) and K m (µM), were determined by a nonlinear regression analysis of the initial velocity versus concentration using JMP IN 7 software. In the presence of inhibitors, the proteins were pre-incubated with the tested inhibitor for 30 s, at 25ºC, before the addition of the substrate. The reaction was followed under the same condition as of in the absence of inhibitors.

118

J. Phenotype Testing

Bacterial strain of wild-type Keio (BW25113) was obtained from the Genomic

Center at Yale University. After growing cultures of wild-type in LB media at 37°C

overnight, cultures were centrifuged at 6000g for 10 min at 4ºC. Cells were re-suspended

and washed twice with M9 medium before inoculation. Cells were grown under

starvation in M9 medium for 24 h before testing. To test the ability of the wild-type strain

to grow on D,L-alanine in the presence and absence of the most potent inhibitors, starved

cells were transferred into M9 medium supplemented with either 22 mM D,L-alanine or

10 mM glucose. Cultures were grown at 37°C for 40 h before OD 600 was measured.

III. Results and Discussion

A. Inhibitors’ Synthesis and Evaluation

Based on the previous findings and knowledge of the general outline of substrate specificity requirements, compounds 1 and 2 (Fig. 1) were designed and synthesized following Scheme 1. Both compounds were titrated into the phosphoramidase assay using the fluorogenic substrate tryptamine 5’-adenosine phosphoramidate (TpAd) (Fig.

1a) as described previously 11 and in the experimental procedures section.

While Hints did not show any appreciable esterase activity against compound 1

and 2, kinetic analysis revealed a significant and dose-dependent decrease in the apparent

Vmax of Hint1 in the presence of either compound 1 or 2 (Fig. 2). In contrast, the K m value for TpAd was unaltered by the presence of the inhibitor. Thus, the mechanism of inhibition was clearly of a noncompetitive type, suggesting that the inhibitor and TpAd

119 substrate do not share a common binding site on the enzyme. Ki values were estimated from the Dixon plot and are summarized in Table 1.

A second series of carbamate analogues was prepared following Schemes 2 and 3.

As shown in Table 1, compound 3, a thymidine carbamate analogue in which the 3- hydroxyl group of the ribose ring was replaced with a primary amine group, inhibited

Hint with a Ki value of 122 µM for ecHinT and 139 µM for hHint1. Replacement of the amine group with an azido moiety at the 3’ position, compound 4, resulted in an inhibitor with a Ki value of 565 µM for ecHinT and 715 µM for hHint1. The 5-fold enhancement in equilibrium binding conferred by the amine group at the 3’ position is most likely explained by the formation of favorable hydrogen bonding interactions between the inhibitor and its binding site in the protein structure. Adenosine carbamate, compound 5, and guanosine carbamate 94 , compound 6, were the most potent inhibitors prepared in this

study. As shown in Table 1, Ki values for the adenosine carbamate analogue ranged

between 73 µM for ecHinT and 103 µM for hHint1 while the Ki values for the guanosine carbamate analogue ranged between 42 µM for ecHinT and 34 µM for hHint1. All compounds exhibited non-competitive inhibition profiles (Fig. 2).

120

Scheme 1. General synthetic scheme of compounds 1 and 2.

H H N O N O R1 R R1 HO 1 O O a n O b n O O O O O O OH OH 2 1

a) indole-3-propionic (n =1) or -butyric (n =2 ) acid, DCC, DMAP, CHCl (0.1% DMSO), RT, 16-48h, 2 2 b) TFA:H2O (4:1), RT, 60 min, 40%

121

Figure 1. Structures of a) Tryptamine 5’-adenosine phosphoramidate and its b) ester and

c) carbamate analogues.

a) NH2 N N H O N N N P O O HN O- OH OH

TpAd

b) H O N R1 n O O

R3 R2

1 n= 1 2 n= 2

c) H N O R1 N O O H

R3 R2

122

Table 1. Inhibition constants determined in HEPES buffer (pH7.2) at 25ºC.

Compound R1 R2 R3 ecHinT hHint Ki (µM) a Ki (µM) a

1 Adenine OH OH 375 (±13) 400 (±27)

2 Adenine OH OH 393 (±17) 457 (±32)

3 Thymine H NH 2 122 (±10) 139 (±17)

4 Thymine H N3 565 (±23) 715 (±35) 5 Adenine OH OH 73 (±4) 103 (±18)

6 Guanine OH OH 42 (±6) b 34 (±4) a Values are means of three experiments, standard deviation is given in parentheses. b Value was taken from reference 94.

123

Scheme 2. General synthetic scheme of compounds 3 and 4.

O O H O NH Cl NH N NH O b O N O a N O N O HO O O N O O O H O N N N3 3 4 3 3

H O c N NH O N O N O H O 5 NH2

a) p-Cl phenyl chloroformate, pyridine, quant. b) tryptamine, anhyd THF, 60ºC, 64% c) 10%Pd-C, H2, MeOH, 72%

124

Scheme 3. General synthetic scheme of compound 5.

125

Figure 2. All compounds exhibited a non-competitive inhibition profile.

a) A representative inhibition profile exhibited by the ester analogue, compound 1.

126 b) A representative inhibition profile exhibited by compound 3.

127 c) A representative inhibition profile exhibited by compound 4.

128 d) A representative inhibition profile exhibited by compound 5.

129

B. Phenotype Testing

Recently, we have been able to connect ecHinT to an established phenotype in E. coli . Using active-site mutants, we demonstrated that the catalytically active ecHinT is required for E. coli to utilize D-alanine as a carbon source. To assess the ability of our compounds to inhibit ecHinT in vivo , we tested the ability of E. coli BW25113 strain to grow on D-alanine in the presence of 100 µM of either compound 3 or 5. Cultures were grown in M9 media, supplemented with either 22 mM D,L-alanine or 10 mM glucose, at

37ºC for 40 h before OD 600 was measured. Consistent with our previous finding, that the

catalytic activity of ecHinT is required for E. coli to grow on D-alanine, inhibition of

ecHinT in BW25113 resulted in its impaired growth when D-alanine is the sole carbon

source (Fig. 3). In addition to the cell permeability properties of compounds 3 and 5, no

toxicity was observed with either inhibitor when glucose is the sole carbon source in the

growth medium.

IV. Summary and Concluding Remarks

In conclusion, we have designed and synthesized the first generation of cell-

permeable Hint inhibitors. Previously, we have shown that ∆hinT E. coli, a mutant strain in which hinT was deleted, exhibited impaired growth when D-alanine is the sole carbon source. 94 Using the developed inhibitors as chemical probes, this work has

demonstrated that inhibition of ecHinT resulted in similar impaired growth observed for

∆hinT in the presence of D-alanine as a sole carbon source. Guanosine carbamate 94 , one

of the best inhibitors reported in this study, may serve as structural template for the

design and development of more potent inhibitors with enhanced inhibition constants.

Ongoing crystallographic studies of the inhibitor-bound form of Hint should provide a

130 structural foundation for understanding the mechanism of inhibition and for the design of selective Hint inhibitors.

131

Figure 3. ecHinT inhibition resulted in impaired phenotype.

*Measurements were carried out in duplicates and the standard deviation was ± 2%.

