A ROLE FOR PROTEIN KINASE G IN FOLATE METABOLISM AND INTRACELLULAR

SURVIVAL IN MYCOBACTERIA

by

KERSTIN ANDREA WOLFF

Submitted in partial fulfillment of the requirements

for the degree of Doctor of Philosophy

Dissertation Adviser: Dr. Liem Nguyen

Department of Molecular Biology and Microbiology

CASE WESTERN RESERVE UNIVERSITY

January 2012

2

Table of Contents

List of Tables 8

List of Figures 9

Acknowledgements 12

List of Abbreviations 13

Abstract 16

Chapter 1: Introduction 18

1.A. and Other Mycobacterial Infections 19

1.A.1. Active and Latent M. tuberculosis Infections 20

1.A.2. Mycobacteria and Intrinsic Drug Resistance 23

1.A.3. MDR and XDR TB 27

1.B. Targeting Resistance Mechanisms to Potentiate Antibiotics 28

1.C. Folate Metabolism and Antifolate Antibiotics 33

1.C.1. Folate Antagonism in Chemotherapies and Antifolate

Resistance in Mycobacterial Infections 37

1.C.2. Potentiation of Antifolates in Mycobacteria 39

1.D. Biofilms 42

1.E. Protein kinase G (PknG) 44

1.F. MutT3 50

1.G. RplM 54

3

Chapter 2: Materials and Methods 57

2.A. Bacterial strains, chemicals, media and growth conditions 58

2.B. Cloning, construction of plasmids and construction of strains 58

2.B.1. Targeted deletion of pknGMsm and mutT3Msm and

T11A T11E replacement of rplMMsm with rplMMsm or rplMMsm

in M. smegmatis by modified recombineering method 59

2.B.2. Targeted deletion of mutT3Mtb and replacement of

T11A T11E rplMMtb with rplMMtb or rplMMtb in M. tuberculosis

by specialized transduction 62

2.B.3. Deletion of ntpA in E. coli by one-step phage λ-Red-based

recombineering method 64

2.B.4. Cloning for in trans expression of proteins 65

2.C. RT-PCR 77

2.D. SDS-PAGE, Coomassie Brilliant Blue staining and Western Blot 77

2.E. Antibiotic Susceptibility Testing 79

2.F. Cellular aggregation, Hydrophobicity Index and Congo red

binding assay 80

2.G. Zeta Potential 80

2.H. Biofilm Assays 80

2.I. Kinase Assays 81

2.J. Radioactive nucleotide binding assays 82

2.K. Macrophage infection, lysosomal trafficking and microcopy 82

4

Chapter 3: Protein Kinase G is Required for Intrinsic Antibiotic

Tolerance in Mycobacteria 84

3.A. Abstract 85

3.B. Introduction 86

3.C. Results 88

3.C.1. Identification of MAR4, an M. smegmatis pknG transposon

mutant displaying multidrug sensitivity 88

3.C.2. Targeted deletion of pknG and complementation

experiments confirm the MDR function of PknG in

pathogenic and non-pathogenic mycobacteria 89

3.C.3. Loss of pknG leads to altered cell surface properties 97

3.D. Discussion 101

3.E. Acknowledgements 103

Chapter 4: Modulation of Folate Metabolism by Protein Kinase G

Determines Biofilm Growth and Subcellular Localization of

Pathogenic Mycobacteria in Macrophages 104

4.A. Abstract 105

4.B. Introduction 106

4.C. Results 109

4.C.1. PknG is required for survival of folate starvation 109

4.C.2. Chemical inhibition of PknG activity potentiates classical

antifolate antibiotics 110

5

4.C.3. PknG is required for biofilm growth 114

4.C.4. PknG is involved in regulation of lysosomal delivery in

macrophages via modulation of mycobacterial folate

metabolism 119

4.D. Discussion 123

4.E. Acknowledgements 126

Chapter 5: Protein Kinase G Regulates Folate Metabolism

Through Phosphorylation of Ribosomal Protein RplM 127

5.A. Abstract 128

5.B. Introduction 129

5.C. Results 132

5.C.1. PknG does not mediate regulation of folate metabolism

through phosphorylation of GarA 132

5.C.2. mutT3 is required for survival of folate starvation and

biofilm growth 133

5.C.3. Preliminary characterization of MutT3 as a Nudix

Hydrolase 137

5.C.4. RplM, a novel substrate for PknG, co-purifies with MutT3 144

5.C.5. Phosphorylation of RplM by PknG on conserved

threonine 11 mediates regulation of folate metabolism 153

5.D. Discussion 158

5.E. Acknowledgements 163

6

Chapter 6: Discussion and Future Directions 164

6.A. Future Experiments 165

6.A.1. Hierarchy of PknG, RplM and MutT3 165

6.A.2. RplM point mutant alleles in M. tuberculosis and effect on

lysosomal trafficking 167

6.A.3. Subcellular localization of RplM/MutT3-containing

complexes 168

6.A.4. Biochemical characterization of MutT3 enzymatic

function 171

6.A.5. MutT3 deletion effect on lysosomal trafficking 173

6.A.6. Intracellular folate levels 174

6.B. Discussion 175

6.B.1. Folate significance during colonization of host

macrophages 175

6.B.1.a. Folate and cellular dormancy 176

6.B.1.b. Immune evasion and cell wall maintenance 178

6.B.2. Regulation of folate metabolism 182

6.B.3. Implications for TB and MDR or XDR TB 187

Bibliography 191

7

List of Tables

Table 1-1. Tuberculosis drugs 26

Table 2-1. Primers used in this study 67

Table 2-2. Plasmids and phasmids used in this study 70

Table 3-1. Susceptibility of M. smegmatis strains to antibiotics 91

Table 3-2. Surface properties of M. smegmatis strains 99

Table 5-1. MutT3 co-purified proteins identified by LC/MS/MS 148

8

List of Figures

Figure 1-1. A simple plateau model for persistent infection in

mice 22

Figure 1-2. Schematic structure of mycobacterial cell wall 24

Figure 1-3. Concept of drug potentiation by targeting resistance 30

Figure 1-4. Folate metabolism and antagonism in bacteria 34

Figure 1-5. Protein kinase G is involved in lysosomal delivery to

colonized phagosomes. 46

Figure 1-6. Chemical inhibition of PknG accelerates phagosome-

lysosome fusion 47

Figure 1-7. PknG-dependent trafficking of pathogenic

mycobacteria 49

Figure 1-8. Scheme of the 7,8 Dihydroneopterin Triphosphate

Pyrophosphatase (DNHTPase) Catalytic Reaction and

Effect of the ntpA Deletion on E. coli's Folate

Concentration 53

Figure 1-9. 3D-model of the E. coli 50S subunit showing the

position of each phosphorylated ribosomal protein. 56

Figure 3-1. Identification of MAR4, a M. smegmatis pknG

transposon mutant 90

Figure 3-2. Construction and complementation of the ΔpknGMsm

strain 95

9

Figure 3-3. Antibiotic susceptibilities of M. tuberculosis strains 96

Figure 3-4. Loss of PknG does not affect growth rate in liquid,

shaking medium 98

Figure 4-1. Requirement of PknGMtb activity for survival of folate

starvation and exploitation for antifolate potentiation. 111

Figure 4-2. Requirement of PknGMsm activity for survival of folate

starvation. 113

Figure 4-3. PknGMtb is required for folate-dependent biofilm

growth in M. tuberculosis. 115

Figure 4-4. PknGMsm is required for folate-dependent biofilm

growth in M. smegmatis. 117

Figure 4-5. PknG is involved in lysosomal delivery to infected

phagosomes through folate levels. 120

Figure 5-1. PknG does not mediate control over folate

metabolism through phosphorylation of GarA. 134

Figure 5-2. Folate de novo biosynthesis pathway. 136

Figure 5-3. mutT3Mtb is required for survival of folate starvation

and biofilm growth. 138

Figure 5-4. mutT3Msm is required for survival of folate starvation

and biofilm growth. 140

10

Figure 5-5. ntpA, a dihydroneopterin-triphosphate

pyrophosphatase, is required for survival of folate

starvation. 143

Figure 5-6. MutT3 binds nucleotides in vitro. 145

Figure 5-7. RplM, a novel substrate of PknG, co-purifies with

MutT3. 149

Figure 5-8. MutT3 co-purified with RplM and is not regulated

translationally. 151

Figure 5-9. RplM is phosphorylated on threonine 11 by PknG in

vitro. 154

Figure 5-10. Phosphorylation of RplMMsm on threonine 11 is

required for survival of folate starvation and biofilm

growth. 156

Figure 5-11. Model of folate biosynthesis regulation via PknG,

RplM and MutT3. 161

11

Acknowledgements

Thank you to my adviser, Dr. Liem Nguyen, for taking a risk with me as his

first graduate student, even though I was the only student in my year that couldn’t

apply for additional funding, for opening the world of genetics to me and for being a

mentor in more than just science.

Hoa Nguyen, who has been with me on the roller coaster that has been this

project from the start, and who has become one of my closest friends, thank you.

Marissa Sherman, who has only recently joined our group but has quickly

become an essential part of my day (and the greatest cat sitter ever!), thank you.

Past and present members of our lab have helped me in many ways, thank you Joe Timpona, Sam Ogwang, Rich Cartabuke, Ajay Singh, and Soumya Gogula.

My committee members have been invaluable in steering me along my

graduate journey; thank you for your advice, your patience and your foresight, Dr.

Arne Rietsch, Dr. W. Henry Boom, Dr. Jonathan Karn, and Dr. Patrick Viollier.

I owe more to my parents than I could express, for letting me cross an ocean, for spending their money on American university fees, and for supporting me in every way. Thank you a million times.

My parents-in-law and their family, thank you for welcoming me with open arms and making me feel at home in the U.S., with or without my gall bladder.

Finally to my wife, Niki Smith, without whom I am nothing, and who, conveniently also solves every Adobe® Illustrator® problem I create. Thank you for

always bearing with me. I love you.

12

List of Abbreviations

AIDS Acquired immune-deficiency syndrome

BCG Bacillus Calmette-Guérrin

BMDMs bone marrow derived macrophages

bp

5-CHO-H4PteGlu folinic acid

5-CH3-H4PteGlu 5-methyl-tetrahydrofolate

CFU colony forming unit

dGTP deoxy-guanosine-5-triphosphate

DHFR dihydrofolate reductase

DHPS dihydropteroate synthethase

DOTS daily observed treatment schedule

EMB ethambutol

ETH ethionamide

FA folinic acid

FPLC fast protein liquid chromatography

GAIT interferon-γ-activated inhibitor of translation

GDH glutamate dehydrogenase

GDP guanosine 5’-disphosphate

GDP-manp guanosine 5’-diphosphate-mannose-1-phosphate

GFP green fluorescent protein

13

GLU glutamate

GTP guoanosine 5’-triphosphate

H2Pte dihyrdopteroate

H2PteGlu dihydrofolate

H4PteGlu tetrahydrofolate

Hcy homocysteine

HIV human immunodeficiency virus

HPLC high performance liquid chromatography

INH isoniazid

KEGG Kyoto Encyclopedia of and Genomes

KGD alpha-ketoglutarate decarboxylase

LAM lipoarabinomannan

LB Luria-Bertani

LTBI latent tuberculosis infection

MAC Mycobacterium avium complex

MDR TB multiple drug resistant tuberculosis

mRNA messenger RNA

MS/MS tandem mass spectrometry

MTB Mycobacterium tuberculosis

MTHFS 5,10-methenyl-tetrahydrofolate synthase

MutT3 mutator protein T3

NTM Non-tuberculous mycobacterial infection

14

NtpA (dihydro)neopterin triphosphate pyrophosphohydrolase

OD# optical density at # nanometers

ODH 2-oxoglutarate dehydrogenase pABA para-aminobenzoic acid

PAS p-aminosalicylic acid pGp guanosine 5’,10’-bisphosphate

PI(3)P phosphoinositol-3-phosphate

PknG protein kinase G ppGpp guanosine 5’,10’-bispyrophosphate

RIF rifampicin

RplM ribosomal protein L13

SAM S-adenosine methionine

SHMT serine hydroxymethyltransferase

SDS-PAGE Sodium dodecyl sulfate polyacrylamide gel electrophoresis

SSDB Sequence Similarity Database

(RT)-PCR (reverse transcription)-polymerase chain reaction

TB tuberculosis

TBS Tris-buffered saline

WHO World Health Organization

XDR TB extensively drug resistant tuberculosis

15

A Role for Protein Kinase G in Folate Metabolism and Intracellular

Survival in Mycobacteria

Abstract

by

KERSTIN ANDREA WOLFF

The eukaryotic-like serine/threonine protein kinase G (PknG) of

Mycobacterium tuberculosis has been shown to be important for survival of the bacillus in host macrophages, presumably by preventing lysosomal delivery. It was therefore suggested that PknG serves as a virulence factor, although the underlying mechanism has remained uncharacterized.

Here we present evidence that PknG regulates de novo folate biosynthesis that is required for static growth conditions including those found in surface biofilms and in host macrophages. Deletion of pknG resulted in defects in folate synthesis, growth under static conditions, as well as resistance of mycobacterial

16

species to classical antifolates and other antibiotics. We have identified a novel substrate for PknG, the ribosomal protein RplM of the large ribosomal subunit. RplM is phosphorylated by PknG at a unique, mycobacterially conserved residue, threonine 11, and co-purified with MutT3, a putative Nudix that exhibits a dihydroneopterin-triphosphate pyrophosphatase activity, catalyzing the first committed step of de novo folate biosynthesis. We demonstrate that inhibition of

PknG has the potential to be exploited for potentiating the anti-mycobacterial activity of classical antifolate drugs and we present evidence that the previously demonstrated role of PknG in blocking phagolysosomal synthesis in infected host macrophages is mediated through its function in mycobacterial folate metabolism.

Our work suggests that folate antagonism could be used to sensitize pathogenic mycobacteria to available chemotherapeutic reagents, as well as the innate bactericidal activity of host phagocytic cells.

17

Chapter 1.

Introduction

18

1.A. Tuberculosis and Other Mycobacterial Infections

Mycobacterial infections are one of the leading causes of death in humans

through disease world-wide 1, encompassing infections such as tuberculosis (TB),

leprosy, Buruli ulcers, and opportunistic non-tuberculosis mycobacterial (NTM)

infections in immune-compromised individuals, especially patients with acquired

immune deficiency syndrome (AIDS) 2. These diseases are caused by Mycobacterium

tuberculosis (MTB), Mycobacterium leprae, Mycobacterium ulcerans, and environmental mycobacteria, such as Mycobacterium avium complex (MAC),

Mycobacterium kansasii, Mycobacterium chelonae and Mycobacterium fortuitum 2; 3, respectively. The World Health Organization (WHO) has estimated that one third of

the world’s population is currently infected with M. tuberculosis, the causative agent

of TB 1. Only about ten percent of those infected will develop active disease, unless

they are also afflicted with immune-compromising conditions such as AIDS that

highly increase the likelihood of active disease development 1. The lung typically is the affected organ, but the brain, kidneys, spine, and other body parts can also be infected. Mycobacterial loads in the affected organs and whether intracellular M. tuberculosis are dormant or actively growing depend on containment that must be successfully mediated by the host immune system (see section 1.A.1) 4; 5. This

process is severely impaired in immune-compromised individuals, and as a result there is greater opportunity for M. tuberculosis to transition from latency to active

TB in AIDS patients 4. Geographical areas with the highest incidence rates of TB are

located in sub-Saharan Africa and Southeast Asia, coinciding with human

19

immunodeficiency virus (HIV) hot spots, where up to 75 percent of HIV-infected

patients also are infected with M. tuberculosis 1. Accordingly, twenty-four percent of

all TB associated deaths occur in people co-infected with HIV 1.

1.A.1. Active and Latent TB Infections

Infections with TB are classified as either active or latent (LTBI) infections that differ significantly in infectiousness, symptoms and appropriate treatment regimens 6.

The vast majority of M. tuberculosis infections are LTBI cases, in which the

infected individual does not feel sick and does not exhibit any symptoms, since they do not develop active TB disease 6. Latent infection is diagnosed by either a positive

reaction to the tuberculin skin test or the QuantiFERON® blood test in combination with a negative chest X-ray. Patients with LTBI cannot spread M. tuberculosis to

others 6; 7. The daily observed treatment schedule (DOTS) for latent, drug susceptible TB infection consists of six to nine months of daily oral INH doses 8.

Latency in the bacillus is associated with reduced metabolic activity and slow or absent cell division and growth 4; 6; 11; 12.

About ten percent of latently infected persons will develop active TB at some time in their lives, with sharply increased likelihood of disease development by up to ten percent per year in immune-compromised patients, such as individuals co- infected with HIV 1; 6; 9. Fifty percent of individuals who develop active TB do so

within the first two years after initial infection 6. Symptoms of active TB disease

include weight loss, loss of appetite, fever and night sweats, fatigue and chills,

20

persistent coughing, hemoptysis, and chest pain 6. Active TB disease is diagnosed by chest X-ray and culturing of M. tuberculosis from patient sputum samples, along with positive tuberculin skin or QuantiFERON® blood tests. Patients with active TB

disease are infectious and spread M. tuberculosis by aerosolization when coughing 6.

The WHO estimates that a person with active TB who does not seek treatment will infect between ten and twelve people per year 1.

M. tuberculosis is internalized to phagosomes by host macrophages (see also section 1.E.), and like all intracellular pathogens must be able to survive and persist within infected host cells 10. The course of TB infections typically occurs in distinct phases 4; 5. The first phase following initial infection is a burst of active replication of

M. tuberculosis in the infected host tissue, followed by a phase of immune-mediated containment, in which a balance established between the host immune system and the pathogen leads to a constant, low bacterial load (Fig. 1-1) 4; 5. During this stage,

M. tuberculosis transitions from active growth to a dormant form, which is

metabolically distinct 4; 6; 11; 12. In this persistent state, M. tuberculosis becomes

highly tolerant to antibiotics compared to actively dividing cells 13; 14; 15; 16.

While the process by which M. tuberculosis establishes successful persistence is poorly understood, it has been shown that both host and mycobacterial factors are involved in phagosome maturation of infected host macrophages 17; 18; 19. More recently, mycobacterial protein phosphatases and kinases have been identified to perform key roles in establishing the survival niche in mature, but non-lysosomally fused phagosomes 19; 20; 21. One such kinase is the mycobacterial eukaryotic-like

21

Figure 1-1. A simple plateau model for persistent infection in mice. Low-dose infection of C57BL/6 mice with M. tuberculosis results in initial active replication in the lung (1st month), followed by immune-mediated containment, and a persistence phase that is characterized by a constant bacterial load. This provides a convenient model to study the molecular basis of mycobacterial and host responses to persistent infection. Parallel studies with human tissues are required to determine the extent to which persistence in mice resembles latent tuberculosis. CFU, colony- forming units.

Reproduced with permission from 4.

22

serine/threonine protein kinase G (PknG), which was suggested to support survival

of pathogenic mycobacteria in the host macrophage by preventing phagolysosomal fusion and therefore acidification of the mycobacterial compartment 20 (see also section 1.E.). Finally, in patients who develop active tuberculosis disease, immune-

mediated containment fails, often due to other factors weakening the immune

response, e.g. HIV infection, other diseases or aggressive immune-suppressing

therapy. As a consequence, the bacilli transition from the latent state to active

growth and expand in the affected tissue, causing symptoms and making spread to

other individuals possible 5. Interestingly, in yet another case of phenotypic

association between antibiotic tolerance and virulence 22; 23; 24, the transition from active division to the persistent state can be triggered by exposing M. tuberculosis to antibiotics 25.

1.A.2. Mycobacteria and Intrinsic Drug Resistance

Although mycobacteria are classified as Gram-positive bacteria, which do not have an outer membrane on top of their peptidoglycan layer, the mycobacterial cell wall is extremely thick and multi-layered, with varied hydrophobicity among the layers, and thereby poses an effective obstacle for the entry of most chemical compounds (Fig. 1-2) 3; 23; 26; 27. The peptidoglycan network is covered by an arabinogalactan layer, and both of these components are hydrophilic and likely limiting for the penetration of hydrophobic compounds (Fig. 1-2) 3. A layer of long chain fatty acids called mycolic acids of about seventy to ninety carbon units in length linked to acyl lipids overlays these two layers and forms a waxy, non-fluid

23

Figure 1-2. Schematic structure of mycobacterial cell wall. (a) Schematic

structure of mycobacterial cell wall depicted according to Minnikin's model 28.

Intercalation of hydrophilic arabinogalactan and hydrophobic mycolate containing layers creates an extremely impermeable envelope for antibiotic penetration. Small molecules and nutrients are transported through porin channels that are deposited through these layers. Modified from 29.

Reproduced with permission from 23.

24

barrier that is very efficient at restricting transport of both hydrophobic and

hydrophilic molecules (Fig. 1-2) 26; 30. As a result, mycobacteria are highly

problematic pathogens resistant to most of the typically used antibiotics and other

chemotherapeutic agents, even to extents greater than seen in Gram-negative

bacteria, simply by virtue of their complex cell envelope 3; 23. For example, the cell

wall permeability for hydrophilic compounds measured in Mycobacterium chelonae

was roughly three orders of magnitude lower than that of the Escherichia coli outer

membrane, and ten times lower than the permeability measured for the outer

membrane of Pseudomonas aeruginosa 31.

While permeability and slow influx across the mycobacterial cell wall is a necessary factor for the high intrinsic resistance of mycobacteria to most common antibiotics, it is not sufficient and usually must be supported by dedicated resistance mechanisms operating in the cytosol 3. Apart from several drug degradation and inactivation processes that occur inside mycobacteria, they also possess over 30 drug efflux transporters that have been shown to be involved in resistance to aminoglycosides, chloramphenicol, fluoroquinolones, isoniazid, linezolid, rifampicin, tetracycline and other toxic compounds 32; 33; 34; 35.

Cumulatively, the extremely high level of intrinsic resistance to most

antimicrobial drug classes exhibited by M. tuberculosis has left us with a severely

limited arsenal of useful anti-tuberculosis drugs (Table 1-1) 36. Complicating matters further, these antimycobacterial drugs must be used in regimens consisting of a minimum combination of four drugs to be effective for treatment of active

25

Table 1-1. Tuberculosis drugs.

Antibiotics Year of discovery

First-line antibiotics Isoniazid 1952 Rifampin 1966 Pyrazinamide 1952

Ethambutol 1961

Rifabutina 1980

Rifapentinea 1965

Second-line antibiotics Cycloserine 1952 Ethionamide 1956 P-aminosalicylic acid 1946

Streptomycin 1944 Amikacin 1972

Kanamycina 1957

Capreomycin 1963

Levofloxacina (Levaquin, DR3355, Daiichi) 1986

Moxifloxacina (Avelox, BAY 12-8039) 1996

Gatifloxacina (Tequin, AM-1155) 1992

Antibiotics Year of discovery

aNot approved by the U.S. Food and Drug Administration (FDA) for TB treatment 11.

Reproduced with permission from 36.

26

tuberculosis disease 36, making TB and NTM infections one of the most difficult

human disease groups that are caused by bacteria to treat at the current time.

1.A.3. MDR and XDR TB

The five first-line drugs used to treat M. tuberculosis infections, isoniazid

(INH), rifampicin (RIF), ethambutol (EMB), pyrazinamide, and streptomycin were

all discovered more than sixty years ago (Table 1-1) 23; 36. They have been used

regularly since then and therefore are encountered by M. tuberculosis strains

constantly 23; 36. Furthermore, the current standard regimen (DOTS) for TB is

comprised of six to nine months of daily antibiotic treatment with a combination of

four out of these five drugs 23; 36. These lengthy regimens often lead to poor patient

adherence and incomplete courses of treatment, resulting in frequent exposure of M.

tuberculosis to sub-optimal doses of drugs 23; 36. This provides ample opportunity for the bacillus to acquire additional resistance by amassing sequential mutations in drug-target encoding genes 23; 36; 37. Due to the rapid mutational frequency

characteristic of bacteria, we now face the increasingly wide spread of multiple drug

resistant (MDR) and extensively drug resistant (XDR) M. tuberculosis strains 23; 36; 37.

MDR TB strains exhibit acquired resistance to at least the two most potent first-line

drugs (RIF and INH) 1; 36; 37; 38. XDR TB strains are resistant to any fluoroquinolones and to at least one of the three injectable second-line drugs (capreomycin, kanamycin, and amikacin) (Table 1-1) in addition to RIF and INH resistance 1; 36; 37; 38.

Infections with such strains require even more prolonged and aggressive treatment

courses consisting of combinations of numerous second-line drugs that often exhibit

27

toxic side effects and are expensive to administer 36; 39. An example can be found in the South-African province KwaZulu-Natal, where extraordinarily high rates of

HIV/AIDS and MDR TB coincide. In this region poor patient compliance to treatments requiring between 18 and 24 months of hospitalization for drug administration has led to an outbreak that accounts for one sixth of the worldwide

XDR TB cases 40. Roughly eighty percent of these active XDR TB patients are also

infected with HIV 40. During the 2006 outbreak, the mean patient survival time from sputum collection amounted to only 16 days 40. Per month, about 30 new cases of active XDR TB are diagnosed in KwaZulu-Natal alone 40.

As the spread of these infections is diminishing our already limited arsenal of effective antibiotics even further, the situation is becoming more and more problematic, and we desperately require new, alternative drugs 37; 41; 42; 43. Indeed,

some XDR M. tuberculosis strains have become virtually untreatable with the

currently available anti-TB medicines 41; 42; 43.

1.B. Targeting Resistance Mechanisms to Potentiate Antibiotics

The current prevalence of drug-resistant mycobacterial strains has created a dire need for alternative TB therapies to use in patients infected with MDR or XDR

TB. However, development of completely novel anti-tubercular drugs is both time- intensive and costly; roughly twelve to fifteen years and US $500 million are required to facilitate development of a new drug from the laboratory stage to its distribution on the market 44. Despite massive investments of effort and capital into

28

this approach, only a few compounds have proceeded into pre-clinical or clinical

stages, and none have been newly approved for treatment.

An alternative approach to this pathway that may help relieve the current crisis is presented by the concept of “targeting resistance” 45. This drug potentiation

approach makes use of our knowledge of molecular resistance mechanisms in order

to (re)sensitize pathogenic bacteria to already available drugs 45. This is achieved by coadministration of existing drugs and inhibitors suppressing resistance mechanisms to these drugs, thereby allowing ineffective drugs to (re)gain their antimicrobial activity (Fig. 1-3) 45; 46; 47. Because we are currently using only a limited number of drugs due to existing resistance in M. tuberculosis, potentiation by targeting resistance may become an important trend in the new era of drug development for TB. Drug potentiation could be used to regain and extend current

TB drug efficacy, as well as to make available for use other approved drugs that are currently inactive against the bacillus 45. The extension of lifespan and utility of

valuable approved antibiotics of known pharmacology, toxicology, and treatment

schedule represents a unique advantage of the drug potentiation approach 45.

Recent findings indicate that several available drug classes encompassing

both TB and non-TB antibiotics are promising candidates for potentiation by

targeting resistance mechanisms 45. Efforts are underway exploring this approach using β-lactams, ethionamide, and classical antifolates 45.

29

Active Resistance ANTIBIOTIC Mechanism

Resistant Bacterium

ANTIBIOTIC

Susceptible Bacterium

POTENTIATOR Inactive Resistance Mechanism

Figure 1-3. Concept of drug potentiation by targeting resistance. An active

resistance mechanism allows survival of bacterial pathogens in the face of

antibiotics. A potentiator that inhibits the resistance mechanism would (re)sensitize

the bacteria to the antibiotic, thus enhancing antibacterial activity.

Reproduced with permission from 45.

30

β-lactams are the most widely used group of antibiotics today. They make up

a broad class of drugs including penicillin and penicillin derivatives, cephalosporins,

monobactams and carbapenems, that targets bacterial cell wall synthesis at the peptidoglycan layer 48; 49. Currently, β-lactams are the only successful clinical

example of potentiation through targeting resistance, but inhibitors of β-lactamases

have prolonged the life of β-lactams for more than thirty years and are commonly co-administered with β-lactams in clinical use 50. As mycobacterial susceptibility to

β-lactams in the absence of β-lactamase activity, in the case of M. tuberculosis this

activity is exhibited by BlaC, is well within clinically achievable levels 51; 52, effective chemical inactivation of β-lactamases should similarly increase β-lactam sensitivity in these bacteria. Indeed, recent in vitro studies have shown three FDA-approved inhibitors to effectively inhibit purified BlaC protein 53; 54. Further in vitro data show

that a combination of meropenem/clavulanate kills both aerobically and

anaerobically grown M. tuberculosis with great efficiency 55, is effective against

thirteen tested XDR M. tuberculosis strains 55, and significantly reduces

mycobacterial counts in mouse peritoneal macrophages infected with M. tuberculosis upon treatment within clinically achievable dose ranges 56; 57 . Another study demonstrated imipenem to double the survival rate of infected mice 58 and one human study using imipenem alone in MDR TB patients with poor predicted

outcomes achieved a 70% cure rate as well 58. Treatment of TB patients with an amoxicillin/clavulanate combination significantly reduced M. tuberculosis counts with early bactericidal activity comparable to patients treated with the frontline

31

drug INH in an independent study 59, and perhaps most significantly, an advanced

XDR TB patient was treated with meropenem/clavulanate in conjunction with other

drugs after all previous other treatments had failed and has since fully recovered 60.

Ethionamide (ETH) is one of the most commonly used drugs in the treatment of MDR TB, but ETH usage is complicated by its low therapeutic index 61; 62. This

index signifies a narrow concentration range in which ETH is therapeutically

effective yet safe to use without eliciting toxic side effects in the patient 61; 62; 63. The

minimal dose required for inhibition of M. tuberculosis growth causes adverse side

effects including hepatitis and gastrointestinal conditions 64. ETH is a structural thioamide analogue of INH that must be metabolically activated by EthA to form adducts in reaction with nicotinamide adenine dinucleotide (NAD) that target synthesis of mycolic acids, the major mycobacterial cell wall component, by inhibiting InhA 65; 66; 67; 68; 69; 70; 71; 72. EthA is transcriptionally repressed by EthR,

leading to lowered activation and efficacy of ETH, and therefore, targeting EthR by inhibitors to achieve derepression of EthA transcription should sensitize mycobacteria and potentiate the antimycobacterial activity of ETH 64; 70; 73; 74.

Hydrophobic ester compounds can bind the DNA-binding domain of EthR and prevent binding of EthR to the ethA operator 73; 74; 75; 76. When tested as potentiators in combination with ETH, these esters were shown to significantly increase the effects of ETH in in vitro cultures 73; 74; 75; 76. An intelligently designed ester based on a lead compound identified in a screen for EthR inhibitors, BDM31343, allows for twentyfold lower ETH doses to inhibit growth of M. tuberculosis in vitro and

32

threefold lower ETH doses to significantly reduce bacterial loads in lungs of infected

mice 64; 74. Potentiation of ETH by hydrophobic ester inhibitors of EthR could therefore not only re-sensitize XDR TB strains to ETH, but also allow for elimination or reduction of toxic side effects in MDR TB patients due to lower required doses 45.

Evidence obtained from numerous studies performed in vitro, in animals, and

in humans, has clearly illustrated that β-lactams potentiated by β-lactamase

inhibitors and ETH potentiated by EthR inhibitors could be used as an effective

addition to the treatment of drug resistant TB. This powerfully demonstrates the

potential value of potentiation by targeting resistance as a tool for relieving the need

for alternative options in TB treatment 45.

1.C. Folate Metabolism and Antifolate Antibiotics

One group of antibiotics that seems promising for potentiation in mycobacterial infections is the antifolates, which target folate biosynthesis and metabolism 45. The term folate is used to refer to a number of B vitamins of similar

but distinct chemistry, all of which are absolutely essential for the existence of

cellular organisms in all kingdoms of life 45. Most naturally occurring folates are

derivatives of tetrahydrofolate (H4PteGlu, Fig. 1-4), a reduced folate compound, but

the commonly used nutritional supplement vitamin B9 or folic acid is chemically

synthesized 45. While all folates consist of three molecular components, namely a

two-ring pteridine nucleus derived from guanosine-5’-triphosphate (GTP), a para-

aminobenzoic acid (pABA) group, and one or more glutamate residues attached via

33

Figure 1-4. Folate metabolism and antagonism in bacteria. (A) Simplified interconversions of folate derivatives in de novo folate synthesis and one-carbon metabolic network. DHFS, dihydrofolate synthase; Gly, glycine; Met, methionine; MT, methionine synthase; Pte, pteroate; Ser, serine. (B) Chemical structure of monoglutamylated tetrahydrofolate and its derivatives carrying C1 groups at various levels of oxidation attached to N-5 or/and N-10.

Reproduced with permission from 45.

34

amide linkages, they are varied in the nature of the one-carbon group attached to

45 the N-5 and/or N-10 positions of H4PteGlu (Fig. 1-4B) .

Folates are critically important for all cellular activities because H4PteGlu derivatives are required as co- in one-carbon unit transfer reactions (Fig. 1-

4A) 45. These reactions are required for the biosynthesis of purines, thymidines, glycine, panthotenate and methionine. Additionally, in bacteria, folates also act as co-enzymes for synthesis of formyl-methionyl-tRNA, critical for initiation of translation 77; 78; 79. Therefore, folate molecules are essential for the synthesis of the building blocks of macromolecules such as nucleic acids and proteins. Consequently, folate deficiency interferes with almost all processes in actively dividing cells and ultimately leads to cell death 45. Yet another effect of folate starvation manifests itself in interference with recycling of both homocysteine (Hcy, Fig. 1-4) and S- adenosine methionine (SAM) 45. This elevates cellular Hcy concentrations, a condition called homocysteinemia, and reduces cellular methylation activities 45. As a result, folates are of particular importance during periods of rapid cell division and growth 77; 78.

Folate metabolism is regarded as bipartite, with a clear division between de novo biosynthesis (upstream) and utilization (downstream) of folates (Fig. 1- 4A).

De novo folate biosynthesis is a process comprised of several steps: (i) synthesis and modification of the pteridine group from guanosine-5’-triphosphate, (ii) synthesis of pABA from chorismate, (iii) condensation of pteridine and pABA to form dihydropteroate (H2Pte), and (iv) glutamylation which adds one or more glutamate

35

45 groups to form dihydrofolate (H2PteGlu) that is reduced to form H4PteGlu . This de novo biosynthesis pathway is present in most bacteria and plants, but absent in many other species, including mammals 45. The downstream utilization of folate, also called one-carbon metabolism, involves co-enzyme action of the various active forms of H4PteGlu in distinct reactions donating or accepting one-carbon units for

the formation of purines, thymidines, glycine, panthotenate, methionine, and

formyl-methionyl tRNA, as mentioned above (Fig. 1-4) 45; 77; 78; 79.

