"Medicine, Faculty of"@en . "Medicine, Department of"@en . "Experimental Medicine, Division of"@en . "DSpace"@en . "UBCV"@en . "Wong, Dennis Dick-Hang"@en . "2012-05-16T22:35:43Z"@en . "2012"@en . "Doctor of Philosophy - PhD"@en . "University of British Columbia"@en . "One of the main mechanisms by which the etiological agent of tuberculosis (TB), Mycobacterium tuberculosis (Mtb), survives in the host macrophage is by its capacity to arrest phagosome acidification and fusion with lysosomes. This Mtb feature is associated with phagosomal exclusion of the Vacuolar H\u00E2\u0081\u00BA-ATPase (V-ATPase) proton pump, which normally drives luminal acidification of membranous organelles. Although this phenomenon has been known for 20 years, the mechanism by which Mtb blocks phagosome acidification remains obscure. Research on Mtb pathophysiology shows that a wide array of Mtb lipid and protein molecules contribute to maintaining the mycobacterial phagosome in an immature state. We have previously found that Mtb protein-tyrosine Phosphatase A (PtpA) is required for Mtb infection of human macrophages. PtpA is secreted into the macrophage cytosol to inactivate the human VPS33B, a component of the Class C VPS Complex that regulates endosomal membrane fusion. VPS33B dephosphorylation by PtpA results in the inhibition of phagosome-lysosome fusion. In this work, we demonstrated that, in addition to its phosphatase activity, PtpA is also capable of binding to subunit H of the macrophage V-ATPase complex, indicating that PtpA can directly disrupt phagosome acidification. Indeed, we found that a Mtb strain expressing a V-ATPase-binding defective mutant PtpA protein failed to inhibit phagosome acidification, and expression of wild-type PtpA protein in E. coli-infected macrophages is sufficient to block acidification. Furthermore, we showed that the Class C VPS complex associates with V-ATPase during phagosome maturation, identifying a novel role for V-ATPase in coordinating endocytic membrane fusion. PtpA interaction with host V-ATPase is required for the previously reported dephosphorylation of VPS33B and subsequent exclusion of V-ATPase from the phagosome during Mtb infection. Taken together, these findings reveal, for the first time, the long-sought mechanism behind the lack of acidification in the mycobacterial phagosome. Interestingly, we found that PtpA is also a substrate for the newly identified Mtb protein tyrosine kinase PtkA, which is encoded within a shared operon with PtpA, indicating a regulatory control of PtpA during Mtb infection. Understanding the pathophysiological importance of PtpA in Mtb infection might contribute to the development of novel antitubercular therapeutics."@en . "https://circle.library.ubc.ca/rest/handle/2429/42331?expand=metadata"@en . "ROLE OF MYCOBACTERIUM TUBERCULOSIS PROTEIN TYROSINE PHOSPHATASE A IN THE PATHOGENESIS OF TUBERCULOSIS by Dennis Dick-Hang Wong B.Sc., The University of British Columbia, 2007 A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY in The Faculty of Graduate Studies (Experimental Medicine) THE UNIVERSITY OF BRITISH COLUMBIA (Vancouver) May 2012 \u00C2\u00A9 Dennis Dick-Hang Wong, 2012 \t\r \u00C2\u00A0 ii ABSTRACT One of the main mechanisms by which the etiological agent of tuberculosis (TB), Mycobacterium tuberculosis (Mtb), survives in the host macrophage is by its capacity to arrest phagosome acidification and fusion with lysosomes. This Mtb feature is associated with phagosomal exclusion of the Vacuolar H+-ATPase (V- ATPase) proton pump, which normally drives luminal acidification of membranous organelles. Although this phenomenon has been known for 20 years, the mechanism by which Mtb blocks phagosome acidification remains obscure. Research on Mtb pathophysiology shows that a wide array of Mtb lipid and protein molecules contribute to maintaining the mycobacterial phagosome in an immature state. We have previously found that Mtb protein-tyrosine Phosphatase A (PtpA) is required for Mtb infection of human macrophages. PtpA is secreted into the macrophage cytosol to inactivate the human VPS33B, a component of the Class C VPS Complex that regulates endosomal membrane fusion. VPS33B dephosphorylation by PtpA results in the inhibition of phagosome-lysosome fusion. In this work, we demonstrated that, in addition to its phosphatase activity, PtpA is also capable of binding to subunit H of the macrophage V-ATPase complex, indicating that PtpA can directly disrupt phagosome acidification. Indeed, we found that a Mtb strain expressing a V- ATPase-binding defective mutant PtpA protein failed to inhibit phagosome acidification, and expression of wild-type PtpA protein in E. coli-infected macrophages is sufficient to block acidification. Furthermore, we showed that the Class C VPS complex associates with V-ATPase during phagosome maturation, \t\r \u00C2\u00A0 iii identifying a novel role for V-ATPase in coordinating endocytic membrane fusion. PtpA interaction with host V-ATPase is required for the previously reported dephosphorylation of VPS33B and subsequent exclusion of V-ATPase from the phagosome during Mtb infection. Taken together, these findings reveal, for the first time, the long-sought mechanism behind the lack of acidification in the mycobacterial phagosome. Interestingly, we found that PtpA is also a substrate for the newly identified Mtb protein tyrosine kinase PtkA, which is encoded within a shared operon with PtpA, indicating a regulatory control of PtpA during Mtb infection. Understanding the pathophysiological importance of PtpA in Mtb infection might contribute to the development of novel antitubercular therapeutics. \t\r \u00C2\u00A0 iv PREFACE Parts of this thesis have been published in the peer-reviewed journals listed below in chronological order: 1. 