132

Chapter 5

Possible Physiological Role of The Histidine Triad Nucleotide Binding Protein in Escherichia coli : Regulation of D-alanine Detoxification

133

I. Introduction

Amino acids are best known as the building blocks of proteins, which themselves form the biological machinery of all cells. The chemical side chains of amino acids can be arranged clockwise or counterclockwise forming D or L amino acids. In all kingdoms of life, cells predominantly use the overwhelmingly “left-handed” form of amino acids,

L-amino acids, as building blocks in protein synthesis. Nevertheless, significant quantities of D-amino acids can be produced in most bacteria. D-Ala and D-Glu are the most widely produced amino acids in bacteria. 95 The two amino acids are incorporated

into peptidoglycan in the cell wall structure. 90 Peptidoglycan is a strong and elastic

heteropolymer that is synthesized and modified by penicillin binding proteins. 90 The

chemistry and structure of the peptidoglycan is dynamic and depends on factors that

regulate its D-amino acid composition and architecture. 95; 96

Recently, it has been shown that D-amino acids are released, by a diverse number

of bacterial species, in the stationary phase as agents that can control cell wall assembly

and modification. 97 The regulatory role of D-amino acids appears to be mediated by

signaling metabolic slowing in cell wall and cytoplasmic compartments when resources

become scarce. 97 Supporting the proposed signaling role of D-amino acids in bacteria, it

has been found that a mixture of D-amino acids produced in bacteria prevents biofilm

formation and signals for their disassembly in certain bacterial populations. 91 Indeed, D- tyrosine, D-leucine, D-tryptophan, and D-methionine were shown to function as active inhibitory factors of biofilm formation. 91 Not only did these D-amino acids prevent biofilm formation, but they also disrupted existing biofilm. 91 Interestingly, the inhibitory

role of D-amino acids on biofilm formation is believed to be mediated by the release of

134 amyloid fibers that linked cells in the biofilm together. 91 Although a growing evidence has emerged that bacteria have taken advantage of amino acid chirality to sense and respond to specific environmental conditions, it is still controversial whether D-amino acids have a potential role in signaling between individual bacteria.

Since bacteria are the major producers of D-amino acids in nature, it is therefore not surprising that bacteria possess the biochemical machinery to synthesize and degrade such components. In bacteria, neutral D-amino acids can be oxidized to the corresponding α-keto acids by a flavoenzyme named D-amino acid dehydrogenase

(DadA). 84; 85; 98 The enzyme has a broad specificity towards D-amino acids, of which D- alanine is the best substrate. 84; 85; 98 DadA is a membrane bound protein that is directly linked to a respiratory chain. 86 To protect bacteria from the harmful reactive oxygen species that could damage cellular components, D-amino acid oxidation is not coupled to

86; 87 O2 reduction in bacteria. E. coli DadA is a heterodimer formed by a small subunit and a large subunit. The small subunit contains FAD that receives two electrons from the substrate and transfers them to a sulfur-iron center located in the large subunit. 86; 87

DadA was shown to play two main roles in bacteria. Initially, it allows bacteria to grow using D-amino acids as the sole carbon, nitrogen, and energy source. 86; 87 Second, it

prevents local accumulation of D-amino acids as some D-amino acid analogues have

been shown to exhibit specific inhibitory effects on certain bacterial processes, such as

germination. 88; 89

It has been shown that the expression of DadA is induced by the presence of either D-alanine or L-alanine in the growth medium. 99 A catabolic alanine racemase

converts the D- into the L-stereoisomer of alanine. However, in mutants lacking alanine

135 racemase, D-alanine does not induce transcription from the dad operon, suggesting that

L-alanine is an internal inducer of the operon. 100 In addition, it has been shown that the

induction of DadA is regulated by catabolite repression and therefore the induced and

non-induced levels of DadA are significantly lower in cells grown in the presence of

glucose. 82; 100; 101 The extent of catabolite repression depends upon the concentration of

c-AMP (3'-5 cyclic adenosine monophosphate) inside the cells. Binding of c-AMP to

CAP (Catabolite Activator Protein) causes it to undergo a transition which allows it to

bind to the DNA in the promotor region and thereby permits transcription. 99

In Chapter 3, we have demonstrated, by E. coli gene disruption studies, that the

ecHinT enzyme is necessary for growth under conditions when alanine is the sole carbon

and nitrogen source. 94 Further investigations, with active-site mutants and ecHinT

inhibitors, revealed a unique connection between the discovered phenotype and the

catalytic activity of ecHinT. 94 In addition, we have demonstrated that the expression of

catalytically-active ecHinT is essential for the activity of the enzyme D-alanine

dehydrogenase. 94

Histidine triad nucleotide binding proteins (Hints) are ubiquitous enzyme

members of the histidine triad (HIT) protein superfamily of nucleotidyl transferases and

hydrolases.2 The Hint branch represents the most ancient branch of this superfamily, and can be found in Archaea, Bacteria, and Eukaryotae. Despite their wide-spread and high conservation among all kingdoms of life, a rationale for the occurrence of Hint remains elusive. In particular, previous to our work, the role of Hint1 in the life cycle of prokaryotes in general, and bacteria, in particular, has not been investigated.

136

In this study, we demonstrate that, in addition to its role in the regulation of D- amino acid metabolism, ecHinT is also involved in D-alanine detoxification in E. coli .

Indeed, we discovered that E. coli mutants lacking dadA or hinT genes are highly susceptible to D-alanine toxicity and that the enzymatic activity of ecHinT is essential to protect E. coli from the D-alanine toxicity.

Recently, the biochemical connection between ecHinT and Lysyl tRNA synthetase has been elucidated. 32 Supported by the results of several radiolabeling and mutagenic experiments, it has been proposed that Hint may function in part to regulate the catalytic activity of LysRS. 32 Interestingly, the catalytic activity of Hint has been

shown to be an essential requirement for ability of Hint to hydrolyze lysyl-AMP

generated by LysRS. 32 The proposed function of Hint provided a possible new

mechanism that regulates protein translation by either pre-transfer editing or scavenging

highly reactive inappropriately released aminoacyl-adenylates. Given the similarity

between the molecular determinants that govern both ecHinT-LysRS and ecHinT-DadA

interactions, we hypothesize that the ability of ecHinT to activate DadA is governed by

the ability of aminoacyl tRNA synthetases to adenylate ecHinT to generate the ecHinT-

AMP intermediate. Once the intermediate formed, DadA becomes functionally active to

catalyze D-alanine degradation and prevent its accumulation in the intracellular

compartments. Failure to activate the metabolic pathway of D-alanine catabolism in

bacteria would probably lead to mis-incorporation of D-alanine in protein structures,

generating functionless proteins and thus toxicity.

137

II. Experimental Procedures

A. Bacterial Strains, Media, and Growth Conditions

All bacterial strains used in this study were obtained from the E. coli Genetic

Resource Center at Yale University. All strains were received as glycerol stocks on

filters, sub-cultured twice, on Luria–Bertani (LB)-agar medium and incubated at 37°C for

48 h before testing. The pH of the medium was adjusted to 7.0 with 1.0 M sodium

hydroxide, and the medium was sterilized at 121°C and 15 psi for 20 min. Liquid cultures

were grown aerobically in LB medium supplemented with kanamycin (kan, 50 µg/µl) at

37°C, with shaking, for 18 h. Strains were stored as 10% glycerol stocks at - 80°C. The

mutations in gene knock-out strains were verified using the Polymerase Chain Reaction

(PCR) technique as described previously in chapter 3.

B. Growth Inhibitory Effects of D-amino Acids on Escherichia coli

IC 50 defined as the D-amino acid concentration that produces 50% of the maximal effect was determined for three E. coli strains, wild-type BW25113, ∆hinT and

∆dadA . A starter culture of 1 µL of each strain was added to 5 ml of LB medium and

allowed to grow overnight at 37ºC while shaking at 220 rpm. The following day, 10 µL

of the starter culture was inoculated into 10 ml LB medium and was allowed to shake at

37ºC. Stock solutions of 1 M of the D-amino acids were prepared in phosphate buffered

saline (pH 7.00), and pH was adjusted using 1 N NaOH and 1 N HCl. The stock solutions

were filter-sterilized and diluted in LB medium to various concentrations ranging

between 0-300 mM. Fresh cultures of each strain were supplemented with the desired D-

amino acid concentration. Optical densities at 600 nm, OD 600 , measurements were

138 obtained, using Varian Cary 50 UV-visible spectrophotometer (Palo Alto, CA, USA), after 24 h incubation at 37ºC with shaking. The concentration of D-amino acid inhibiting bacterial growth by 50% (IC 50 ) was estimated by plotting the millimolar concentration of

D-amino acid versus OD 600 and the concentration that causes 50% growth inhibition was determined.

C. Growth Kinetics

Growth curves of E. coli strains metabolizing D-alanine were constructed by

monitoring OD 600 of cultures growing in LB medium in presence of 65 mM of D-alanine.

The test flasks were shaken at 37ºC and samples were drawn at each time point for OD 600 readings until the stationary phase was reached.