It is not surprising that compounds as essential to the vast majority of cellular processes as the folates were soon recognized as promising candidates to target in chemotherapeutic treatments of many conditions and diseases and were one of the first antibiotics invented 45. Folate antagonism has been exploited

successfully for therapies of cancers, malaria, psoriasis, rheumatoid arthritis, graft- versus-host disease, and bacterial infections 80; 81; 82. Folate antagonists, called antifolates, were initially used extensively to treat a variety of infectious diseases, but their use has been scaled back since the 1960s 80; 83. This is due in part to emergence of resistant strains, to the cytotoxicity of antifolates that can cause side effects in the patient, and most importantly due to the introduction of new drugs more effective against certain pathogens and diseases 80; 83. Other infectious

diseases, for example urinary tract infections, pneumonia caused by Pneumocystis

jiroveci, and shigellosis, are still commonly treated with a combination therapy

consisting of trimethoprim and sulfonamide antifolates that makes use of the

synergy between the two drug classes 83; 84; 85. Furthermore, the same combination

36

regimen is also used effectively as a prophylactic treatment for recurring and drug-

resistant infections 83; 84; 85.

While there are a multitude of well-studied proteins contributing to the

processes of folate metabolism, currently used folate antagonists seem to

exclusively target biosynthesis and reduction of folate 86; 87; 88. Whereas trimethoprim and folate analogs such as methotrexate inhibit the reduction step through inhibition of dihydrofolate reductases (DHFR), sulfonamides and sulfone drugs are pABA analogs that outcompete pABA in the condensation with the pteridin group, catalyzed by dihydropteroate synthase (DHPS) (Fig. 1-4) 86; 87; 88. A factor that makes antifolates a particularly attractive group of antibiotics is the absence of these targeted enzymes acting in de novo folate biosynthesis in humans and other mammals that must rely on their diet for folate intake, making antifolates a group of drugs that will specifically affect the bacterial pathogens, but not infected host cells 86.

1.C.1. Folate Antagonism in Chemotherapies and Antifolate Resistance in

Mycobacterial Infections

Since both steps in de novo folate biosynthesis and downstream folate utilization have been targeted by successful antimicrobial and antineoplastic

antifolate drugs since the 1940s, acquired resistance to antifolates in pathogenic

bacteria is by now wide-spread 45. This resistance is typically acquired by amassing

mutations that alter either expression levels or protein structures of the targeted

enzymes 80; 83. DHFR can become resistant to trimethoprim through point mutations

37

of active-site residues, thus altering its affinity for trimethoprim 89; 90. While clinical

resistant strains frequently show a variety of mutations, mutations of residues

crucial for trimethoprim affinity are highly conserved among the isolates 89. For

example, a point mutation in the DHFR that changes a conserved Isoleucine

residue (Isoleucine 94 in M. tuberculosis DHFR) to Leucine confers 50-fold higher

trimethoprim resistance in Streptococcus pneumoniae 89. This same mutation is commonly found in DHFR from mammalian, parasitic and bacterial resistant isolates

89; 90. Likewise, single point mutations at the Serine 53 or Proline 55 residues within

DHPS are found in resistant isolates of M. leprae and render the enzyme resistant to

sulfonamides and sulfone drugs 91; 92. The two residues are located in the drug

binding region of M. tuberculosis DHPS and are highly conserved throughout bacteria and protozoa 91; 92. Since only these two enzymes are currently being

targeted, combined mutations in DHFR and DHPS encoding genes confer resistance

to all available antifolates 86; 87; 88. Most studies of trimethoprim and sulfonamide resistance thus far use bacteria distantly related to M. tuberculosis. Due to recent increases in the use of antifolates as prophylactic treatments in HIV patients 93, many of whom are co-infected with M. tuberculosis, the bacillus is bound to encounter antifolates regularly and therefore a better understanding of antifolate resistance mechanisms in M. tuberculosis is urgently needed 94.

Despite this lack of intricate knowledge concerning resistance mechanisms in

mycobacteria, antifolate drugs have been used in the treatment of mycobacterial

infections. For example, PAS (p-aminosalicylic acid) is currently used as a second-

38

line drug for TB 95 and the sulfone drug Dapsone has long been effective in

monodrug regimens for the treatment of leprosy 96. More recently, a study

suggested that the frontline TB-drug INH, which is not a classical antifolate

antibiotic, may also target folate metabolism through DHFR inhibition by its adducts

97. With the current desperate need for alternative treatment options of MDR and

XDR TB infections, interest in the potential use of combination regimens of classical antifolates against M. tuberculosis has been renewed 98; 99; 100. As a result, several

recent in vitro studies and a case report suggest antifolate combinations such as

those of co-trimoxazole (trimethoprim plus sulfamethoxazole) could be effective

against TB 98; 99; 100. M. tuberculosis clinical strains isolated from TB patients were

shown to be widely susceptible to clinically achievable concentrations of co-

trimoxazole 98, and even to sulfamethoxazole alone 99. However, since exposure of M.

tuberculosis to antifolates is likely to rise due to the World Health Organization’s call

for widespread use of co-trimoxazole in the prophylactic treatment of HIV-AIDS

patients to prevent opportunistic infections, selection for resistant strains may soon

occur and prematurely render this promising class of drugs useless 82; 93. It would

therefore be prudent to seek ways to potentiate classical antifolate antibiotics by

targeting resistance mechanisms, similar to the approach used for β-lactams, so that

M. tuberculosis strains that acquire antifolate resistance could be re-sensitized and

remain treatable by this group of drugs 45.

1.C.2. Potentiation of Antifolates in Mycobacteria

39

A method for boosting antifolate efficacy by utilizing combinations of drugs

that target individual steps in folate biosynthesis is already in place (see section

1.C.1). Trimethoprim is commonly coadministered with sulfonamides, for example

sulfamethoxazole in the co-trimoxazole combination, in order to achieve a

synergistic effect and increase efficacy 83 (Fig. 1-4). However, in many cases

including that of M. tuberculosis, the synergistic effect of trimethoprim on

sulfonamides remains questionable and has not been proven conclusively 98; 99; 101.

More problematically, bacterial strains have already acquired resistance mutations, making them untreatable with both trimethoprim and sulfonamides 86; 87; 88.

Therefore, novel potentiation approaches targeting resistance mechanisms not rooted in point mutations of DHFR and DHPS might be more effective in both potentiating available antifolates and preventing additional emergence of resistant strains.

A recent study aimed at targeting intrinsic antifolate resistance in mycobacteria has the potential to reveal valuable targets for such resistance- targeted potentiation approaches 102. To systematically identify novel antifolate

resistance determinants, a genetic screen was first employed using a saturated

transposon-insertion library of M. smegmatis 102. The mutants are tested for

increases in susceptibility to a number of classical folate antagonists to identify

genes that encode for antifolate tolerance contributing proteins 102. Hits are confirmed by chemical complementation using folate derivatives of both the de novo synthesis and the one-carbon interconversion network 102. This chemogenomic

40

profiling approach allows for identification of novel determinants previously

unknown to function in mycobacterial intrinsic antifolate resistance and identified fifty such components of the mycobacterial antifolate resistome (unpublished data)

102. One of these identified genes was later confirmed to encode for a protein that acts as a 5,10-methenyl-tetrahydrofolate synthase (MTHFS) homolog, and disruption of this gene resulted in up to 64-fold lower resistance to various of the tested trimethoprim/sulphonamide combinations 102. A series of genetic knockout

and complementation studies, including the human MTHFS allele, indicated that

MTHFS activity is required for mycobacterial intrinsic antifolate resistance, despite

there being no function for MTHFS in the typically targeted pathways of folate de

novo biosynthesis or folate reduction 102. The role of MTHFS in intrinsic antifolate

resistance was confirmed in M. tuberculosis and Escherichia coli 102; 103, suggesting

that this determinant functions ubiquitously among bacteria. Pharmaceutical

inactivation of MTHFS activity should therefore result in sensitization of M.

tuberculosis to classical antifolates, potentially including current anti-TB drugs that

target folate metabolism (PAS, INH, etc.). Potentiation of the anti-TB efficacy of

classical antifolates may also allow for reduction of effective therapeutic doses,

thereby minimizing their cytotoxicity that has been a clinical problem.

Considering these promising reports concerning the potential of making

classical antifolate antibiotics a useful, readily available alternative to the aggressive

and often not effective treatments required in MDR and XDR TB infections 36; 39, fundamental studies of molecular mechanisms conferring both acquired and

41

intrinsic antifolate resistance in M. tuberculosis and related mycobacteria need to be conducted 45. The knowledge obtained from these studies will be essential for

strategic implementations of antifolate use for TB, and will reveal valid targets for

the resistance-targeted potentiation of classical antifolates 45.

1.D. Biofilms

Biofilms are surface-bound communities of microbial cells that can consist of

one or multiple bacterial or fungal species and occur both in natural environments,

such as lake water, and in clinical settings, for example on catheters in patients or on

the surface of implants 104; 105; 106; 107. Bacteria growing in biofilms display enhanced

tolerance for antibiotics, a phenomenon that can be ascribed to the fact that biofilm

growth is metabolically distinct from planktonic growth, resembling static-phase or

persistent growth stages 104; 105; 106; 107. In accordance with the idea of separate metabolic requirements, there is evidence that the stringent response, a stress response to amino acid starvation that is triggered by exposure to antifolates, affects biofilm formation 108; 109.

Biofilm development is typically modeled in four distinct phases, consisting of (a)

attachment of planktonically growing cells to a surface, (b) secretion of polysaccharides,

protein and extracellular DNA resulting from autolysis that then form an extracellular

matrix in which cells aggregate, (c) biofilm maturation, and (d) biofilm dispersal 104.

While the most extensively studied and best understood bacterial biofilm

models pertain to Escherichia coli, Vibrio cholerae, Bacillus subtilis, Pseudomonas

aeruginosa and Staphylococcus epidermis 104; 110, mycobacteria, including M.

42

tuberculosis, have been shown to form biofilms both in vivo and in vitro 111; 112; 113; 114.

Mycobacteria differ from most other biofilm-forming bacterial species in that they

do not externalize polysaccharides to form an extracellular matrix, but rather

aggregate through short-chain and free mycolic acids of their waxy cell envelope 112;

113. In other aspects however, mycobacterial biofilms behave just like those of other

pathogenic bacteria and have been shown to play a role in the pathogenesis of

Mycobacterium chelonae and Mycobacterium ulcerans 114; 115; 116; 117. Mycobacterial

biofilms also exhibit higher antibiotic tolerance than planktonic mycobacteria,

which is consistent with the idea that biofilm growth may represent a late static

growth stage similar to persistence, as M. tuberculosis in its persistent state during

latent infection also exhibits higher levels of resistance than actively growing M.

tuberculosis 22. Conversely, exposure to antibiotics is one of the factors inducing the

transition from active growth to persistence 22, and indeed a recent study in M. tuberculosis has revealed a correlation of biofilm growth and the development of persisters with increased antibiotic tolerance that provide an important reservoir of cells that can repopulate colonized sites upon removal of drug treatment 108; 113; 118.

The details of mycobacterial biofilms are not well understood, but it has been

shown that alterations in cell wall characteristics and composition, as well as in

sliding motility, result in defects in biofilm formation in both M. tuberculosis and M.

smegmatis 113; 119; 120. Furthermore it has been demonstrated that M. tuberculosis

requires OdhA, the E1 subunit of the 2-oxoglutarate dehydrogenase (ODH) complex,

also known as the α-ketoglutarate dehydrogenase (KGD) complex, an enzyme

43

complex of the citric acid cycle, for biofilm formation 113. Interestingly, the enzymatic activity of ODH is controlled by the phosphorylation state of OdhI and its homolog GarA in Corynebacterium glutamicum and M. tuberculosis, respectively 121;

122; 123. OdhI/GarA can be phosphorylated by the eukaryotic-like serine/threonine

kinases PknG and PknB in both M. tuberculosis and in the related actinobacterium

Corynebacterium glutamicum 121; 122; 124. Deletion of Rv2454c and Rv2455c, the two

genes encoding for the subunits of ODH complex, leads to defects in biofilm

formation, which could be bypassed by exogenous supplementation of CO2 113.

1.E. Protein kinase G (PknG)

Protein kinases can be sub-divided into 3 groups defined by the residue to which the phosphate group is transferred. Histidine kinases typically are part of two-component signaling systems, where they first autophosphorylate on a histidine residue upon reception of a stimulus, and then proceed to transfer the phosphate group onto the aspartic acid residue of a response regulator 125; 126. These two-component systems are mainly found in bacteria and eukaryotes outside the

animal kingdom 125; 126; 127. Tyrosine kinases have evolved separately and use

distinct mechanisms of phosphorylation in bacteria and eukaryotes 128. Third,

serine/threonine kinases exist both in eukaryotes 129 and in bacteria, where they are

called eukaryote-like or eukaryotic-type serine/threonine kinases 130. While their importance in signaling has been extensively studied for many years in eukaryotes

131, our understanding of their roles in bacteria has only recently begun to expand.

Eukaryotic-type serine/threonine kinases have been found to play roles in

44

pathogenicity, stress responses and developmental processes in various bacteria 130;

132.

M. tuberculosis contains 11 putative eukaryotic-type serine/threonine

protein kinase-encoding genes in its genome 133; 134. For many of these kinases either function or substrates remain unknown 10, but it was soon found that kinase

inhibitors prevented uptake by and infection of murine macrophages by

Mycobacterium leprae, indicating crucial roles in pathogenicity 135.

While the vast majority of mycobacterial serine/threonine kinases exhibit typical receptor-kinase structures with a trans-membrane domain and an intracellular kinase domain, two kinases, PknG and PknK, appear to be soluble rather than membrane-associated 132; 133; 134.

PknG is conserved throughout the mycobacterial genus, including M. leprae,

which indicates that it plays an important role in bacterial survival in the host cell,

since M. leprae is dependent on its host and therefore has reduced its genome to a

minimum set of genes that are absolutely required for intracellular survival 136. It

has been shown that PknG plays a critical role in M. tuberculosis’ ability to establish

the phagolysosomal fusion block inside host macrophages, although its mechanism

of action and targets involved in this process remain unknown 20; 137. Both genetic deletion of pknG and chemical inhibition of PknG by a specific inhibitor 20; 138 in M.

bovis BCG resulted in increased fusion of lysosomes to phagosomes containing

mycobacteria, correlating with decreased survival of internalized mycobacteria (Fig.

1-5, 1-6) 20; 23. It was therefore suggested that PknG may act as a virulence factor

45

Figure 1-5. Protein kinase G is involved in lysosomal delivery to colonized phagosomes. Macrophages were infected with M. bovis BCG or M. bovis BCG/ΔpknG

expressing green fluorescent protein (GFP) for 1 hour, followed by a 2-hour chase.

Cells were processed as described in supplementary materials online and stained

with antibody to LAMP followed by Alexa Fluor 568 (LAMP-568). Scale bar, 10 μm.

Reproduced with permission from 20.

46

Figure 1-6. Chemical inhibition of PknG accelerates phagosome-lysosome fusion. Macrophages were infected with M. bovis BCG for 1 hour, followed by a 2- hour chase, in the absence (left) or presence (right) of AX20017 (20 μM), fixed, permeabilized, and stained with antibody to LAMP followed by Alexa Fluor 568.

Scale bar, 10 μm.

Reproduced with permission from 20. 47

that is secreted into the host cytosol and actively modifies host molecules to

establish the phagolysosomal fusion block 20; 23; 139. It has also been suggested that

this virulence function may have been acquired in pathogenic mycobacteria that

regulate PknG expression translationally to levels dramatically elevated over those

seen in non-pathogenic mycobacterial species, and that virulence function depends

upon expression levels 140.

Additionally, recent findings confirmed that, as had been shown for

Corynebacterium 121; 123, PknG is involved in the tricarboxylic acid cycle as well as glutamine utilization in Mycobacterium, where it phosphorylates GarA on threonine

21, directly adjacent to threonine 22, which is phosphorylated by PknB 122; 124; 141.

The two phosphorylation events are mutually exclusive 121; 122; 124; 141. GarA in its unphosphorylated state inhibits action of ODH complex and, depending on substrate concentration, either inhibits or activates glutamate dehydrogenase (GDH) 122. ODH converts 2-oxoglutarate to succinate, generating CO2, which has been shown to play a role in biofilm formation in mycobacteria 113. GDH converts 2-oxoglutarate to glutamate. Glutamine and its derived molecules, such as polyglutamines, may serve as components that are required for the integrity of the mycobacterial cell wall, thereby helping to pose an effective barrier to diffusion of antibiotics of diverse size and polarity, and are also important for static-phase growth 122; 132; 141.

PknG can autophosphorylate on four N-terminally located threonine residues, threonines 23, 32, 63 and 64, which are distinct from the typically

48

Figure 1-7. PknG-dependent trafficking of pathogenic mycobacteria. (a) Model of PknG-mediated phagosome-lysosome fusion blockage by pathogenic mycobacteria. PknG (purple) is secreted by pathogenic mycobacteria into the phagosomal lumen and cytosol of macrophages to block the fusion of phagosomes to lysosomes. Deletion of the pknG gene leads to rapid lysosomal transfer and destruction of the resulting mutant.

Reproduced with permission from 23.

49

autophosphorylated residues located in a conserved activation loop in other

mycobacterial serine/threonine kinases 142. Consistent with this observation,

autophosphorylation of PknG is not required for activation of its kinase activity,

setting it further apart from the trans-membrane domain containing kinases in

mycobacteria 142. Strikingly, autophosphorylation is essential for establishing the phagolysosomal fusion block in infected host macrophages 142.

Cumulatively, while no exclusive substrate for PknG has been identified, it is evident that PknG is required for the ability of M. tuberculosis to prevent lysosomal fusion to the colonized phagosome 20 (Fig. 1-7) 23, a function that is dependent on

autophosphorylation of PknG 142 and potentially based on expression levels 140.

Furthermore, PknG may influence survival inside the host through regulation of glutamine and glutamate levels, which are crucial for static-phase growth in the persistent state 121; 122; 124; 141.

1.F. MutT3

MutT motif or Nudix hydrolases are pyrophosphatases that catalyze the hydrolysis of a nucleoside diphosphate linked to a variable moiety x 143;

144. They are characterized by the presence of about 10 conserved amino acids within a 20 amino acid sequence and occur throughout all classes of life 143; 144.

Substrates however are very diverse in nature and include nucleoside triphosphates, dinucleoside polyphosphates, nucleotide sugars, and co-enzymes 143; 144.

Furthermore, substrate specificity typically is determined by a preference for

certain substrates, but most members of the Nudix family are promiscuous and can

50

hydrolyze a variety of substrates 143; 144. It follows that these enzymes function in a number of metabolic pathways as diverse as the Nudix substrates 143; 144. For

example, the prototype Nudix , E. coli MutT is mainly involved in hydrolysis of 8-oxoguanine, a potentially mutagenic form of deoxy-guanosine-5’- triphosphate (dGTP), and therefore helps to prevent mutations in newly synthesized

DNA 143; 144; 145. Most Nudix hydrolases seem to be involved in the clearance of

potentially deleterious or hazardous molecules or in balancing metabolites and metabolic intermediates, preventing their build-up and acting as “housekeeping” enzymes 143; 144; 146. Recently, however, it has been shown that some Nudix hydrolases serve as enzymes in intermediate steps of biosynthetic pathways 146; 147;

148 and as regulators of cellular growth in response to specific environmental stresses 149.

In particular, the Ndx8 Nudix hydrolase of Thermus thermophilus has been shown to be involved in regulation of guanosine 3',5'-bispyrophosphate (ppGpp) levels independently of the RelA/SpoT synthesis and degradation cycles that tightly regulate this bacterial alarmone 149. ppGpp is synthesized by RelA in response to amino acid starvation and initiates a growth arrest to ensure cell survival until sufficient amino acid levels are encountered or re-established, a process called the stringent response 150; 151. Once the triggers for this response are removed, SpoT

degrades ppGpp to guanosine-5-diphosphate (GDP) and the bacterial cell can re-

enter the growth phase 151; 152. Ndx8 on the other hand degrades ppGpp to 3’,5’-

bisphosphate (pGp), and may act in a relay system with SpoT, thereby helping to

51

control the transition from growth arrest to active growth after amino acid

starvation conditions have been removed 149. Interestingly, one specific trigger of amino acid starvation that induces the ppGpp-mediated stringent response is the action of classical antifolate antibiotics 153.

Three Nudix enzymes from E. coli (NtpA), Lactococcus lactis (YlgG) and

Arabidopsis (AtNUDT1) have been shown to catalyze the removal of pyrophosphate

from dihydroneopterin-triphosphate (DHN-3P) to create dihydroneopterin (DHN)

in an Mg2+-dependent manner 147; 148. DHN-3P is one of the earliest intermediates in folate de novo biosynthesis, and its conversion to DHN, which is eventually turned into folate, constitutes the first committed step in this process (Fig. 1-8) 86; 147; 148.

For over fifty years, this reaction had previously been thought to occur by

spontaneous dephosphorylation 148; 154.

While no homologous DHN-3Pase enzyme catalyzing this reaction has been so far identified in M. tuberculosis, SSDB (Sequence Similarity Database) gene cluster search at the Kyoto Encyclopedia of Genes and Genomes (KEGG:

http://ssdb.genome.jp/ssdb-bin/ssdb_gclust?org_gene=mtu:Rv0410c&type=g; Date

of access: September 02, 2011) suggested the putative Nudix enzyme (MutT3) as a

likely candidate DHN-3Pase. The amino acid sequences of MutT3 from M. smegmatis (MutT3Msm) and M. tuberculosis (MutT3Mtb) show 78% homology

(http://tuberculist.epfl.ch/blast_output/Rv0413.fasta.out; Date of access:

September 02, 2011), and no significant homology with the prototypic E. coli MutT3,

52

Figure 1-8. Scheme of the 7,8 Dihydroneopterin Triphosphate

Pyrophosphatase (DNHTPase) Catalytic Reaction and Effect of the ntpA

Deletion on E. coli's Folate Concentration. (Left) Reaction catalyzed by the

product of the gene ntpA, a DHNTPase. (Right) Folic acid concentration of parental

cells and cells lacking DHNTPase activity. Each bar represents the mean of four separate cultures, two of which were grown in M9 minimal medium and two in

Luria Broth. Cells were harvested and assayed microbiologically using Lactobacillus casei (ATTC 7469) 155. Bar A: parent cells containing plasmid without insert

(MG1655 harboring pTrc 99A); bar B: parent cells harboring plasmid with ntpA

(MG1655-pTrc 99A-NtpA); bar C: ntpA deletion mutant harboring plasmid without

insert (MG1655ntpA− − pTrc 99A); bar D: ntpA deletion mutant harboring plasmid with ntpA insert (MG1655ntpA− − pTrc 99A-NtpA).

Reproduced with permission from 147. 53

but in a 33 amino acid region, 8 residues that define the Nudix motif 143; 144 are

perfectly conserved 156. Mycobacterial MutT3 does not appear to function as either a mutator protein that hydrolyzes 8-oxo-guanine and dGTP, like MutT1 and MutT4, or as an 8-oxo-guanosine triphosphatase like MutT1 and MutT2 156. Strikingly, mutT3 is

located three genes upstream of pknG, with highly conserved synteny of the locus

across the mycobacterial genus and even some related actinobacteria. Proximity of

gene location often indicates similar function or involvement in the same pathways,

suggesting that MutT3 and PknG may act in the same pathway or regulate each

other.

1.G. RplM

The ribosomal protein RplM, also known as L13, is encoded by rplM, which is

part of a single transcriptional unit together with rpsI, a gene encoding ribosomal

protein RpsI or S9 157; 158; 159. RplM is an essential protein and constitutes a component of the large ribosomal subunit 157, but can also act in sub-ribosomal particles 160 or outside the ribosomal context 161. On the large ribosomal subunit, the

C-terminus of RplM is located near the factor between ribosomal

proteins L3 and L6, towards the back of the 50S subunit, while the N-terminal end

interacts with ribosomal proteins L20 and L21 (Fig. 1-9) 162; 163. Interestingly, RplM

has been found to directly interact with Obg, an essential GTP-binding protein that

is required for regulation of the σB-mediated stress response in B. subtilis 164; 165.

Obg co-fractionated with ribosomal particles and stress response regulators, indicating a role for ribosomal function mediated by RplM interaction in induction

54

of the σB-mediated stress response in B. subtilis 165. While interactions of mycobacterial Obg with RplM have not been assayed, Obg does associate with fully assembled ribosomes as well as both ribosomal subunits in M. tuberculosis 166. RplM

has also been detected in active sub-ribosomal particles that had bound both 23S

and 5S rRNA in Thermus aquaticus 160. In a non-ribosomal setting, RplM has been

demonstrated to be involved in anti-termination together with ribosomal proteins

of the small subunit, S4, and large subunit, L3 and L4, in E. coli 161. As part of anti-

terminating transcription complexes, RplM was shown to be responsible for an 11-

fold increase in terminator read-through in in vitro transcription reactions 161.

Comprehensive analysis of phosphorylated proteins in ribosomes of E. coli

identified RplM to be phosphorylated on 5 residues, tyrosines 16, 44 and 53 and

threonines 45 and 50 167. The protein kinases responsible for this phosphorylation remain unknown, as does function of phosphorylation of these residues. However, almost all of the phosphorylated residues are located on the solvent interface of RplM, suggesting that they may be important for interactions with ribosome-associated proteins (Fig1-9) 167.

55

Figure 1-9. 3D-model of the E. coli 50S subunit showing the position of each phosphorylated ribosomal protein. The phosphorylated ribosomal proteins in the

50S subunit are represented by green color, while the phosphorylated residues mapped by mass spectrometry and unphosphorylated proteins are colored red and pink, respectively. Coordinates of the E. coli 50S subunit were obtained from the

Protein Data Bank (Acc. # 2AW4). (A) Depicting the 50S subunit from the solvent side, (B) gives a view of the large subunit from the 30S interface, and (C) is the side view of the large subunit showing the phosphorylated proteins around the SRL.

Specific regions, peptidyl center (PTC), central protuberance (CP), sarcin-ricin loop (SRL), L1 and L7/L12 stalks and the 50S ribosomal proteins were labeled in the model generated by PyMol software (DeLano Scientific LLC). Red arrow indicates position of L13 on the 50S subunit.

Reproduced with permission from 167. 56

Chapter 2.

Materials and Methods

57

2.A. Bacterial strains, chemicals, media and growth conditions.

Depending on the experiments, M. smegmatis strains were grown using 7H9

(Difco), biofilm 112 liquid medium, 7H10 (Difco), NE 168, or Luria-Bertani (LB) agar medium supplemented with 0.5% Tween 80 at 37˚C. M. tuberculosis strains were grown using 7H9 (Difco), Sauton’s medium without Tween 80 113, or 7H10 (Difco)

agar medium supplemented with 0.05% Tween 80 and OADC (BD) at 37˚C. All E. coli strains were grown in LB medium or on LB agar medium at 37˚C. Kanamycin was used at 50µg/ml; hygromycin at 75µg/ml for mycobacteria and at 100µg/ml for E. coli; and ampicillin was used at 100µg/ml for E. coli.

Media and supplements were obtained from Sigma (St. Louis, MI) unless otherwise indicated, AX20017 was synthesized by GLSynthesis Inc (Worcester, MA), chelerythrine was obtained from Sigma (St. Louis, MI), [γ-32P]-ATP was from

PerkinElmer (Waltham, MA), and Guanosine 5’-triphosphate-[8-3H] ([8-3H]-GTP)

was from MP Biomedicals (Solon, OH).

2.B. Cloning, construction of vectors and construction of strains.

All primers used in this study and their sequences are listed in table 2-1.

Primers were synthesized by Fisher Scientific (Pittsburgh, PA) or Eurofins MWG

Operon (Huntsville, Alabama). All PCRs were performed using the Expand Long

Template PCR kit (Roche Molecular Biochemicals). Sequences were verified by

cloning PCR products into pGEM-T Easy vector (Promega, Madison, WI) and

sequencing (Biotic Solutions, New York, NY or ACGT Inc., Wheeling, IL). Plasmids

used in this study are listed in table 2-3. Restriction endonucleases and T4

58

were obtained from New England Biolabs (Beverly, MA). Preparation of competent cells and transformation were performed as previously described 169.

2.B.1. Targeted deletion of pknGMsm and mutT3Msm and replacement of rplMMsm

T11A T11E with rplMMsm or rplMMsm in M. smegmatis by modified recombineering

method.

Targeted gene deletion or replacement in M. smegmatis was performed by

recombineering method 170 with modifications 171.

The temperature-sensitive plasmid pVN701B, which expresses

mycobacteriophage Che9c recombination proteins 170, kanamycin resistance genes

and the counter-selectable sacB marker gene 172; 173, was constructed by removing

the SpeI-XbaI fragment of pJV53 170 and inserting it into the XbaI site of pPR27 172.

pknGMsm flanking regions were amplified using primers MS-PknG-del1 and

MS-PknG-del2 (Table 2-1) for the 5’-pknG-flanking region (551bp) and primers MS-

MutT3del3 and MS-MutT3del4 (Table 2-1) for the 3’-pknG-flanking region (528bp), respectively. pVN740 was constructed by cloning the 5’-pknGMsm-flanking region

directionally into pYUB854 adjacent to the hygromycin-resistance gene using

SpeI/HindIII sites and cloning the 3’-pknGMsm-flanking region directionally into pYUB854 adjacent to the hygromycin-resistance gene using XbaI/KpnI, respectively.

Wild-type M. smegmatis mc2155 carrying plasmid pVN701B were induced to

express the recombineering system from the vector and transformed with the linear

pknG deletion cassette fragment removed from pVN740 by digestion with SpeI/KpnI.

Selection for replacement of chromosomal DNA with the Ωhyg-containing

59

deletion/replacement cassette by homologous recombination was performed on

7H10 agar medium containing hygromycin and kanamycin. Removal of pVN701B

was performed by shifting cultures of ∆pknGMsm to 39ºC in 7H9 medium containing

only hygromycin and confirming loss of kanamycin resistance as well as gain of

ability to grow on 7H10 agar medium supplemented with 1% (wt/vol) sucrose.

pknGMsm-Ωhyg junctions were amplified using primers P1 + P2 (Table 2-1) for the 5’-

pknGMsm-Ωhyg junction and primers P3 + P4 (Table 2-1) for the 3’-pknGMsm-Ωhyg junction, respectively.

mutT3Msm flanking regions were amplified using primers MS-MutT3del1 and

MS-MutT3del2 (Table 2-1) for the 5’-mutT3-flanking region (606bp) and primers

MS-MutT3del3 and MS-MutT3del4 (Table 2-1) for the 3’-mutT3-flanking region

(500bp), respectively. pVN755 was constructed by cloning the 5’-mutT3Msm-flanking region or 5’-PrplM-flanking region directionally into pYUB854 adjacent to the hygromycin-resistance gene using SpeI/HindIII sites and cloning the 3’-mutT3Msm- flanking region directionally into pYUB854 adjacent to the hygromycin-resistance gene using XbaI/KpnI, respectively. Wild-type M. smegmatis mc2155 carrying

plasmid pVN701B were induced to express the recombineering system from the

vector and transformed with the linear pknG deletion, mutT3 deletion or rplM

replacement cassette fragment removed from pVN755 by digestion with SpeI/KpnI.

Selection for replacement of chromosomal DNA with the Ωhyg-containing

deletion/replacement cassette by homologous recombination was performed on

7H10 agar medium containing hygromycin and kanamycin. Removal of pVN701B

60

was performed by shifting cultures of ∆mutT3Msm to 39ºC in 7H9 medium containing

only hygromycin and confirming loss of kanamycin resistance as well as gain of

ability to grow on 7H10 agar medium supplemented with 1% (wt/vol) sucrose.

mutT3Msm-Ωhyg junctions were amplified using primers MMsm1 + P2 (Table 2-1) for the 5’-mutT3Msm-Ωhyg junction and primers P3 + MMsm2 (Table 2-1) for the 3’- mutT3Msm-Ωhyg junction, respectively.

The 872bp-region of the upstream of the rplMMsm

promotor area was amplified using primers MS-RplMrep1 and MS-RplMrep2 (Table

T11A T11E 3-1), and mutant alleles (715bp) of rplMMsm encoding RplMMsm or RplMMsm

were generated by a two-stage PCR procedure. Primers MS-RplMT11Arev and MS-

RplMT11Afwd (Table 2-1) were used together with primers MS-RplMrep3 and MS-

RplMrep4 (Table 2-1), respectively, and PCR products let anneal. Primers MS-

RplMrep3 and MS-RplMrep4 (Table 2-1) were then added to amplify the mutant

T11A T11E rplM or rplM alleles after the native promotor region (PrplM). pVN895 and pVN896 were constructed by cloning the 5’-PrplM-flanking region directionally into pYUB854 adjacent to the hygromycin-resistance gene using SpeI/HindIII sites and

T11A T11E cloning the PrplM- rplM region or the PrplM- rplM region directionally into pYUB854 adjacent to the hygromycin-resistance gene using XbaI/BspHI (pVN895 and pVN896). Wild-type M. smegmatis mc2155 carrying plasmid pVN701B were

induced to express the recombineering system from the vector and transformed

with the linear rplM replacement cassette fragment removed from pVN895 or

pVN896 by digestion with SpeI/BspH1. Selection for replacement of chromosomal

61

DNA with the Ωhyg-containing deletion/replacement cassette by homologous

recombination was performed on 7H10 agar medium containing hygromycin and

kanamycin. Removal of pVN701B was performed by shifting cultures of

mc2155/rplMT11A and mc2155/rplMT11E to 39ºC in 7H9 medium containing only

hygromycin and confirming loss of kanamycin resistance as well as gain of ability to

grow on 7H10 agar medium supplemented with 1% (wt/vol) sucrose. Insertion of

Ωhyg was confirmed by PCR using primers RMsm1 and RMsm2 (Table 2-1) to observe according molecular weight shift. Primer MS-RplMrep3 (Table 2-1) was used to confirm sequence had changed to mutant allele.

2.B.2. Targeted deletion of mutT3Mtb and replacement of rplMMtb with

T11A T11E rplMMtb or rplMMtb in M. tuberculosis by specialized transduction.