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(5): 316-322. In this study, I was responsible for performing several experiments, contributing to the development of work shown in Figure 2 (A, B, C, and E) and Figure 4 (A and C) and opening a lead into the work described in this thesis. H. Bach was responsible for the experimental design, performing experiments and writing the manuscript. K. G. Papavinasasundaram constructed the \u00CE\u0094ptpA deletion Mtb mutant. Y. Av-Gay designed the study and wrote the manuscript. This published work is located in Section 1.3.2. 2. Bach, H., Wong, D., & Av-Gay, Y. (2009) Mycobacterium tuberculosis PtkA is a novel protein tyrosine kinase whose substrate is PtpA. Biochem J 420(2): 155-160. In this study, I was responsible for performing most of the experiments, contributing to the development of work shown in Figure 1, Figure 2 (A, B and D), Table 1 and Table 2. H. Bach was responsible for the experimental design and writing the manuscript. Y. Av-Gay wrote the manuscript. This published work is located in Section 3.2. 3. 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. Biochim Biophys Acta 1804(3): 620-627. In this review article, I was responsible for writing part of the manuscript, contributing to the sections titled \u00E2\u0080\u009C3. Protein tyrosine phosphorylation\u00E2\u0080\u009D, \u00E2\u0080\u009C4. \t\r \u00C2\u00A0 v Protein tyrosine phosphatase A (PtpA), \u00E2\u0080\u009C5. Protein tyrosine phosphatase B (PtpB)\u00E2\u0080\u009D, \u00E2\u0080\u009C6. Inhibitors of Mtb protein tyrosine phosphatases\u00E2\u0080\u009D and \u00E2\u0080\u009C7. Protein tyrosine kinase A (PtkA)\u00E2\u0080\u009D. J. Chao and X. Zheng contributed to the session titled \u00E2\u0080\u009C2. PknG\u00E2\u0080\u009D, Figure 1 and Table 1. V. Poirier contributed to the session titled \u00E2\u0080\u009C4. Protein tyrosine phosphatase A (PtpA)\u00E2\u0080\u009D and Figure 2. Y. Av-Gay wrote the \u00E2\u0080\u009CIntroduction\u00E2\u0080\u009D and \u00E2\u0080\u009CConcluding Remarks\u00E2\u0080\u009D. J. Chao edited and finalized the manuscript. This published work is located in Sections 1.3.2 and 4.3. 4. Mascarello, A., Chiaradia, L. D., Vernal, J., Villarino, A., Guido, R. V., Perizzolo, P., Poirier, V., Wong, D., Martins, P. G., Nunes, R. J., Yunes, R. A., Andricopulo, A. D., Av-Gay, Y., & Terenzi, H. (2010) Inhibition of Mycobacterium tuberculosis tyrosine phosphatase PtpA by synthetic chalcones: kinetics, molecular modeling, toxicity and effect on growth. Bioorg Med Chem 18(11): 3783-3789. In this study, I was responsible for performing the in vivo analysis of the efficacy of a family of synthetic chalcones as a novel drug against Mtb in the THP-1 macrophage infection model. My work contributed to Figure 4 and Section 2.4 \u00E2\u0080\u009CInhibitors of MPtpA reduced mycobacterial survival in macrophages\u00E2\u0080\u009D. Y. Av-Gay designed the study and participated in writing the manuscript. This published work is located in Section 4.3. 5. Wong, D., Bach, H., Sun, J., Hmama, Z., Av-Gay, Y. (2011) Mycobacterium tuberculosis protein tyrosine phosphatase (PtpA) excludes host vacuolar-H+-ATPase to inhibit phagosome acidification. Proc Natl Acad Sci U S A 108(48): 19371-6. In this study, I designed and performed all experiments. J. Sun and Z. Hmama assisted in work related to Supplemental Figure S7A and S8. I wrote the manuscript. Hmama, Z. helped with editing of the manuscript. Bach, H. was involved in experimental design. Y. Av-Gay participated in \t\r \u00C2\u00A0 vi designing the study and writing the manuscript. This published work is located in Section 3.1. 6. Wong, D. and Av-Gay, Y. Mycobacterium tuberculosis Phosphatase Interference with Human Signaling Pathways. Manuscript in Preparation. In this work, I was responsible for writing a review article on the current knowledge of phosphatases in Mycobacterium tuberculosis. Y. Av-Gay assisted in writing and editing the manuscript. This work is located in Section 1.3.2. The work in this dissertation is conducted with approval from University of British Columbia Office of Research and in accordance with the University of British Columbia Polices and Procedures, Biosafety Practices and Public Health Agency of Canada Guidelines. Biohazard Approval Certificate: B10-0112. Work with radioactive materials is approval by University of British Columbia Committee on Radioisotopes and Radiation Hazards. Radioisotope License: MEDI-3175-13. \t\r \u00C2\u00A0 vii TABLE OF CONTENTS ABSTRACT ........................................................................................................... ii PREFACE ............................................................................................................. iv TABLE OF CONTENTS ...................................................................................... vii LIST OF TABLES ................................................................................................. xi LIST OF FIGURES .............................................................................................. xii LIST OF ABBREVIATIONS ............................................................................... xiv ACKNOWLEDGEMENTS ................................................................................ xviii CHAPTER 1: INTRODUCTION ............................................................................. 1 1.1 Tuberculosis .......................................................................................................... 1 1.1.1 Disease epidemiology ...................................................................................... 