D. D-alanine Toxicity in Presence of ecHinT Inhibitors

The synthesis and evaluation of ecHinT inhibitors were described in Chapter 3 and 4. Cultures of E. coli wild-type BW25113 strain in LB medium were prepared as

described previously in the experimental procedures. Stock solutions of 10 mM of the

TpGc inhibitor were prepared in molecular grade DMSO and diluted with distilled water.

100 µM of the inhibitor was added to each culture, with or without D-alanine, with the

control culture having no inhibitor. Cultures were incubated at 37ºC with shaking.

Absorbance readings at 600 nm were obtained after 24 hours.

139

E. Mammalian Cell Lines and Cell Culturing

The cytotoxicity effects of D-alanine, in mammalian cells, were tested in the

following five human cancer cell lines: the human breast cancer cell lines MCF-7 and

MDA-MB-231, the human pancreatic carcinoma cell line MIA PaCa-2, the T-cell

leukemia cell line HPB-MLT and the human Burkitt's lymphoma cell line Raji. Cell lines

were cultured in either high glucose Dulbecco’s Modified Eagle Medium (DMEM)

(Invitrogen, Carlsbad, CA) or Roswell Park Memorial Institute Medium (RPMI)

(Invitrogen, Carlsbad, CA) containing 10% Heat Inactivated Fetal Bovine Serum (HI-

FBS), 2 mM L-glutamine, 50 U/ml penicillin and 50 µg/ml streptomycin. Cell lines were

maintained at 37°C in a 5% CO 2 atmosphere with 95% humidity. Cells were passaged

weekly, and the culture medium was changed twice a week. According to their growth

profiles, the optimal plating densities were determined. To ensure exponential growth

throughout the experimental period and to ensure a linear relationship between

absorbance at 492 nm and cell number when analyzed by the MTS assay, the following

densities were used for each cell line: 5×10 3 cells/well for the MCF-7, MDA-MB-231

and MIA PaCa-2 cell lines, and 15×10 3 cells/well for the HPB-MLT and Raji cell lines.

F. Cell Viability Assay

Cultured cells were counted using trypan blue exclusion method and seeded to 96- well plates at the desired densities. D-alanine and L-alanine stock solutions at a concentration of 1 M were prepared in phosphate buffer and further diluted in medium to produce the following concentrations: 50, 100, 150, 200 and 300 mM. D-alanine or L- alanine at various concentrations was added to each well in triplicate. Treated cells were incubated in a 37°C 5% CO 2 incubator for 24 h. In order to assess cell viability after

140 incubation with various concentrations of D-alanine or L-alanine, MTS assays were carried out using CellTiter 96 Aqueous One solution cell proliferation assay according to the manufacturer’s protocol. The absorbance at 492 nm was read on a Tecan Plate Reader

(Tecan Group Ltd., Mannedorf, Switzerland).

G. D-alanine Cytotoxicity in MCF-7, MDA-MB-231 and MIA PaCa-2 Cell Lines

Cells were detached with the Non-Enzymatic Cell Dissociation Complex (Sigma

Chemical Co.) to make single-cell suspension, and viable cells were counted and diluted with DMEM medium to give a final density of 5×10 3 cells/well. 100 µL per well of cell

suspension was seeded in 96-well plates and incubated to allow for cell attachment. After

24 h, the cells were treated with either D-alanine or L-alanine. Each concentration of D-

alanine or L-alanine was added to each well in triplicate. The plates were incubated for

24 h at 37°C 5% CO 2. Cell viability was measured using MTS assay. Control wells without D-alanine or L-alanine treatment were prepared under the same conditions.

H. D-alanine Cytotoxicity in HPB-MLT and Raji Cell Lines

Viable cells were counted and diluted with RPMI medium to give a final density of 15×10 3 cells/well. 100 µL per well of cell suspension was seeded in 96-well plates.

After 24 h, the cells were treated with either D-alanine or L-alanine. Each concentration

of D-alanine or L-alanine was added to each well in triplicate. The plates were incubated

for 24 h at 37°C 5% CO 2. Cell viability was measured using MTS assay. Control wells without D-alanine or L-alanine treatment were prepared under the same conditions.

141

III. Results

A. D-amino Acids Toxicity in E. coli

The concentration of D-amino acid that inhibits the growth of half of an inoculum of E. coli (IC 50 ) was estimated for three E. coli strains (Table 1). As shown in Figure 1

(A-E), wild-type BW25113 appears to be the most resistant strain to D-alanine treatment, with an IC 50 value that is 2.6-fold and 3.8-fold higher than the IC 50 value of ∆hinT and

∆dadA strains respectively (Table 1). Compared to D-alanine, none of the tested strains show preferential sensitivity to treatment with other tested D-amino acids. All measurements were carried out in duplicates with SD less than 5%. To confirm that the toxicity observed with D-alanine is specific to the D form and not to the L form of alanine, growth of the three strains, under conditions where LB medium was supplemented with various concentrations of L-alanine, was studied. Consistent with the literature reports, our results show that the D-isomer, but not the L-isomer, of alanine was associated with the observed toxicity in E. coli (Fig. 2). The toxicity of the D-isomers of several amino acids for Gram-positive and Gram-negative bacteria has been clearly established. 102; 103; 104 In addition, high concentrations of glycine were reported to

inhibit bacterial growth by replacing both D- and L-alanine residues in peptidoglycan

subunits 104 thereby impairing transpeptidization within the cell wall and leading to

subsequent lysis of bacteria.

142

Figure 1. D-amino acids IC 50 determination in E. coli strains: wild-type BW25113,

∆hinT and ∆dadA were grown in LB medium at 37°C for 24 h in presence of various concentrations of D-amino acids.

a) D-alanine IC 50 .

2

1.5

600 1 wild-type OD ∆dadA ∆hinT 0.5

0 0 50 100 150 200 250 300

[D-alanine] mM

143 b) D-serine IC 50 .

2

1.5

1 wild-type 600

OD ∆hinT

0.5 ∆dadA

0 0 50 100 150 200 250 300 [D-serine] mM

c) D-proline IC 50 .

2

1.5

1 wild-type 600 ∆hinT OD 0.5 ∆dadA

0 0 50 100 150 200 250 300 [D-proline] mM

144 d) D-lysine IC 50 .

2

1.5

600 1 OD wild-type ∆hinT 0.5 ∆dadA

0

0 50 100 150 200 250 300 D-[lysine ] mM

e) Glycine IC 50 .

2

1.5

1 wild-type

600 ∆hinT ∆dadA OD 0.5

0

0 50 100 150 200 250 300

[Glycine] mM

145

Figure 2. Sensitivity of E. coli strains: wild-type BW25113, ∆hinT and ∆dadA , to L- alanine treatment.

2

1.5

1 wild-type

600 ∆hinT

OD 0.5 ∆dadA

0 0 50 100 150 200 250 300

[L-alanine] mM

146

Table 1. Summary of IC 50 values in millimolar.

Amino acid wild-type ∆hinT ∆dadA

L-alanine > 200 > 200 > 200

D-alanine 230 ± 15 90 ± 5 60 ± 5

D-serine 170 ± 10 170 ± 10 170 ± 10

D-proline > 300 > 300 > 300

D-lysine 60 ± 5 60 ± 5 60 ± 5

glycine 130 ± 10 130 ± 10 130 ± 10

147

The growth kinetic experiments were conducted to examine the inhibitory effect of D-alanine on the growth rate of D-alanine-metabolizing E. coli stains. Relative to wild-type BW25113 strain, both ∆hinT and ∆dadA strains show reduced growth rate in

the exponential phase when grown in LB medium in the presence of 65 mM of D-alanine

(Fig 3).

The lethal effects of toxic amino acids were shown to be additive in several

studies . 105 To study the potential for synergitic toxicity between different D-amino acids, the IC 50 of D-alanine in presence of 30 and 60 mM D-lysine was estimated. As

illustrated in Figure 4, 3.8-fold decrease in the IC 50 value, in the presence of 60 mM D-

lysine, and 1.5-fold decrease in the IC 50 value, in the presence of 30 mM D-lysine, were

observed. Although specific parallel pathways may be affected by different amino acids,

it is also possible that a common target is altered, which leads to an additive effect in

toxicity.