Targeted gene deletion or replacement in M. tuberculosis H37Rv was performed by specialized transduction 174.

mutT3Mtb flanking regions were amplified using primers TB-MutT3del1 and

TB-MutT3del2 (Table 2-1) for the 5’-mutT3Mtb-flanking region (502bp) and primers

TB-MutT3del3 and TB-MutT3del4 (Table 2-1) for the 3’-mutT3Mtb-flanking region

(570bp), respectively. pVN755 was constructed by cloning the 5’-mutT3Mtb-flanking directionally into pYUB854 adjacent to the hygromycin-resistance gene using

SpeI/HindIII sites and cloning the 3’-mutT3Mtb-flanking region directionally into pYUB854 adjacent to the hygromycin-resistance gene using XbaI/KpnI, respectively.

After PacI-l digestion to linearize the recombinant cosmids, they were ligated to shuttle phasmid phAE87 digested with PacI, and the ligation packaged with the

62

GIGAPackIII GOLD system (Stratagene). Phages containing the recombinant

phasmid ph∆mutT3Mtb were used to transduce E. coli NM554 and transductants were selected on hygromycin. Phasmid DNA was extracted from pooled transductants and used to transform M. smegmatis mc2155 by electroporation.

Plaques obtained at 30ºC were isolated and used to transduce M. tuberculosis H37Rv and CDC1551. Transductants were selected on hygromycin and gene deletion/replacement confirmed by PCR using primers MMtb1 + MMtb2 (Table 2-1) to observe according molecular weight shift. Strains were designated ∆mutT3Mtb.

The region of the chromosome upstream of the rplMMtb promotor area

(716bp) was amplified using primers TB-RplMrep1 and TB-RplMrep2 (Table 2-1),

and mutant alleles of rplMMtb (676bp) in which the phosphorylated residue threonine 11 (T11) was replaced by an alanine (T11A) or a glutamate residue

(T11E) were generated by a two-stage PCR procedure. Primers TB-RplMT11Arev and

TB-RplMT11Afwd (Table 2-1) were used together with primers TB-RplMrep3 and TB-

RplMrep4 (Table 2-1), respectively, and PCR products let anneal. Primers TB-

RplMrep3 and TB-RplMrep4 (Table 2-1) were then added to amplify the mutant

T11A T11E rplM or rplM alleles after the native promotor region (PrplM). pV897 and pVN898 were constructed by cloning the 5’-PrplM-flanking region directionally into pYUB854 adjacent to the hygromycin-resistance gene using SpeI/XhoI sites and

T11A T11E cloning the PrplM-rplM region or the PrplM-rplM region directionally into pYUB854 adjacent to the hygromycin-resistance gene using XbaI/KpnI, respectively.

After PacI-l digestion to linearize the recombinant cosmids, they were ligated to

63

shuttle phasmid phAE87 digested with PacI, and the ligation was packaged with the

GIGAPackIII GOLD system (Stratagene). Phages containing recombinant phasmids

T11A T11E phrplMMtb or phrplMMtb were used to transduce E. coli NM554 and transductants were selected on hygromycin. Phasmid DNA was extracted from pooled transductants and used to transform M. smegmatis mc2155 by electroporation. Plaques obtained at 30ºC were isolated and used to transduce M. tuberculosis H37Rv and CDC1551. Transductants were selected on hygromycin and gene deletion/replacement confirmed by PCR using primers RMsm1 + RMsm2 (Table

2-1) to observe according molecular weight shift. Primer TB-RplMrep3 was used to

confirm sequence had changed to the mutant allele. Strains were designated

H37Rv/rplMT11A, H37Rv/rplMT11E, CDC1551/rplMT11A and CDC1551/rplMT11E.

2.B.3. Deletion of ntpA in E. coli by one-step phage λ-Red-based recombineering method.

The chromosomal gene encoding the E. coli mutT3 homolog (ntpA) was deleted using a one-step phage λ-Red-based recombineering method as previously described 175; 176. The KanR cassette was amplified using primers NtpA-Kan1 and

Ntpa-Kan2, which carried extensions homologous to the flanking DNA regions of

ntpA (Table 2-1). PCR products were gel-purified and directly electroporated into E.

coli TB10 cells (MG1655, nadA::Tn10 λcI857 Δ(cro-bioA)) 177 that had been heat- induced to express the λ-Red system 175; 176. Kanamycin-resistant transformants were selected and correct replacement of ntpA by KanR verified by PCR using primers annealing adjacent to homologous sequences, NtpA1 + NtpA2 (Table 2-1).

64

The obtained ∆ntpA locus was transferred to the wild type E .coli strain MG1655 by

P1 phage-mediated transduction as previously described 177.

2.B.4. Cloning for in trans expression of proteins.

pVN578 and pVN579 were constructed as previously described. 20

pknGMtb was amplified using primers TB-PknG-1 and TB-PknG-2 (Table 2-1)

and cloned directionally into pET15b using NdeI/XhoI to fuse the N-terminus to a

6xHis-tag, creating pVN792.

garAMsm was amplified using primers MS-GarA-1 + MS-GarA-2 (Table 2-1)

and cloned directionally into pMV361 using EcoRI/HindIII to be expressed from

Phsp60, creating plasmid pVN821.

mutT3Msm was amplified using primers MS-MutT3-1 + MS-MutT3-2 (Table 2-

1) and cloned directionally into pMV361 using EcoRI/HindIII to be expressed from

Phsp60, creating plasmid pVN753. For addition of a C-terminal 6xHis-tag, mutT3Msm

was amplified using primers MS-MutT3-1 + MS-MutT3-2-6xHis (Table 2-1) and cloned directionally into pVN747 or pET11c using NdeI/HindIII or NdeI/BamHI to be expressed from PSOD or PT7, creating pVN823 and pVN835, respectively. For expression from its native promotor PmutT3, mutT3Msm was amplified using primers

MS-pro-MutT3 + MS-MutT3-2-6xHis (Table 2-1) and cloned directionally into pCV125 using NdeI/HindIII, creating pVN866.

mutT3Mtb was amplified using primers TB-MutT3-1 + TB-MutT3-2 (Table 2-

1) and cloned directionally into pMV361 using EcoRI/HindIII to be expressed from

Phsp60, creating plasmid pVN771. For expression from its native promotor PmutT3,

65

mutT3Mtb and its upstream promotor region was amplified using primers TB-pro-

MutT3 + TB-MutT3-2 (Table 2-1) and cloned directionally into pVN839 using

NdeI/HindIII, creating pVN840.

ntpA was amplified using primers NtpA-1 + NtpA-2 (Table 2-1) and cloned

directionally into pMV361 using EcoRI/HindIII to be expressed from Phsp60, creating plasmid pVN773.

ylgG sequence was optimized for codon usage in mycobacteria, synthesized by GenScript (Piscataway, NJ) with added restriction sites, and cloned directionally into pMV361 using EcoRI/HindIII to be expressed from Phsp60, creating plasmid pVN776.

rplMMsm was amplified using primers MS-RplM-1 and MS-RplM-2 (Table 2-1) and cloned directionally into pET15b using NdeI/BamHI to be expressed from T7, creating pVN844.

rplMMtb was amplified using primers TB-RplM-1 and TB-RplM-2 and cloned

directionally into pET15b using NdeI/BamHI to be expressed from T7, creating

pVN885. Mutant alleles of rplMMtb in which the residues threonine 11 (T11),

threonine 12 (T12) and/or serine 14 (S14) were replaced by an alanine (T11A,

T12A, S14A) were generated by a two-stage PCR procedure. Primers TB-RplM3T-fwd,

TB-RplMT11A-fwd, TB-RplMT12A-fwd or TB-RplMS14A-fwd (Table 2-1) were used together with primer TB-RplM-2 (Table 2-1) and PCR products let anneal with primers TB-RplM3T-1, TB-RplMT11A-1, TB-RplMT12A-1 or TB-RplMS14A-1 (Table 2-1),

respectively. Primers TB-RplM-1 and TB-RplM-2 (Table 2-1) were then added to

66

Table 2-1. Primers used in this study.

Primer Sequence

MS-PknG-1 gaattccatatgacttcacccgagaacc

MS-PknG-2 aagcttccatgaaagcacggtcgacgtg

P1 atcggcagcaacctgttctcgttc

P2 ctgcacgacttcgaggtgttcga

P3 gcgattcagcctggtatgatcagc

P4 ccgtcttctacagtcgtcgtagg

PknG-del1 actagtaccaggtggacatcgtcgtcaaga

PknG-del2 aagcttgatcagggttctcgggtgaagtca

PknG-del3 tctagactggtcgacctggccaacagc

PknG-del4 ggtacccaatgggaattgacgggtcgatcc

RT-PknG1 gccaccgacatctacaccgt

RT-PknG2 ggtgtgcgccaccagcag

TB-PknG-1 CATatggccaaagcgtcagagaccgaa

TB-PknG-2 CTCGAGttagaacgtgctggtgggccgga

MMsm1 gatctgcgtctgcatcatcgacga

MMsm2 acctgatccagctgcgtgacaag

MMtb1 gcctctgacagatatgacgacc

MMtb2 ggcaaggctgtatctgtgcacc

P2 ctgcacgacttcgaggtgttcga

P3 ccgattcagcctggtatgatcagc

MS-MutT3del1 ACTAGTgcaactggatcagcgcctcgtc

MS-MutT3del2 AAGCTTccgttctccgacatcacccagc

MS-MutT3del3 TCTAGAtgctcaaccggcagcgctgatc

MS-MutT3del4 GGTACCcccgtcaggaactcgaagcgc

67

TB-MutT3del1 ACTAGTcccagcaattgcaccaacgattcg

TB-MutT3del2 AAGCTTcgtatgcctaacaccatctgtcgg

TB-MutT3del3 TCTAGAcggatcagttcgctgctgtaagc

TB-MutT3del4 GGTACCggagtggacatcatccagttgcg

MS-MutT3-1 CATATGTCTAGAgcaactggatcagcgcctcgtc

MS-MutT3-2 AAGCTTtcagcgctgccggttgagcaac

MS-MutT3-2-6xHis GGATCCAAGCTTTCAATGATGATGATGATGGTGGCTGCTGCCGCT GCCGCGCGGCACCAGgcgctgccggttgagcaacgg

MS-pro-MutT3 CATATGcgatggcctgctcgtagcgct

TB-MutT3-1 GAATTCCATATGccgacagatggtgttaggcatacg

TB-pro-MutT3 TCTAGAcaccgacagcgtggtgtacagc

TB-MutT3-2 AAGCTTgcttacagcagcgaactgatccg

NtpA-1 GAATTCCATAtgaaggataaagtgtataagcgtcc

NtpA-2 GGATCCAAGCTTtcaggcagcgttaattacaaactgt

NtpA-Kan1 tcgatcttagtggtcatctacgcacaagatacgaaacgggtccggggatccgtcgacc

NtpA-Kan2 cgcctgccggttgctccaggacttagtgagcgccgccgcgtgtaggctggagctgcttcg

MS-GarA-1 GAATTCCATATGgatgtgacggtagagaccacatcg

MS-GarA-2 AAGCTTGGATCCccatgggtcaggcgttcgagc

RMsm1 ccgaaaatctcgcacacaagccaac

RMsm2 cgctctcggtcaccacctcg

MS-RplMrep1 ACTAGTcagtgcgtcagtagtagggagc

MS-RplMrep2 AAGCTTcgtcctattcatatgccgtttctgg

MS-RplMrep3 TCTAGAgtgcgtaccggttccctctatg

MS-RplMrep4 TCATGAgtttcggtcacatcggtcactgg

TB-RplMrep1 ACTAGTccatgaagacctcctcagggagg

TB-RplMrep2 CTCGAGcggaactagcgtgcccaagct

TB-RplMrep3 TCTAGAgcgcaacagacccctgtgtggt

68

TB-RplMrep4 GGTACCggctggggtggtttcggtcat

MS-RplMT11Arev taccacgaacgcgtggcgtca

MS-RplMT11Afwd tgacgccacgcgttcgtggta

MS-RplMT11Erev gtaccacgaacgcgtctcgtca

MS-RplMT11Efwd tgacgagacgcgttcgtggtac

TB-RplMT11Arev cacgatcgcgtggcgtcaccc

TB-RplMT11Afwd gggtgacgccacgcgatcgtg

TB-RplMT11Erev cacgatcgcgtctcgtcaccc

TB-RplMT11Efwd gggtgacgagacgcgatcgtg

MS-RplM-1 GAATCCCATATGgtgcctacttacacgccgaagg

MS-RplM-2 AAGCTTGGATCCgtttcggtcacatcggtcactgg

TB-RplM-1 CATATGagcgctgtgcccacgtacgc

TB-RplM-2 GGATCCggtggtttcggtcattgcgcca

TB-RplM3T-fwd acgccgcgcgagcgtggta

TB-RplM3T-1 taccacgctcgcgcggcgtcacccgccttgggcgcgtacgtgggcacagcgctCATATG

TB-RplMT11A-fwd acgccacgcgatcgtggta

TB-RplMT11A-1 taccacgatcgcgtggcgtcacccgccttgggcgcgtacgtgggcacagcgctCATAG

TB-RplMT12A-fwd acaccgcgcgatcgtggta

TB-RplMT12A-1 taccacgatcgcgcggtgtcacccgccttgggcgcgtacgtgggcacagcgctCATAG

TB-RplMS14A-fwd acaccacgcgagcgtggta

TB-RplMS14A-1 taccacgctcgcgtggtgtcacccgccttgggcgcgtacgtgggcacagcgctCATATG

Capital letters represent restriction sites or tags added to genetic sequence (lower case).

69

Table 2-2. Plasmids and phasmids used in this study.

Plasmid Relevant Features References / Company

pCV125 KanR, mycobacterial integrative vector carrying 178 lacZ pET11c AmpR, replicative E. coli vector (OripMB1) carrying Novagen a built-in PT7 promotor followed by a lac operator and a multiple cloning site. Carries lacI.

pET15b AmpR, replicative E. coli vector (OripMB1) carrying Novagen a built-in PT7 promotor followed by a lac operator, an N-terminal 6xHis-tag sequence, a thrombin site and a multiple cloning site. Carries lacI.

pJV53 KanR, mycobacterial replicative vector expressing 179 the Che9c 60-61 genes under control of an acetamide-inducible promotor

pPknGK181M KanR, pSD5 -derived, expressing M. tuberculosis 20 pknGK181M provided by Jean Pieters

pPR27 GenR, mycobacterial vector, replicated from a 173

temperature sensitive Ori and expresses a

sucrose counter selectable marker sacB

pMV361 KanR, mycobacterial integrative vector carrying a 180 built-in heat shock promotor (Phsp60) for translational fusion

pVN578 KanR, pMV361-derived, expressing M. tuberculosis This study 171 pknG from Phsp60

pVN579 KanR, pMV361-derived, expressing M. smegmatis This study 171 pknG from Phsp60

pVN701B KanR GenR, pJV53-derived vector replicating from This study 171 a temperature-sensitive OriTs and expressing a sucrose counter selectable marker sacB as well as mycobacteriophage Che9c recombination proteins

pVN740 HygR, pYUB854 -derived vector with M. This study 171 smegmatis pknG flanking-regions cloned orientationally flanking the hygromycin cassette

70

pVN747 HygR, pMS2–derived replicative vector carrying a 181 built-in PSOD promotor

pVN753 KanR, pMV361-derived, expressing M. smegmatis This study mutT3 from Phsp60

pVN755 HygR, pYUB854-derived vector with M. smegmatis This study mutT3 flanking-regions cloned orientationally flanking the hygromycin cassette

pVN771 KanR, pMV361-derived, expressing M. tuberculosis This study mutT3 from Phsp60

pVN773 KanR, pMV361-derived, expressing E. coli ntpA This study from Phsp60

pVN776 KanR, pMV361-derived, expressing L. lactis ylgG This study from Phsp60

pVN791 HygR, pYUB854-derived vector with M. This study tuberculosis mutT3 flanking-regions cloned orientationally flanking the hygromycin cassette

pVN792 AmpR, pET15b-derived, expressing n-terminally This study 6x-His-tagged M. tuberculosis pknG from PT7

pVN798 AmpR, pET15b-derived, expressing n-terminally This study 6x-His-tagged E.coli ntpA from PT7

pVN821 KanR, pMV361-derived, expressing M. smegmatis This study garA from Phsp60

pVN823 HygR, pVN747-derived, expressing c-terminally This study 6xHis-tagged M. smegmatis mutT3 from PSOD

pVN835 AmpR, pET11c-derived, expressing c-terminally This study 6xHis-tagged M. smegmatis mutT3 from PT7

pVN839 KanR, integrative mycobacterial vector carrying This study OriC, KanR, and integration sites from pMV361 and PSOD and multiple cloning sites from pVN747

pVN840 KanR, pVN839-derived, expressing M. tuberculosis This study mutT3 from it native promotor PmutT3

71

pVN844 AmpR, pET15b-derived, expressing n-terminally This study 6x-His-tagged M. smegmatis rplM from PT7

pVN866 KanR, pCV125-derived, expressing c-terminally This study 6xHis-tagged M. smegmatis mutT3 from its native promotor PmutT3

pVN885 AmpR, pET15b-derived, expressing n-terminally This study 6x-His-tagged M. tuberculosis rplM from PT7

pVN889 AmpR, pET15b-derived, expressing n-terminally This study 6x-His-tagged M. tuberculosis rplMT11A,T12A,S14A from PT7

pVN890 AmpR, pET15b-derived, expressing n-terminally This study 6x-His-tagged M. tuberculosis rplMT11A from PT7

pVN891 AmpR, pET15b-derived, expressing n-terminally This study 6x-His-tagged M. tuberculosis rplMT12A from PT7

pVN892 AmpR, pET15b-derived, expressing n-terminally This study 6x-His-tagged M. tuberculosis rplMS14A from PT7

pVN895 HygR, pYUB854-derived vector with M. smegmatis This study rplMT11A-containing flanking-regions cloned orientationally flanking the hygromycin cassette

pVN896 HygR, pYUB854-derived vector with M. smegmatis This study rplMT11E-containing flanking-regions cloned orientationally flanking the hygromycin cassette

pVN897 HygR, pYUB854-derived vector with M. This study tuberculosis rplMT11A-containing flanking-regions cloned orientationally flanking the hygromycin cassette

pVN898 HygR, pYUB854-derived vector with M. This study tuberculosis rplMT11E-containing flanking-regions cloned orientationally flanking the hygromycin cassette

pYUB854 HygR, E. coli replicative vector carrying a 174 hygromycin cassette and lambda cos sites

72

phAE87 TM4-derived mycobacterial shuttle phasmid, 182 conditionally replicative at 30°C

ph∆mutT3Mtb HygR, phAE87 –derived phasmid carrying PacI- This study digested mutT3Mtb-deletion cassette of pVN791

phrplMMtbT11A HygR, phAE87 –derived phasmid carrying PacI- This study digested rplMMtbT11A-replacement cassette of pVN987 phrplMMtbT11E HygR, phAE87 –derived phasmid carrying PacI- This study digested rplMMtbT11E-replacement cassette of pVN989 HygR, hygromycin resistance. KanR, kanamycin resistance. GenR, gentamycin resistance. AmpR, ampicyclin resistance

73

T11A,T12A,S14A T11A T12A S14A amplify the mutant rplMMtb , rplMMtb , rplMMtb or rplM alleles and products cloned directionally into pET15b using NdeI/BamHI to be expressed from

T7, creating pVN889, pVN890, pVN891 and pVN892, respectively.

For expression and purification of PknGMtb-6xHis, BL21 transformed with

pVN792 were grown at 37ºC until OD600 had reached 0.65 and induced with 0.1mM

IPTG at 22ºC, 200rpm for 16 hours, as previously described 20. Soluble fraction of cell lysates was prepared in TBS buffer containing protease inhibitor cocktail by sonication (15 cycles of 10 seconds on ice with chilling intervals) and diluted 1:1 with wash buffer (50mM sodium phosphate, 300mM NaCl, and 45mM imidazole), loaded onto a cobalt metal affinity spin column (Pierce) pre-equilibrated with the same buffer. The column was washed 7 times with 2 column volumes of wash buffer and bound protein was eluted 4 times with 1 column volume of elution buffer (50 mM sodium phosphate, 300 mM NaCl, and 250 mM imidazole). Eluted fractions were pooled and stored at -80ºC with 50% glycerol.

For expression and purification of MutT3Msm-6xHis from E.coli, BL21 transformed with pVN835 were grown at 37ºC until OD600 had reached 0.5 and

induced with 0.5mM IPTG at 30ºC, 250rpm for 3 hours. Soluble fraction of cell

lysates were prepared in lysis buffer (50mM sodium phosphate, 300mM NaCl, and

10mM imidazole) containing protease inhibitor cocktail by sonication (15 cycles of

10 seconds on ice with chilling intervals), loaded onto a nickel metal affinity spin

column (Qiagen) preequilibrated with the same buffer. The column was washed 6

times with 2 column volumes of wash buffer (50mM sodium phosphate, 300mM

74

NaCl, and 20mM imidazole) and bound protein was eluted 4 times with 1 column

volume of elution buffer (50 mM sodium phosphate, 300 mM NaCl, and 500 mM

imidazole). Eluted fractions were pooled and stored at -80ºC with 10% glycerol.

For purification of MutT3Msm-6xHis from M. smegmatis, 7H9 medium

2 containing hygromycin was inoculated with 5 OD600 units/L of mc 155 transformed

with pVN823 and cultures grown at 37ºC, 240rpm for 72 hours. Cell pellets were

harvested at 4ºC, 4000rpm for 15 minutes and washed twice with TBS buffer

containing a protease inhibitor cocktail (Roche Molecular Biochemicals). After

resuspension in 1/10 culture volume of TBS buffer + protease inhibitors, cells were

disrupted by sonication (25 times for 10 seconds on ice with 1 minute cooling

intervals), and lysate spun for 20 minutes at 10,000rpm, 4ºC. Supernatant was

filtered at 4ºC through 0.22µm filters (Denville) and diluted 1:1 with wash buffer

(50mM sodium phosphate, 300mM NaCl, and 20mM imidazole), loaded onto a

cobalt metal affinity spin column (Pierce) preequilibrated with the same buffer. The

column was washed 10 times with 2 column volumes of wash buffer and bound

protein was eluted 3 times with 1 column volume of elution buffer (50 mM sodium

phosphate, 300 mM NaCl, and 150 mM imidazole). Eluted fractions were pooled and

exchanged into 20mM Tris-HCl buffer, pH8.0 using PD-10 desalting columns (GE

Healthcare) before removing remaining contaminants using a strong anion

exchange column (HiTrap Q FF, GE Healthcare) and a 0-1M NaCl gradient. Fractions

were analyzed for presence of MutT3Msm-6xHis by SDS-PAGE stained with

Coomassie Brilliant Blue (see Methods section 3.C.7.) and fractions containing

75

protein (20mM Tris-HCL, 100-150mM NaCl) were pooled and concentrated using 9

kDa molecular weight cut-off spin concentrators (Pierce). Obtained purified fraction

was stored at -80ºC with 10% glycerol.

For expression and purification of RplMMsm-6xHis, RplMMtb-6xHis,

T11A,T12A,S14A T11A T12A S14A RplMMtb , RplMMtb RplMMtb , and RplMMtb , BL21 was transformed with pVN844, pVN885, pVN890, pVN891, and pVN892, respectively,

and grown at 37ºC until OD600 had reached 1. These cultures served as a 1/100 seed

for LB medium supplemented with ampicillin and the cultures were then grown at

37ºC, 240rpm overnight. Soluble fraction of cell lysates were prepared in lysis

buffer (50mM sodium phosphate, 300mM NaCl, and 10mM imidazole, pH9.0)

containing protease inhibitor cocktail by sonication (15 cycles of 10 seconds on ice

with chilling intervals), loaded onto a nickel metal affinity spin column (Qiagen) pre-

equilibrated with the same buffer. The column was washed 10 times with 2 column

volumes of wash buffer (50mM sodium phosphate, 300mM NaCl, and 20mM

imidazole, pH9.0) and bound protein was eluted 3 times with 1 column volume of

elution buffer (50 mM sodium phosphate, 300 mM NaCl, and 500 mM imidazole,

pH9.0). Eluted fractions were pooled, protein exchanged into thrombin buffer

(20mM Tris-HCl, pH8.4, 0.15M NaCl, 2.5mM CaCl2) using PD-10 desalting columns

(GE Healthcare) and 6xHis tag was removed by thrombin cleavage (Novagen) at a

5:8 ratio of recombinant protein (mg) : thrombin (U) at 4ºC for 20 hours. Cleavage

reactions were concentrated using 9 kDa molecular weight cut-off spin

concentrators (Pierce), and the cleaved tag and other contaminants were removed

76

by fast protein liquid chromatography (FPLC) using a HiLoad 16/60 Superdex 75

prep grade column (GE Healthcare). Fraction containing purified, untagged protein

was concentrated using 9 kDa molecular weight cut-off spin concentrators (Pierce) and stored at -80ºC with 10% glycerol.

2.C. RT-PCR.

Primers (Table 2-1) were synthesized by Fisher Scientific (Pittsburgh, PA).

RNA from M. smegmatis strains was isolated according to manufacturers protocol using the RNeasy minikit (Qiagen) and reverse transcribed as previously published

140 using the Omniscript RT kit (Qiagen). RT-PCRs were performed using the Expand

Long Template PCR kit (Roche Molecular Biochemicals). cDNA of pknG was

amplified using primers RTpknG1 and RTpknG2 (Table 2-1), which anneal to

conserved sequences present in both pknGMsm and pknGMtb.

2.D. SDS-PAGE, Coomassie Blue staining and Western blots.

For preparation of whole cell extracts, mycobacterial cells were washed

three times in Tris-HCl-buffered saline (TBS) containing protease inhibitors (Roche

Molecular Biochemicals) and disrupted by glass bead beating using a Fastprep 24

(MP Biomedicals, Solon, OH) for 6 cycles of 40 seconds at frequency 60s-1 with chilling intervals. Glass beads, unbroken cells and insoluble material were removed by centrifugation (20,000rpm, 20 min) and the total cell lysate treated with SDS- sample buffer and heated at 95ºC for 10 minutes.

SDS-PAGE was performed using the Biorad Protean III system. Lysates, protein purification fractions or kinase assay samples were heated for 10 min at

77

95°C in SDS-sample buffer, centrifuged briefly and loaded on 10 or 15 %

polyacrylamide SDS-PAGE gels and resolved at 20mA per gel.

Gels used for kinase assay or Western blot were blotted by semi-dry transfer

(BioRad, Hercules, CA) onto PVDF membranes for 4 hours at 0.8mA/cm2 of membrane.

Coomassie Brilliant Blue G-250 (Fisher Bioreagents, Pittsburgh, PA) staining of SDS-PAGE gels or PVDF membranes was done for 30 minutes at room temperature.

PknG Western Blot was performed using 35µg total protein of M. smegmatis strains (except ∆pknGMsm/pVN578 and ∆pknGMsm/pVN579, where 10 µg total protein

were loaded) or M. bovis BCG per lane. Membranes were incubated with 1/5000

20 polyclonal antibody raised against PknGMtb in 5% milk in TBS supplemented with

1% Tween 80 (TBS-T) (Sigma) for 3 hours at room temperature. Secondary

antibody was horseradish-peroxidase conjugated goat anti-rabbit IgG antibody

(Pierce) in 5% milk in TBS-T. Detection was done using Pierce ECL Western blotting

kit and signals were recorded on film.

His-tag Western blot was performed using 20µg of mycobacterial cell extract

or 0.1µg of purified 6x-His-tagged protein per lane. Membranes were incubated with

1/1000 penta-His antibody (Qiagen) in 3% BSA (Sigma) in TBS supplemented with

1% Tween 80 (TBS-T) (Sigma) overnight at 4ºC. Secondary antibody was

horseradish-peroxidase conjugated goat anti-mouse IgG antibody (Pierce) in 5%

milk in TBS-T. Detection was done using Pierce ECL Western blotting kit and signals

78

were recorded on film.

2.E. Antibiotic susceptibility testing.

Antibiotic susceptibility of M. smegmatis strains was estimated using a diffusion assay with antibiotic-containing filter paper discs (BD Diagnostic Systems) as previously described 168. Minimal antibiotic inhibitory concentrations (MIC) were

® determined using E-test (AB Biodisk) strips on NE agar medium with 0.04 OD600

units of cells and read out after 48-72 hour growth at 37°C as described previously

168.

Antibiotic susceptibility of E. coli strains was estimated using an adapted

168 filter paper disc assay where cultures were grown to OD600 = 1, NE medium replaced by LB agar medium and plates read out after 24 hours.

In M. tuberculosis strains, erythromycin susceptibility was assessed by

streaking 10µl of cultures at OD600 = 1 onto 7H10-OADC medium with or without

erythromycin (100mg/L). Growth was assessed after 4 week incubation at 37°C.

Sulfachloropyridazine susceptibility on solid medium was assessed either by streaking 10µl of cultures at OD600 =1 or by replicating (96-well layout) or spotting

2µl volumes of serial dilutions of cultures of defined OD600 onto 7H10-OADC

medium containing titrations of sulfachloropyridazine and folinic acid. Growth was compared after 4 week incubation at 37°C. Sulfachloropyridazine susceptibility in

liquid medium was assessed by daily monitoring of growth (OD600) in 7H9-OADC

medium containing increasing amounts of antibiotic (0-100µg/ml) over a 7 day

incubation period at 37°C and 200rpm (n=3).

79

2.F. Cellular aggregation, Hydrophobicity Index and Congo red binding assays.

Cellular aggregation assays were performed as previously described 183.

Hydrophobicity assays were performed as previously described 183. Briefly,

M. smegmatis strains were grown to late exponential phase in 7H9 medium and single cell suspensions were prepared and washed with PBS. OD600 was adjusted to

1.0 and mixed with hexadecane (Sigma). After 15 minute incubations at room temperature allowing for phase separation, the aqueous phase was harvested and

OD600 measurements used to assay percentage of cells transitioning from hydrophilic to hydrophobic phase.

Congo Red binding assay was performed as previously described 183. Briefly,

M. smegmatis strains were grown in the presence of 20µg/ml Congo red for 72 hours in 7H9 medium, washed until supernatant was completely colorless, cell resuspended in acetone to extract bound Congo Red and OD at 488nm monitored.

Cells were then pelleted and dried and Congo red binding expressed as OD488/g dry cell mass.

2.G. Zeta Potential.

Cell surface charge of M. smegmatis strains was measured by zeta potential

183 determination as previously described . Briefly, 1 OD600 units of PBS-washed single cell suspensions were used for measurements in a ZetaMaster zetameter (Malvern Instruments) and expressed in mV.

2.H. Biofilm assays.

Biofilm assays were done as described 112; 113. Briefly, strains were grown in

80

7H9 liquid medium supplemented with glucose (M. smegmatis) or OADC (M.

tuberculosis) and appropriate antibiotics until OD600 had reached 0.5-0.8 (M. smegmatis) or >1 (M. tuberculosis). Growing cultures (10 µl for M. smegmatis, 100 µl for M. tuberculosis) were used to inoculate 60x15mm polystyrene petri plates containing 10 ml of biofilm medium (M. smegmatis) 112 or Sauton’s medium without

Tween 20 113. Depending on assay, medium was supplemented with antifolates (0-

50 µM) (Sigma), kinase inhibitors, folic acid (0-1mM) (Schircks Laboratories) or

folinic acid (0-1mM) (Schircks Laboratories). Standing cultures were incubated at

30ºC and observed for 7 days (M. smegmatis) or at 37ºC under humidified

conditions and wrapped in parafilm and observed for 6 weeks (M. tuberculosis). For

quantitation of biofilm growth, the liquid medium and any planktonic cells were

removed with a syringe connected to a sterile needle from beneath the biofilm layer and the remaining biofilm biomass harvested. Viable bacteria counts were determined by colony forming unit (CFU) assay or total protein per plate determined by Bradford method (Bio-Rad Laboratories, CA).

2.I. Kinase assays.

In 20µl volumes, 0.5µg of purified PknGMtb-6xHis (depending on assay pre- treated for 15 minutes at 37°C with 5mM AX20017) were incubated with 1.5µg of purified substrate (RplM or MutT3 forms, with or without 6x-His-tag, purified from

E. coli or M. smegmatis) in 10mM HEPES, 20µM DTT, 40µM MnCl2 with 10µCi of [γ-

32P]-ATP (PerkinElmer) for 30 minutes at 37°C. Reactions were stopped by addition

of SDS-sample buffer and boiling at 95°C for 10 minutes and resolved on 15%

81

polyacrylamide SDS-PAGE gels (see section 2.E.). Gels were transferred by semi-dry

blot (BioRad) onto PVDF membranes (BioRad) and stained with Coomassie Brilliant

Blue R-250 (Fisher) before exposing to autoradiography imaging screens for 6 days.

Screen were read on a Storm 820 PhosphoImager (GE Healthcare) and analyzed with

ImageQuant software (GE Healthcare).

2.J. Radioactive nucleotide binding assays.

Radioactive nucleotide binding assays were adapted from previously

184; 185 described protocols . Briefly, in triplicates, purified MutT3Msm-6xHis was incubated with [8-3H]-GTP (MP Biomedicals) in the presence or absence of

unlabeled GTP (Sigma) in buffer (50mM Tris-HCl, pH 7.4, 100mM NH4Cl, 10mM

MgCl2, 1mM dithiothreitol) in 200µl total volumes for 20 minutes at 37°C before stopping reactions by addition of 3ml chilled stopping solution (50mM Tris-HCl, pH

7.4, 10mM NH4Cl, 10mM MgCl2). Reactions were immediately filtered through

0.45um pore size nitrocellulose membrane discs (25mm) and membranes washed with three 3ml volumes of chilled stopping solution before soaking in scintillation fluid (PerkinElmer). Released radioactive signal that had been retained on membranes was read by a Beckman LS6500 liquid scintillation counter.

3 For titration of MutT3Msm-6xHis, [8- H]-GTP was kept constant at 6.4µCi

(2.5µM). For titration of labeled GTP, MutT3Msm-6xHis was kept constant at 10 or

20µg. For outcompetition of [8-3H]-GTP by unlabeled GTP, [8-3H]-GTP and

MutT3Msm-6xHis were kept constant at 6.4µCi (2.5µM) and 20µg, respectively.

2.K. Macrophage infection, lysosomal trafficking and microscopy.