B. Effect of ecHinT Inhibition on E. coli Sensitivity to D-alanine

Growth of the wild-type BW25113 strain in LB medium supplemented with 150

mM D-alanine in presence of ecHinT inhibitor, TpGc, was studied. Synthesis and

evaluation of the inhibitor used in this study were described earlier in Chapter 3. As

shown in Figure 5, compared to the control culture, where the medium was supplemented

with 150 mM D-alanine with no inhibitor added, the inhibition of ecHinT results in 2.7-

fold reduction in the growth level observed for the wild-type BW25113 strain. Similar

growth levels in control cultures, in the presence or absence of 100 µM of TpGc, suggest

that the inhibitor itself is not responsible for the observed toxicity (Fig. 5).

148

Figure 3. Growth kinetic curves of three E. coli strains: wild-type BW25113, ∆hinT and

∆dadA were grown in LB medium at 37°C in presence of 65 mM D-alanine.

149

Figure 4. D-alanine and D-lysine synergism: wild-type BW25113 strain was grown in

LB medium at 37°C in presence of various concentrations of D-alanine and D-lysine.

2

1.5

600 1 0 mM D-lys OD 30 mM D-lys 0.5 60 mM D -lys

0 0 50 100 150 200 250 300

[D-alanine] mM

150

Figure 5. Effect of ecHinT inhibition on E. coli sensitivity to D-alanine.

151

C. D-alanine Toxicity in Mammalian Cell Lines

MCF-7, MDA-MB-231, MIA PaCa-2, HPB-MLT and Raji cells were incubated with different concentrations of either D-alanine or L-alanine for up to 24 h. The degree of cell viability after treatment was obtained by conducting MTS assay. Optical densities at 492 nm was used as an indicator of cell viability and expressed as a percentage of cell viability. For the MCF-7 and MDA-MB-231 breast cancer cell lines, a dose-dependent reduction in the percentage of viable cells when treated with D-alanine, but not with L- alanine, was observed (Fig. 6a, 6b). However, D-alanine toxicity was not observed for the human pancreatic carcinoma cell line MIA PaCa-2, the T-cell leukemia cell line

HPB-MLT and the human Burkitt’s lymphoma cell line Raji (Fig. 6c, 6d, 6e). In addition to other possible contributing pathways of toxicity, the potential toxicity of D-alanine on

MCF-7, MDA-MB-231 might be explained by the oxidative damage to cells by H 2O2, which is formed upon the oxidation of D-alanine by D-amino acid oxidase. 106

IV. Discussion

In E. coli, D-alanine might be utilized in two main pathways that prevent its accumulation in cellular compartments. First, D-alanine is incorporated in the peptidoglycan, the structural heteropolymer found in bacterial cell walls. The incorporation of D-alanine into peptidoglycan precursors is catalyzed via the sequential action of alanine racemase and several ligases. Although this is a dynamic process that is well-regulated by enzymatic activity, less is known about the factors that regulate alterations in the composition and architecture of the peptidoglycan heteropolymer (Fig.

7).

152

Figure 6. D-alanine toxicity on mammalian cell lines. a) MCF-7 cell line.

153 b) MDA-MB-231 cell line.

c) MIA PaCa-2 cell line.

154 d) Raji cell line.

e) HPB-MLT cell line.

155

Figure 7. The structure of peptidoglycan. Peptidoglycan is built from prefabricated units that contain a peptide chain of three to five amino acids attached to the N-Acetyl- muramic acid. The peptide chain contains a terminal d-alanine-d-alanine unit.

156

The second pathway is a metabolic pathway that involves the conversion of D-

alanine into pyruvate, by DadA; thus serving as an energy precursor for cellular

processes. Under basal conditions, the two pathways are optimized to regulate and

maintain certain levels of intracellular pools of D-alanine for cell wall biosynthesis and

energy utilization. When E. coli is exposed to an environment that is rich in D-alanine,

these pathways probably get activated to protect the cells against D-alanine toxicity that

is induced by exceeding the capacity of the cellular and metabolic machinery.

Although potential toxicity has been linked to D-amino acids, the fundamental

mechanism by which toxicity occurs is not fully understood. In animal studies,

administration of D-amino acids to rats and chicks resulted in growth inhibition of the

animals. 107 In addition, D-amino acids were found to accumulate in certain tissues

resulting in serious damage, e.g., suppression of the synthesis of glutamate oxaloacetate

transaminase, glutamic pyruvic transaminase, and lactate dehydrogenase. 107

Several mechanisms were proposed to explain the toxicity associated with D- amino acids in both eukaryotes and prokaryotes. The proposed pathway of D-amino acids toxicity in eukaryotes appears to be related to the oxidative damage to cells by

H2O2, which is formed upon the oxidation of D-amino acids by D-amino acid oxidase

and/or D-aspartate oxidase. 106 In E. coli , however, D-amino acids oxidation by DadA is not coupled to O 2. Hence, the toxicity of D-amino acids in E. coli could not be explained

by the previous mechanism.

Another pathway has been proposed to explain the toxicity of D-amino acids in

both eukaryotes and prokaryotes. Based on several pieces of evidence, it has been

suggested that D-amino acids toxicity is directly related to the formation of D-aminoacyl

157 tRNA. D-Amino acids are usually prevented from being incorporated into proteins because aminoacyl tRNA synthetases have preferential binding sites for L-amino acids.

108; 109; 110; 111 However, it was observed that Escherichia coli and Bacillus subtilis tyrosyl-tRNA synthetases can potentially catalyze the formation of D-tyrosyl-tRNA Tyr to the same extent found for the L-enantiomer. 108; 109 Interestingly, extracts of E. coli , yeast, rabbit reticulocytes, and rat liver were shown to have an enzyme activity capable of accelerating the hydrolysis of the ester linkage of D-Tyr-tRNA, thus releasing the tRNA from D-tyrosine. 112 In E. coli , this enzyme activity has been assigned to D-Tyrosyl-

tRNA Tyr deacylase which is encoded by the yihZ gene. D-Tyrosyl-tRNA Tyr deacylase has

a broad range of substrate specificity and can also hydrolyze D-aspartyl-tRNA Asp and D-

tryptophanyl-tRNA Trp into the corresponding D-amino acid and free tRNA. An E. coli strain, in which the yihZ gene was disrupted, was found to be more sensitive to D- tyrosine, D-tryptophan, D-aspartate, D-glutamate, and D-serine than the wild-type strain.

112 Consistent with this finding, about 40% of the cellular tRNA Tyr in E. coli lacking D- tyrosyltRNA Tyr deacylase was demonstrated to bear D-tyrosyl-tRNA Tyr when the cells were grown in the presence of 2.4 mM D-tyrosine. 111 Over-expression of tRNA Tyr , tRNA Trp , and tRNA Asp protects the E. coli yihZ –knock–out strain against the toxic effect of D-tyrosine, D-tryptophan, and D-aspartate, respectively. 111

The fate of D-aminoacyl tRNA in E. coli has never been studied previously.

However, Dedkova et al . demonstrated that D-methionine and D-phenylalanine charged

on aminoacyl tRNA were incorporated into proteins in vitro 113 , suggesting that the

toxicity of D-amino acids might arise from the mis-incorporation of D-amino acids in

protein structure resulting in the formation of functionless protein.

158

In previous work, we demonstrated that lysyl-AMP generated by LysRS is a

physiological substrate for Hints. 32 In addition, we showed that the ability of Hint to hydrolyze lysyl-AMP depends on its enzymatic activity. Based on these findings, we proposed that Hints might function as potential regulators of protein translation processes. Interestingly, the molecular determinants governing this regulation appear to be overlapping with the determinants of ecHinT-DadA interaction regulation (Table 2). 94

Based on these observations, we hypothesize that there might be a possible correlation between the ability of aminoacyl tRNA synthetases to adenylate ecHinT and the activation of DadA by ecHinT (Fig. 8). Once DadA is functionally active, the metabolic pathway of D-alanine catabolism is also activated; thus preventing D-alanine accumulation in the intracellular compartment. We propose that accumulation of D- alanine would probably result in charging tRNA molecules with D-alanine instead of the

L-isomers of amino acids. Consequently, D-alanine might be mis-incorporated in protein structures generating mis-folded proteins, which would probably lead to the observed toxicity of D-alanine in E. coli . The ability of Hint to cross-talk to aminoacyl tRNA synthetases and hydrolyze aminoacyl-AMP intermediates suggest that Hint has a role in pre-transfer editing or scavenging the inappropriately released D-aminoacyl-adenylates; thus preventing D-amino acids mis-incorporation in protein structures. Ongoing studies attempting to unravel the Hint interaction network in E. coli should shed light on this hypothesis.