82

Macrophage infections and lysosomal trafficking assays were adapted from previously published methods 20; 186. Bone marrow–derived macrophages were generated by incubating bone marrow monocytes from C57BL/6 mice (provided by

Kurt Lu) for 7 days in high-glucose-Dulbecco’s Modified Eagle Medium containing

25% L-929 conditioned medium, 1% penicillin, 1% streptomycin, 4.5g/L glucose

(HyClone) supplemented with 4 mM L-glutamine (Sigma), 15% heat-inactivated fetal calf serum (Sigma), and 0.02mg/L macrophage-stimulating growth factor

(Sigma) at 37°C and 10% CO2. Non-activated macrophages were seeded on MatTek

(Ashland, MA) glass bottom 14mm microwell dishes with 1.5 coverglass and let

adhere for 2 hours (37°C, 10%CO2) prior to infection with M. bovis BCG or ∆pknGBCG

expressing GFP (provided by Jean Pieters) 20; 186 (MOI 625) for 1 hour (37°C,

10%CO2). After removal of non-phagocytosed bacteria, infected macrophages were

subjected to a 16-hour chase with 30µM sulfachloropyridazine (Sigma) and/or 2mM

folinic acid (Schircks Laboratories). After the chase period, lysosomal compartments

were labelled for 30 minutes (37°C, 10%CO2) with 1µM neutral red (Invitrogen), an acidic organelle–selective cell-permeant fluorescent probe, and fixed with 4% formaldehyde (Fisher) for 20 minutes before staining with DAPI (Invitrogen) for 5 minutes. Slides were mounted in ProLong Gold antifade reagent (Invitrogen). Assays were analysed on a Zeiss LSM510 confocal microscope using LSM510 REL3.5

software provided by the manufacturer. 100 events were counted for each condition

from triplicate slides in 2 independent experiments.

83

Chapter 3.

Protein Kinase G is Required for Intrinsic Antibiotic Tolerance in

Mycobacteria

Kerstin A. Wolff,† Hoa T. Nguyen,† Richard Cartabuke, Ajay Singh, Sam Ogwang,

and Liem Nguyen

Department of Molecular Biology and Microbiology, School of Medicine, Case Western

Reserve University, Cleveland, Ohio 44106

† These authors contributed equally to this research.

84

3.A. Abstract.

The eukaryote-like serine/threonine protein kinase G (PknG) of M.

tuberculosis has previously been reported to be involved in survival of the bacillus in infected host macrophages, contributing to its virulence and pathogenesis. We

identified pknG in a transposon mutagenesis based screen for genome wide

determinants of multiple drug resistance in mycobacteria, and demonstrate that

PknG is required for intrinsic multiple drug tolerance (MDR) in both pathogenic and

non-pathogenic mycobacteria. Deletion of pknG leads to altered cell surface

properties and increased susceptibility to a wide variety of antibiotics in both M.

smegmatis and M. tuberculosis, indicating that PknG is a genetic factor underlying

the commonly observed phenotypic association between antibiotic resistance and

virulence.

85

3.B. Introduction.

Traditionally, host persistence and antibiotic resistance have been regarded

as separate phenomena, but phenotypic association of the two is commonly

observed both in vivo and in vitro 22; 23; 24. For example, during persistence in the

host, M. tuberculosis exhibits reduced antibiotic susceptibility when compared to

actively dividing bacteria 13; 14; 15; 16. Conversely, the transition from active growth to persistence can be induced by supplementation of antibiotics 25.

Like for all intracellular pathogens, it is crucial for M. tuberculosis to be able to survive and persist within infected host cells 10. While we do not have a complete understanding of persistence in the host, which is established by colonizing the early phagosome and arresting its maturation, several factors from both host and M. tuberculosis have been shown to play crucial roles in this process 4; 5; 17; 18; 19. Among

the mycobacterial factors are glycolipids and the RD1 specialized secretion system

17; 18; 19. In addition, mycobacterial protein phosphatases and kinases have recently been identified to perform key roles in establishing the survival niche within host phagosomes 19; 20; 21. One such kinase is the mycobacterial eukaryotic-like

serine/threonine protein kinase G (PknG), which was suggested to support survival

of pathogenic mycobacteria in the host macrophage by preventing phagolysosomal

fusion and therefore acidification of the mycobacterial compartment 20. Thus, PknG has been thought to function as a virulence factor of M. tuberculosis 20; 187. However,

genes encoding PknG are ubiquitously conserved throughout the Mycobacterium genus, including non-pathogenic mycobacteria such as M. smegmatis 122; 140.

86

Sequence identity between the M. smegmatis and M. tuberculosis homologs of PknG

is ~80%, and the gene locus is also highly conserved in both bacteria 140.

Cumulatively, these facts suggest that PknG performs a physiologically relevant function in all mycobacterial species besides the previously proposed virulence function in the pathogenic species.

M. tuberculosis displays a high level of intrinsic antibiotic tolerance to most available antibiotics, making treatment of M. tuberculosis infections difficult and

creating a requirement for long treatment schedules with a limited arsenal of useful

drugs 23. This intrinsic tolerance of antibiotics is mediated by a variety of factors including several drug degradation and inactivation processes that occur inside mycobacterial cells. Additionally, over 30 drug efflux transporters that have been shown to be involved in resistance to aminoglycosides, chloramphenicol, fluoroquinolones, isoniazid, linezolid, rifampicin, tetracycline and other toxic compounds also contribute to intrinsic drug resistance 32; 33; 34; 35. Finally, the extremely thick and multi-layered mycobacterial cell wall, which is composed of multiple layers of varied hydrophobicity and thereby poses an effective obstacle for the entry of most chemical compounds 3; 23; 26; 27.

Here we demonstrate a role for PknG in mycobacterial multiple drug

resistance, which appears to be conserved in both non-pathogenic and pathogenic

mycobacteria, thereby making PknG a direct link between antibiotic susceptibility

and intracellular persistence.

87

3.C. Results.

3.C.1. Identification of MAR4, a M. smegmatis pknG transposon mutant

displaying multidrug sensitivity.

A library of more than 5,000 transposon mutants was generated from wild

type M. smegmatis mc2155 179 by Himar1-mediated mutagenesis 168, and deposited in 96 well-plates. The library was subjected to exhaustive screens to identify

mutants sensitive to one or more antibiotics. This screening process has identified

genome-wide drug resistance determinants in mycobacteria, and a subgroup of the

obtained drug sensitive mutants encompassing those that displayed a decreased

Multiple Antibiotic Resistance phenotype was designated as the MARs. One of the

MAR mutants, MAR4, exhibited increased sensitivity to many currently clinically

important antibiotics when compared to wild-type M. smegmatis (Table 3-1). These

antibiotics are of diverse chemical structures and function, and include drugs

targeting cell wall synthesis (imipenem, vancomycin, and ethambutol), protein

synthesis (erythromycin), folate synthesis (sulfachloropyridazine), and

transcription (rifampicin) (Table 3-1). Of these, two (rifampicin and ethambutol)

are frontline antibiotics that are part of the standard regimen to treat tuberculosis.

E-test® drug sensitivity testing (AB Biodisk, Sweden) was performed 168 to obtain

representative minimal inhibitory concentrations (MICs) that determined that

MAR4 was 8-, 12-, 15-, 4-, and >8-fold more susceptible to erythromycin,

vancomycin, imipenem, ethambutol, and rifampin, respectively, than wild-type M.

smegmatis mc2155 (Table 3-1).

88

An arbitrary PCR method 168 first localized the transposon insertion in MAR4 to msmeg_0786, which encodes for a homolog of protein kinase G (PknGMsm), with

140 almost 80% identity to the M. tuberculosis homolog PknGMtb (Fig. 3-1). The

disruption of pknG in M. smegmatis MAR4 was further confirmed by PCR amplification using primers flanking the putative open reading frame (MS-PknG1 and MS-PknG2, Table 2-1 and arrows in Fig. 3-1A). The mutant gene generated a larger fragment corresponding to the inserted 1825 base pairs of the Himar1- transposon, which showed decreased mobility on an agarose gel (Fig. 3-1B).

Sequencing of the junction region from this PCR product identified the insertion site

at the dinucleotide TA1015-1016, introducing a stop codon after the triplet encoding

residue Ile338 (Fig. 3-1A, bottom). To confirm loss of the pknG gene product, Western

20 blots were done using a polyclonal antibody raised against PknGMtb . The antibody

recognized a protein band of ~ 78 kDa corresponding to the molecular weight of

2 PknGMsm in cell extract of wild type M. smegmatis mc 155 but failed to do so in

protein extract of MAR4, confirming the deletion of the pknG gene and loss of its

gene product (Fig. 3-1C).

3.C.2. Targeted deletion of pknG and complementation experiments confirm the MDR function of PknG in pathogenic and non-pathogenic mycobacteria.

Since secondary mutations or polar effects to downstream genes can cause false attribution of the drug susceptibility phenotype, a targeted pknG mutant (the

∆pknGMsm strain) was constructed by a modified version of the previously published recombineering approach 170 to independently assess if the multi-drug sensitive

89

Figure 3-1. Identification of MAR4, a M. smegmatis pknG transposon mutant.

(A) Synteny of the pknG loci in M. smegmatis and M. tuberculosis genomes and the

Himar1 transposon insertion in MAR4. Homologous genes are uniformly illustrated.

The nucleotide sequence at the pknG-Himar1 junction shows truncation of PknGMsm

338 at Ile . (B) PCR amplification using primers flanking pknGMsm (Table 2-1, arrows in panel A) from genomic DNA of wild-type M. smegmatis mc2155 and MAR4. (C)

Western blot assay using PknGMtb antibody that recognizes PknG in cell extracts of M.

bovis BCG and wild-type M. smegmatis mc2155 but not in MAR4 extract.

Reproduced with permission from 171. 90

Table 3-1. Susceptibility of M. smegmatis strains to antibiotics.

MIC (mg/liter)a Strain EM VA IP EB RI mc2155 32 48 6 0.25 >32 MAR4 4 4 0.38 0.064 4 mc2155/pknGK181M 12 16 1 0.064 4

ΔpknGMsm strain 3 4 0.38 0.064 4

K181M ΔpknGMsm/pknG strain 3 4 0.38 ND ND

ΔpknGMsm/pknGMtb strain 32 48 6 0.25 >32

ΔpknGMsm/pknGMsm strain 32 48 6 0.25 >32 aAbbreviations: EM, erythromycin; VA, vancomycin; IP, imipenem; EB, ethambutol; RI, rifampin; ND, not determined.

Reproduced with permission from 171.

91

phenotype of MAR4 was due to pknG disruption. Vector pVN701B expressing the mycobacteriophage Che9c recombination proteins gp60 and gp61 (which encode for homologs of RecE and RecT, respectively) from pJV53 170 and replicating from a

173 temperature-sensitive origin of replication (Orits) from pPR27 was constructed

(Fig. 3-2A). Chromosomal DNA sequences encompassing 551bp and 528bp of the 5’

and 3’-flanking regions, respectively, adjacent to pknGMsm were cloned into vector pYUB854 174 to flank the hygromycin-resistance cassette Ωhyg. The resulting pknG

deletion substrate (5’-pknG-flanking-region_ Ωhyg_3’-pknG-flanking-region) was

excised by SpeI/KpnI digestion (Fig. 3-2B). Homologous recombination promoted

by the induced mycobacteriophage system replaced pknGMsm with Ωhyg on the

chromosome, leading to hygromycin resistance in the resulting mutants (Fig. 3-

2B,D). Besides being resistant to both hygromycin and kanamycin (KanR carried by pVN701B), these transformants showed reduced growth on medium supplemented with sucrose (Fig. 3-2D, left panel), indicating the presence of pVN701B. Shifting cultures to 39°C in the absence of kanamycin removed pVN701B to prevent further undesired random recombination events (Fig. 3-2D).

Homologous recombination replacing chromosomal pknG with the hygromycin cassette in ∆pknGMsm mutant candidates was tested by PCR and sequencing using primers annealing to chromosomal regions adjacent to the homologous sequences of the pknGMsm flanking regions: PCR A (primers P1 and P2)

recovered the 5’-ΔpknG-Himar1 transition; PCR B (primers P3 and P4) recovered

the Himar1-3’-ΔpknG transition (Table 2-1 and Fig. 3-2B). All tested hygromycin

92

resistant transformants showed correct homologous recombination leading to

deletion of pknGMsm (Fig. 3-2C).

For complementation, plasmids expressing the M. smegmatis (pVN579) or M. tuberculosis (pVN578) pknG gene from the heat shock promotor (Table 2-2) 20 were

transformed into ΔpknGMsm, generating strains ΔpknGMsm/pknGMsm and

ΔpknGMsm/pknGMtb.

Next, reverse transcription PCR (Fig. 3-2E) was used to analyze the

transcription of pknG in all generated M. smegmatis strains. RNA was isolated and

reverse transcribed, followed by amplification of pknG cDNA using primers

RTpknG1 and RTpknG2 (Table 2-1) that recognized both M. smegmatis and M.

tuberculosis pknG sequences and thus allowed for monitoring of transcription of

either pknG gene into mRNA. Whereas transcripts of pknG were readily detected in

RNA samples isolated from wild type M. smegmatis mc2155, transcripts of pknG in

ΔpknGMsm were not detectable. Transformation of pVN578 or pVN579 restored transcripts of pknG (Fig. 3-2E).

Similarly, at the protein level, PknG was present in protein extracts of wild

type M. smegmatis and the complemented strains ΔpknGMsm/pknGMsm and

ΔpknGMsm/pknGMtb, but absent in ΔpknGMsm (Fig. 3-2F). Notably, compared to M. bovis

BCG (far left, Fig. 3-2F), a dramatic reduction of PknG expression was observed in wild type M. smegmatis, in agreement with previously published observations 140.

The reduced signal is likely caused by lower PknGMsm production in M. smegmatis

rather than by reduced specificity of the antibody (generated using PknGMtb), as in

93

trans expression of PknGMsm or PknGMtb from the same vector yielded comparable

signals on Western blots (lanes 4 and 5, Fig. 3-2F). The low expression of PknGMsm in

M. smegmatis was completely abolished in the ΔpknGMsm mutant (lane 3, Fig. 3-2F),

just as in the transposon mutant, MAR4 (Fig. 3-1C). Strains ΔpknGMsm/pknGMsm and

ΔpknGMsm/pknGMtb overproduced PknGMsm and PknGMtb, respectively (Fig. 3-2F).

Cumulatively, these results confirmed targeted deletion and complementation of pknG.

The generated strains allowed for analysis of PknGMsm function in antibiotic resistance. As before, antibiotic susceptibility was tested using E-test® diffusion

strips (AB Biodisk, Sweden). The antibiotic susceptibility level of ΔpknGMsm was identical to MAR4 for all tested antibiotics (Table 3-1). Importantly, in trans

expression of pknGMsm from vector pVN579 completely restored wild type levels of

antibiotic resistance to ΔpknGMsm (Table 3-1), confirming that the observed

phenotype is a direct function of pknGMsm expression. Expression of pknGMtb also

rescued ΔpknGMsm, suggesting that PknG provides a physiological, multiple drug tolerance function to not only non-pathogenic, but also to pathogenic mycobacteria

(Table 3-1).

Next, multiple drug susceptibility was assayed in M. tuberculosis H37Rv,

20 where phenotypes for ΔpknGMtb are similar to those observed in M. smegmatis, with increased susceptibility to the representative drugs erythromycin (Fig. 3-3A) and sulfachloropyridazine (Fig. 3-3B). Again, expression of pknGMtb restored wild-

94

Figure 3-2. Construction and complementation of the ΔpknGMsm strain. (A)

Construction of pVN701B. (B) The pknGMsm deletion cassette and pknG locus in wild-

2 type M. smegmatis mc 155 and its derived ΔpknGMsm strain. (C) PCR confirmation of

pknG deletion in eight random mutant candidates. (D) Removal of pVN701B by a

temperature shift resulted in kanamycin sensitivity and sucrose resistance of the

ΔpknGMsm strain. (E) Reverse transcription (RT)-PCR detection of pknG transcription

in M. smegmatis strains. (F) Western blot detection of PknG in M. smegmatis strains.

Reproduced with permission from 171. 95

Figure 3-3. Antibiotic susceptibilities of M. tuberculosis strains. (A) Growing

cultures (10 μl; optical density at 600 nm, 1.0) of wild-type M. tuberculosis H37Rv,

its derived ΔpknGMtb mutant, and the complemented ΔpknGMtb/pknGMtb (pVN578)

strain were plated on solid 7H10-OADC medium 19 in the absence (−) or presence

(+) of erythromycin (100 mg/liter) and incubated at 37°C for 4 weeks. (B) M. tuberculosis H37Rv (circles), the ΔpknGMtb strain (triangles), and the

ΔpknGMtb/pknGMtb strain (squares) were grown at 37°C for 7 days with shaking (200 rpm) in liquid 7H9-OADC 168 supplemented with increasing concentrations of

sulfachloropyridazine. Growth was recorded as optical density at 600 nm 19

normalized to the absorbance of control (no antibiotic).

Reproduced with permission from 171. 96

type levels of antibiotic resistance to the deletion mutant (Fig. 3-3). These results

confirm that PknG indeed performs similar functions in multiple drug tolerance of

pathogenic as well as non-pathogenic mycobacterial species.

Finally, to test whether the kinase activity of PknG is required for its function

K181M 181 20 in antibiotic tolerance, a kinase-dead form of PknGMtb, PknG (Lys → Met) , was introduced into wild type M. smegmatis (plasmid pPknGK181M, Table 2-2).

Antibiotic susceptibility assays showed that overproduction of this dominant negative PknGK181M 20 form resulted in reduction of antibiotic resistance of wild type

M. smegmatis, similar to the effect caused by pknG deletion (Table 3-1). This result strongly supports the role of PknG phosphorylation activity in intrinsic multiple drug resistance of mycobacteria.

3.C.3. Loss of pknG leads to altered cell surface properties.

Proliferation of wild-type M. smegmatis and ΔpknGMsm in vitro was monitored by optical density measurements at a wavelength of 600nm. While no differences in growth rates in either 7H9 or biofilm liquid medium 112 could be observed between wild type and ΔpknGMsm mutant in non-static growth conditions

(Fig. 3-4A), increased cell-cell aggregation compared to wild type was observed for the mutant strain. Quantification of cellular aggregation showed an increase by 53%

(Table 3-2) in ΔpknGMsm compared to wild type. In trans expression of PknGMsm

rescued the aggregation phenotype of the mutant.

97

Figure 3-4. Loss of PknG does not affect growth rate in liquid, shaking medium.

2 In vitro growth of wild type M. smegmatis mc 155 (squares) and ΔpknGMsm (circles) in 7H9 (filled symbols) and biofilm liquid medium 112 (open symbols). Saturated

cultures growing in 7H9 medium were used to inoculate fresh medium to initial

optical density at 600nm (OD600) of 0.2 and incubated at 37ºC with shaking

(250rpm). At indicated time intervals, growth of the cultures was monitored by measuring OD600.

98

Table 3-2. Surface properties of M. smegmatis strains.

Strain

2 Surface Property mc 155 ΔpknGMsm ΔpknGMsm/pknGMsm

Cellular 18.8 +/- 3.6 72.0 +/- 2.5 24.7 +/- 1.3 Aggregationˆ

27.3 +/- 2.41 13.63 +/- 5.34 29.79 +/- 2.96 Hydrophobicity Index˜

Congo Red 224 +/- 45 119 +/- 16 257 +/- 48 Binding˚

Zeta Potential¯ -20.47 +/- 1.26 -27.13 +/- 1.57 -23.67 +/- 0.71

ˆ Percentage of aggregated cells ˜Arbitrary units, expressed as the percentage of cells recovered in the organic phase. ˚10−4*A488 ¯ Expressed in mV

99

To assess whether this was due to alterations in hydrophobicity of the cell

wall, the hydrophobicity index of the strains was measured by hexadecane partition

procedure 183. Single cell suspensions adjusted to units of 1 OD were mixed with

various volumes of hexadecane and the hydrophilic phase recovered and assessed

for reduction in optical density. The results indicated that the surface of ΔpknGMsm is

more hydrophilic than that of wild type, and ΔpknGMsm/pknGMsm was rescued completely (Table 3-2). Further confirmation of altered hydrophobicity was obtained by Congo Red binding 183. Congo Red is a hydrophobic diazo-dye that binds

lipids, lipoproteins and other macromolecules found in the mycobacterial cell wall

and has been used as an indicator of phenotypic drug susceptibility 188. Consistent

with the observed decrease in hydrophobicity index of ΔpknGMsm, ΔpknGMsm bound only half as much Congo red as wild type or complemented cells (Table 3-2).

To assess whether the loss of PknG also results in charge alterations on the cell surface, zeta potential measurements were performed on single cell

183 suspensions as previously described using a Malvern ZetaMaster . ΔpknGMsm

showed an increase in negative cell surface charge over wild type by 34.8%, and in

trans expression of PknGMsm restored wild type charge to ΔpknGMsm (Table 3-2).

100

3.D. Discussion.

The presented work constitutes a comprehensive set of evidence suggesting

that PknG plays a role in intrinsic mycobacterial resistance to multiple antibiotics of diverse structures, targets and functions. Both transposon-mediated (MAR4) and targeted mutagenesis (ΔpknGMsm and ΔpknGMtb) resulted in mutant strains with increased multiple antibiotic susceptibility (Table 2-3) in both non-pathogenic and pathogenic mycobacterial species. The non-specific sensitivity of MAR4, ΔpknGMsm

and ΔpknGMtb to antibiotics with diverse chemical structures and mechanisms of action (Table 2-3, Fig. 3-3) suggests that PknG controls either (i) the activities of dedicated multiple drug resistance mechanisms such as efflux pumps, (ii) the integrity of the mycobacterial cell envelope through PknG involvement in essential cellular functions, or (iii) the physiology and fitness of mycobacteria. The results obtained in this study support the latter hypothesis, as strain ΔpknGMsm showed altered cell wall characteristics (Table 2-4). Loss of PknGMsm resulted in a cell wall both more hydrophilic and more negatively charged than that of wild type M.

smegmatis (Table 2-4). Reduced cell wall hydrophobicity has previously been

reported to be important for multiple antibiotic tolerance in mycobacteria 168; 189.

Additionally, recent findings that PknG is involved in the tricarboxylic acid cycle as well as glutamine utilization in Corynebacterium and Mycobacterium also favor this possibility 121; 122; 124. Glutamine and its derived molecules, such as polyglutamines, may serve as components that are required for the integrity of the mycobacterial cell wall, thereby helping to pose an effective barrier to diffusion of antibiotics of

101

diverse size and polarity. Furthermore, the likelihood that PknG does not directly

control dedicated MDR systems is supported by the fact that increased expression of

neither PknGMtb nor PknGMsm in the ΔpknGMsm strain (lanes 4 and 5 in Fig. 3-2F)

leads to an increase in intrinsic resistance to any of the tested antibiotics compared

to wild-type M. smegmatis (Table 2-3).

Expression analyses revealed that the amount of PknG protein detectable in

the cytoplasm of M. smegmatis is dramatically reduced in comparison to those found

in M. tuberculosis or M. bovis BCG (Fig. 3-2F). These results are in agreement with a

recently published observation that PknG expression is translationally regulated in

mycobacteria 140, suggesting the existence of regulatory mechanisms to control

expression levels of PknG in mycobacterial species. A dramatic increase of PknGMtb

expression may have provided pathogenic mycobacteria with an additional amount of the protein required for its new function in virulence 20. This possibility was also suggested in a recent study that addressed differential expression of PknG in various mycobacterial species in relation to its virulence function 140.

Together with previous reports 20; 124; 140, our data suggest that signaling

circuitries surrounding PknG may serve as a phenogenetic network that controls

persistence in the infected host and intrinsic antibiotic resistance to multiple drugs

in mycobacteria 20.

102

3.F. Acknowledgements

We thank Jean Pieters, Julia van Kessel and Graham F. Hatfull for providing

materials; Abram Stavitsky and Kien Nguyen for critical reading of the manuscript.

Support was provided by start-up funds from the School of Medicine, a STERIS

Infectious Diseases Research Grant and a CFAR Developmental Award from Case

Center For AIDS Research. Sam Ogwang is a recipient of a Forgarty AIDS

International Training and Research Program (AITRP) Fellowship.

103

Chapter 4.

Modulation of Folate Metabolism by Protein Kinase G Determines

Biofilm Growth and Subcellular Localization of Pathogenic

Mycobacteria in Macrophages

Kerstin A. Wolff, Hoa T. Nguyen and Liem Nguyen

Department of Molecular Biology and Microbiology, School of Medicine, Case Western

Reserve University, Cleveland, Ohio 44106

104

4.A. Abstract

The eukaryotic-like serine/threonine protein kinase G (PknG) of M. tuberculosis has been shown to be involved in preventing lysosomal fusion to

phagosomes containing the bacilli in infected host macrophages, thereby creating an

intracellular niche. It was therefore suggested that PknG serves as a virulence factor

that directly modifies host signaling cascade proteins. Other reports however have

shown involvement of PknG, which is highly conserved in Actinobacteria, in

glutamate metabolism, cell surface properties, and intrinsic multiple antibiotic

tolerance in both non-pathogenic and pathogenic mycobacteria.

Here we present evidence that PknG regulates mycobacterial folate

metabolism in late stage growth conditions, and is required for survival of folate

starvation and biofilm growth. We also demonstrate that inhibition of PknG has the

potential to be exploited to potentiate classical antifolate drugs that target folate

biosynthesis in other steps for use against M. tuberculosis. Lastly, we present

evidence suggesting that the previously demonstrated role of PknG in inhibition of

phagolysosomal fusion is mediated through regulation of mycobacterial folate levels

rather than modification of host signaling cascades.

105

4.B. Introduction

Upon infection of host macrophages, M. tuberculosis colonizes the host

phagosome by preventing acidification of this compartment 10; 12; 190 and modulation

of other antimicrobial mechanisms of the macrophage 191. Acidification normally

occurs via fusion of lysosomal compartments to the bacteria-containing phagosomes,

and colonization of the phagosome is achieved by arresting its maturation and

preventing lysosomal delivery 10; 12; 190; 192. The underlying mechanisms of this arrest

have not yet been fully elucidated, but both host and mycobacterial factors play

crucial roles in the process 4; 5; 17; 18; 19. One such factor is the soluble eukaryote-like

serine/threonine kinase PknG of M. tuberculosis that has been shown to play a

critical role in M. tuberculosis’ ability to establish the phagolysosomal fusion block,

though its mechanism of action and targets remain unknown 20; 137. In mouse

macrophage infections, both genetic deletion of pknG and chemical inhibition of

PknG by a specific inhibitor, AX20017 20; 138, in M. bovis BCG resulted in increased

fusion of lysosomes to phagosomes containing mycobacteria, and this correlated

with decreased survival of internalized mycobacteria (Fig. 1-5, 1-6) 20; 23. It has therefore been proposed that PknG is actively secreted into the macrophage cytosol where it acts as a virulence factor and modifies eukaryotic host substrates that are active in signalling cascades necessary for establishment of lysosomal trafficking to the colonized phagosome 20.

Recently it has been shown that PknG is also involved in bacterial glutamate

metabolism 121; 122; 124, maintenance of mycobacterial cell surface properties 171, as

106

well as intrinsic multiple drug resistance 171, determining PknG as a genetic factor underlying the often observed phenotypic association of virulence and antibiotic resistance. One example of this association is biofilm growth, which has been suggested to represent a growth condition similar to static growth during persistence of mycobacteria in the host 104; 105; 106; 107; 112; 113. Bacteria growing in biofilms, like persistent mycobacteria during latent infection, display enhanced tolerance for antibiotics, a phenomenon that is likely due to the metabolic differences between biofilm growth and planktonic growth 104; 105; 106; 107.

Mycobacteria, including M. tuberculosis, have been shown to form biofilms both in vivo and in vitro 111; 112; 113; 114, and biofilms of Mycobacterium chelonae and

Mycobacterium ulcerans have been shown to play a role in pathogenesis 114; 115; 116;

117. Consistent with the idea that biofilm growth represents a late static growth stage similar to host persistence, mycobacterial biofilms also exhibit higher antibiotic tolerance than planktonic mycobacteria, similar to persistent M. tuberculosis during latent infection 22, and biofilm of M. tuberculosis has been linked

to the formation of drug-tolerant persisters 113.

Here we present further evidence suggesting that the role of PknG in

phagolysosome biosynthesis is mediated through the physiology of the bacillus

rather than a role in host signaling control. We show that PknG regulates folate

metabolism in the bacterium, specifically in late stage growth conditions that mirror

latent infection and persistence inside the host macrophage. We demonstrate that

regulation of folate metabolism by PknG is required for mycobacterial survival of

107

folate starvation induced by classical antifolate antibiotics and for biofilm growth.

We also present evidence that inhibition of PknG phosphorylation has the potential

to be exploited for boosting the antimycobacterial activity of classical antifolates.

Lastly, we show that the previously demonstrated inhibition of phagosome- lysosome fusion in infected macrophages by PknG 20 is mediated through regulation of bacterial folate levels, rather than through active modification of eukaryotic host proteins in signaling cascades.

108

4.C. Results

4.C.1. PknG is required for survival of folate starvation.

We had previously identified pknG, the gene encoding for the eukaryotic-like serine/threonine protein kinase G in mycobacteria, in a transposon mutagenesis screen for genome-wide drug resistance determinants in mycobacteria. We had confirmed the requirement for PknG kinase activity for intrinsic tolerance of M. tuberculosis to many clinically important antibiotics. This observation suggested a physiological function of PknG based on modification of mycobacterial substrates rather than a virulence function based on phosphorylation of host signaling factors

171. The identified drug classes included antifolates 171, a group of antibiotics that encompasses both sulfadrugs and trimethoprim and inhibits de novo folate biosynthesis. In a later, independent transposon mutagenesis screen that looked for antifolate resistance determinants in mycobacteria 102 pknG was identified again, further indicating that PknG plays a role in susceptibility of mycobacterial species to

this group of antibiotics. The previously observed multi-drug tolerance function

thus seemed to be a consequence of a defect in folate metabolism 171.

171 As previously shown , deletion of pknGMtb led to increased susceptibility to the sulfadrug sulfachloropyridazine when compared to wild type M. tuberculosis

H37Rv (Fig. 4-1A, upper row). To exclude polar effects on downstream genes, we

expressed pknGMtb in trans from plasmid pVN578 (Table 2-2) in the ∆pknGMtb mutant strain (strain ∆pknGMtb/pknGMtb) and could completely restore wild-type levels of

sulfachloropyridazine resistance (Fig. 4-1A, upper row). Notably, supplementation

109

of exogenous folinic acid, a storage form of folate, bypassed the requirement for

PknG (Fig. 4-1A, bottom row), indicating that the function of PknG in antifolate

resistance is directly related to de novo folate biosynthesis or folate metabolism.

Similar results were observed in M. smegmatis (Fig. 4-2A), indicating that this role of

PknG is conserved in both pathogenic and non-pathogenic mycobacteria.

4.C.2. Chemical inhibition of PknG activity potentiates classical antifolate

antibiotics.

Potentiation of currently available TB drugs by targeting resistance mechanisms to

these agents has recently gained attention because of the rapid spread of MDR and

XDR TB spread that has limited the current options for TB chemotherapy 44. We sought to explore whether the kinase activity of PknG mediates its function in antifolate resistance, and as a result, whether chemical inhibition of PknG activity could be exploited to potentiate the antimycobacterial activity of classical antifolates. For this purpose, the PknG-specific kinase inhibitor AX20017 20 that

specifically inhibits PknG activity (Fig. 4-1B) was titrated along with increasing

concentrations of sulfachloropyridazine (Fig. 4-1C). While inhibition of PknG in the

absence of antibiotics did not affect growth of wild-type M. tuberculosis H37Rv,

addition of AX20017 increased the antimycobacterial activity of

sulfachloropyridazine in a concentration-dependent manner (Fig. 4-1C). These data

indicate that not only the kinase activity of PknG is required for intrinsic

mycobacterial tolerance to antifolates, but also that this mechanism of resistance

110

Figure 4-1. Requirement of PknGMtb activity for survival of folate starvation and exploitation for antifolate potentiation. (A) Sulfachloropyridazine susceptibility of M. tuberculosis strains. Growing cultures (10μl; optical density at

600 nm, 1.0) of wild-type M. tuberculosis H37Rv, its derived ΔpknGMtb mutant, and

111

the complemented ΔpknGMtb/pknGMtb (pVN578) strain were plated on solid 7H10-

OADC medium 19 containing increasing amounts (0-5µM) of sulfachloropyridazine

(SL) and either 0 or 1mM folinic acid (FA) and incubated at 37°C for 4 weeks. (B) In vitro kinase assay using radioactive [γ-32P]-ATP to assess autophosphorylation activity of purified PknGMtb-6xHis (0.5µg). 15 minute pre-incubation with 5mM

AX20017 markedly reduced inhibited kinase activity of PknG. (C) Potentiation of sulfachloropyridazine activity by chemical inhibition of PknG. Serial dilution of growing H37Rv culture (2μl volumes) were spotted in triplicate onto solid 7H10-

OADC medium containing increasing amounts of sulfachloropyridazine (SL) and

AX20017 (0-2mM). Growth was recorded after 3 weeks of incubation at 37°C.

112

Figure 4-2. Requirement of PknGMsm activity for survival of folate starvation. (A)

Sulfachloropyridazine susceptibility of M. smegmatis strains. Growing cultures (50µl;

2 optical density at 600 nm, 0.6) of wild-type M. smegmatis mc 155, its derived ΔpknGMsm

mutant, and the complemented ΔpknGMsm/pknGMsm (pVN579) and ΔpknGMsm/pknGMtb

(pVN578) strains were plated on solid NE medium containing 0 or 1mM folic acid (FA),

paper discs containing 0.25µg of sulfachloropyridazine (SL) applied and incubated at

37°C for 4 days. (B) Potentiation of sulfachloropyridazine activity by chemical inhibition

of PknG. Growing cultures (50µl; optical density at 600 nm, 0.6) of wild-type M.

smegmatis mc2155 were plated onto solid NE medium and paper discs containing 0.25µg

sulfachloropyridazine (SL) and 0.1 µg kinase inhibitor chelerythrine (Che) and incubated

at 37°C for 4 days.

113

can be targeted by chemical inhibition in order to potentiate classical antifolate

antibiotics against mycobacteria.

4.C.3. PknG is required for biofilm growth.

We observed a greater increase in susceptibility of ∆pknGMtb to antifolates and other antibiotics in late-stage growth than in log-phase 171, suggesting importance of PknG involvement in folate regulation in late-stage or latent infection- like conditions. Since biofilm growth has been shown to be correlated with the ability to form persisters displaying increased antibiotic tolerance in M. tuberculosis

113, we sought to explore whether biofilm behavior might be altered upon loss of

pknGMtb.

Biofilm assays using a liquid-air interface model 112; 113 indeed showed that the ∆pknGMtb strain showed a delay in formation of mature biofilms compared to wild-type M. tuberculosis H37Rv (Fig. 4-3A, left), corresponding with decreased biofilm biomass (Fig. 4-3A, right). In trans expression of pknGMtb rescued the mutant

phenotype (Fig. 4-3A). Strikingly, supplementation of exogenous folinic acid in the

medium reestablished normal biofilm growth in the ∆pknGMtb strain (Fig. 4-3A).

Similar results were obtained in M. smegmatis (Fig. 4-4A). These results suggest that

PknG plays a role in biofilm formation through regulation of folate biosynthesis.

To assay whether kinase activity of PknG was required, AX20017 was added

to biofilm cultures of both wild type and ∆pknGMtb/pknGMtb strains, which resulted in

immature biofilms closely resembling that of the ∆pknGMtb strain (Fig. 4-3B).

Likewise, inhibition of PknGMsm by the Pkcα inhibitor chelerythrine yielded similar

114

115

Figure 4-3. PknGMtb is required for folate-dependent biofilm growth in M. tuberculosis. (A) Biofilm growth of wild type M. tuberculosis H37Rv, its derived

ΔpknGMtb mutant, and the complemented ΔpknGMtb/pknGMtb (pVN578) strain was

assayed in the presence or absence of 1mM folinic acid (FA). Biofilm was pictured at

6 weeks (left). Surface biofilm growth was quantified at week 6 by measuring total

protein per biofilm plate (right). Supplementation of folinic acid bypassed the

requirement for PknG. (B) Surface biofilm growth of M. tuberculosis H37Rv strains

was assayed in the absence or presence of 1mM AX20017 (PknG inhibitor). Growth

was recorded at week 6 after inoculation. (C) Effect of folate starvation on biofilm formation. Biofilm growth of wild type M. tuberculosis H37Rv was assayed in the presence of increasing amounts of the antifolate drug sulfachloropyridazine (SL) and either 0 or 1mM folinic acid. Growth was quantified at week 6 after inoculation by measuring total protein per biofilm plate.

116

Figure 4-4. PknGMsm is required for folate-dependent biofilm growth in M. smegmatis. (A) Biofilm growth of wild type M. smegmatis mc2155 and its derived

ΔpknGMsm mutant was assayed in the presence or absence of 1mM folic acid (FA).

Biofilm was pictured after 1 week (left). Surface biofilm was quantified after 1 week of growth by measuring total protein per biofilm plate (right). Supplementation of folate bypassed the requirement for PknGMsm. (B) Surface biofilm growth of M. smegmatis mc2155 was assayed in the absence or presence of 1mM chelerythrine

117

(Che). Growth was recorded at week 1 after inoculation (left) and quantified by colony forming unit (CFU) measurement of surface biofilm.

118

results in an M. smegmatis background (Fig. 4-4B). PknG kinase activity is therefore

required for normal biofilm growth.

To further clarify whether folate availability mediates normal biofilm

formation, biofilm assays were conducted in the presence of sulfachloropyridazine.

Folate starvation in wild-type M. tuberculosis H37Rv led to decreased biofilm

formation in a dose-dependent manner, and could be overcome by supplementation of exogenous folinic acid (Fig. 4-3C).

Cumulatively, these data support a model in which biosynthesis and availability of folate are crucial in late-stage growth such as in the biofilm setting. In this condition, PknG kinase activity is required to establish this required folate level by an unknown mechanism.

4.C.4. PknG is involved in regulation of lysosomal delivery in macrophages via modulation of mycobacterial folate metabolism.

In light of our data and the implied importance of mycobacterial substrates of PknG, we sought to determine whether an alternate explanation to the proposed virulence role of PknG in phagolysosomal fusion 20 might be the requirement for mycobacterial folate biosynthesis in late-stage growth of M. tuberculosis infections.

Failure to produce sufficient amounts of folate would have led to decreased survival, which would also account for increased processing and clearance by infected macrophages upon loss of PknG function, as observed with heat-killed bacteria 20.

The increase in phagolysosomal fusion we observed in bone marrow derived

murine macrophages infected with ∆pknGBCG strain was comparable to previously

119

120

Figure 4-5. PknG is involved in lysosomal delivery to infected phagosomes through folate levels. Bone marrow derived murine macrophages were infected with GFP-expressing M. bovis BCG or its derived ΔpknGBCG mutant (green) and

chased with either 2mM folinic acid (FA) or 30µM sulfachloropyridazine (SL) before

staining with DAPI (blue) and Neutral Red (red). Fixed macrophages were pictured

at 170x magnification (top) and colocalization quantified (bottom). Folinic acid

reestablished failure of macrophages to deliver lysosomes to phagosomes in cells

infected with ΔpknGBCG and in cells infected with BCG that was folate-starved by classical antifolate antibiotics.

121

published observations 20 (Fig. 4-5). Supplementation of exogenous folinic acid in the medium reestablished the inability of infected macrophages to deliver lysosomes to phagosomes containing mycobacteria (Fig. 4-5), bypassing the requirement for PknG. Furthermore, treatment of infected macrophages with sulfachloropyridazine, which targets the bacteria, but not the host cells, led to increased lysosomal delivery, similar to that observed in the absence of PknG (Fig.

4-5).

These results indicate that latent growth conditions such as those of biofilm or macrophage infections require folate biosynthesis that is involved in lysosomal trafficking to phagosomes containing pathogenic mycobacteria, and that PknG is necessary for this process. It also indicated that substrates for PknG might be prokaryotic in nature.

122

4.D. Discussion

In this report we present evidence suggesting a novel function of PknG in the

physiology and metabolism of mycobacterial species. We demonstrate that PknG

regulates mycobacterial folate metabolism, which is of particular importance in static growth conditions. We also show that this regulation of folate is required for mycobacterial persistence to classical antifolate antibiotics, biofilm growth, and, most importantly, for successful prevention of lysosomal delivery of pathogenic mycobacteria after internalization by host macrophages 20. It was previously

suggested that PknG may be actively secreted into the host cell, where it interferes

with host signalling cascades, possibly involving Pkcα 127, thereby mediating the

control over phagolysosomal fusion and aiding the survival of pathogenic

mycobacteria in macrophages 20. Our results, on the other hand, indicate a mycobacterial function for PknG instead that is in agreement with previous reports proposing roles for PknG in glutamate metabolism, multiple antibiotic tolerance and cell wall maintenance 121; 122; 124; 171.

While the intracellular bacilli that persist during latent TB infections are generally regarded to be non-replicating and growing very slowly 193; 194; 195, recent reports have shown that they are metabolically active 196 and should therefore require folate for the operation of metabolically essential one-carbon transfer

reactions 45; 77; 78; 79. Our results indicating that modulation of mycobacterial folate

metabolism by PknG is most important in conditions of static growth, are in agreement with the observations that pknG is upregulated during latent infection in

123

mice 196, and that is plays a role in stationary phase metabolism 196; 197. Dependence on folate levels in an intracellular setting would also account for the presence of pknG in the obligate intracellular pathogen M. leprae, which has reduced its genome to the bare minimum. It may also be the reason for the generally higher expression levels of PknG in pathogenic mycobacterial species 140. The involvement of PknG in

multiple antibiotic resistance and cell wall integrity of both non-pathogenic and

pathogenic bacteria 157 is mediated through bacterial folate levels. This may partially account for the association of stationary phase growth, latency, and biofilm

growth with higher levels of antibiotic resistance to a variety of drugs 22; 13; 14; 15; 16.

Non-pathogenic mycobacteria do not normally infect eukaryotic hosts. The

conservation of pknG in such typically environmental, saprophytic species may be accounted for by a requirement to frequently form biofilms, for which PknG activity is also required.

We also present evidence that, chemical inhibition of PknG leads to potentiation of classical antifolate antibiotics against M. tuberculosis. Therefore,

PknG provides a potential target for development of potentiators that boost activity of classical antifolates 45. Considering the role of PknG in regulation of folate metabolism that is required for preventing lysosomal delivery of pathogenic mycobacteria, co-administration of PknG inhibitors with antifolates might not only result in potentiation of antimycobacterial activity of TB drugs, but also promote

phagosome-lysosome fusion, leading to improved killing and clearance of M.

tuberculosis by host phagocytic cells. It is therefore critical to further study the

124

pathways that PknG is involved in, as well as to identify its targets and molecular mechanism of action.

125

4.E. Acknowledgements.

We thank Jean Pieters (BCG/GFP, ∆pknGBCG, ∆pknGMtb), Julia van Kessel and

Graham F. Hatfull (pJV53), Bryan Doreian and Kurt Lu (mouse bone-marrow derived macrophages). Support was provided by an NIH R01 grant (AI 087903), start-up funds from the School of Medicine, a STERIS Infectious Diseases Research

Grant, and a CFAR Developmental Award from Case Center For AIDS Research.

126

Chapter 5.

Protein Kinase G Regulates Folate Metabolism in Mycobacteria

Through Phosphorylation of Ribosomal Protein RplM

Kerstin A. Wolff and Liem Nguyen

Department of Molecular Biology and Microbiology, School of Medicine, Case Western

Reserve University, Cleveland, Ohio 44106

127

5.A. Abstract

We have previously demonstrated that the eukaryotic-like serine/threonine protein kinase G (PknG) of M. tuberculosis modulates folate metabolism, cell wall

integrity, intrinsic multiple drug resistance, biofilm formation, and lysosomal

trafficking to phagosomes, allowing for persistence of pathogenic mycobacteria in

host macrophages. This modulation of folate metabolism is mediated through its

kinase activity.

In this study, we sought to determine the mycobacterial substrate(s) and

pathway(s) targeted by PknG. We identify a novel substrate for PknG, the RplM

protein of the large ribosomal subunit. RplM is phosphorylated by PknG on

threonine 11 that is conserved in mycobacterial species. RplM co-purifies with

MutT3, a putative Nudix hydrolase identified as an effector of PknG function in

folate-dependent biofilm growth. We hypothesize that MutT3 functions as a

dihydroneopterin-triphosphate pyrophosphatase that catalyzes the first committed

step of folate de novo biosynthesis.

128

5.B. Introduction

In previous chapters, we demonstrated that the prokaryotic-like

serine/threonine kinase G (PknG) is required for antibiotic tolerance in both M.

tuberculosis and non-pathogenic mycobacteria 171, where it also regulates folate

metabolism that is required for static growth resembling M. tuberculosis persistence inside the host macrophage. PknG-mediated regulation of folate metabolism is required for mycobacterial survival of folate starvation triggered by classical antifolate antibiotics, as well as for biofilm growth and the successful prevention of

phagolysosome biosynthesis in infected macrophages 20. These results suggest that

the substrate(s) of PknG is likely a mycobacterial protein, rather than host proteins

acting in signaling cascades 20; 139. We therefore sought to elucidate the bacterial pathways and substrates that are affected by PknG, and, more specifically, how PknG modulates folate metabolism in mycobacteria.

In this chapter, we identify RplM as a novel substrate for PknG. RplM, also known as L13, is an essential ribosomal protein and is a component of the large ribosomal subunit 157. It has also been shown to act in sub-ribosomal particles 160

and outside the ribosomal context 161. RplM has been found to directly interact with

Obg, an essential GTP-binding protein that is required for induction and regulation

of the σB-mediated stress response in B. subtilis 164; 165. RplM has also been detected

in active sub-ribosomal particles that bind both 23S and 5S rRNA in Thermus

aquaticus 160. In a non-ribosomal setting, RplM has been demonstrated to be

129

involved in anti-termination together with ribosomal proteins of the small and large subunits in E. coli 161. RplM is phosphorylated on several residues in E. coli, although

the involved kinases remain unknown 167. These phosphorylated residues are located on

the solvent interface of RplM, suggesting that they may be important for interactions with

ribosome-associated proteins 167. We show here that mycobacterial RplM is

phosphorylated by PknG on threonine 11, which is exclusively conserved

throughout the Mycobacterium genus.

RplM is co-purified with MutT3, a putative Nudix hydrolase and an effector of

PknG. MutT motif hydrolases, also known as Nudix hydrolases, catalyze the

hydrolysis of a nucleoside diphoshate linked to a variable moiety x, removing

pyrophosphate 143; 144. Their manifold substrates include nucleoside triphosphates,

dinucleoside polyphosphates, nucleotide sugars and co-enzymes 143; 144. Most Nudix hydrolases seem to be involved in the clearance of potentially deleterious or hazardous molecules, acting as “housekeeping” enzymes 143; 144; 146. Some Nudix

hydrolases however have been shown to act in metabolic pathways, such as ppGpp

regulation in response to amino acid starvation 149 and the catalysis of the removal

of pyrophosphate from dihydroneopterin-triphosphate in de novo folate biosynthesis 147; 148. We hypothesize that mycobacterial MutT3 functions as a dihydroneopterin-triphosphate pyrophosphatase (DHN-3Pase) catalyzing this first committed step in the pathway, in accordance with previous observations that this

130

Nudix hydrolase does not appear to function as either a mutator protein or an 8-

oxo-guanosine triphosphatase 156.

We also present evidence that phosphorylation of threonine 11 on RplM and

the presence of MutT3 are required for antifolate resistance and biofilm formation.

These results indicate a signaling cascade involving PknG, RplM and MutT3 that is required for modulation of folate metabolism. This signaling cascade is important for static growth conditions including biofilm surfaces and the phagosomal niche within host phagocytic cells.

131

5.C. Results

5.C.1. PknG does not mediate regulation of folate metabolism through phosphorylation of GarA.

Only one bacterial substrate has so far been identified for PknG. PknG of C.

glutamicum phosphorylates OdhI, which then inhibits 2-oxoglutarate

dehydrogenase (ODH) complex 121; 123. A later study confirmed that, along with

PknB, PknG can phosphorylate GarA, the mycobacterial homolog of OdhI, in vitro 122.

GarA, in its unphosphorylated state inhibits action of ODH complex, as is the case in

C. glutamicum, and, depending on substrate concentration, either inhibits or

activates glutamate dehydrogenase (GDH) 122.

To address regulation of folate biosynthesis by PknG through

phosphorylation of GarA, both limitations in CO2 and glutamate had to be addressed.

Since ODH complex is inhibited by unphosphorylated GarA 122, pknG deletion should lead to inhibition of ODH and therefore decreased levels of CO2. We therefore performed biofilm assays in the presence of 5% CO2 to determine whether the

defect in biofilm formation of ∆pknGMsm could be bypassed by exogenous

supplementation of CO2, as had been shown with an ODH mutant 113. While biofilm formation overall was accelerated, loss of PknG still led to a significant delay in the mutant (data not shown). It therefore appears that CO2 is important for biofilm formation, but not sufficient to compensate for the lack of PknG in mycobacteria.

132

As GarA also acts as an inhibitor of GDH 122, lack of PknG should lead to decreased levels of glutamate. Since glutamate is a building block of folate, we quantified biofilm formation of ∆pknGMsm in the presence and absence of exogenous

2 folinic acid. While GarA overexpression (strain mc 155/garAMsm) indeed leads to a

biofilm formation defect similar to that resulting from loss of PknG, supplementation of

exogenous folate does not at all alleviate the biofilm defect of the GarA overexpressing

strain (Fig. 5-1). Therefore, phosphorylation of GarA by PknG is not how PknG controls folate biosynthesis.

5.C.2. mutT3 is required for survival of folate starvation and biofilm growth.

To examine mycobacterial proteins involved in folate biosynthesis that might be regulated by PknG, this pathway was investigated by SSDB (Sequence Similarity

Database) gene cluster search at Kyoto Encyclopedia of Genes and Genomes (KEGG: http://ssdb.genome.jp/ssdb-bin/ssdb_gclust?org_gene=mtu:Rv0410c&type=g). The

only step in folate biosynthesis in mycobacteria for which the catalytic enzyme

remains unknown is the conversion of dihydroneopterin-triphosphate (DHN-3P) to

dihydroneopterin (DHN) (Fig. 5-2). This conversion had been thought to occur in a

spontaneous chemical reaction until 2007, when a dedicated Nudix hydrolase was

identified as the catalytic enzyme in L. lactis and Arabidopsis 148. KEGG suggested that this reaction is catalyzed in mycobacteria by the putative Nudix enzyme MutT3.

MutT3 is encoded only three genes upstream of pknG by Rv0413 and msmeg_0790 in

133

Figure 5-1. PknG does not mediate control over folate metabolism through phosphorylation of GarA. Biofilm growth of M. smegmatis strains. Surface biofilm

2 growth of M. smegmatis mc 155, its derived ΔpknGMsm mutant, and the

2 overexpressing mc 155/garAMsm strain was assayed in the absence or presence of

134

1mM folinic acid (FA). Surface biofilm growth was recorded at week 1 after inoculation and quantified by measuring total protein per biofilm plate.

135

Figure 5-2.Folate de novo biosynthesis pathway. De novo folate biosynthesis pathway in M. tuberculosis. Intermediates are shown in green boxes, classical antifolate antibiotics in orange ellipses, enzymes in black.

136

M. tuberculosis and M. smegmatis respectively, and shares the same intergenic region in a locus conserved across the mycobacterial genus and some related actinobacteria (Fig. 5-3A and Fig. 5-4A). Furthermore, proximity of mutT3 and pknG indicates similar function or involvement in the same pathways.

Strain ∆mutT3Mtb showed increased susceptibility to classic antifolate

antibiotics when compared to wild-type M. tuberculosis H37Rv, similar to the phenotype observed in ∆pknGMtb, and in trans expression of mutT3Mtb from plasmid

pVN771 (Table 2-2) restored wild-type levels of resistance (Fig 5-3C). Similar to

deletion of pknGMtb, supplementation of exogenous folate in the form of folinic acid bypassed the requirement for mutT3Mtb (Fig 5-3C). Similar results were obtained in

M. smegmatis assays (Fig. 5-4C), indicating that MutT3 might be involved in folate biosynthesis and be regulated by PknG.

Biofilm assays also showed similar phenotypes for ∆mutT3Mtb and ∆pknGMtb, and again genetic or chemical complementation could be achieved by in trans expression of mutT3PknG or exogenous addition of folinic acid, respectively (Fig. 5-

3D). M. smegmatis experiments yielded similar results (Fig. 5-4D).

5.C.3. Preliminary characterization of MutT3 as a Nudix Hydrolase

Nudix hydrolases are pyrophosphatases that catalyze the hydrolysis of a

nucleoside diphoshate linked to a variety of moieties. While most members of this

family are involved in clearance of potentially deleterious metabolites from the cell

137

138

Figure 5-3. mutT3Mtb is required for survival of folate starvation and biofilm growth. (A) The mutT3Mtb deletion cassette and mutT3 loci in wild-type M. tuberculosis H37RV and its derived ΔmutT3Mtb strain. (B) PCR confirmation of

mutT3Mtb deletion using primers MMtb1 + MMtb2 (Table 2-1) shows shift corresponding to replacement of mutT3Mtb with the hygromycin resistance cassette.

Sequencing also confirmed the replacement of mutT3Mtb gene by the hygromycin cassette (not shown) (C) Biofilm growth of M. tuberculosis strains. Biofilm growth of wild type M. tuberculosis H37Rv, its derived ΔpknGMtb and ΔmutT3Mtb mutants, and

the complemented ΔmutT3Mtb/mutT3Mtb and ΔmutT3Mtb/ylgG strains was assayed in

the presence of either 0 or 1mM folinic acid. Growth was quantified at week 6 after

inoculation by measuring total protein per biofilm plate. (D) Sulfachloropyridazine

susceptibility of M. tuberculosis strains. Serial dilutions of growing cultures of wild-

type M. tuberculosis H37Rv, its derived ΔpknGMtb and ΔmutT3Mtb mutants, and the complemented ΔmutT3Mtb/mutT3Mtb and ΔmutT3Mtb/ylgG strains were replicated

onto solid 7H10-OADC medium containing sulfachloropyridazine (SL) and either 0

or 1mM folinic acid (FA) and incubated at 37°C for 3 weeks.

139

140

Figure 5-4. mutT3Msm is required for survival of folate starvation and biofilm growth. (A) The mutT3Msm deletion cassette and mutT3Msm loci in wild-type M.

2 smegmatis mc 155 and its derived ΔmutT3Msm strain. (B) PCR confirmation of mutT3Msm deletion using primers MMtsm1 + P2 and P3 + MMsm2 (Table 2-1) amplify the 5’- and 3’-transitions from genomic DNA to the inserted hygromycin cassette, respectively. (C) Biofilm growth of M. smegmatis strains. Surface biofilm growth of M.

2 smegmatis mc 155, its derived ΔpknGMsm and ΔmutT3Msm mutants, and the complemented ΔmutT3Msm/mutT3Msm, ΔmutT3Msm/mutT3Mtb, ΔmutT3Msm/ntpA and

ΔmutT3Mtb/ylgG strains was assayed in the absence or presence of 1mM folic acid

(FA). Growth was recorded at week 1 after inoculation. (D) Sulfachloropyridazine susceptibility of M. smegmatis strains. Growing cultures of wild-type M. smegmatis

2 mc 155, its derived ΔpknGMsm and ΔmutT3Msm mutants, and the complemented

ΔmutT3Msm/mutT3Msm, ΔmutT3Msm/mutT3Mtb, and ΔmutT3Mtb/ylgG strains were plated onto solid NE medium containing either 0 or 1mM folic acid (FA), paper discs containing 1mg of sulfachloropyridazine (SL) applied to the center and plates incubated at 37°C for 4 days.

141

and prevent build-up of metabolic intermediates, some Nudix hydrolases serve as enzymes in intermediate steps of biosynthetic pathways 146; 147; 148. In particular,

Nudix enzymes of E. coli (ntpA), Lactococcus lactis (ylgG) and Arabidopsis have only recently been shown to catalyze the removal of pyrophosphate from dihydroneopterin-triphosphate (DHN-3P) 147; 148, one of the earliest intermediates in folate biosynthesis, to create dihydroneopterin (DHN), which is eventually turned into folate 86; 147; 148 (Fig. 1-8). This reaction had been thought to occur by spontaneous chemical dephosphorylation for over 50 years 148; 154.

To determine whether loss of DHN-3P pyrophosphatase activity yielded similar sensitivity profiles, as had been observed in mycobacterial backgrounds, we created strain ∆ntpA in E. coli MG1665 by one-step phage λ-Red -based

recombineering method 176; 177; 184 and performed a sulfamethoxazole sensitivity

assay (Fig. 5-5). Deletion of ntpA led to increased susceptibility to the sulfadrug (Fig.

5-5), indicating a general requirement of DHN-3Pase for bacterial antifolate

resistance.

If mutT3 indeed encodes a DHN-3Pase activity, deletion of mutT3 should be complemented by in trans expression of other DHN-3Pases. ∆mutT3Mtb regained wild-type levels of resistance to sulfachloropyridazine when L. lactis ylgG (pVN776) was expressed in trans (Fig. 5-3D). Similar results were obtained in M. smegmatis mc2155 by in trans expression of ntpA (pVN773) or ylgG (pVN776) (Fig. 5-4D).

142

Fig. 5-5. ntpA, a dihydroneopterin-triphosphate pyrophosphatase, is required for antifolate resistance in E. coli. Sulfamethoxazole sensitivity of E.coli strains.

Growing culture of wild-type E. coli MG1665, its derived ∆ntpA or the complemented

∆ntpA/ntpA strains (50µl, OD600=1.0) were plated onto solid LA medium and paper discs containing 1mg sulfamethoxazole were applied. Chemical complementation was achieved by exogenous supplementation of 1mM folinic acid (FA). Plates were incubated at 37ºC overnight.

143

Expression of DHN-3Pases also restored surface biofilm growth to normal levels in

strain ∆mutT3Mtb/ylgG (Fig. 5-3C), further strengthening the hypothesis that MutT3 performs DHN-3Pase activity in mycobacteria, and that folate levels underlie the biofilm and antifolate resistance phenotypes.

Next we tested whether MutT3 can bind GTP by radioactive nucleotide

184; 185 binding assay . A titration of purified MutT3Msm at saturating amounts of tritium-labeled GTP showed binding occurring in a MutT3-dependent manner (Fig.

5-6A). When the amount of tritiated GTP was titrated for defined amounts of purified MutT3Msm, binding was GTP-dependent and leveled out at plateaus (Fig. 5-

6B), indicating that saturation of binding sites was occurring. Finally, when purified

3 MutT3Msm was incubated with saturating levels of [8- H]-GTP and increasing

amounts of unlabeled GTP, cold GTP could outcompete labeled GTP (Fig. 5-6C).

These results confirmed specific binding of GTP to MutT3 and strengthened the

hypothesis that mutT3 encodes a Nudix hydrolase.

5.C.4. RplM, a novel substrate of PknG, co-purifies with MutT3.

To investigate whether MutT3 was directly regulated by PknG phosphorylation, in vitro radioactive kinase assays were performed using purified

proteins. PknGMtb autophosphorylates, (Fig. 5-7A, lane 1), which could be inhibited

by addition of the specific inhibitor AX20017 (Fig. 5-AB, lane 2). While no substrate

phosphorylation was observed when using MutT3Msm-6xHis that had been recombinantly expressed in E. coli (Fig 5-AB, lane 9), a phosphorylated substrate

144

145

Figure 5-6. MutT3 binds nucleotides in vitro. Radioactive nucleotide binding

assay using tritium-labeled GTP and purified MutT3Msm-6xHis. Radioactivity retained on membranes binding protein but not unbound nucleotides was measured in counts per minute (CPM) detected by liquid scintillation counting. Error bars represent standard deviations from triplicate experiments. (A) Saturating levels of tritium-labeled GTP (2.5µM) were incubated with 0 to 50µg purified MutT3. (B) 10 or 20µg MutT3 were incubated with 0 to 25µM tritium-labeled GTP. (C) 20µg MutT3 were incubated with 25µM labeled and 0 to 50µM unlabeled GTP.

146

could be detected by autoradiograph when using MutT3Msm-6xHis purified from M. smegmatis (Fig. 5-7A, lane 7). This band however displayed a lower molecular weight compared to MutT3Msm-6xHis (Fig. 5-7A, lane 7 top panel vs. bottom panel).

When large amounts of the purified fraction were loaded, it could be detected by

Coomassie Blue stain (data not shown), suggesting that it was a peptide that co- purified with MutT3 from M. smegmatis lysate. Since no band was observed for

comparable fractions of M. smegmatis extract not expressing 6x-His-tagged

MutT3Msm, this co-purification seemed to occur in a specific, MutT3-dependent manner rather than through non-specific binding to the Nickel resin (Fig. 5-7A, lane

8). LC/MS/MS analysis of the MutT3Msm-6xHis fraction obtained from M. smegmatis identified 15 proteins associated with MutT3, 5 of which were ribosomal (Table 5-1).

The two proteins that were predicted to have a similar molecular weight as the detected PknG-phosphorylated band on the autoradiograph, RplM and SmpB, were selected for recombinant expression and purification in E. coli. While no phosphorylation was detectable for in vitro radioactive kinase assays using

SmpBMsm-6xHis (Fig. 5-7B, lane 4), RplMMsm-6xHis was phosphorylated by PknG (Fig.

5-7A, lane 6 and Fig. 5-7B, lane 2). Removal of the 6xHis-tag by thrombin cleavage

resulted in perfect alignment with the original band observed as co-purified with

MutT3Msm-6xHis (Fig. 5-7A, lane 4). Inhibition of PknG activity abolished RplM phosphorylation, indicating specific phosphorylation of RplMMsm by PknG (Fig. 5-7A,

lane 5). RplMMtb is likewise phosphorylated by PknGMtb (Fig. 5-9B, lane 3) and PknG

147

Table 5-1. MutT3 co-purified proteins identified by LC/MS/MS.

Peptides N Name (95%) Cupin domain protein OS=Mycobacterium smegmatis (strain ATCC 1 700084 / mc(2)155) GN=MSMEG_5707 PE=4 SV=1 25 30S ribosomal protein S19 OS=Mycobacterium sp. (strain MCS) GN=rpsS 3 PE=3 SV=1 6 Pseudouridine synthase OS=Mycobacterium smegmatis (strain ATCC 4 700084 / mc(2)155) GN=rluB PE=3 SV=1 5 5 Putative uncharacterized protein OS=Plasmid pAL5000 PE=4 SV=1 4 Urease accessory protein UreE 1 OS=Mycobacterium smegmatis (strain 6 ATCC 700084 / mc(2)155) GN=MSMEG_1091 PE=4 SV=1 4 Pantothenate kinase OS=Mycobacterium smegmatis (strain ATCC 700084 7 / mc(2)155) GN=coaA PE=3 SV=1 3 Transcriptional regulator, TetR family protein OS=Mycobacterium smegmatis (strain ATCC 700084 / mc(2)155) GN=MSMEG_2153 PE=4 8 SV=1 4 50S ribosomal protein L27 OS=Mycobacterium smegmatis (strain ATCC 9 700084 / mc(2)155) GN=rpmA PE=1 SV=1 4 SsrA-binding protein OS=Mycobacterium smegmatis (strain ATCC 10 700084 / mc(2)155) GN=smpB PE=1 SV=1 3 11 Putative transposase OS=Mycobacterium smegmatis PE=4 SV=1 2 IS1549, transposase OS=Mycobacterium smegmatis (strain ATCC 700084 / 11 mc(2)155) GN=MSMEG_0074 PE=4 SV=1 2 Putative uncharacterized protein OS=Mycobacterium smegmatis (strain 12 ATCC 700084 / mc(2)155) GN=MSMEG_4029 PE=4 SV=1 2 50S ribosomal protein L16 OS=Mycobacterium smegmatis (strain ATCC 13 700084 / mc(2)155) GN=rplP PE=1 SV=1 1 50S ribosomal protein L13 OS=Mycobacterium smegmatis (strain ATCC 14 700084 / mc(2)155) GN=rplM PE=1 SV=1 1 50S ribosomal protein L21 OS=Mycobacterium smegmatis (strain ATCC 15 700084 / mc(2)155) GN=rplU PE=1 SV=1 1

Bolded proteins were of comparable size to observed band and were therefore recombinantly expressed in

E. coli and tested in hot in vitro kinase assays as substrates of PknG.

148

149

Figure 5-7. RplM, a novel substrate of PknG, co-purifies with MutT3. (A) In vitro

kinase assay using radioactive [γ-32P]-ATP to assess phosphorylation by PknG. 0.5µg

purified PknGMtb-6xHis (with or without 15 minute pre-incubation with 5mM

AX20017) were incubated with 1.5µg of purified substrate candidate. Lane 1:

PknGMtb-6xHis, Lane 2: PknGMtb-6xHis + AX20017, Lane 3: RplMMsm-6xHis (purified

from E. coli (pVN844), Lane 4: PknGMtb-6xHis + RplMMsm (purified from E. coli

(pVN844)), Lane 5: PknGMtb-6xHis + RplMMsm (purified from E. coli (pVN844)) +

AX20017, Lane 6: PknGMtb-6xHis + RplMMsm-6xHis (purified from E. coli (pVN844)),

Lane 7: PknGMtb-6xHis + MutT3Msm-6xHis (purified from M. smegmatis (pVN823)),

Lane 8: PknGMtb-6xHis + control purification from M. smegmatis, Lane 9: PknGMtb-

6xHis + MutT3Msm-6xHis (purified from E. coli (pVN835)). (B) In vitro kinase assay

using radioactive [γ-32P]-ATP to assess phosphorylation by PknG. 0 or 0.5µg purified

PknGMtb-6xHis were incubated with 1.5µg of purified RplMMsm-6xHis or SmpBMsm-

6xHis. (C) Kinase assay using radioactive [γ-32P]-ATP and whole cell extract to assess specificity of RplM phosphorylation by PknG. Coomassie Blue stain (left) and autoradiograph (right) of hot kinase assay using 1mg whole cell extract of mc2155,

∆pknGMsm, ∆pknGMsm/pknGMtb or ∆mutT3Msm and 1.5µg RplM-6xHis followed by

6xHis-pulldown to enrich for 6xHis-MutT3.

150

Figure 5-8. MutT3 co-purified with RplM and is not regulated translationally.

(A) Interaction model of RplMMsm and MutT3Msm-6xHis in vivo. RplM either directly interacts with MutT3 or their association is mediated by a third-party partner. (B)

Western blot assaying mutT3 expression in M. smegmatis strains. Total protein

2 extract of strains mc 155/mutT3Msm-6xHis and ∆pknGMsm/mutT3Msm-6xHis were

separated into soluble and insoluble fractions and probed with penta-His antibody

(Qiagen). No difference in expression levels was observed in the absence of PknG.

151

inhibition also abolished phosphorylation (Fig. 5-8B, lane 4). RplM, the large

ribosomal subunit protein L13, is therefore phosphorylated by PknG in both

pathogenic and non-pathogenic mycobacterial species in vitro. These data show that

MutT3 associates with RplM in a complex involving other proteins, and that RplM is

a novel substrate of PknG in mycobacteria.

To investigate if MutT3 levels in the cell are regulated at a translational level,

2 Western blot assays were performed using strains mc 155/mutT3Msm-6xHis and

∆pknGMsm/mutT3Msm-6xHis in which 6xHis-tagged MutT3Msm is produced under the control of the native mutT3 promotor. No difference in protein levels was observed upon deletion of pknG, thereby excluding translational control exerted by PknG (Fig.

5-8B).

In order to test whether phosphorylation of RplM is PknG-specific, total cell

2 extract was prepared from M. smegmatis mc 155, the ∆pknGMsm, ∆mutT3Msm and

∆pknGMsm/pknGMtb strains and used in hot kinase assays with purified RplMMsm-

6xHis followed by 6xHis pull-down. A faint band at the correct molecular weight for

RplMMsm-6xHis was observed on autoradiographs with all extracts except the

∆pknGMsm mutant (Fig. 5-7C), indicating that native PknG at natural expression

levels can phosphorylate RplM. 6xHis-tagged RplM expressed in M. smegmatis from

the heat-shock promotor could not be detected, most likely due to too low

expression levels (data not shown). This result indicates that phosphorylation of

RplM is PknG-specific.

152

5.C.5. Phosphorylation of RplM by PknG on threonine 11 mediates regulation of folate metabolism.

Of five potentially phosphorylated serine or threonine sites identified by

LC/MS/MS in the sequence of RplMMsm, three were conserved across the

mycobacterial genus (Fig. 5-9A, asterisks). A triple point mutant of RplMMtb in

which the three conserved residues threonine 11, threonine 12 and serine 14 were

167; 179; 180 mutated to alanines, which cannot accept phosphate groups . This RplMMsm

T11A,T12A,S14A could not be phosphorylated by PknGMtb, indicating that phosphorylation occurs on one or more of these residues (Fig. 5-9B, lane 5). Single

point mutants were therefore cloned and tested in hot kinase assays, showing

complete loss of phosphorylated signal upon mutation of threonine 11 to alanine

(Fig. 5-9B, lane 6), and partial loss upon mutation of threonine 12 (Fig. 5-9B, lane 7), likely due to steric interactions interfering with PknG access to the immediately adjacent threonine 11. Mutation of serine 14 did not affect phosphorylation levels

(Fig. 5-9B, lane 8). These data indicate that phosphorylation of RplM by PknG occurs specifically on the conserved threonine 11 residue.