159

Table 2. Summary of the ecHinT-LysRS and ecHinT-DadA studies.

Protein Activity LysRS-Adenylation DadA Activity

ecHinT Yes Yes Yes hHint1 Yes No No ecHinT-H101A No No No ∆114-119 Yes No No ∆117-119 Yes No No ec /Hs Yes Yes Yes Hs/ ec Yes Yes No

160

Figure 8. D,L-alanine metabolism and possible ecHinT-LysRS regulation.

YanJ alanine ACSS CycA Serine/Alanine Transporter Glycine APC transporter

O Alanine O Racemase -O -O

NH3+ NH3+ D-Alanine L-Alanine ‘

H2O dadA-Active hinT LysRS•Lys-AMP Pi ? NH4+ dadA-Inactive hinT-AMP LysRS•Lys ATP

O -O Glycolysis 1 O Pyruvate

161

BIBLIOGRAPHY

1. Kijas, A. W., Harris, J. L., Harris, J. M. & Lavin, M. F. (2006). Aprataxin forms a

discrete branch in the HIT (histidine triad) superfamily of proteins with both

DNA/RNA binding and nucleotide hydrolase activities. Journal of Biological

Chemistry 281, 13939-13948.

2. Brenner, C. (2002). Hint, Fhit, and GalT: Function, structure, evolution, and

mechanism of three branches of the histidine triad superfamily of nucleotide

hydrolases and transferases. Biochemistry 41, 9003-9014.

3. Liu, H. D., Rodgers, N. D., Jiao, X. & Kiledjian, M. (2002). The scavenger

mRNA decapping enzyme DcpS is a member of the HIT family of

pyrophosphatases. EMBO Journal 21, 4699-4708.

4. Kwasnicka, D. A., Krakowiak, A., Thacker, C., Brenner, C., and Vincent, S. R.

(2003). Coordinate expression of NADPH-dependent flavin reductase, Fre-1, and

Hint-related 7meGMP-directed hydrolase, DCS-1. J. Biol. Chem. 278, 39051-

39058.

5. Lima, C. D., Klein, M. G., Weinstein, I. B., and Hendrickson, W. A. (1996).

Three-dimensional structure of human protein kinase C interacting protein 1, a

member of the HIT family of proteins. Proceedings of the National Academy of

Sciences of the United States of America 93, 5357-5362.

6. Brenner, C., Bieganowski, P., Pace, H. C. & Huebner, K. (1999). The histidine

triad superfamily of nucleotide-binding proteins. Journal of Cellular Physiology

181, 179-187.

162

7. Brenner, C., Garrison, P., Gilmour, J., Peisach, D., Ringe, D., Petsko, G. A. &

Lowenstein, J. M. (1997). Crystal structures of HINT demonstrate that histidine

triad proteins are GalT-related nucleotide-binding proteins. Nature Structural

Biology 4, 231-238.

8. Chou, T. F., Bieganowski, P., Shilinski, K., Cheng, J. L., Brenner, C. & Wagner,

C. R. (2005). P-31 NMR and genetic analysis establish hinT as the only

Escherchia coli purine nucleoside phosphoramidase and as essential for growth

under high salt conditions. Journal of Biological Chemistry 280, 15356-15361.

9. Bieganowski, P., Garrison, P. N., Hodawadekar, S. C., Faye, G., Barnes, L. D. &

Brenner, C. (2002). Adenosine monophosphoramidase activity of Hint and Hnt1

supports function of Kin28, Ccl1, and Tfb3. Journal of Biological Chemistry 277,

10852-10860.

10. Krakowiak, A., Pace, H. C., Blackburn, G. M., Adams, M., Mekhalfia, A.,

Kaczmarek, R., Baraniak, J., Stec, W. J. & Brenner, C. (2004). Biochemical,

crystallographic, and mutagenic characterization of hint, the AMP-lysine

hydrolase, with novel substrates and inhibitors. Journal of Biological Chemistry

279, 18711-18716.

11. Chou, T.-F., Baraniak, J., Kaczmarek, R., Zhou, X., Cheng, J., Ghosh, B., and

Wagner, C. R. (2007). Phosphoramidate Pronucleotides: A Comparison of the

Phosphoramidase Substrate Specificity of Human and E. coli Histidine Triad

Nucleotide Binding Proteins (Hint1). Mol. Pharmaceutics 4, 208-217.

12. Pace, H. C., Garrison, P. N., Robinson, A. K., Barnes, L. D., Draganescu, A.,

Rosler, A., Blackburn, G. M., Siprashvili, Z., Croce, C. M., Huebner, K. &

163

Brenner, C. (1998). Genetic, biochemical, and crystallographic characterization of

Fhit-substrate complexes as the active signaling form of Fhit. Proceedings of the

National Academy of Sciences of the United States of America 95, 5484-5489.

13. Ohta, M., Inoue, H., Cotticelli, M. G., Kastury, K., Baffa, R., Palazzo, J.,

Siprashvili, Z., Mori, M., McCue, P., Druck, T., Croce, C. M., and Huebner, K.

(1996). The FHIT gene, spanning the 3p14.2 fragile site and renal

carcinoma-associated t(3;8) breakpoint, is abnormal in digestive tract cancers.

Cell 84, 587-597.

14. Ishii H, O. K., Furukawa Y. (2003). Alteration of the fragile histidine triad gene

early in carcinogenesis: an update. J. Exp. Ther. Oncol. 3, 291-296.

15. Huebner, K., Croce, CM. (2003). Cancer and the FRA3B/FHIT fragile locus: it's a

HIT. Br. J. Cancer 88, 1501-1506.

16. Pekarsky, Y., Zanesi, N., Palamarchuk, A., Huebner, K., Croce, CM. (2002).

FHIT: from gene discovery to cancer treatment and prevention. Lancet. Oncol 3,

748-754.

17. Pekarsky, Y., Garrison, P. N., Palamarchuk, A., Zanesi, N., Aqueilan, R. I.,

Huebner, K., Barnes, L. D., Croce, C. M. (2004). Fhit is a physiological target of

the protein kinase Src. Proc Natl Acad Sci USA 101, 3775-3779.

18. Semba S, T. F., Fabbri, M., McCorkell, K. A., Volinia, S., Druck, T., Iliopoulos,

D., Pekarsky, Y., Ishii, H., Garrison, P. N., Barnes, L. D., Croce, C. M. and

Huebner, K. (2006). Fhit modulation of the Akt-survivin pathway in lung cancer

cells: Fhit-tyrosine 114 (Y114) is essentialFhit modulation of the Akt-survivin

pathway. Oncogene 25, 2860-2872.

164

19. Siprashvili, Z., Sozzi, G., Barnes, L. D., McCue, P., Robinson, A. K., Eryomin,

V., Sard, L., Tagliabue, E., Greco, A., Fusetti, L., Schwartz, G., Pierotti, M. A.,

Croce, C. M. & Huebner, K. (1997). Replacement of Fhit in cancer cells

suppresses tumorigenicity. Proceedings of the National Academy of Sciences of

the United States of America 94, 13771-13776.

20. Trapasso, F., Krakowiak, A., Cesari, R., Arkles, J., Yendamuri, S., Ishii, H.,

Vecchione, A., Kuroki, T., Bieganowski, P., Pace, H. C., Huebner, K., Croce, C.

M. & Brenner, C. (2003). Designed FHIT alleles establish that Fhit-induced

apoptosis in cancer cells is limited by substrate binding. Proceedings of the

National Academy of Sciences of the United States of America 100, 1592-1597.

21. Frey, P. A., Wong, L. J., Sheu, K. F. & Yang, S. L. (1982). Galactose-1-

Phosphate Uridylyltransferase - Detection, Isolation, and Characterization of the

Uridylyl Enzyme. Methods in Enzymology 87, 20-36.