To test whether phosphorylation of RplM mediates PknG regulation of folate metabolism and biofilm growth, the gene encoding RplM was replaced with alleles in which threonine 11 was changed to either alanine (T11A), which could not accept phosphate groups 167; 179; 180, or glutamate (T11E), mimicking constitutive phosphorylation 167; 179; 180. This was achieved by recombineering method in M.

153

Figure 5-9. RplM is phosphorylated on threonine 11 by PknG in vitro. (A)

Sequence alignment of RplM homologs. (B) In vitro kinase assay using radioactive

32 [γ- P]-ATP to assess phosphorylation by PknG. 0.5µg purified PknGMtb-6xHis (with

or without 15 minute pre-incubation with 5mM AX20017) were incubated with

1.5µg of purified RplMMtb and its derived point mutants.

154

smegmatis mc2155, generating strains mc2155/rplMT11A, mc2155/rplMT11E (Fig. 5-

10A, B). Replacement of rplM with the point mutant alleles was confirmed by sequencing and PCR (Fig. 5-10B).

2 M. smegmatis mc 155, its derived ∆pknGMsm mutant and strains

mc2155/rplMT11A and mc2155/rplMT11E were assayed biofilm growth assays (Fig. 5-

10C). Replacement with the non-phosphorylateable point mutant allele (strain

mc2155/rplMT11A) led to decreased surface biofilm formation, which mimicked the

defect observed in ∆pknGMsm. The requirement for native rplM could be bypassed by providing folinic acid in the medium (Fig. 5-10C). However, mc2155/rplMT11E also showed slighlty inhibited growth (Fig. 5-10C), indicating that while phosphorylation of RplM on threonine 11 is absolutely required for late stage growth, increased phosphorylation levels may be harmful for biofilm growth. These results strongly indicate that PknG mediates its regulation of folate control by balancing phosphorylation of RplM on threonine 11.

155

156

Figure 5-10. Phosphorylation of RplMMsm on threonine 11 is required for

survival of folate starvation and biofilm growth. (A) The rplMMsm replacement

2 cassette and rplMMsm locus in wild-type M. smegmatis mc 155 and its derived

2 T11A 2 T11E mc 155/rplM and mc 155/rplM strains. (B) PCR confirmation of rplMMsm

replacement using primers RMsm1 + RMsm2 (Table 3-1) shows shift corresponding to

insertion of the hygromycin resistance cassette. (C) M. smegmatis biofilm growth.

2 Biofilm growth of wild type M. smegmatis mc 155, its derived ΔpknGMsm, mc2155/rplMT11A and mc2155/rplMT11E strains were assayed in the presence or absence of 1mM folinic acid (FA). Biofilm was pictured at 1 week (top). Surface biofilm growth was quantified at 1 week by measuring total protein per biofilm plate (bottom).

157

5.D. Discussion

We have previously demonstrated that the eukaryotic-like serine/threonine

kinase G (PknG) of pathogenic and non-pathogenic mycobacterial species mediates antibiotic tolerance in both non-pathogenic and pathogenic mycobacteria 171. In this study, we found that PknG is involved in regulation of de novo folate biosynthesis, especially in late stage growth similar to the persistence of pathogenic mycobacteria inside the host macrophage. Furthermore, we have shown that PknG-dependent failure of host macrophages to deliver lysosomes to phagosomes 20 is mediated by

PknG through regulation of mycobacterial folate metabolism. These findings

question the proposed function of PknG as a virulence factor that modifies host

signaling networks in phagolysosome synthesis 20. We have identified a novel,

mycobacterial substrate of PknG, RplM, an essential protein that is a component of

the large ribosomal subunit 157. We showed that PknG phosphorylation of RplM

occurs on the mycobacterially conserved residue threonine 11 is important for a

continuation of the signaling cascade leading to MutT3. We provide preliminary

evidence that MutT3 functions as a nucleoside diphosphate hydrolase (Nudix)

enzyme. We hypothesize that MutT3 functions as a dihydroneopterin-triphosphate

pyrophosphatase (DHN-3Pase) that catalyzes the first committed step of folate de

novo biosynthesis. Our data suggested that this signaling axis is required for

mycobacterial resistance to antifolates and biofilm formation.

158

Based on our data, we propose a model in which PknG, upon receiving

signals for a folate requirement such as seen in growth in macrophages or biofilms, phosphorylates RplM (Fig. 5-11), which forms a complex with MutT3. The phosphorylation event affects MutT3 activity, possibly by either promoting or preventing efficient assembly of the MutT3 and RplM involving complex (Fig. 5-11).

The nature of this complex remains unclear, but based on LC/MS/MS analysis of proteins co-purified with MutT3, this complex involves several ribosomal proteins

that may localize on the ribosome itself or on a sub-ribosomal complex (Model A, Fig.

5-11), or on an independent complex (Model B, Fig. 5-11) 160. MutT3 likely encodes a

Nudix pyrophosphohydrolase activity 143; 144 directly or indirectly affecting folate

metabolism that results in changes in folate availability. Since the absence of MutT3

can be complemented by in trans expression of YlgG (Fig. 5-3 and Fig. 5-4),

previously shown to catalyze dihydroneopterin-triphosphatase activity, and since deletion of DHN-3Pase (NtpA) in E. coli mimics phenotypes observed in mycobacteria, MutT3 likely acts as a DHN-3Pase in mycobacterial de novo folate biosynthesis 86; 147; 148.

Our previous results suggested that regulation of folate metabolism by PknG

is involved in lysosomal localization in infected macrophages (Chapter 4).

Knowledge of substrates and precise molecular mechanisms underlying this

observation may provide a basis for identification of inhibitors to this pathway. Such

inhibitors, along with inhibitors of PknG, may also be used to potentiate classical

159

antifolates as well as to promote the innate mycobactericidal of host macrophages 45.

This signaling axis thus represents a novel target for development of antifolate potentiators that, when used together with antifolates, could also enhance the innate immunity of host macrophages against M. tuberculosis.

160

Figure 5-11. Model of folate biosynthesis regulation via PknG, RplM and MutT3.

(Model A) PknG phosphorylates RplM, localized on the large subunit of the ribosome, on threonine 11. Phosphorylated RplM is released from the ribosome and associates 161

with a complex that includes MutT3. Interaction of RplM and this complex in turn

leads to alterations in de novo folate biosynthesis via MutT3 Nudix hydrolase

activity. (Model B) PknG phosphorylates RplM, localized on the large subunit of the ribosome, on threonine 11. Phosphorylated RplM signals to MutT3, also localized on

the large ribosomal subunit, which in turn leads to alterations in de novo folate

biosynthesis via MutT3 Nudix hydrolase activity.

162

5.E. Acknowledgements.

We thank Jean Pieters, Julia van Kessel, Graham F. Hatfull, Magdalena

Taracila, Robert Bonomo, Bing Liu, Piet deBoer, Anne Walburger, Cathy Carlin, and

Hoa Nguyen for providing materials and expertise. Support was provided by an NIH

R01 grant (AI087903), start-up funds from the School of Medicine, a STERIS

Infectious Diseases Research Grant, and a CFAR Developmental Award from Case

Center For AIDS Research.

163

Chapter 6.

Discussion and Future Directions

164

6.A. Future Experiments

6.A.1. Hierarchy of PknG, RplM and MutT3

We have demonstrated that PknG, RplM, and MutT3 all are required for

survival of folate starvation and for biofilm formation, and thus likely act together in a folate metabolic pathway. While preliminary inferences can be made from our data supporting this hierarchy, more evidence will be needed to determine the

precise relationship among them. We proposed in our model (Fig. 5-11) that PknG

acts first, phosphorylating RplM, which then modulates MutT3 activity in some way.

In favor of our model, we demonstrated that the kinase activity of PknG itself

(Fig. 4-2B, Fig. 4-3B and Fig. 4-4B) and its phosphotransfer to threonine 11 of RplM

(Fig. 5-10 C, D) are specifically required for normal folate metabolism to occur in the

cell. However, a feedback loop in which phosphorylation of RplM as part of the

ribosome 157, or even as an anti-terminator 161, may regulate PknG levels and therefore PknG activity on RplM or other substrates, by way of translational or transcriptional control, respectively, cannot be excluded.

MutT3 likely acts downstream of PknG phosphorylating RplM, as it is not modified by PknG directly (Fig. 5-7A) but associates with RplM (Fig. 5-8A, Table 5-1).

MutT3 is not required for phosphorylation of RplM by PknG to occur in cellular extracts (Fig. 5-8B), further strengthening the hierarchy proposed in our model.

Likewise, MutT3 is not regulated transcriptionally or translationally by PknG (Fig. 5-

165

8B), indicating that MutT3 activity is regulated by formation of a complex with RplM

rather that ribosomal or anti-termination activity of RplM.

To definitively determine the order of events and interactions of these three

proteins in the regulation of folate biosynthesis, we propose experiments in which

2 T11A 2 T11E strains ∆mutT3Msm, mc 155/rplM and mc 155/rplM will be tested in

antifolate susceptibility (see section 2.F.) and biofilm assays (see section 2.I.) alone or in combination with the specific PknG kinase inhibitor AX20017 20. If PknG kinase

action is the first event in the pathway as we propose, and if no feedback loop exists,

inhibition of PknG should not aggravate or otherwise alter the levels of

susceptibility to folate starvation or impairment of biofilm in comparison to

untreated strains. If however either MutT3 or RplM act before PknG and mediate

feedback control over PknG, PknG inhibition should lead to more severe phenotypes

2 T11A in ∆mutT3Msm or mc 155/rplM that mimics constitutively non-phosphorylated

RplM 181; 198; 199. On the other hand, strain mc2155/rplMT11E, which mimics constitutively phosphorylated RplM 181; 198; 199, should not exhibit any difference in phenotype regardless of treatment.

Additionally, we will assay total protein extracts of H37Rv and ∆mutT3Mtb for

expression levels of PknG to address either a feedback loop or a constitutive

mechanism in which MutT3 Nudix hydrolase activity impacts transcriptional or

translational regulation of pknG.

166

6.A.2. RplM point mutant alleles in M. tuberculosis and effect on lysosomal trafficking

We have demonstrated that phosphorylation of RplM on threonine 11 by

PknG is required for survival of folate starvation and biofilm growth in M. smegmatis

(Fig 5-10 C, D). To confirm that this is also true for pathogenic mycobacteria, we are

T11A T11E currently working on generating strains H37Rv/rplM and ∆pknGMtb/rplM in

M. tuberculosis by specialized transduction 174 (see section 2.C.2.). Once these strains are obtained, they will be subjected to antifolate susceptibility (see section 2.E.) and biofilm assays (see section 2.H) to assess their phenotypes and whether they correspond to what we observed for M. smegmatis.

We propose that the phosphorylation event mediates folate metabolic control in latent growth conditions and that this in turn regulates the ability of M. tuberculosis to establish its intracellular localization 20 (see section 2.K. and Fig. 4-5).

T11A T11E Strains BCG/GFP/rplM and ∆pknGBCG/GFP/rplM constructed by specialized transduction (see section 2.C.2.) will be used in a lysosomal trafficking assay using murine bone-marrow derived macrophages (see section 2.K.) that would allow us to determine whether phosphorylation states of RplM are required for phagosomal localization of pathogenic mycobacteria independently of PknG. Strain

BCG/GFP/rplMT11A should exhibit increased lysosomal delivery in the presence of

T11E PknG, whereas ∆pknGBCG/GFP/rplM should remain localized in the phagosome in the absence of PknG. Furthermore, folate supplementation should rectify the

167

T11A phagosomal localization defect of BCG/GFP/rplM , as seen with ∆pknGBCG/GFP,

T11E whereas sulfonamide treatment should send ∆pknGBCG/GFP/rplM to lysosomes.

6.A.3. Subcellular localization of RplM/MutT3-containing complexes

Since loss of PknG does not lead to alterations in MutT3 expression levels

(Fig. 5-7 D), phosphorylation of RplM does not likely affect MutT3 expression levels.

We failed to observe direct interaction between purified MutT3Msm and RplMMsm in vitro (data not shown). Taken together with the identification of a number of other co-purified proteins in MutT3 purification fractions by LC/MS/MS, several of which, like RplM, are ribosomal, it is likely that MutT3 localizes on a larger complex on which it comes into (in)direct contact with RplM, the phosphorylation status of which regulates MutT3 activity. The nature of this complex remains elusive and will have to be further investigated. The complex on which MutT3 performs its Nudix hydrolase activity could be ribosomal or sub-ribosomal in nature. We proposed two possible models (Fig. 5-11) that involve different complexes.

Model A proposes that the phosphorylation of threonine 11 on RplM leads to a specific release of phospho-RplM from the ribosome, followed by association with a complex involving MutT3 that regulates de novo folate biosynthesis (Fig. 5-11).

This model is based on the interferon-γ-activated inhibitor of translation (GAIT) complex found in eukaryotes, in which L13a, the human homolog of RplM, is phosphorylated and thereby released from the ribosome, making possible association with the GAIT system 200; 201; 202. The GAIT system proposes that L13a is

168

not required for translation by the ribosome, and that the ribosome serves as a

depot for L13a until it is phosphorylated on a specific serine residue by death-

associated kinase-1 201; 202. Free phospho-L13a can then bind specifically to the

other components of the GAIT complex and repress translation of specific genes 200;

202. Similarly, RplM may be “stored” on the ribosome in mycobacteria until PknG

phosphorylates its threonine 11 in response to conditions requiring folate synthesis.

L13 is exclusively surface-associated without burrowing into the large ribosomal subunit in archaea 162; 203, and RplM of E. coli does not localize closely to the areas of the ribosome that bind message or are catalytically active (Fig. 1-9) 162; 163. It is likely

that RplM does not structurally support the ribosome, and can therefore be released

upon phosphorylation to perform extra-ribosomal functions, such as being an

essential component of the GAIT complex in eukaryots 200; 201; 202, acting as part of an anti-termination complex in bacteria 161, or as part of our proposed folate- synthesizing complex involving MutT3 (Fig 5-11).

Model B on the other hand proposes that both MutT3 and RplM localize on the large ribosomal subunit either as part of the ribosome or a sub-ribosomal complex (Fig 5-11). Some have been shown to co-fractionate with the ribosome where they regulate growth and stress responses 164; 165 and even show increased hydrolytic activity in the presence of ribosomal particles 204. This suggests that MutT3 may require ribosomal contact for performance of its activity, and this contact may only be possible upon phosphorylation of RplM, an event that may alter

169

the ease with which MutT3 binds to the complex. Alternatively, MutT3 activity could

be activated by release from the complex, in which setting the phosphorylation

event of RplM may facilitate this release by altering affinities of neighboring

proteins or binding pockets that directly interact with MutT3.

To confirm whether MutT3 indeed localizes on sub-ribosomal particles, and

whether the ability to localize is determined by PknG phosphorylation, we propose

2 to perform subcellular fractionations of strains mc 155/pVN866, ∆pknGMsm/pVN866,

mc2155/rplMT11A/pVN866 and mc2155/rplMT11E/pVN866 that express 6x-His-

tagged MutT3Msm from the native MutT3 promotor. Ribosomal fractions will be prepared by standard protocols for mycobacterial samples using ultracentrifugation at 104,500 x g 205; 206 and assayed by Western blot for MutT3-6xHis (see section 2.E),

PknG (see section 2.E), and soluble as well as ribosomal controls. This assay will

allow us to determine whether MutT3 can associate with the ribosome or sub-

ribosomal complexes at all and whether association is promoted, decreased or

unaffected by phosphorylation of threonine 11 on RplM by PknG.

To test the hypothesis that RplM is released from the large ribosomal subunit,

2 2 subcellular fractionations of mc 155, mc 155 treated with AX20017, ∆pknGMsm, mc2155/rplMT11A and mc2155/rplMT11E will be prepared and ribosomal as well as

soluble fractions probed for presence of RplM by Western blot. This will allow us to

determine whether RplM can be released from the ribosome (wild-type), whether

PknG activity is required for the release (AX20017 treatment, ∆pknGMsm), and

170

whether phosphorylation of threonine 11 is the essential release trigger

(mc2155/rplMT11A, mc2155/rplMT11E).

Should we find that RplM indeed is released from the ribosome to associate with MutT3 on a folate-synthesizing complex, we will seek to identify this complex by cross-linking experiments followed by pull-down of 6xHis-tagged MutT3 and submitting the pulled down materials to LC/MS/MS analysis.

6.A.4. Biochemical characterization of MutT3 enzymatic function

MutT3 contains the characteristic Nudix motif 156;143; 144, but does not appear

to function as a mutator protein or an 8-oxo-guanosine triphosphatase 156. The fact

that mutT3 is required for antifolate resistance and biofilm growth and that this

requirement could be bypassed by exogenous folate supplementation or expression of DHN-3Pases from other bacteria (Fig. 5-3 and Fig. 5-4) suggest that MutT3

performs the pyrophosphatase reaction converting dihydroneopterin-triphosphate

to dihydroneopterin-phosphate.

We have established a collaborative effort with the Amzel laboratory (Johns

Hopkins University, Baltimore, MD) that initially identified and characterized E. coli

NtpA 147. This collaborator will help us to characterize the DHN-3Pase activity of

MutT3 using their established methodology 147; 148; 207; 208; 135; 136. It is expected that several substrates will show moderate activity in this assay, as Nudix hydrolases tend to be promiscuous 143; 144, but a preference may be observed, as was the case

171

for E. coli NtpA 147 and L. lactis YlgG 147. This study will further substantiate the role

of MutT3 in folate metabolism (Fig. 5-11).

We previously tested GTP and ATP hydrolysis by MutT3, followed by

LC/MS/MS analysis to detect breakdown products and did not observe significant

activity of MutT3 with these substrates (data not shown). It is possible that contact

with the ribosome may stimulate GTPase or DHNTP pyrophosphatase activity, as

has been observed for the E. coli GTPase YjeQ 204. Phosphorylation of RplM may lead to improved binding or other changes facilitating MutT3 function, and loss of MutT3 or PknG would therefore lead to similar phenotypes. This may be a regulation mechanism specific to mycobacteria, since PknG is not widely conserved throughout more distantly related bacteria such as E. coli 140.

If the previously proposed in vitro characterization fails to detect activity, we will repeat hydrolysis assays adding purified ribosomes to stimulate pyrophosphatase activity and monitor hydrolysis by LC/MS/MS with the help of Dr.

Guofang Zhang in the Nutrition Department (see also section 6.A.6). If ribosomal addition indeed leads to activity increases, we will try to modify the above- described colorimetric assay accordingly.

Furthermore, the Amzel laboratory will make efforts to obtain crystal structures of MutT3Msm and MutT3Mtb, which will allow for comparison and

superimposition of conformations and substrate binding pockets to E. coli

NtpA147 and other crystallized Nudix hydrolases to help elucidate possible

172

substrates and mechanisms. Crystal structures could also help attempts to intelligently design inhibitors aimed at potentiation of classical antifolate antibiotics for treatment of MDR and XDR TB (see section 1.C.2.)

6.A.5. MutT3 deletion effects on lysosomal trafficking

We have demonstrated that mutT3Msm and mutT3Mtb are required for survival of folate starvation induced by classical antifolate antibiotics and biofilm growth (Fig 5-3 and Fig. 5-4). We further postulate that MutT3 is a Nudix hydrolase likely involved in folate de novo biosynthesis by acting as a DHN-3Pase (see Fig. 5-11 and summary in section 6.A.4), or otherwise mediates folate level control through

Nudix function with different substrates, e.g. by controlling growth via GTP or ppGpp level modulation (see section 1.F.), as part of an epigenetic network also involving PknG and RplM. We propose that MutT3 action mediates folate metabolic control in latent growth conditions and that this in turn regulates the lysosomal delivery to phagosomes containing M. tuberculosis in infected macrophages 20 (see section 2.L. and Fig. 4-5). We propose to generate strain ∆mutT3BCG/GFP by

specialized transduction (see section 2.C.2.) for use in a lysosomal trafficking assay

using murine bone-marrow derived macrophages (see section 2.L.) that would

allow us to determine whether mutT3BCG is involved in lysosomal trafficking and whether this is due to folate requirements of the bacterium in the macrophage. We predict that ∆mutT3BCG/GFP will exhibit increased lysosomal delivery comparable to

∆pknGBCG/GFP and that wild type characteristics will be restored when exogenous

173

folate is provided. This would also allow us to determine whether inhibition of

MutT3 may be a valuable concept to consider for potentiation of classical antifolate antibiotics (see. Section 1.C.2.).

6.A.6. Intracellular folate levels

We have observed that supplementation of exogenous folate bypasses requirements for PknG, RplM phosphorylation and MutT3 for survival of folate starvation (Fig. 4-1, Fig. 4-2, Fig. 5-3, Fig. 5-4 and Fig. 5-10) and for biofilm formation (Fig. 4-3, Fig. 4-4, Fig. 5-3, Fig. 5-4 and Fig. 5-10), and the requirement for

PknG to prevent lysosomal delivery to phagosomes in infected macrophages (Fig. 4-

5). We have also shown that folate starvation leads to a similar increase in lysosomal delivery as observed for pknGBCG deletion (Fig. 4-5). We have therefore

proposed that PknG, RplM and MutT3 act in a pathway regulating intracellular folate

levels in static growth conditions in mycobacteria, likely by regulation of MutT3

DHN-3Pase activity in folate de novo biosynthesis. Loss of any of the three proposed

components of this regulatory pathway would therefore lead to deregulation and

reduction of folate levels in static growth conditions such as biofilm, and we propose

to confirm altered intracellular folate levels in strains ∆pknGMsm, ∆mutT3Msm, and

mc2155/rplMT11A.

Total folate levels will be quantified by a microbiological assay employing a

strain of L. casei that cannot synthesize folate de novo and that is therefore

dependent on exogenous supplementation of folate 209; 210.

174

To analyze individual folate species levels inside the mycobacterial cell upon

loss of pknG or mutT3 or the ability to phosphorylate threonine 11 on RplM, the

above-listed strains will be grown in biofilm assays (see section 2.I.) and used for

extraction of intracellular folate species and analyis by LC/MS/MS.

6.B. Discussion

6.B.1. Significance of folate during mycobacterial colonization of host macrophages

Folates are essential co-enzymes in one-carbon unit transfer that, among other things, are involved in the biosynthesis of purines, thymidines, glycine, panthotenate and methionine, as well as bacterial formyl-methionyl-tRNA critical for initiation of transcription 77; 78; 79; 45 (see section 1.C.). Consequently, folate

molecules are essential for the synthesis of macromolecules such as nucleic acids

and proteins, and both deregulation of folate levels and folate starvation affect

almost all cell division 45. Yet another effect of folate starvation manifests itself in interference with recycling of both homocysteine (Hcy, Fig. 1-4) and S-adenosine methionine (SAM) 45. Folates are usually considered to be of particular importance during periods of rapid cell division and growth 77; 78 and a requirement for folate

biosynthesis during late-stage rather than exponential growth phases may appear

counter-intuitive at first. However, while intracellular mycobacteria in latent TB

infections are thought to be non-replicating and exhibit only very slow growth 193;

194; 195, they are metabolically active 196 and should therefore require folates for the

175

essential one-carbon transfer reactions 45; 77; 78; 79. In fact, our results have indicated that modulation of mycobacterial folate biosynthesis and availability by PknG are of particular importance in such conditions of late stage or stationary growth. Support for this hypothesis is provided by observations that pknG is among the genes sharply upregulated during immune suppression and reactivation in mice infected with M. tuberculosis, and that pknG is likely required for stationary phase metabolism 196; 197. The requirement for folate biosynthesis and PknG regulation

thereof in the context of the macrophage may be rooted in (a) mycobacterial

dormancy and persister formation requiring specific folate forms, (b) the stringent

response, triggered by folate starvation, to facilitate survival in the face of anti-

bacterial immune mechanisms, (c) immune evasion mechanisms requiring

metabolic adjustments, or in (d) cell wall and cell surface adjustments involved in

adaptation to the phagosomal environment.

6.B.1.a. Folate and cell dormancy

Other than the requirement for folate forms in basic cell cycle mechanisms

such as DNA synthesis, DNA replication and protein synthesis that occur in any

metabolically active cell, specific folate forms appear to be important for the

establishment of dormant, non-replicating cellular states 211; 212; 213. In particular, 5-

CHO-H4PteGlu (see Fig. 1-4), the most stable naturally occurring form of reduced

folate, has been suggested to function as a folate storage form that is necessary for

induction and survival of dormancy in Neurospora and soy beans 211; 212; 213.

176

Interestingly, lack of MTHFS in mycobacterial cells leads to elevation of

102 polyglutamylated forms of 5-CHO-H4PteGlu , and its homolog in E. coli, YgfA, is required for the formation of persisters that are non-replicating and display increased multiple antibiotic resistance 214, similar to what is thought to occur in TB

latency 15; 195. A role for MTHFS in intrinsic antifolate resistance was also confirmed in M. tuberculosis and Escherichia coli 102; 103, suggesting that this determinant functions ubiquitously among bacteria. Folate regulation by PknG may therefore

influence levels of particular folate forms that are required for M. tuberculosis to transition into a dormant, metabolically active, but not actively dividing, state that facilitates survival inside the host macrophage in chronic TB infections.

The current Wayne model of M. tuberculosis dormancy postulates that the cell progresses through two non-replicating stages, the first of which involves activation of glycine dehydrogenase 15; 194; 195. Since glycine biosynthesis requires folate co-enzymatic action, folate deficiency leads to glycine deficiency 77; 78; 79; and

the reaction catalysed by glycine dehydrogenase cannot occur, thus making the

transition from active growth to persistence ineffective or potentially even

impossible, and accounting for another reason that folate biosynthesis regulation by

PknG is necessary for successful colonization of macrophages by M. tuberculosis.

Another mechanism that has been shown to be important in MTB latency and

is triggered by antifolates 153 or otherwise caused folate deficiency is the stringent response, a stress response to amino acid starvation 108; 109. In this stress response,

177

the alarmone ppGpp is synthesized by RelA/SpoT to trigger growth arrest and

ensure cell survival until sufficient amino acid levels are encountered or

reestablished 150; 151. In mycobacteria, RelA synthesis of ppGpp is activated by the presence of uncharged tRNAs 215, which are formed during amino acid starvation, as in the case of folate starvation leading to non-functional glycine and methionine synthesis 77; 78; 79. It has been shown that relA is 5-fold upregulated in nutritionally

starved M. tuberculosis cultures that mimic dormancy inside the macrophage 216.

Interestingly, the Ndx8 Nudix hydrolase of Thermus thermophilus has been shown to

be involved in regulation of guanosine 3',5'-bispyrophosphate (ppGpp) levels independently of the RelA/SpoT synthesis and degradation cycles, degrading ppGpp to 3’,5’-bisphosphate (pGp), and potentially acting in a relay system with SpoT 149.

Ndx8 may thereby provide an additional mechanism to control the transition from growth arrest to active growth after amino acid starvation conditions have been removed 149. While we have proposed a DHN-3Pase function for MutT3, we cannot exclude alternative substrates for MutT3, and rather than directly controlling folate metabolism, a response to altered folate biosynthesis and folate starvation may be controlled through the PknG-RplM-MutT3 cascade by MutT3 modulation of ppGpp levels in the mycobacterial cell, also providing important control over persistence in host macrophages.

6.B.1.b. Immune evasion and cell wall maintenance

While PknG is expressed in non-pathogenic mycobacterial species and in the

178

related corynebacteria, expression levels are dramatically increased in pathogenic mycobacteria. There is evidence that PknG expression is translationally regulated

140, suggesting the existence of dedicated regulatory mechanisms to control

expression levels of PknG in mycobacterial species. It has been shown that PknG is

involved in the tricarboxylic acid cycle as well as glutamine utilization in

Corynebacterium glutamicum, which is used to produce and isolate large amounts of

glutamate industrially, and in vitro studies have shown that mycobacterial PknG can

also phosphorylate GarA, the glutamate regulator found to be its target in C.

glutamicum 121; 122; 124. However, glutamate levels in mycobacteria do not appear to be altered to significant amounts upon loss of pknG 141, and GarA does not play a role

in the folate mediated late stage growth functions we have observed (see Chapter 4).

It had therefore been suggested that elevated expression levels may have provided

the means for additional virulence functions of PknG in pathogenic mycobacteria 12;

20. This possibility was also proposed in a recent study that addressed differential

expression of PknG in various mycobacterial species in relation to its virulence

function 140. In light of our results, we also favor the hypothesis that GarA and

glutamate regulation do not represent the primary target of PknG in mycobacteria,

but we think it is probable that responses to folate requirements are a more likely

function for mycobacterial PknG. Glutamate may however have some impact, as it is

required for folate de novo biosynthesis (see Fig. 5-2).

Interestingly, it has been demonstrated that many bacteria that have adapted

179

to survival inside host macrophages or other immune cells induce stress responses

upon host cell infection, thereby inducing degradation and repair or replacement of

bacterial proteins damaged by antimicrobial immune mechanisms 217; 218. Since

folate starvation induces the ppGpp-mediated stringent response 153, and RplM has

been shown to associate with Obg in B. subtilis 165 and M. tuberculosis 166 to mediate

the σB-mediated stress response in Bacillus subtilis 164; 165, the PknG-RplM-MutT3 cascade that we propose to regulate folate metabolism could contribute to intracellular survival and immune evasion by means of stress response mediation.

As folate availability is essential for synthesis of proteins and other molecules that impact cellular building blocks, it is not surprising that regulation of folate metabolism and biosynthesis by PknG is involved with multiple antibiotic resistance and cell wall integrity of both non-pathogenic and pathogenic bacteria 157.

Loss of PknG function results in altered cell surface properties, likely due to changes in cell wall composition 157. Stationary phase growth, TB latency and biofilm growth have all been shown to exhibit higher levels of antibiotic resistance to a variety of drugs 22; 13; 14; 15; 16, which is likely due to decreased permeability of these

compounds through a thick, intact, potentially specifically adapted cell envelope 168;

189. Glutamine and its derived molecules, such as polyglutamines, may serve as

components that are required for the integrity of the mycobacterial cell wall as well,

thereby helping to pose an effective barrier to diffusion of antibiotics of diverse size

and polarity. PknG-mediated inhibition of glutamate reactions thus may also

180

contribute in ways not directly related to folate. It is possible that folate molecules

are simply necessary to maintain a normal cell wall, and therefore contribute to the

high intrinsic resistance to antibiotics found in mycobacteria, and likewise to lower

penetration rates of oxidative stress molecules generated in macrophages aiming to

degrade phagocytosed bacteria 217; 219. However, it has been shown that several

mycobacterial cell wall components actively modulate and suppress immune-

mediated mechanisms aimed at degradation of M. tuberculosis 217, and it is possible

that folates are required for synthesis of building blocks necessary to maintain the

needed amounts of these components as well as their delivery and proper

conformation on the cell surface.

More specifically, it has been demonstrated that one mechanism of immune

evasion in TB infection consists of early phagosomes containing M. tuberculosis

failing to recruit and activate hVPS34 220; 221, an effector kinase required for synthesis and accumulation of phosphatidylinositol-3-phosphate (PI(3)P) 222. PI(3)P signaling anchors later effectors that are required for phagosome maturation and

fusion to lysosomes in the phagosomal membrane 223; 224. M. tuberculosis actively modulates this failure to activate hPSV34 by shedding immunologically active lipoarabinomannan (LAM) and related lipoglycans, a lipid cell wall component that becomes distributed throughout the endocytic network of infected macrophages and prevents accumulation of Ca2+ in the host cytosol 225; 226. Lack of Ca2+

accumulation leads to failure to activate hVPS34 through calmodulin, and therefore

181

leads to failure to generate PI(3)P and an arrest in maturation at an early phagosome state 222; 225. Synthesis of LAM and related lipoglycans require the building block mannose, a sugar that is converted to GDP-mannose-1-phosphate

(GDP-manp) 227. If mannose cannot be taken up from extracellular sources, as is likely the condition inside host macrophages, mycobacteria obtain mannose-6-

phosphate from the glycolytic pathway and then convert it in a series of reactions to

GDP-manp 227. Interestingly, enzymatic activity of glycolytic enzymes in humans,

mice and rats has been shown to be regulated by folate availability 228. Folate level

regulation by the proposed PknG-RplM-MutT3 cascade may therefore be necessary

for synthesis of LAM and other immunologically active lipoglycans through

regulation of glycolysis, thereby mediating immune evasion and maturation arrest

of mycobacterially infected phagosomes.

6.B.2. Regulation of folate metabolism

We have proposed a model in which PknG phosphorylates RplM, likely upon receiving cues regarding folate requirements in stationary phase growth like persistence in the macrophage or biofilm formation (Fig. 5-11). In our model, RplM co-localizes on a multi-protein complex with MutT3, and the phosphorylation event leads to a change in the shared environment of these two proteins and consequently leads to altered MutT3 activity, possibly by either promoting or preventing efficient binding of MutT3 to the complex (Fig. 5-11). The complex may be ribosomal or sub- ribosomal in nature, and we propose that MutT3 acts as a Nudix

182

pyrophosphohydrolase 143; 144 to directly or indirectly affect folate metabolism and

lead to changes in folate availability. We believe the most likely function for MutT3

to be that of a DHN-3Pase in mycobacterial de novo folate biosynthesis 86; 147; 148, but we cannot exclude alternative substrates such as GTP or ppGpp.

While colonization of host macrophages represents a growth environment fairly specific to M. tuberculosis and other pathogenic mycobacteria, biofilm growth is not, and general folate regulation in late stage growth phases may be required in other bacteria as well. The investigated pathway involving phosphorylational control of a ribosomal protein and downstream regulation of pyrophosphohydrolase activity in the context of a multi-protein complex appears to function in both pathogenic and non-pathogenic mycobacterial species, and that this mode of control may not be a concept exclusive to the mycobacterial genus.

Beginning with PknG, which acts first in the hierarchy we have proposed, phosphorylational control of folate metabolism has to be considered. The eukaryotic-like serine/threonine kinases found in mycobacteria are involved in a variety of functions, including metabolic processes 132; 133; 134. PknG itself is

conserved throughout the entire mycobacterial genus, including M. leprae 140; 171, and a homolog has been identified in Corynebacterium glutamicum 121. It is

noteworthy that deletion of pknG in C. glutamicum leads to a set of phenotypes

distinct from those observed in mycobacteria; PknG seems to be primarily involved

in glutamate metabolic control in C. glutamicum 121 that remains mostly unaffected

183

in mycobacteria 122; 141. Its main function in mycobacteria likely involves regulation of folate synthesis and metabolism, thereby affecting cell wall maintenance 159, survival of folate starvation, biofilm growth, multiple antibiotic tolerance 159 and, in

the case of M. tuberculosis, it affects inhibition of phagolysosomal fusion in infected host macrophages 20. Sequence identity of PknG from C. glutamicum and M.

tuberculosis is only 46%

(http://tuberculist.epfl.ch/quicksearch.php?gene+name=pkng&submit=Search),

while sequence identity within the mycobacterial genus amounts to at least 80% 140;

171. Therefore, substrate panels may not be identical between these two genera,

accounting for the observed phenotypic discrepancies. Also in favor of this theory is

the fact that the phosphorylated residue threonine 11 of RplM is conserved only

within the mycobacterial genus, and is not present in corynebacteria or in any bacteria lacking a PknG homolog (see Fig. 5-10). Regulation of folate metabolism by

PknG therefore appears to be a mechanism specific to mycobacteria, which may represent specific growth and infection modes in response to stimuli specific to mycobacterial molecules affected by folate biosynthesis, such as cell wall components exclusive to the mycobacterial cell envelope.

General phosphorylational control and involvement of ribosomal proteins or complexes in regulation of folate metabolic processes, on the other hand, are much more likely to occur in bacteria other than the mycobacterial genus. While further studies concerning the nature of the multi-protein complex on which RplM and

184

MutT3 appear to colocalize are required, it is likely that this complex is ribosomal or sub-ribosomal in its composition, although RplM has been shown to be active in non-ribosomal contexts as well 161. While dormant, non-culturable mycobacterial

cells are metabolically active 196, a decrease in ribosome numbers has been reported in M. smegmatis 229 that could free up ribosomal subunits, sub-ribosomal complexes

or single ribosomal protein components for regulation of other metabolic processes

such as folate biosynthesis.

Interestingly, phosphorylation of ribosomal proteins occurs in distinct profiles in vegetatively growing Dictyostelium discoideum, and also in response to starvation of these amoeba compared to actively growing, non-starved cells 230.

Among these phosphorylated ribosomal proteins in Dictyostelium was L13 230, a homolog of RplM 145. Phosphorylation of ribosomal components including RplM has also been identified in Escherichia coli, although no analysis of involved kinases and effects of phosphorylation has been performed 167. Phosphorylation of ribosomal

proteins similarly has been shown to occur in mammalian cells. It thus appears that

L13 is among proteins that can be phosphorylated both in isolated subunits or

mono-ribosomes as well as in poly-ribosomes, which markedly decrease

phosphorylation of other ribosomal proteins 231.

Cumulatively, it would appear that RplM phosphorylation is required in both actively dividing cells as well as dormant cells, though different residues may be phosphorylated by distinct kinases, depending on required modulation in the

185

different growth stages. Protein kinase activities have been shown to co-purify with

ribosomal fractions of Streptomyces collinus, an actinobacterium related to

mycobacteria 232. A number of ribosomal proteins were phosphorylated primarily

on serine and threonine residues, and this seemed to result in decreased

translational activity of the isolated ribosomes 232. By similar mechanisms, phosphorylation of RplM by PknG may lower the translational activity of ribosomes, which may become more accessible to MutT3, or alternatively free up complexes for association with and function in other metabolic processes such as folate metabolism.

Folate-dependent enzymes can act as translational regulators and modulate ribosomal activity, as is the case for thymidilate synthase, which can bind specific messenger RNAs (mRNA) directly, including its cognate mRNA as well as p53 mRNA, in eukaryotes 233. Binding of mRNA leads to translational repression of the message

233, demonstrating that folates influence basic cellular processes in multi-faceted

ways.

While ribosomal protein phosphorylation and ribosomal complex

involvement have not previously been described in the context of the de novo folate biosynthesis, non-ribosomal multi-enzyme complexes have been described to

associate with 5,6,7,8-tetrahydrofolate in microorganisms capable of folate

biosynthesis, and interestingly, intermediary enzymes of the pathway associate with

this same complex 234. Likewise, enzymes, whose activity is controlled by folate

186

biosynthesis via the requirement for folates as co-enzymes, have been reported to

occur in multi-protein complexes. Such is the case for example with the glycine

decarboxylase multi-enzyme system of plant mitochondria that is involved in

degradation of glycine 235. Glycine is a molecule that is essential for transition of intracellular M. tuberculosis into dormancy and therefore successful persistence of

TB 15; 194; 195, and glycine synthesis is also folate-dependent (see also section 6.B.1.).

Clearly, folate reactions required for both folate biosynthesis and

consumption can occur in large multi-protein complexes and regulate various

metabolic and other cellular pathways. Ribosomal involvement or binding of

ribosomal proteins to other protein complexes could therefore represent a widely

used strategy of tight control of folate levels inside the cell, which in turn modulate almost all cellular processes. Phosphorylation of ribosomal proteins occurs across all kingdoms of life and therefore may be a convenient control mechanism for such folate level modulation, although it is likely that the involved kinases, phosphorylated residues, and their resulting effects on folate levels will vary among organisms as well as across the growth or stress conditions that require folate adjustments.

6.B.3. Implications for TB and MDR or XDR TB

Due to the growing spread of MDR and XDR TB infections 23; 36; 37 that cannot be treated with the standard first-line TB drugs 1; 36; 37; 38, we face an urgent need for

alternative treatment options.

187

One such alternative approach is the concept of “targeting resistance” 45, which seeks to use current knowledge of molecular resistance mechanisms employed by bacteria in order to (re)sensitize pathogens to already available, approved drugs 45. This is achieved by coadministration of existing drugs and

inhibitors suppressing resistance mechanisms to these drugs, so that currently

ineffective drugs (re)gain antimicrobial activity (Fig. 1-3) 45; 46; 47. Classical antifolate

antibiotics attack folate de novo biosynthesis in only two steps (Fig. 1-4), and no

potentiators are available, although drugs targeting both steps are commonly

combined to achieve synergy and appear to be a promising group of drugs to use in

TB infection 98; 99; 100. However, resistance can readily be gained by point mutations

in one or both of the enzymes targeted 80; 83, and therefore identification of

potentiators that inhibit intrinsic antifolate tolerance mechanisms in steps not

involving these enzymes could present a valuable strategy for alternative

antimycobacterial treatment development 45; 102.

We have presented here evidence in favour of a regulatory pathway in mycobacteria involving the serine/threonine protein kinase PknG, the ribosomal protein RplM and the putative Nudix hydrolase MutT3 (Fig. 5-11), the disruption of which leads to increased susceptibility to classical antifolate antibiotics. Inhibitors targeting either of these three proteins or other steps that may occur in the proposed pathway should therefore act as potentiators of antifolate antimycobacterial activity, and we have shown here that, at least in vitro, chemical

188

inhibition of PknG can potentiate sulfadrug action against M. tuberculosis (Fig. 4-1).

Pharmaceutical inactivation of this regulatory pathway as well as other pathways regulating folate availability could possibly also potentiate the effects of current anti-TB drugs that target folate metabolism (PAS, INH, etc.) 45; 102. Potentiation of the anti-TB efficacy of classical antifolates may also allow for reduction of effective therapeutic doses, thereby minimizing the cytotoxicity that has been a clinical problem 45; 102.

Detailed studies concerning RplM modulation of MutT3 activity, the nature of the multi-protein complex involved, as well as MutT3 enzymatic activity and substrates should yield further targets for inhibition of folate level control and antifolate potentiator development. Such inhibitors could be obtained by rational design or small molecule screens 236; 237; 238, and would then have to be tested for

potentiation in in vitro mycobacterial cultures and in in vivo settings, progressing from macrophage experiments through mice studies and clinical trials 44; 45; 239.

Clinical trials may be significantly shortened in comparison to those involving entirely new drug classes, however, since potentiator targets in folate de novo

biosynthesis are not present in humans and these molecules should therefore not

affect patients, unless they exhibit off-target effects 86; 87; 88. Antifolates have been

extensively studied 80; 83, are used on a standard basis for human treatment of

various diseases 80; 81; 82, and already have been characterized for side effects 240; 241;

189

242, which might even be further reduced by potentiation-enabled lower antifolate

doses necessary to obtain sufficient antimycobacterial activity of these drugs 45.

We have demonstrated the requirement for PknG-mediated folate metabolism regulation in preventing lysosomal delivery to mycobacteria containing phagosomes (Fig. 4-5), which indicates that co-administration of inhibitors to PknG and other key players in this regulatory pathway together with antifolate antibiotics could also promote phagosome-lysosome fusion, leading to improved killing and clearance of M. tuberculosis by host immune system mechanisms. Targeting PknG, its substrate RplM, and downstream proteins involved in folate metabolic control such as MutT3 may therefore effectively be doubly valuable and facilitate both drug- and immune-mediated antimycobacterial mechanisms, providing a very promising alternative MDR and XDR TB treatment option.

190

Bibliography

1. WHO, W. H. O.-. (2010). Tuberculosis Fact Sheet, Vol. 2011.

2. Cook, J. L. (2010). Nontuberculous mycobacteria: opportunistic

environmental pathogens for predisposed hosts. British medical

bulletin 96, 45-59.

3. Brennan, P. J. & Nikaido, H. (1995). The envelope of mycobacteria. Annu

Rev Biochem 64, 29-63.

4. Stewart, G. R., Robertson, B. D. & Young, D. B. (2003). Tuberculosis: a

problem with persistence. Nature reviews. Microbiology 1, 97-105.

5. Lin, P. L. & Flynn, J. L. (2010). Understanding latent tuberculosis: a

moving target. Journal of immunology 185, 15-22.

6. CDC, C. f. D. C. a. P.-. (2010). Fact Sheet - The Difference Between Latent

TB Infection and Active TB Disease Vol. 2011.

7. Diel, R., Loddenkemper, R. & Nienhaus, A. (2010). Evidence-based

comparison of commercial interferon-gamma release assays for

detecting active TB: a metaanalysis. Chest 137, 952-68.

8. (2000). Targeted tuberculin testing and treatment of latent

tuberculosis infection. This official statement of the American Thoracic

Society was adopted by the ATS Board of Directors, July 1999. This is a

Joint Statement of the American Thoracic Society (ATS) and the Centers

191

for Disease Control and Prevention (CDC). This statement was endorsed

by the Council of the Infectious Diseases Society of America. (IDSA),

September 1999, and the sections of this statement. American journal of

respiratory and critical care medicine 161, S221-47.

9. Selwyn, P. A., Alcabes, P., Hartel, D., Buono, D., Schoenbaum, E. E., Klein,

R. S., Davenny, K. & Friedland, G. H. (1992). Clinical manifestations and

predictors of disease progression in drug users with human

immunodeficiency virus infection. The New England journal of medicine

327, 1697-703.

10. Nguyen, L. & Pieters, J. (2005). The Trojan horse: survival tactics of

pathogenic mycobacteria in macrophages. Trends in cell biology 15,

269-76.

11. Blumberg, H. M., Leonard, M. K., Jr. & Jasmer, R. M. (2005). Update on

the treatment of tuberculosis and latent tuberculosis infection. JAMA :

the journal of the American Medical Association 293, 2776-84.

12. Pieters, J. (2008). Mycobacterium tuberculosis and the macrophage:

maintaining a balance. Cell host & microbe 3, 399-407.

13. Wallis, R. S., Palaci, M. & Eisenach, K. (2007). Persistence, not resistance,

is the cause of loss of isoniazid effect. The Journal of infectious diseases

195, 1870-1; author reply 1872-3.

192

14. Wallis, R. S., Patil, S., Cheon, S. H., Edmonds, K., Phillips, M., Perkins, M. D.,

Joloba, M., Namale, A., Johnson, J. L., Teixeira, L., Dietze, R., Siddiqi, S.,

Mugerwa, R. D., Eisenach, K. & Ellner, J. J. (1999). Drug tolerance in

Mycobacterium tuberculosis. Antimicrobial agents and chemotherapy

43, 2600-6.

15. Wayne, L. G. & Hayes, L. G. (1996). An in vitro model for sequential

study of shiftdown of Mycobacterium tuberculosis through two stages

of nonreplicating persistence. Infection and immunity 64, 2062-9.

16. Mitchison, D. A., Jindani, A., Davies, G. R. & Sirgel, F. (2007). Isoniazid

activity is terminated by bacterial persistence. The Journal of infectious

diseases 195, 1871-2; author reply 1872-3.

17. Chua, J., Vergne, I., Master, S. & Deretic, V. (2004). A tale of two lipids:

Mycobacterium tuberculosis phagosome maturation arrest. Current

opinion in microbiology 7, 71-7.

18. Tan, T., Lee, W. L., Alexander, D. C., Grinstein, S. & Liu, J. (2006). The

ESAT-6/CFP-10 secretion system of Mycobacterium marinum

modulates phagosome maturation. Cellular microbiology 8, 1417-29.

19. Vergne, I., Chua, J., Lee, H. H., Lucas, M., Belisle, J. & Deretic, V. (2005).

Mechanism of phagolysosome biogenesis block by viable

Mycobacterium tuberculosis. Proceedings of the National Academy of

Sciences of the United States of America 102, 4033-8.

193

20. Walburger, A., Koul, A., Ferrari, G., Nguyen, L., Prescianotto-Baschong, C.,

Huygen, K., Klebl, B., Thompson, C., Bacher, G. & Pieters, J. (2004).

Protein kinase G from pathogenic mycobacteria promotes survival

within macrophages. Science 304, 1800-4.

21. Bach, H., Papavinasasundaram, K. G., Wong, D., Hmama, Z. & Av-Gay, Y.

(2008). Mycobacterium tuberculosis virulence is mediated by PtpA

dephosphorylation of human vacuolar protein sorting 33B. Cell host &

microbe 3, 316-22.

22. McKinney, J. D. (2000). In vivo veritas: the search for TB drug targets

goes live. Nature medicine 6, 1330-3.

23. Nguyen, L. & Pieters, J. (2009). Mycobacterial subversion of

chemotherapeutic reagents and host defense tactics: challenges in

tuberculosis drug development. Annual review of pharmacology and

toxicology 49, 427-53.

24. Sacchettini, J. C., Rubin, E. J. & Freundlich, J. S. (2008). Drugs versus

bugs: in pursuit of the persistent predator Mycobacterium tuberculosis.

Nature reviews. Microbiology 6, 41-52.

25. Michele, T. M., Ko, C. & Bishai, W. R. (1999). Exposure to antibiotics

induces expression of the Mycobacterium tuberculosis sigF gene:

implications for chemotherapy against mycobacterial persistors.

Antimicrobial agents and chemotherapy 43, 218-25.

194

26. Liu, J., Barry, C. E., 3rd & Nikaido, H. (1999). Cell wall: physical structure

and permeability. In Mycobacteria Molecular Biology and Virulence

(Ratledge, C. & Dale, J., eds.), pp. 220-239. Blackwell Sciences Ltd.,

Oxford.

27. Brennan, P. J. (2003). Structure, function, and biogenesis of the cell wall

of Mycobacterium tuberculosis. Tuberculosis 83, 91-7.

28. Minnikin, D. (1982). Complex lipids: their chemistry, biosynthesis, and

roles. In The Biology of Mycobacteria (Ratledge, C., Standford, J., ed.), pp.

94-184. Academic, London.

29. Gokhale, R. S., Saxena, P., Chopra, T. & Mohanty, D. (2007). Versatile

polyketide enzymatic machinery for the biosynthesis of complex

mycobacterial lipids. Natural product reports 24, 267-77.

30. Liu, J., Rosenberg, E. Y. & Nikaido, H. (1995). Fluidity of the lipid domain

of cell wall from Mycobacterium chelonae. Proc Natl Acad Sci U S A 92,

11254-8.

31. Jarlier, V. & Nikaido, H. (1990). Permeability barrier to hydrophilic

solutes in Mycobacterium chelonei. J Bacteriol 172, 1418-23.

32. Li, X. Z. & Nikaido, H. (2004). Efflux-mediated drug resistance in

bacteria. Drugs 64, 159-204.

33. Li, X. Z. & Nikaido, H. (2009). Efflux-mediated drug resistance in

bacteria: an update. Drugs 69, 1555-623.

195

34. Louw, G. E., Warren, R. M., Gey van Pittius, N. C., McEvoy, C. R., Van

Helden, P. D. & Victor, T. C. (2009). A balancing act: efflux/influx in

mycobacterial drug resistance. Antimicrobial agents and chemotherapy

53, 3181-9.

35. Li, X. Z., Zhang, L. & Nikaido, H. (2004). Efflux pump-mediated intrinsic

drug resistance in Mycobacterium smegmatis. Antimicrobial agents and

chemotherapy 48, 2415-23.

36. Nguyen, L. & Thompson, C. J. (2006). Foundations of antibiotic

resistance in bacterial physiology: the mycobacterial paradigm. Trends

in microbiology 14, 304-12.

37. Borrell, S. & Gagneux, S. (2011). Strain diversity, epistasis and the

evolution of drug resistance in Mycobacterium tuberculosis. Clinical

microbiology and infection : the official publication of the European

Society of Clinical Microbiology and Infectious Diseases 17, 815-20.

38. WHO, W. H. O.-. (2006). Questions : XDR-TB, Vol. 2011.

39. Dye, C. (2000). Tuberculosis 2000-2010: control, but not elimination.

The international journal of tuberculosis and lung disease : the official

journal of the International Union against Tuberculosis and Lung Disease

4, S146-52.

40. Singh, J. A., Upshur, R. & Padayatchi, N. (2007). XDR-TB in South Africa:

no time for denial or complacency. PLoS medicine 4, e50.

196

41. Gandhi, N. R., Moll, A., Sturm, A. W., Pawinski, R., Govender, T., Lalloo, U.,

Zeller, K., Andrews, J. & Friedland, G. (2006). Extensively drug-resistant

tuberculosis as a cause of death in patients co-infected with

tuberculosis and HIV in a rural area of South Africa. Lancet 368, 1575-

80.

42. Jassal, M. & Bishai, W. R. (2009). Extensively drug-resistant tuberculosis.

Lancet Infect Dis 9, 19-30.

43. LoBue, P. (2009). Extensively drug-resistant tuberculosis. Curr Opin

Infect Dis 22, 167-73.

44. Bolten, B. M. & DeGregorio, T. (2002). From the analyst's couch. Trends

in development cycles. Nature reviews. Drug discovery 1, 335-6.

45. Wolff, K. A., Sherman, M. B. and Nguyen, L. (2011). Potentiation of

Available Antibiotics by Targeting Resistance - An Emerging Trend in

Tuberculosis Drug Development. In Drug Development - A Case Study

Based Insight into Modern Strategies (Rundfeldt, C., ed.). InTech.

46. Wright, G. D. (2000). Resisting resistance: new chemical strategies for

battling superbugs. Chem Biol 7, R127-32.

47. Wright, G. D. & Sutherland, A. D. (2007). New strategies for combating

multidrug-resistant bacteria. Trends Mol Med 13, 260-7.

197

48. Koch, A. L. (2003). Cell wall-deficient (CWD) bacterial pathogens: could

amylotrophic lateral sclerosis (ALS) be due to one? Critical reviews in

microbiology 29, 215-21.

49. Waxman, D. J., Yocum, R. R. & Strominger, J. L. (1980). Penicillins and

cephalosporins are active site-directed acylating agents: evidence in

support of the substrate analogue hypothesis. Philosophical

transactions of the Royal Society of London. Series B, Biological sciences

289, 257-71.

50. Drawz, S. M. & Bonomo, R. A. (2010). Three decades of beta-lactamase

inhibitors. Clinical microbiology reviews 23, 160-201.

51. Flores, A. R., Parsons, L. M. & Pavelka, M. S., Jr. (2005). Genetic analysis

of the beta-lactamases of Mycobacterium tuberculosis and

Mycobacterium smegmatis and susceptibility to beta-lactam antibiotics.

Microbiology 151, 521-32.

52. Quinting, B., Reyrat, J. M., Monnaie, D., Amicosante, G., Pelicic, V., Gicquel,

B., Frere, J. M. & Galleni, M. (1997). Contribution of beta-lactamase

production to the resistance of mycobacteria to beta-lactam antibiotics.

FEBS letters 406, 275-8.

53. Hugonnet, J. E. & Blanchard, J. S. (2007). Irreversible inhibition of the

Mycobacterium tuberculosis beta-lactamase by clavulanate.

Biochemistry 46, 11998-2004.

198

54. Tremblay, L. W., Hugonnet, J. E. & Blanchard, J. S. (2008). Structure of

the covalent adduct formed between Mycobacterium tuberculosis beta-

lactamase and clavulanate. Biochemistry 47, 5312-6.

55. Hugonnet, J. E., Tremblay, L. W., Boshoff, H. I., Barry, C. E., 3rd &

Blanchard, J. S. (2009). Meropenem-clavulanate is effective against

extensively drug-resistant Mycobacterium tuberculosis. Science 323,

1215-8.

56. Chambers, H. F., Moreau, D., Yajko, D., Miick, C., Wagner, C., Hackbarth,

C., Kocagoz, S., Rosenberg, E., Hadley, W. K. & Nikaido, H. (1995). Can

penicillins and other beta-lactam antibiotics be used to treat

tuberculosis? Antimicrob Agents Chemother 39, 2620-4.

57. Prabhakaran, K., Harris, E. B. & Randhawa, B. (1999). Bactericidal

action of ampicillin/sulbactam against intracellular mycobacteria.

International journal of antimicrobial agents 13, 133-5.

58. Chambers, H. F., Turner, J., Schecter, G. F., Kawamura, M. & Hopewell, P.

C. (2005). Imipenem for treatment of tuberculosis in mice and humans.

Antimicrob Agents Chemother 49, 2816-21.

59. Chambers, H. F., Kocagoz, T., Sipit, T., Turner, J. & Hopewell, P. C. (1998).

Activity of amoxicillin/clavulanate in patients with tuberculosis. Clin

Infect Dis 26, 874-7.

199

60. Dauby, N., Muylle, I., Mouchet, F., Sergysels, R. & Payen, M. C. (2011).

Meropenem/Clavulanate and Linezolid Treatment for Extensively Drug-

Resistant Tuberculosis. The Pediatric infectious disease journal.

61. Sood, A. & Panchagnula, R. (2003). Design of controlled release delivery

systems using a modified pharmacokinetic approach: a case study for

drugs having a short elimination half-life and a narrow therapeutic

index. International journal of pharmaceutics 261, 27-41.

62. Zimmerman, T. J., Kooner, K. S., Kandarakis, A. S. & Ziegler, L. P. (1984).

Improving the therapeutic index of topically applied ocular drugs.

Archives of ophthalmology 102, 551-3.

63. Burns, M. (1999). Management of narrow therapeutic index drugs.

Journal of thrombosis and thrombolysis 7, 137-43.

64. Flipo, M., Desroses, M., Lecat-Guillet, N., Dirie, B., Carette, X., Leroux, F.,

Piveteau, C., Demirkaya, F., Lens, Z., Rucktooa, P., Villeret, V., Christophe,

T., Jeon, H. K., Locht, C., Brodin, P., Deprez, B., Baulard, A. R. & Willand, N.

(2011). Ethionamide boosters: synthesis, biological activity, and

structure-activity relationships of a series of 1,2,4-oxadiazole EthR

inhibitors. Journal of medicinal chemistry 54, 2994-3010.

65. Morlock, G. P., Metchock, B., Sikes, D., Crawford, J. T. & Cooksey, R. C.

(2003). ethA, inhA, and katG loci of ethionamide-resistant clinical

200

Mycobacterium tuberculosis isolates. Antimicrobial agents and

chemotherapy 47, 3799-805.

66. Zhang, Y. (2005). The magic bullets and tuberculosis drug targets.

Annual review of pharmacology and toxicology 45, 529-64.

67. Vilcheze, C., Av-Gay, Y., Attarian, R., Liu, Z., Hazbon, M. H., Colangeli, R.,

Chen, B., Liu, W., Alland, D., Sacchettini, J. C. & Jacobs, W. R., Jr. (2008).

Mycothiol biosynthesis is essential for ethionamide susceptibility in

Mycobacterium tuberculosis. Molecular microbiology 69, 1316-29.

68. Vannelli, T. A., Dykman, A. & Ortiz de Montellano, P. R. (2002). The

antituberculosis drug ethionamide is activated by a flavoprotein

monooxygenase. The Journal of biological chemistry 277, 12824-9.

69. Brossier, F., Veziris, N., Truffot-Pernot, C., Jarlier, V. & Sougakoff, W.

(2011). Molecular investigation of resistance to the antituberculous

drug ethionamide in multidrug-resistant clinical isolates of

Mycobacterium tuberculosis. Antimicrobial agents and chemotherapy

55, 355-60.

70. Frenois, F., Engohang-Ndong, J., Locht, C., Baulard, A. R. & Villeret, V.

(2004). Structure of EthR in a ligand bound conformation reveals

therapeutic perspectives against tuberculosis. Molecular cell 16, 301-7.

71. DeBarber, A. E., Mdluli, K., Bosman, M., Bekker, L. G. & Barry, C. E., 3rd.

(2000). Ethionamide activation and sensitivity in multidrug-resistant

201

Mycobacterium tuberculosis. Proceedings of the National Academy of

Sciences of the United States of America 97, 9677-82.

72. Wang, F., Langley, R., Gulten, G., Dover, L. G., Besra, G. S., Jacobs, W. R., Jr.

& Sacchettini, J. C. (2007). Mechanism of thioamide drug action against

tuberculosis and leprosy. The Journal of experimental medicine 204, 73-

8.

73. Frenois, F., Baulard, A. R. & Villeret, V. (2006). Insights into mechanisms

of induction and ligands recognition in the transcriptional repressor

EthR from Mycobacterium tuberculosis. Tuberculosis 86, 110-4.

74. Willand, N., Dirie, B., Carette, X., Bifani, P., Singhal, A., Desroses, M.,

Leroux, F., Willery, E., Mathys, V., Deprez-Poulain, R., Delcroix, G.,

Frenois, F., Aumercier, M., Locht, C., Villeret, V., Deprez, B. & Baulard, A.

R. (2009). Synthetic EthR inhibitors boost antituberculous activity of

ethionamide. Nature medicine 15, 537-44.

75. Fraaije, M. W., Kamerbeek, N. M., Heidekamp, A. J., Fortin, R. & Janssen,

D. B. (2004). The prodrug activator EtaA from Mycobacterium

tuberculosis is a Baeyer-Villiger monooxygenase. The Journal of

biological chemistry 279, 3354-60.

76. Weber, W., Schoenmakers, R., Keller, B., Gitzinger, M., Grau, T., Daoud-El

Baba, M., Sander, P. & Fussenegger, M. (2008). A synthetic mammalian

202

gene circuit reveals antituberculosis compounds. Proceedings of the

National Academy of Sciences of the United States of America 105, 9994-8.

77. Blakley, R. L. (1969). The biochemistry of folic acid and related pteridines.

Frontiers of biology (Amsterdam), v. 13, North-Holland Publishing Co.,

Amsterdam, London,.

78. Green, J., Nichols, B. & Matthews, R. (1996). Folate biosynthesis,

reduction, and polyglutamylation. In Escherichia coli and Salmonella

typhimurium - Cellular and Molecular Biology. (Neidhardt, F., Curtiss III,

R., Ingraham, J., Lin, E., Low, J. K., Magasanik, B., Reznikoff, W., Riley, M.,

Schaechter, M. & Umbarger, H., eds.), pp. 665-673. ASM Press,

Washington, DC.

79. Selhub, J. (2002). Folate, vitamin B12 and vitamin B6 and one carbon

metabolism. J Nutr Health Aging 6, 39-42.

80. Bertino, J. R. & New York Academy of Sciences. Section of Biological and

Medical Sciences. (1971). Folate antagonists as chemotherapeutic

agents. Annals of the New York Academy of Sciences,, New York

Academy of Sciences, [New York].

81. Gorlick, R., Goker, E., Trippett, T., Waltham, M., Banerjee, D. & Bertino, J.

R. (1996). Intrinsic and acquired resistance to methotrexate in acute

leukemia. N Engl J Med 335, 1041-8.

203

82. Vinetz, J. M. (2010). Intermittent preventive treatment for malaria in

sub-saharan African: a halfway technology or a critical intervention?

The American journal of tropical medicine and hygiene 82, 755-6.

83. Libecco, J. A. & Powell, K. R. (2004). Trimethoprim/sulfamethoxazole:

clinical update. Pediatr Rev 25, 375-80.

84. Grim, S. A., Rapp, R. P., Martin, C. A. & Evans, M. E. (2005).

Trimethoprim-sulfamethoxazole as a viable treatment option for

infections caused by methicillin-resistant Staphylococcus aureus.

Pharmacotherapy 25, 253-64.

85. Proctor, R. A. (2008). Role of folate antagonists in the treatment of

methicillin-resistant Staphylococcus aureus infection. Clin Infect Dis 46,

584-93.

86. Bermingham, A. & Derrick, J. P. (2002). The folic acid biosynthesis

pathway in bacteria: evaluation of potential for antibacterial drug

discovery. BioEssays : news and reviews in molecular, cellular and

developmental biology 24, 637-48.

87. Gangjee, A., Jain, H. D. & Kurup, S. (2007). Recent advances in classical

and non-classical antifolates as antitumor and antiopportunistic

infection agents: part I. Anticancer Agents Med Chem 7, 524-42.

204

88. Gangjee, A., Jain, H. D. & Kurup, S. (2008). Recent advances in classical

and non-classical antifolates as antitumor and antiopportunistic

infection agents: Part II. Anticancer Agents Med Chem 8, 205-31.

89. Adrian, P. V. & Klugman, K. P. (1997). Mutations in the dihydrofolate

reductase gene of trimethoprim-resistant isolates of Streptococcus

pneumoniae. Antimicrobial agents and chemotherapy 41, 2406-13.

90. Volpato, J. P. & Pelletier, J. N. (2009). Mutational 'hot-spots' in

mammalian, bacterial and protozoal dihydrofolate reductases

associated with antifolate resistance: sequence and structural

comparison. Drug resistance updates : reviews and commentaries in

antimicrobial and anticancer chemotherapy 12, 28-41.

91. Baca, A. M., Sirawaraporn, R., Turley, S., Sirawaraporn, W. & Hol, W. G.

(2000). Crystal structure of Mycobacterium tuberculosis 7,8-

dihydropteroate synthase in complex with pterin monophosphate: new

insight into the enzymatic mechanism and sulfa-drug action. Journal of

molecular biology 302, 1193-212.

92. Kai, M., Matsuoka, M., Nakata, N., Maeda, S., Gidoh, M., Maeda, Y.,

Hashimoto, K., Kobayashi, K. & Kashiwabara, Y. (1999).

Diaminodiphenylsulfone resistance of Mycobacterium leprae due to

mutations in the dihydropteroate synthase gene. FEMS microbiology

letters 177, 231-5.

205

93. Date, A. A., Vitoria, M., Granich, R., Banda, M., Fox, M. Y. & Gilks, C. (2010).

Implementation of co-trimoxazole prophylaxis and isoniazid

preventive therapy for people living with HIV. Bull World Health Organ

88, 253-9.

94. Koser, C. U., Veerapen-Pierce, R. N., Summers, D. K. & Archer, J. A. (2010).

Role of mutations in dihydrofolate reductase DfrA (Rv2763c) and

thymidylate synthase ThyA (Rv2764c) in Mycobacterium tuberculosis

drug resistance. Antimicrobial agents and chemotherapy 54, 4522-3;

author reply 5423-5.

95. Rengarajan, J., Sassetti, C. M., Naroditskaya, V., Sloutsky, A., Bloom, B. R.

& Rubin, E. J. (2004). The folate pathway is a target for resistance to the

drug para-aminosalicylic acid (PAS) in mycobacteria. Mol Microbiol 53,

275-82.

96. Doull, J. A. (1963). Sulfone Therapy of Leprosy. Background, Early

History and Present Status. Int J Lepr 31, 143-60.

97. Argyrou, A., Vetting, M. W., Aladegbami, B. & Blanchard, J. S. (2006).

Mycobacterium tuberculosis dihydrofolate reductase is a target for

isoniazid. Nat Struct Mol Biol 13, 408-13.

98. Forgacs, P., Wengenack, N. L., Hall, L., Zimmerman, S. K., Silverman, M. L.

& Roberts, G. D. (2009). Tuberculosis and trimethoprim-

sulfamethoxazole. Antimicrob Agents Chemother 53, 4789-93.

206

99. Ong, W., Sievers, A. & Leslie, D. E. (2010). Mycobacterium tuberculosis

and sulfamethoxazole susceptibility. Antimicrob Agents Chemother 54,

2748; author reply 2748-9.

100. Young, L. S. (2009). Reconsidering some approved antimicrobial agents

for tuberculosis. Antimicrob Agents Chemother 53, 4577-9.

101. Suling, W. J., Reynolds, R. C., Barrow, E. W., Wilson, L. N., Piper, J. R. &

Barrow, W. W. (1998). Susceptibilities of Mycobacterium tuberculosis

and Mycobacterium avium complex to lipophilic deazapteridine

derivatives, inhibitors of dihydrofolate reductase. The Journal of

antimicrobial chemotherapy 42, 811-5.

102. Ogwang, S., Nguyen, H. T., Sherman, M., Bajaksouzian, S., Jacobs, M. R.,

Boom, W. H., Zhang, G. F. & Nguyen, L. (2011). Bacterial conversion of

folinic acid is required for antifolate resistance. The Journal of biological

chemistry 286, 15377-90.

103. Nichols, R. J., Sen, S., Choo, Y. J., Beltrao, P., Zietek, M., Chaba, R., Lee, S.,

Kazmierczak, K. M., Lee, K. J., Wong, A., Shales, M., Lovett, S., Winkler, M.

E., Krogan, N. J., Typas, A. & Gross, C. A. (2011). Phenotypic landscape of

a bacterial cell. Cell 144, 143-56.

104. Fey, P. D. (2010). Modality of bacterial growth presents unique targets:

how do we treat biofilm-mediated infections? Current opinion in

microbiology 13, 610-5.

207

105. Roberts, M. E. & Stewart, P. S. (2005). Modelling protection from

antimicrobial agents in biofilms through the formation of persister cells.

Microbiology 151, 75-80.