22. Frey, P. A. (1996). The Leloir pathway: A mechanistic imperative for three

enzymes to change the stereochemical configuration of a single carbon in

galactose. FASEB Journal 10, 461-470.

23. Moreira, M. C., Barbot, C., Tachi, N., Kozuka, N., Uchida, E., Gibson, T.,

Mendonca, P., Costa, M., Barros, J., Yanagisawa, T., Watanabe, M., Ikeda, Y.,

Aoki, M., Nagata, T., Coutinho, P., Sequeiros, J. & Koenig, M. (2001). The gene

mutated in ataxia-ocular apraxia 1 encodes the new HIT/Zn-finger protein

aprataxin. Nature Genetics 29, 189-193.

24. Date, H., Onodera, O., Tanaka, H., Iwabuchi, K., Uekawa, K., Igarashi, S., Koike,

R., Hiroi, T., Yuasa, T., Awaya, Y., Sakai, T., Takahashi, T., Nagatomo, H.,

165

Sekijima, Y., Kawachi, I., Takiyama, Y., Nishizawa, M., Fukuhara, N., Saito, K.,

Sugano, S. & Tsuji, S. (2001). Early-onset ataxia with ocular motor apraxia and

hypoalbuminemia is caused by mutations in a new HIT superfamily gene. Nature

Genetics 29, 184-188.

25. Ahel, I., Rass, U., El-Khamisy, S. F., Katyal, S., Clements, P. M., McKinnon, P.

J., Caldecott, K. W. & West, S. C. (2006). The neurodegenerative disease protein

aprataxin resolves abortive DNA ligation intermediates. Nature 443, 713-716.

26. Wang, Z. R. & Kiledjian, M. (2001). Functional link between the mammalian

exosome and mRNA decapping. Cell 107, 751-762.

27. Guranowski, A. (2000). Specific and nonspecific enzymes involved in the

catabolism of mononucleoside and dinucleoside polyphosphates. Pharmacology

& Therapeutics 87, 117-139.

28. Guranowski, A., Starzynska, E., Bojarska, E., Stepinski, J. & Darzynkiewicz, E.

(1996). Dinucleoside 5',5'''-P-1,P-3-triphosphate hydrolase from yellow lupin

(Lupinus luteus) seeds: Purification to homogeneity and hydrolysis of mRNA 5'-

cap analogs. Protein Expression and Purification 8, 416-422.

29. Gilmour, J., Liang, N. C. & Lowenstein, J. M. (1997). Isolation, cloning and

characterization of a low-molecular-mass purine nucleoside- and nucleotide-

binding protein. Biochemical Journal 326, 471-477.

30. Shuman, S. & Lima, C. D. (2004). The polynucleotide ligase and RNA capping

enzyme superfamily of covalent nucleotidyltransferases. Current Opinion in

Structural 14, 757-764.

166

31. Fankhauser, H., Schiff, J. A. & Garber, L. J. (1981). Purification and Properties of

Adenylyl Sulfate - Ammonia Adenylyltransferase from Chlorella Catalyzing the

Formation of Adenosine 5'-Phosphoramidate from Adenosine 5'-Phosphosulfate

and Ammonia. Biochemical Journal 195, 545-560.

32. Chou, T. F. & Wagner, C. R. (2007). Lysyl-tRNA synthetase-generated lysyl-

adenylate is a substrate for histidine triad nucleotide binding proteins. Journal of

Biological Chemistry 282, 4719-4727.

33. Lima, C. D., Klein, M. G. & Hendrickson, W. A. (1997). Structure-based analysis

of catalysis and substrate definition in the HIT . Science 278, 286-

290.

34. Lee, Y. N., Nechushtan, H., Figov, N. & Razin, E. (2004). The function of Lysyl-

tRNA synthetase and Ap4A as signaling regulators of MITF activity in Fc epsilon

RI-activated mast cells. Immunity 20, 145-151.

35. Barnes, L. D., Garrison, P. N., Siprashvili, Z., Guranowski, A., Robinson, A. K.,

Ingram, S. W., Croce, C. M., Ohta, M. & Huebner, F. (1996). Fhit, a putative

tumor suppressor in humans, is a dinucleoside 5',5'''-P-1,P-3-triphosphate

hydrolase. Biochemistry 35, 11529-11535.

36. Geeganage, S. & Frey, P. A. (1998). Transient kinetics of formation and reaction

of the uridylyl-enzyme form of galactose-1-P uridylyltransferase and its Q168R-

Variant: Insight into the molecular basis of galactosemia. Biochemistry 37, 14500-

14507.

167

37. Wedekind, J. E., Frey, P. A. & Rayment, I. (1996). The structure of

nucleotidylated Histidine-166 of galactose-L-phosphate uridylyltransferase

provides insight into phosphoryl group transfer. Biochemistry 35, 11560-11569.

38. Abend, A., Garrison, P. N., Barnes, L. D. & Frey, P. A. (1999). Stereochemical

retention of the configuration in the action of Fhit on phosphorus-chiral

substrates. Biochemistry 38, 3668-3676.

39. Huang, K., Arabshahi, A., Wei, Y. & Frey, P. A. (2004). The mechanism of

action of the fragile histidine triad, Fhit: Isolation of a covalent adenylyl enzyme

and chemical rescue of H96G-Fhit. Biochemistry 43, 7637-7642.

40. Huang, K. S., Arabshahi, A. & Frey, P. A. (2005). pH-dependence in the

hydrolytic action of the human fragile histidine triad. European Journal of

Organic Chemistry , 5198-5206.

41. McDonald, J. R. & Walsh, M. P. (1985). Ca-+2-Binding Proteins from Bovine

Brain Including a Potent Inhibitor of Protein Kinase-C. Biochemical Journal 232,

559-567.

42. Pearson, J. D., Dewald, D. B., Mathews, W. R., Mozier, N. M., Zurcherneely, H.

A., Heinrikson, R. L., Morris, M. A., McCubbin, W. D., McDonald, J. R., Fraser,

E. D., Vogel, H. J., Kay, C. M. & Walsh, M. P. (1990). Amino-Acid-Sequence

and Characterization of a Protein Inhibitor of Protein Kinase-C. Journal of

Biological Chemistry 265, 4583-4591.

43. Klein, M. (1997). Studies on the b isoform of protein kinase C and a putative

protein kinase C inhibitor which is a member of the highly conserved histidine

triad protein family. New York, Columbia University, Ph.D. thesis.

168

44. Mozier, N. M., Walsh, M. P. & Pearson, J. D. (1991). Characterization of a Novel

Zinc-Binding Site of Protein-Kinase-C Inhibitor-1. FEBS Letters 279, 14-18.

45. Chou, T. F., Tikh, I. B., Horta, B. A. C., Ghosh, B., De Alencastro, R. B. &

Wagner, C. R. (2007). Engineered monomeric human histidine triad nucleotide-

binding protein 1 hydrolyzes fluorogenic acyl-adenylate and lysyl-tRNA

synthetase-generated lysyl-adenylate. Journal of Biological Chemistry 282,

15137-15147.

46. Korsisaari, N. & Makela, T. P. (2000). Interactions of Cdk7 and Kin28 with

Hint/PKCI-1 and Hnt1 histidine triad proteins. Journal of Biological Chemistry

275, 34837-34840.

47. Korsisaari, N., Rossi, D. J., Luukko, K., Huebner, K., Henkemeyer, M. & Makela,

T. P. (2003). The histidine triad protein Hint is not required for murine

development or Cdk7 function. Molecular and Cellular Biology 23, 3929-3935.

48. Weiske, J. & Huber, O. (2005). The histidine triad protein Hint1 interacts with

Pontin and Reptin and inhibits TCF-beta-catenin-mediated transcription. Journal

of Cell Science 118, 3117-3129.

49. Li, H., Zhang, Y., Su, T., Santella, R. M. & Weinstein, I. B. (2006). Hint1 is a

haplo-insufficient tumor suppressor in mice. Oncogene 25, 713-721.

50. Yuan, B. Z., Jefferson, A. M., Popescu, N. C. & Reynolds, S. H. (2004). Aberrant

gene expression in human non small cell lung carcinoma cells exposed to

demethylating agent 5-aza-2 '-deoxycytidine. Neoplasia 6, 412-419.

169

51. Weiske, J. & Huber, O. (2006). The histidine triad protein Hint1 triggers

apoptosis independent of its enzymatic activity. Journal of Biological Chemistry

281, 27356-27366.

52. Wang, L., Zhang, Y. J., Li, H. Y., Xu, Z. H., Santella, R. M. & Weinstein, I. B.

(2007). Hint1 inhibits growth and activator protein-1 activity in human colon

cancer cells. Cancer Research 67, 4700-4708.

53. Li, H. Y., Balajee, A. S., Su, T., Cen, B., Hei, T. K. & Weinstein, I. B. (2008).

The HINT1 tumor suppressor regulates both gamma-H2AX and ATM in response

to DNA damage. Journal of Cell Biology 183, 253-265.

54. Liu, Q., Puche, A. C. & Wang, J. B. (2008). Distribution and expression of

protein kinase c interactive protein (PKCI/HINT1) in mouse central nervous

system (CNS). Neurochemical Research 33, 1263-1276.

55. Su, T., Suzui, M., Wang, L., Lin, C. S., Xing, W. Q. & Weinstein, I. B. (2003).

Deletion of histidine triad nucleotide-binding protein 1/PKC-interacting protein

growth in mice enhances cell growth and carcinogenesis. Proceedings of the

National Academy of Sciences of the United States of America 100, 7824-7829.

56. Rodriguez-Munoz, M., de la Torre-Madrid, E., Sanchez-Blazquez, P., Wang, J. B.

& Garzon, J. (2008). NMDAR-nNOS generated zinc recruits PKC gamma to the

HINT1-RGS17 complex bound to the C terminus of Mu-opioid receptors.

Cellular Signalling 20, 1855-1864.

57. Martin, J., Magnino, F., Schmidt, K., Piguet, A. C., Lee, J. S., Semela, D., St-

Pierre, M. V., Ziemiecki, A., Cassio, D., Brenner, C., Thorgeirsson, S. S. &

170

Dufour, J. F. (2006). Hint2, a mitochondrial apoptotic sensitizer down-regulated

in hepatocellular carcinoma. Gastroenterology 130, 2179-2188.

58. Lenglet, S., Antigny, F., Vetterli, L., Dufour, J. F. & Rossier, M. F. (2008). Hint2

Is Expressed in the Mitochondria of H295R Cells and Is Involved in

Steroidogenesis. Endocrinology 149, 5461-5469.

59. Chou, T. F., Cheng, J. L., Tikh, I. B. & Wagner, C. R. (2007). Evidence that

human histidine triad nucleotide binding protein 3 (Hint3) is a distinct branch of

the histidine triad (HIT) superfamily. Journal of Molecular Biology 373, 978-989.

60. Butland, G., Peregrin-Alvarez, J. M., Li, J., Yang, W. H., Yang, X. C., Canadien,

V., Starostine, A., Richards, D., Beattie, B., Krogan, N., Davey, M., Parkinson, J.,

Greenblatt, J. & Emili, A. (2005). Interaction network containing conserved and

essential protein complexes in Escherichia coli. Nature 433, 531-537.

61. Hopfe, M., Hoffmann, R. & Henrich, B. (2004). P80, the HinT interacting

membrane protein, is a secreted antigen of Mycoplasma hominis. Bmc

Microbiology 4.

62. Kitzerow, A. & Henrich, B. (2001). The cytosolic HinT protein of Mycoplasma

hominis interacts with two membrane proteins. Molecular Microbiology 41, 279-

287.

63. Chou, T. F., Sham, Y. Y. & Wagner, C. R. (2007). Impact of the C-terminal loop

of histidine triad nucleotide binding protein1 (Hint1) on substrate specificity.

Biochemistry 46, 13074-13079.

171

64. Bardaweel, S., Pace, J., Chou, T.-F., Cody, V., Wagner, C. R. (2010). Probing the

impact of the ecHinT C-terminal domain on structure and catalysis. Journal of

Molecular Biology 404, 627-638.

65. Barbier, E., Zapata, A., Oh, E., Liu, Q., Zhu, F., Undie, A., Shippenberg, T. &

Wang, J. B. (2007). Supersensitivity to amphetamine in protein kinase-C

interacting protein/HINT1 knockout mice. Neuropsychopharmacology 32, 1774-

1782.

66. Lee, Y. N. & Razin, E. (2005). Nonconventional involvement of LysRS in the

molecular mechanism of USF2 transcriptional activity in Fc epsilon RI-activated

mast cells. Molecular and Cellular Biology 25, 8904-8912.

67. McPhillips, T. M., McPhillips, S. E., Chiu, H. J., Cohen, A. E., Deacon, A. M.,

Ellis, P. J., Garman, E., Gonzalez, A., Sauter, N. K., Phizackerley, R. P., Soltis, S.

M. & Kuhn, P. (2002). Blu-Ice and the Distributed Control System: software for

data acquisition and instrument control at macromolecular crystallography

beamlines. Journal of Synchrotron Radiation 9, 401-406.

68. Cohen, A. E., Ellis, P. J., Miller, M. D., Deacon, A. M. & Phizackerley, R. P.

(2002). An automated system to mount cryo-cooled protein crystals on a

synchrotron beamline, using compact sample cassettes and a small-scale robot.

Journal of Applied Crystallography 35, 720-726.

69. Gonzalez, A., Moorhead, P., McPhillips, S. E., Song, J., Sharp, K., Taylor, J. R.,

Adams, P. D., Sauter, N. K. & Soltis, S. M. (2008). Web-Ice: integrated data

collection and analysis for macromolecular crystallography. Journal of Applied

Crystallography 41, 176-184.

172

70. Otwinowski, Z. M., W. (1997). Processing of X-ray Diffraction Data Collected in

Oscillation Mode. In: Methods in Enzymology. Edited by C.W. Carter, Jr. & R.

M. Sweet, 276, Part A, pp 224-225. New York: Academic Press .

71. Vagin, A. & Teplyakov, A. (1997). MOLREP: an automated program for

molecular replacement. Journal of Applied Crystallography 30, 1022-1025.

72. Emsley, P. & Cowtan, K. (2004). Coot: model-building tools for molecular

graphics. Acta Crystallographica Section D-Biological Crystallography 60, 2126-

2132.

73. Collaborative Computational Project, Number 4. (1994). The CCP4 Suite:

Programs for Protein Crystallography. Acta Crystallogr. D50, 760-763.

74. Laskowski, R. A., Macarthur, M. W., Moss, D. S. & Thornton, J. M. (1993).

Procheck - a Program to Check the Stereochemical Quality of Protein Structures.

Journal of Applied Crystallography 26, 283-291.

75. Steer, B. A. & Schimmel, P. (1999). Different adaptations of the same peptide

motif for tRNA functional contacts by closely homologous tRNA synthetases.

Biochemistry 38, 4965-4971.

76. Shiba, K., Stello, T., Motegi, H., Noda, T., MusierForsyth, K. & Schimmel, P.

(1997). Human lysyl-tRNA synthetase accepts nucleotide 73 variants and rescues

Escherichia coli double-defective mutant. Journal of Biological Chemistry 272,

22809-22816.

77. Francklyn, C. S., First, E. A., Perona, J. J. & Hou, Y. M. (2008). Methods for

kinetic and thermodynamic analysis of aminoacyl-tRNA synthetases. Methods 44,

100-118.

173

78. Ito, M., Baba, T. & Mori, H. (2005). Functional analysis of 1440 Escherichia coli

genes using the combination of knock-out library and phenotype microarrays.

Metabolic Engineering 7, 318-327.

79. Baba, T., Ara, T., Hasegawa, M., Takai, Y., Okumura, Y., Baba, M., Datsenko, K.

A., Tomita, M., Wanner, B. L. & Mori, H. (2006). Construction of Escherichia

coli K-12 in-frame, single-gene knockout mutants: the Keio collection. Molecular

Systems Biology 2.

80. Bochner, B. R., Gadzinski, P. & Panomitros, E. (2001). Phenotype MicroArrays

for high-throughput phenotypic testing and assay of gene function. Genome

Research 11, 1246-1255.