106. Balaban, N. Q., Merrin, J., Chait, R., Kowalik, L. & Leibler, S. (2004).

Bacterial persistence as a phenotypic switch. Science 305, 1622-5.

107. Costerton, J. W., Stewart, P. S. & Greenberg, E. P. (1999). Bacterial

biofilms: a common cause of persistent infections. Science 284, 1318-22.

108. Singh, P. K., Schaefer, A. L., Parsek, M. R., Moninger, T. O., Welsh, M. J. &

Greenberg, E. P. (2000). Quorum-sensing signals indicate that cystic

fibrosis lungs are infected with bacterial biofilms. Nature 407, 762-4.

109. Boehm, A., Steiner, S., Zaehringer, F., Casanova, A., Hamburger, F., Ritz,

D., Keck, W., Ackermann, M., Schirmer, T. & Jenal, U. (2009). Second

messenger signalling governs Escherichia coli biofilm induction upon

ribosomal stress. Molecular microbiology 72, 1500-16.

110. Lopez, D., Vlamakis, H. & Kolter, R. (2010). Biofilms. Cold Spring Harbor

perspectives in biology 2, a000398.

111. Bardouniotis, E., Ceri, H. & Olson, M. E. (2003). Biofilm formation and

biocide susceptibility testing of Mycobacterium fortuitum and

Mycobacterium marinum. Current microbiology 46, 28-32.

208

112. Ojha, A., Anand, M., Bhatt, A., Kremer, L., Jacobs, W. R., Jr. & Hatfull, G. F.

(2005). GroEL1: a dedicated chaperone involved in mycolic acid

biosynthesis during biofilm formation in mycobacteria. Cell 123, 861-73.

113. Ojha, A. K., Baughn, A. D., Sambandan, D., Hsu, T., Trivelli, X., Guerardel,

Y., Alahari, A., Kremer, L., Jacobs, W. R., Jr. & Hatfull, G. F. (2008). Growth

of Mycobacterium tuberculosis biofilms containing free mycolic acids

and harbouring drug-tolerant bacteria. Molecular microbiology 69, 164-

74.

114. Schulze-Robbecke, R., Janning, B. & Fischeder, R. (1992). Occurrence of

mycobacteria in biofilm samples. Tubercle and lung disease : the official

journal of the International Union against Tuberculosis and Lung Disease

73, 141-4.

115. Kressel, A. B. & Kidd, F. (2001). Pseudo-outbreak of Mycobacterium

chelonae and Methylobacterium mesophilicum caused by

contamination of an automated endoscopy washer. Infection control

and hospital epidemiology : the official journal of the Society of Hospital

Epidemiologists of America 22, 414-8.

116. Marsollier, L., Severin, T., Aubry, J., Merritt, R. W., Saint Andre, J. P.,

Legras, P., Manceau, A. L., Chauty, A., Carbonnelle, B. & Cole, S. T. (2004).

Aquatic snails, passive hosts of Mycobacterium ulcerans. Applied and

environmental microbiology 70, 6296-8.

209

117. Marsollier, L., Stinear, T., Aubry, J., Saint Andre, J. P., Robert, R., Legras,

P., Manceau, A. L., Audrain, C., Bourdon, S., Kouakou, H. & Carbonnelle, B.

(2004). Aquatic plants stimulate the growth of and biofilm formation by

Mycobacterium ulcerans in axenic culture and harbor these bacteria in

the environment. Applied and environmental microbiology 70, 1097-103.

118. Anderson, G. G., Dodson, K. W., Hooton, T. M. & Hultgren, S. J. (2004).

Intracellular bacterial communities of uropathogenic Escherichia coli

in urinary tract pathogenesis. Trends in microbiology 12, 424-30.

119. Arora, K., Whiteford, D. C., Lau-Bonilla, D., Davitt, C. M. & Dahl, J. L.

(2008). Inactivation of lsr2 results in a hypermotile phenotype in

Mycobacterium smegmatis. Journal of bacteriology 190, 4291-300.

120. Recht, J. & Kolter, R. (2001). Glycopeptidolipid acetylation affects

sliding motility and biofilm formation in Mycobacterium smegmatis.

Journal of bacteriology 183, 5718-24.

121. Niebisch, A., Kabus, A., Schultz, C., Weil, B. & Bott, M. (2006).

Corynebacterial protein kinase G controls 2-oxoglutarate

dehydrogenase activity via the phosphorylation status of the OdhI

protein. The Journal of biological chemistry 281, 12300-7.

122. O'Hare, H. M., Duran, R., Cervenansky, C., Bellinzoni, M., Wehenkel, A. M.,

Pritsch, O., Obal, G., Baumgartner, J., Vialaret, J., Johnsson, K. & Alzari, P.

210

M. (2008). Regulation of glutamate metabolism by protein kinases in

mycobacteria. Molecular microbiology 70, 1408-23.

123. Schultz, C., Niebisch, A., Gebel, L. & Bott, M. (2007). Glutamate

production by Corynebacterium glutamicum: dependence on the

oxoglutarate dehydrogenase inhibitor protein OdhI and protein kinase

PknG. Applied microbiology and biotechnology 76, 691-700.

124. Cowley, S., Ko, M., Pick, N., Chow, R., Downing, K. J., Gordhan, B. G., Betts,

J. C., Mizrahi, V., Smith, D. A., Stokes, R. W. & Av-Gay, Y. (2004). The

Mycobacterium tuberculosis protein serine/threonine kinase PknG is

linked to cellular glutamate/glutamine levels and is important for

growth in vivo. Molecular microbiology 52, 1691-702.

125. Wolanin, P. M., Thomason, P. A. & Stock, J. B. (2002). Histidine protein

kinases: key signal transducers outside the animal kingdom. Genome

biology 3, REVIEWS3013.

126. Grebe, T. W. & Stock, J. B. (1999). The histidine protein kinase

superfamily. Advances in microbial physiology 41, 139-227.

127. Koretke, K. K., Lupas, A. N., Warren, P. V., Rosenberg, M. & Brown, J. R.

(2000). Evolution of two-component signal transduction. Molecular

biology and evolution 17, 1956-70.

128. Lee, D. C. & Jia, Z. (2009). Emerging structural insights into bacterial

tyrosine kinases. Trends in biochemical sciences 34, 351-7.

211

129. Cock, J. M., Vanoosthuyse, V. & Gaude, T. (2002). Receptor kinase

signalling in plants and animals: distinct molecular systems with

mechanistic similarities. Current opinion in cell biology 14, 230-6.

130. Pereira, S. F., Goss, L. & Dworkin, J. (2011). Eukaryote-like

serine/threonine kinases and phosphatases in bacteria. Microbiology

and molecular biology reviews : MMBR 75, 192-212.

131. Yarden, Y. & Ullrich, A. (1988). Growth factor receptor tyrosine kinases.

Annual review of biochemistry 57, 443-78.

132. Chao, J., Wong, D., Zheng, X., Poirier, V., Bach, H., Hmama, Z. & Av-Gay, Y.

(2010). Protein kinase and phosphatase signaling in Mycobacterium

tuberculosis physiology and pathogenesis. Biochimica et biophysica acta

1804, 620-7.

133. Cole, S. T., Brosch, R., Parkhill, J., Garnier, T., Churcher, C., Harris, D.,

Gordon, S. V., Eiglmeier, K., Gas, S., Barry, C. E., 3rd, Tekaia, F., Badcock,

K., Basham, D., Brown, D., Chillingworth, T., Connor, R., Davies, R., Devlin,

K., Feltwell, T., Gentles, S., Hamlin, N., Holroyd, S., Hornsby, T., Jagels, K.,

Krogh, A., McLean, J., Moule, S., Murphy, L., Oliver, K., Osborne, J., Quail,

M. A., Rajandream, M. A., Rogers, J., Rutter, S., Seeger, K., Skelton, J.,

Squares, R., Squares, S., Sulston, J. E., Taylor, K., Whitehead, S. & Barrell,

B. G. (1998). Deciphering the biology of Mycobacterium tuberculosis

from the complete genome sequence. Nature 393, 537-44.

212

134. Av-Gay, Y. & Everett, M. (2000). The eukaryotic-like Ser/Thr protein

kinases of Mycobacterium tuberculosis. Trends in microbiology 8, 238-

44.

135. Prabhakaran, K., Harris, E. B. & Randhawa, B. (2000). Regulation by

protein kinase of phagocytosis of Mycobacterium leprae by

macrophages. Journal of medical microbiology 49, 339-42.

136. Vissa, V. D. & Brennan, P. J. (2001). The genome of Mycobacterium

leprae: a minimal mycobacterial gene set. Genome biology 2,

REVIEWS1023.

137. Tiwari, D., Singh, R. K., Goswami, K., Verma, S. K., Prakash, B. &

Nandicoori, V. K. (2009). Key residues in Mycobacterium tuberculosis

protein kinase G play a role in regulating kinase activity and survival in

the host. The Journal of biological chemistry 284, 27467-79.

138. Scherr, N., Honnappa, S., Kunz, G., Mueller, P., Jayachandran, R., Winkler,

F., Pieters, J. & Steinmetz, M. O. (2007). Structural basis for the specific

inhibition of protein kinase G, a virulence factor of Mycobacterium

tuberculosis. Proceedings of the National Academy of Sciences of the

United States of America 104, 12151-6.

139. Chaurasiya, S. K. & Srivastava, K. K. (2009). Downregulation of protein

kinase C-alpha enhances intracellular survival of Mycobacteria: role of

PknG. BMC microbiology 9, 271.

213

140. Houben, E. N., Walburger, A., Ferrari, G., Nguyen, L., Thompson, C. J.,

Miess, C., Vogel, G., Mueller, B. & Pieters, J. (2009). Differential

expression of a virulence factor in pathogenic and non-pathogenic

mycobacteria. Molecular microbiology 72, 41-52.

141. Nguyen, L., Walburger, A., Houben, E., Koul, A., Muller, S., Morbitzer, M.,

Klebl, B., Ferrari, G. & Pieters, J. (2005). Role of protein kinase G in

growth and glutamine metabolism of Mycobacterium bovis BCG. Journal

of bacteriology 187, 5852-6.

142. Scherr, N., Muller, P., Perisa, D., Combaluzier, B., Jeno, P. & Pieters, J.

(2009). Survival of pathogenic mycobacteria in macrophages is

mediated through autophosphorylation of protein kinase G. Journal of

bacteriology 191, 4546-54.

143. McLennan, A. G. (1999). The MutT motif family of nucleotide

phosphohydrolases in man and human pathogens (review).

International journal of molecular medicine 4, 79-89.

144. Bessman, M. J., Frick, D. N. & O'Handley, S. F. (1996). The MutT proteins

or "Nudix" hydrolases, a family of versatile, widely distributed,

"housecleaning" enzymes. The Journal of biological chemistry 271,

25059-62.

145. Bhatnagar, S. K. & Bessman, M. J. (1988). Studies on the mutator gene,

mutT of Escherichia coli. Molecular cloning of the gene, purification of

214

the gene product, and identification of a novel nucleoside

triphosphatase. The Journal of biological chemistry 263, 8953-7.

146. Espinosa, J. C., Tercero, J. A., Rubio, M. A. & Jimenez, A. (1999). The pur7

gene from the puromycin biosynthetic pur cluster of Streptomyces

alboniger encodes a nudix hydrolase. Journal of bacteriology 181, 4914-

8.

147. Gabelli, S. B., Bianchet, M. A., Xu, W., Dunn, C. A., Niu, Z. D., Amzel, L. M. &

Bessman, M. J. (2007). Structure and function of the E. coli

dihydroneopterin triphosphate pyrophosphatase: a Nudix enzyme

involved in folate biosynthesis. Structure 15, 1014-22.

148. Klaus, S. M., Wegkamp, A., Sybesma, W., Hugenholtz, J., Gregory, J. F., 3rd

& Hanson, A. D. (2005). A nudix enzyme removes pyrophosphate from

dihydroneopterin triphosphate in the folate synthesis pathway of

bacteria and plants. The Journal of biological chemistry 280, 5274-80.

149. Ooga, T., Ohashi, Y., Kuramitsu, S., Koyama, Y., Tomita, M., Soga, T. &

Masui, R. (2009). Degradation of ppGpp by nudix pyrophosphatase

modulates the transition of growth phase in the bacterium Thermus

thermophilus. The Journal of biological chemistry 284, 15549-56.

150. Somerville, C. R. & Ahmed, A. (1979). Mutants of Escherichia coli

defective in the degradation of guanosine 5'-triphosphate, 3'-

diphosphate (pppGpp). Molecular & general genetics : MGG 169, 315-23.

215

151. Hogg, T., Mechold, U., Malke, H., Cashel, M. & Hilgenfeld, R. (2004).

Conformational antagonism between opposing active sites in a

bifunctional RelA/SpoT homolog modulates (p)ppGpp metabolism

during the stringent response [corrected]. Cell 117, 57-68.

152. Johnson, G. S., Adler, C. R., Collins, J. J. & Court, D. (1979). Role of the

spoT gene product and manganese ion in the metabolism of guanosine

5'-diphosphate 3'-diphosphate in Escherichia coli. The Journal of

biological chemistry 254, 5483-7.

153. Kwon, Y. K., Higgins, M. B. & Rabinowitz, J. D. (2010). Antifolate-induced

depletion of intracellular glycine and purines inhibits thymineless

death in E. coli. ACS chemical biology 5, 787-95.

154. De Saizieu, A., Vankan, P. & van Loon, A. P. (1995). Enzymic

characterization of Bacillus subtilis GTP cyclohydrolase I. Evidence for a

chemical dephosphorylation of dihydroneopterin triphosphate. The

Biochemical journal 306 ( Pt 2), 371-7.

155. Wilson, S. D. & Horne, D. W. (1982). Use of glycerol-cryoprotected

Lactobacillus casei for microbiological assay of folic acid. Clinical

chemistry 28, 1198-200.

156. Dos Vultos, T., Blazquez, J., Rauzier, J., Matic, I. & Gicquel, B. (2006).

Identification of Nudix hydrolase family members with an antimutator

216

role in Mycobacterium tuberculosis and Mycobacterium smegmatis.

Journal of bacteriology 188, 3159-61.

157. Dabbs, E. R. (1982). The gene for ribosomal protein L13, rplM, is located

near argR, at about 70 minutes on the Escherichia coli chromosomal

linkage map. Journal of bacteriology 149, 779-82.

158. Sanchez, C., Blanco, G., Mendez, C. & Salas, J. A. (1997). Cloning,

sequencing and transcriptional analysis of a Streptomyces coelicolor

operon containing the rplM and rpsI genes encoding ribosomal proteins

ScoL13 and ScoS9. Molecular & general genetics : MGG 257, 91-6.

159. Isono, S., Thamm, S., Kitakawa, M. & Isono, K. (1985). Cloning and

nucleotide sequencing of the genes for ribosomal proteins S9 (rpsI) and

L13 (rplM) of Escherichia coli. Molecular & general genetics : MGG 198,

279-82.

160. Khaitovich, P., Mankin, A. S., Green, R., Lancaster, L. & Noller, H. F.

(1999). Characterization of functionally active subribosomal particles

from Thermus aquaticus. Proceedings of the National Academy of

Sciences of the United States of America 96, 85-90.

161. Torres, M., Condon, C., Balada, J. M., Squires, C. & Squires, C. L. (2001).

Ribosomal protein S4 is a transcription factor with properties

remarkably similar to NusA, a protein involved in both non-ribosomal

and ribosomal RNA antitermination. The EMBO journal 20, 3811-20.

217

162. Ban, N., Nissen, P., Hansen, J., Moore, P. B. & Steitz, T. A. (2000). The

complete atomic structure of the large ribosomal subunit at 2.4 A

resolution. Science 289, 905-20.

163. Mende, L. (1978). The primary structure of protein L13 from the large

subunit of Escherichia coli ribosomes. FEBS letters 96, 313-6.

164. Lee, Y., Bang, W. Y., Kim, S., Lazar, P., Kim, C. W., Bahk, J. D. & Lee, K. W.

(2010). Molecular modeling study for interaction between Bacillus

subtilis Obg and Nucleotides. PloS one 5, e12597.

165. Scott, J. M., Ju, J., Mitchell, T. & Haldenwang, W. G. (2000). The Bacillus

subtilis GTP binding protein obg and regulators of the sigma(B) stress

response transcription factor cofractionate with ribosomes. Journal of

bacteriology 182, 2771-7.

166. Sasindran, S. J., Saikolappan, S., Scofield, V. L. & Dhandayuthapani, S.

(2011). Biochemical and physiological characterization of the GTP-

binding protein Obg of Mycobacterium tuberculosis. BMC microbiology

11, 43.

167. Soung, G. Y., Miller, J. L., Koc, H. & Koc, E. C. (2009). Comprehensive

analysis of phosphorylated proteins of Escherichia coli ribosomes.

Journal of proteome research 8, 3390-402.

168. Nguyen, L., Chinnapapagari, S. & Thompson, C. J. (2005). FbpA-

Dependent biosynthesis of trehalose dimycolate is required for the

218

intrinsic multidrug resistance, cell wall structure, and colonial

morphology of Mycobacterium smegmatis. Journal of bacteriology 187,

6603-11.

169. Braunstein, M., Bardarov, S. S. & Jacobs, W. R., Jr. (2002). Genetic

methods for deciphering virulence determinants of Mycobacterium

tuberculosis. Methods in enzymology 358, 67-99.

170. van Kessel, J. C. & Hatfull, G. F. (2007). Recombineering in

Mycobacterium tuberculosis. Nature methods 4, 147-52.

171. Wolff, K. A., Nguyen, H. T., Cartabuke, R. H., Singh, A., Ogwang, S. &

Nguyen, L. (2009). Protein kinase G is required for intrinsic antibiotic

resistance in mycobacteria. Antimicrobial agents and chemotherapy 53,

3515-9.

172. Pelicic, V., Reyrat, J. M. & Gicquel, B. (1996). Generation of unmarked

directed mutations in mycobacteria, using sucrose counter-selectable

suicide vectors. Molecular microbiology 20, 919-25.

173. Pelicic, V., Jackson, M., Reyrat, J. M., Jacobs, W. R., Jr., Gicquel, B. &

Guilhot, C. (1997). Efficient allelic exchange and transposon

mutagenesis in Mycobacterium tuberculosis. Proceedings of the

National Academy of Sciences of the United States of America 94, 10955-

60.

219

174. Bardarov, S., Bardarov Jr, S., Jr., Pavelka Jr, M. S., Jr., Sambandamurthy,

V., Larsen, M., Tufariello, J., Chan, J., Hatfull, G. & Jacobs Jr, W. R., Jr.

(2002). Specialized transduction: an efficient method for generating

marked and unmarked targeted gene disruptions in Mycobacterium

tuberculosis, M. bovis BCG and M. smegmatis. Microbiology 148, 3007-

17.

175. Datsenko, K. A. & Wanner, B. L. (2000). One-step inactivation of

chromosomal genes in Escherichia coli K-12 using PCR products.

Proceedings of the National Academy of Sciences of the United States of

America 97, 6640-5.

176. Yu, D., Ellis, H. M., Lee, E. C., Jenkins, N. A., Copeland, N. G. & Court, D. L.

(2000). An efficient recombination system for chromosome engineering

in Escherichia coli. Proceedings of the National Academy of Sciences of

the United States of America 97, 5978-83.

177. Johnson, J. E., Lackner, L. L., Hale, C. A. & de Boer, P. A. (2004). ZipA is

required for targeting of DMinC/DicB, but not DMinC/MinD, complexes

to septal ring assemblies in Escherichia coli. Journal of bacteriology 186,

2418-29.

178. Alland, D., Steyn, A. J., Weisbrod, T., Aldrich, K. & Jacobs, W. R., Jr. (2000).

Characterization of the Mycobacterium tuberculosis iniBAC promoter, a

220

promoter that responds to cell wall biosynthesis inhibition. Journal of

bacteriology 182, 1802-11.

179. Snapper, S. B., Melton, R. E., Mustafa, S., Kieser, T. & Jacobs, W. R., Jr.

(1990). Isolation and characterization of efficient plasmid

transformation mutants of Mycobacterium smegmatis. Molecular

microbiology 4, 1911-9.

180. Stover, C. K., de la Cruz, V. F., Fuerst, T. R., Burlein, J. E., Benson, L. A.,

Bennett, L. T., Bansal, G. P., Young, J. F., Lee, M. H., Hatfull, G. F. & et al.

(1991). New use of BCG for recombinant vaccines. Nature 351, 456-60.

181. Nguyen, H. T., Wolff, K. A., Cartabuke, R. H., Ogwang, S. & Nguyen, L.

(2010). A lipoprotein modulates activity of the MtrAB two-component

system to provide intrinsic multidrug resistance, cytokinetic control

and cell wall homeostasis in Mycobacterium. Molecular microbiology 76,

348-64.

182. Bardarov, S., Kriakov, J., Carriere, C., Yu, S., Vaamonde, C., McAdam, R. A.,

Bloom, B. R., Hatfull, G. F. & Jacobs, W. R., Jr. (1997). Conditionally

replicating mycobacteriophages: a system for transposon delivery to

Mycobacterium tuberculosis. Proceedings of the National Academy of

Sciences of the United States of America 94, 10961-6.

183. Etienne, G., Villeneuve, C., Billman-Jacobe, H., Astarie-Dequeker, C.,

Dupont, M. A. & Daffe, M. (2002). The impact of the absence of

221

glycopeptidolipids on the ultrastructure, cell surface and cell wall

properties, and phagocytosis of Mycobacterium smegmatis.

Microbiology 148, 3089-100.

184. Ertel, R., Redfield, B., Brot, N. & Weissbach, H. (1968). Role of GTP in

protein synthesis: interaction of GTP with soluble transfer factors from

E. coli. Archives of biochemistry and biophysics 128, 331-8.

185. Moon, H. M., Redfield, B. & Weissbach, H. (1972). Interaction of

eukaryote EF 1 with guanosine nucleotides and

aminoacyl-tRNA. Proceedings of the National Academy of Sciences of the

United States of America 69, 1249-52.

186. Gatfield, J. & Pieters, J. (2000). Essential role for cholesterol in entry of

mycobacteria into macrophages. Science 288, 1647-50.

187. Koul, A., Choidas, A., Tyagi, A. K., Drlica, K., Singh, Y. & Ullrich, A. (2001).

Serine/threonine protein kinases PknF and PknG of Mycobacterium

tuberculosis: characterization and localization. Microbiology 147, 2307-

14.

188. Cangelosi, G. A., Palermo, C. O., Laurent, J. P., Hamlin, A. M. & Brabant, W.

H. (1999). Colony morphotypes on Congo red agar segregate along

species and drug susceptibility lines in the Mycobacterium avium-

intracellulare complex. Microbiology 145 ( Pt 6), 1317-24.

222

189. Liu, J. & Nikaido, H. (1999). A mutant of Mycobacterium smegmatis

defective in the biosynthesis of mycolic acids accumulates

meromycolates. Proceedings of the National Academy of Sciences of the

United States of America 96, 4011-6.

190. Russell, D. G. (2001). Mycobacterium tuberculosis: here today, and here

tomorrow. Nature reviews. Molecular cell biology 2, 569-77.

191. Mueller, P. & Pieters, J. (2006). Modulation of macrophage

antimicrobial mechanisms by pathogenic mycobacteria. Immunobiology

211, 549-56.

192. Rohde, K., Yates, R. M., Purdy, G. E. & Russell, D. G. (2007).

Mycobacterium tuberculosis and the environment within the

phagosome. Immunological reviews 219, 37-54.

193. Chao, M. C. & Rubin, E. J. (2010). Letting sleeping dos lie: does dormancy

play a role in tuberculosis? Annual review of microbiology 64, 293-311.

194. Parrish, N. M., Dick, J. D. & Bishai, W. R. (1998). Mechanisms of latency

in Mycobacterium tuberculosis. Trends in microbiology 6, 107-12.

195. Wayne, L. G. (1994). Dormancy of Mycobacterium tuberculosis and

latency of disease. European journal of clinical microbiology & infectious

diseases : official publication of the European Society of Clinical

Microbiology 13, 908-14.

223

196. Talaat, A. M., Ward, S. K., Wu, C. W., Rondon, E., Tavano, C., Bannantine, J.

P., Lyons, R. & Johnston, S. A. (2007). Mycobacterial bacilli are

metabolically active during chronic tuberculosis in murine lungs:

insights from genome-wide transcriptional profiling. Journal of

bacteriology 189, 4265-74.

197. Camus, J. C., Pryor, M. J., Medigue, C. & Cole, S. T. (2002). Re-annotation

of the genome sequence of Mycobacterium tuberculosis H37Rv.

Microbiology 148, 2967-73.

198. Dissmeyer, N. & Schnittger, A. (2011). Use of phospho-site substitutions

to analyze the biological relevance of phosphorylation events in

regulatory networks. Methods in molecular biology 779, 93-138.

199. Parkinson, J. S. & Kofoid, E. C. (1992). Communication modules in

bacterial signaling proteins. Annual review of genetics 26, 71-112.

200. Mazumder, B., Sampath, P., Seshadri, V., Maitra, R. K., DiCorleto, P. E. &

Fox, P. L. (2003). Regulated release of L13a from the 60S ribosomal

subunit as a mechanism of transcript-specific translational control. Cell

115, 187-98.

201. Mukhopadhyay, R., Ray, P. S., Arif, A., Brady, A. K., Kinter, M. & Fox, P. L.

(2008). DAPK-ZIPK-L13a axis constitutes a negative-feedback module

regulating inflammatory gene expression. Molecular cell 32, 371-82.

224

202. Mukhopadhyay, R., Jia, J., Arif, A., Ray, P. S. & Fox, P. L. (2009). The GAIT

system: a gatekeeper of inflammatory gene expression. Trends in

biochemical sciences 34, 324-31.

203. Moore, P. B. & Steitz, T. A. (2003). The structural basis of large

ribosomal subunit function. Annual review of biochemistry 72, 813-50.

204. Daigle, D. M. & Brown, E. D. (2004). Studies of the interaction of

Escherichia coli YjeQ with the ribosome in vitro. Journal of bacteriology

186, 1381-7.

205. Yamada, T., Masuda, K., Shoji, K. & Hori, M. (1972). Analysis of

ribosomes from viomycin-sensitive and -resistant strains of

Mycobacterium smegmatis. Journal of bacteriology 112, 1-6.

206. Yamada, T., Nagata, A., Ono, Y., Suzuki, Y. & Yamanouchi, T. (1985).

Alteration of ribosomes and RNA polymerase in drug-resistant clinical

isolates of Mycobacterium tuberculosis. Antimicrobial agents and

chemotherapy 27, 921-4.

207. Ames, B. N. & Dubin, D. T. (1960). The role of polyamines in the

neutralization of bacteriophage deoxyribonucleic acid. The Journal of

biological chemistry 235, 769-75.

208. O'Handley, S. F., Frick, D. N., Bullions, L. C., Mildvan, A. S. & Bessman, M. J.

(1996). Escherichia coli orf17 codes for a nucleoside triphosphate

pyrophosphohydrolase member of the MutT family of proteins. Cloning,

225

purification, and characterization of the enzyme. The Journal of

biological chemistry 271, 24649-54.

209. Sybesma, W., Starrenburg, M., Kleerebezem, M., Mierau, I., de Vos, W. M.

& Hugenholtz, J. (2003). Increased production of folate by metabolic

engineering of Lactococcus lactis. Applied and environmental

microbiology 69, 3069-76.

210. Horne, D. W. & Patterson, D. (1988). Lactobacillus casei microbiological

assay of folic acid derivatives in 96-well microtiter plates. Clinical

chemistry 34, 2357-9.

211. Kruschwitz, H. L., McDonald, D., Cossins, E. A. & Schirch, V. (1994). 5-

Formyltetrahydropteroylpolyglutamates are the major folate

derivatives in Neurospora crassa conidiospores. J Biol Chem 269,

28757-63.

212. Shin, Y. S., Kim, E. S., Watson, J. E. & Stokstad, E. L. (1975). Studies of folic

acid compounds in nature. IV. Folic acid compounds in soybeans and

cow milk. Can J Biochem 53, 338-43.

213. Stover, P. & Schirch, V. (1993). The metabolic role of leucovorin. Trends

Biochem Sci 18, 102-6.

214. Hansen, S., Lewis, K. & Vulic, M. (2008). Role of global regulators and

nucleotide metabolism in antibiotic tolerance in Escherichia coli.

Antimicrobial agents and chemotherapy 52, 2718-26.

226

215. Jain, V., Saleem-Batcha, R., China, A. & Chatterji, D. (2006). Molecular

dissection of the mycobacterial stringent response protein Rel. Protein

science : a publication of the Protein Society 15, 1449-64.

216. Betts, J. C., Lukey, P. T., Robb, L. C., McAdam, R. A. & Duncan, K. (2002).

Evaluation of a nutrient starvation model of Mycobacterium

tuberculosis persistence by gene and protein expression profiling.

Molecular microbiology 43, 717-31.

217. Flannagan, R. S., Cosio, G. & Grinstein, S. (2009). Antimicrobial

mechanisms of phagocytes and bacterial evasion strategies. Nature

reviews. Microbiology 7, 355-66.

218. Raivio, T. L. (2005). Envelope stress responses and Gram-negative

bacterial pathogenesis. Molecular microbiology 56, 1119-28.

219. Cooper, A. M. (2009). Cell-mediated immune responses in tuberculosis.

Annual review of immunology 27, 393-422.

220. Fratti, R. A., Backer, J. M., Gruenberg, J., Corvera, S. & Deretic, V. (2001).

Role of phosphatidylinositol 3-kinase and Rab5 effectors in phagosomal

biogenesis and mycobacterial phagosome maturation arrest. The

Journal of cell biology 154, 631-44.

221. Singh, S. B., Davis, A. S., Taylor, G. A. & Deretic, V. (2006). Human IRGM

induces to eliminate intracellular mycobacteria. Science 313,

1438-41.

227

222. Vieira, O. V., Botelho, R. J., Rameh, L., Brachmann, S. M., Matsuo, T.,

Davidson, H. W., Schreiber, A., Backer, J. M., Cantley, L. C. & Grinstein, S.

(2001). Distinct roles of class I and class III phosphatidylinositol 3-

kinases in phagosome formation and maturation. The Journal of cell

biology 155, 19-25.

223. Gaullier, J. M., Simonsen, A., D'Arrigo, A., Bremnes, B., Stenmark, H. &

Aasland, R. (1998). FYVE fingers bind PtdIns(3)P. Nature 394, 432-3.

224. Kanai, F., Liu, H., Field, S. J., Akbary, H., Matsuo, T., Brown, G. E., Cantley,

L. C. & Yaffe, M. B. (2001). The PX domains of p47phox and p40phox

bind to lipid products of PI(3)K. Nature cell biology 3, 675-8.

225. Vergne, I., Chua, J. & Deretic, V. (2003). Tuberculosis toxin blocking

phagosome maturation inhibits a novel Ca2+/calmodulin-PI3K hVPS34

cascade. The Journal of experimental medicine 198, 653-9.

226. Beatty, W. L., Rhoades, E. R., Ullrich, H. J., Chatterjee, D., Heuser, J. E. &

Russell, D. G. (2000). Trafficking and release of mycobacterial lipids

from infected macrophages. Traffic 1, 235-47.

227. Mishra, A. K., Driessen, N. N., Appelmelk, B. J. & Besra, G. S. (2011).

Lipoarabinomannan and related glycoconjugates: structure, biogenesis

and role in Mycobacterium tuberculosis physiology and host-pathogen

interaction. FEMS microbiology reviews.

228

228. Rosensweig, N. S. (1975). Diet and intestinal enzyme adaptation:

implications for gastrointestinal disorders. The American journal of

clinical nutrition 28, 648-55.

229. Salina, E. G., Zhogina, Y. A., Shleeva, M. O., Sorokoumova, G. M.,

Selishcheva, A. A. & Kaprelyants, A. S. (2010). Biochemical and

morphological changes in dormant ("Nonculturable") Mycobacterium

smegmatis cells. Biochemistry. Biokhimiia 75, 72-80.

230. Ramagopal, S. (1991). Covalent modifications of ribosomal proteins in

growing and aggregation-competent dictyostelium discoideum:

phosphorylation and methylation. Biochemistry and cell biology =

Biochimie et biologie cellulaire 69, 263-8.

231. Francis, T. A. & Roberts, S. (1982). Influence of the state of ribosome

association on the phosphorylation of ribosomal proteins in isolated

ribosome--protein kinase systems from rat cerebral cortex. The

Biochemical journal 208, 289-300.

232. Mikulik, K. & Janda, I. (1997). Protein kinase associated with ribosomes

phosphorylates ribosomal proteins of Streptomyces collinus.

Biochemical and biophysical research communications 238, 370-6.

233. Liu, J., Schmitz, J. C., Lin, X., Tai, N., Yan, W., Farrell, M., Bailly, M., Chen, T.

& Chu, E. (2002). Thymidylate synthase as a translational regulator of

cellular gene expression. Biochimica et biophysica acta 1587, 174-82.

229

234. Toth-Martinez, B. L., Hernadi, F. J. & Dinya, Z. (1978). Assembly of 7,8-

dihydrofolic acid in de novo folate biosynthesizing microorganisms+.

Intermediary enzymes as partial heterologous protein associates of the

5,6,7,8-tetrahydrofolate multienzyme complex. Pharmacological

research communications 10, 391-430.

235. Douce, R., Bourguignon, J., Neuburger, M. & Rebeille, F. (2001). The

glycine decarboxylase system: a fascinating complex. Trends in plant

science 6, 167-76.

236. Korn, K. & Krausz, E. (2007). Cell-based high-content screening of small-

molecule libraries. Current opinion in chemical biology 11, 503-10.

237. Gaither, L. A. (2007). Chemogenomics approaches to novel target

discovery. Expert review of proteomics 4, 411-9.

238. Walters, W. P. & Namchuk, M. (2003). Designing screens: how to make

your hits a hit. Nature reviews. Drug discovery 2, 259-66.

239. Koul, A., Arnoult, E., Lounis, N., Guillemont, J. & Andries, K. (2011). The

challenge of new drug discovery for tuberculosis. Nature 469, 483-90.

240. Castaldo, P., Magi, S., Nasti, A. A., Arcangeli, S., Lariccia, V., Alesi, N.,

Tocchini, M. & Amoroso, S. (2011). Clinical pharmacogenetics of

methotrexate. Current drug metabolism 12, 278-86.

241. Graham, W. R., Jr. (1975). Adverse effects of dapsone. International

journal of dermatology 14, 494-500.

230

242. Harrap, K. R. (1976). Anti-folate toxicity and its reversal. Biochemical

Society transactions 4, 856-9.

231