81. Sambrook, J. F., E.F. and Maniatis, T. (1989) Molecular Cloning: a laboratory

manual. 2nd ed. N.Y., Cold Spring Harbor Laboratory, Cold Spring Harbor

Laboratory Press, 1659 p. ISBN 0-87969-309-6.

82. Franklin, F. C. H. & Venables, W. A. (1976). Biochemical, Genetic, and

Regulatory Studies of Alanine Catabolism in Escherichia-Coli-K12. Molecular &

General Genetics 149, 229-237.

83. Abrahamson, J. L. A., Baker, L. G., Stephenson, J. T. & Wood, J. M. (1983).

Proline Dehydrogenase from Escherichia-Coli-K12 - Properties of the Membrane-

Associated Enzyme. European Journal of Biochemistry 134, 77-82.

84. Olsiewski, P. J., Kaczorowski, G. J. & Walsh, C. (1980). Purification and

Properties of D-Amino-Acid Dehydrogenase, an Inducible Membrane-Bound

Iron-Sulfur Flavoenzyme from Escherichia-Coli-B. Journal of Biological

Chemistry 255, 4487-4494.

174

85. Pollegioni, L., Piubelli, L., Sacchi, S., Pilone, M. S. & Molla, G. (2007).

Physiological functions of D-amino acid oxidases: from yeast to humans. Cellular

and Molecular Life Sciences 64, 1373-1394.

86. Raunio, R. P. a. J., W. T. (1973). D-alanine oxidase from Escherichia coli:

localization and induction by Lalanine. J. Bacteriol. 115, 560 - 566.

87. Deutch, C. E. (2004). Oxidation of 3,4-dehydro-D-proline and other D-amino acid

analogues by D-alanine dehydrogenase from Escherichia coli . FEMS

Microbiology Letters 238, 383-389.

88. Lobocka, M., Hennig, J., Wild, J. & Klopotowski, T. (1994). Organization and

Expression of the Escherichia-Coli K-12 Dad Operon Encoding the Smaller

Subunit of D-Amino-Acid Dehydrogenase and the Catabolic Alanine Racemase.

Journal of Bacteriology 176, 1500-1510.

89. Yasuda, Y. & Tochikubo, K. (1985). Germination-Initiation and Inhibitory

Activities of L-Alanine and D-Alanine Analogs for Bacillus-Subtilis Spores -

Modification of Methyl-Group of L-Alanine and D-Alanine. Microbiology and

Immunology 29, 229-241.

90. Walsh, C. T. (1989). Enzymes in the D-Alanine Branch of Bacterial-Cell Wall

Peptidoglycan Assembly. Journal of Biological Chemistry 264, 2393-2396.

91. Kolodkin-Gal, I., Romero, D., Cao, S. G., Clardy, J., Kolter, R. & Losick, R. D-

Amino Acids Trigger Biofilm Disassembly. Science 328, 627-629.

92. Razin, E., Zhang, Z. C., Nechushtan, H., Frenkel, S., Lee, Y. N., Arudehandran,

R. & Rivera, J. (1999). Suppression of microphthalmia transcriptional activity by

175

its association with protein kinase C-interacting protein 1 in mast cells. Journal of

Biological Chemistry 274, 34272-34276.

93. Yannay-Cohen, N. & Razin, E. (2006). Translation and transcription: the dual

functionality of LysRS in mast cells. Molecules and Cells 22, 127-132.

94. Bardaweel, S., Ghosh, B., Chou, T. F., Sadowsky, M., Wagner, C. R. (2010).

HINT, a histidine triad nucleotide binding protein, is essential for alanine

metabolism in escherichia coli. Journal of Biological Chemistry, Submitted.

95. Holtje, J. V. (1998). Growth of the stress-bearing and shape-maintaining murein

sacculus of Escherichia coli. Microbiology and Molecular Biology Reviews 62,

181-189.

96. Vollmer, W., Blanot, D. & de Pedro, M. A. (2008). Peptidoglycan structure and

architecture. Fems Microbiology Reviews 32, 149-167.

97. Lam, H., Oh, D. C., Cava, F., Takacs, C. N., Clardy, J., de Pedro, M. A. &

Waldor, M. K. (2009). D-Amino Acids Govern Stationary Phase Cell Wall

Remodeling in Bacteria. Science 325, 1552-1555.

98. Olsiewski, P. J., Kaczorowski, G. J., Walsh, C. T. & Kaback, H. R. (1981).

Reconstitution of Escherichia-Coli Membrane-Vesicles with D-Amino-Acid

Dehydrogenase. Biochemistry 20, 6272-6279.

99. Wild, J. & Obrepalska, B. (1982). Regulation of Expression of the Dada Gene

Encoding D-Amino-Acid Dehydrogenase in Escherichia-Coli - Analysis of Dada-

Lac Fusions and Direction of Dada Transcription. Molecular & General Genetics

186, 405-410.

176

100. Wild, J., Hennig, J., Lobocka, M., Walczak, W. & Klopotowski, T. (1985).

Identification of the Dadx Gene Coding for the Predominant Isozyme of Alanine

Racemase in Escherichia-Coli -K12. Molecular & General Genetics 198, 315-

322.

101. Wild, J. & Klopotowski, T. (1981). D-Amino-Acid Dehydrogenase of

Escherichia-Coli -K12 - Positive Selection of Mutants Defective in Enzyme-

Activity and Localization of the Structural Gene. Molecular & General Genetics

181, 373-378.

102. Fling, M., and Fox, S. W. (1945). Antipodal specificity in the inhibition of growth

of L. arabinosus by amino acids. J. Biol. Chem. 160, 329-336.

103. Yaw, K. E., and Kakavas, J. C. (1952). Studies on the effects of D-amino acids on

Brucella abortus. J. Bacteriol . 63, 263-268.

104. Hammes, W., Schle1fer, K. H. and Kandler, O. (1973). Mode of action of glycine

on the biosynthesis of peptidoglycan. J. Bacteriol . 116, 1029-1053.

105. Beardsley, R. (1962). Amino acid cross resistance in agrobacterium tumefaciens.

J. Bacteriol. 84, 1237-1240.

106. Ercal, N., Luo, X., Matthews, R. H. & Armstrong, D. W. (1996). In vitro study of

the metabolic effects of D-amino acids. Chirality 8, 24-29.

107. Daniello, A., Donofrio, G., Pischetola, M., Daniello, G., Vetere, A., Petrucelli, L.

& Fisher, G. H. (1993). Biological Role of D-Amino-Acid Oxidase and D-

Aspartate Oxidase - Effects of D-Amino Acids. Journal of Biological Chemistry

268, 26941-26949.

177

108. Calender, R., Berg, P. (1966). The catalytic properties of tyrosyl ribonucleic acid

synthetases from Escherichia coli and Bacillus subtilis . Biochemistry 5, 1690-

1695.

109. Calender, R., Berg, P. (1966). Purification and physical characterization of tyrosyl

ribonucleic acid synthetases from Escherichia coli and Bacillus subtilis .

Biochemistry 5, 1681-1690.

110. Soutourina, J., Plateau, P. & Blanquet, S. (2000). Metabolism of D-aminoacyl-

tRNAs in Escherichia coli and Saccharomyces cerevisiae cells. Journal of

Biological Chemistry 275, 32535-32542.

111. Soutourina, O., Soutourina, J., Blanquet, S. and Plateau, P. (2004). Formation of

D-tyrosyl-tRNATyr accounts for the toxicity of D-tyrosine toward Escherichia

coli . J. Biol. Chem. 279, 42560-42565.

112. Ferri-Fioni, M. L., Fromant, M., Bouin, A. P., Aubard, C., Lazennec, C., Plateau,

P. & Blanquet, S. (2006). Identification in archaea of a novel D-Tyr-tRNA(Tyr)

deacylase. Journal of Biological Chemistry 281, 27575-27585.

113. Dedkova, L. M., Fahmi, N. E., Golovine, S. Y. & Hecht, S. M. (2003). Enhanced

D-amino acid incorporation into protein by modified ribosomes. Journal of the

American Chemical Society 125, 6616-6617.

178