1 1.1.2 Origin of human Tuberculosis .......................................................................... 2 1.1.3 Mycobacterium tuberculosis ............................................................................ 3 1.1.4 Infection cycle .................................................................................................. 5 1.2 Phagocytosis ......................................................................................................... 8 1.2.1 Phagosome formation ...................................................................................... 8 1.2.2 Phagosome maturation .................................................................................. 10 1.2.2.1 Vacuolar H+-ATPase ............................................................................................ 14 1.2.2.2 Class C VPS ......................................................................................................... 17 1.3 Mtb arrest of phagosome maturation ............................................................... 19 1.3.1 Overview ........................................................................................................ 19 1.3.2 Mtb phosphatase interference with host signaling pathways ......................... 22 1.3.2.1 PtpB ...................................................................................................................... 24 1.3.2.2 SapM .................................................................................................................... 26 1.3.2.3 PtpA ...................................................................................................................... 28 1.4 Aims of the study ................................................................................................ 32 CHAPTER 2: MATERIALS AND METHODS ...................................................... 34 2.1 Reagents and chemicals .................................................................................... 34 2.1.1 Commercial reagents ..................................................................................... 34 2.1.2 Antibodies ...................................................................................................... 35 \t\r \u00C2\u00A0 viii 2.2 Bacteria ................................................................................................................ 35 2.2.1 Strain maintenance and generation ............................................................... 35 2.2.2 Bacterial preparation for infection .................................................................. 36 2.2.3 Mycobacterial lysis ......................................................................................... 37 2.2.4 Transformation of mycobacteria .................................................................... 37 2.3 Cell culture .......................................................................................................... 37 2.3.1 Maintenance of tissue culture and differentiation ........................................... 37 2.3.2 Single transfection of differentiated THP-1 .................................................... 38 2.3.3 Double transfection of differentiated THP-1 ................................................... 38 2.4 Vectors and DNA manipulation ......................................................................... 38 2.4.1 RNA isolation and cDNA synthesis ................................................................ 38 2.4.2 Cloning ........................................................................................................... 39 2.4.3 Site-directed mutagenesis ............................................................................. 41 2.5 Protein expression and purification .................................................................. 42 2.5.1 Expression of recombinant proteins in E. coli ................................................ 42 2.5.2 Expression of recombinant proteins in M. smegmatis ................................... 43 2.6 In vitro pull down assay ..................................................................................... 44 2.7 ALPHAScreen protein-protein interaction assay ............................................. 44 2.8 Mycobacterial \u00E2\u0080\u009CSplit-Trp\u00E2\u0080\u009D protein-protein interaction assay .......................... 45 2.9 Infection ............................................................................................................... 46 2.9.1 Infection of THP-1 .......................................................................................... 46 2.9.2 Phagocytosis assay ....................................................................................... 46 2.10 Immunoprecipitation ........................................................................................ 46 2.11 Measurement of phagosomal pH .................................................................... 47 2.11.1 Overview ...................................................................................................... 47 2.11.2 Phagosomal pH of Mtb phagosomes ........................................................... 48 2.11.3 Phagosomal pH of E. coli phagosomes ....................................................... 49 2.12 Digital confocal microscopy ............................................................................ 49 2.12.1 Intracellular staining ..................................................................................... 49 2.12.2 Microscopy ................................................................................................... 50 2.13 In vitro phosphatase activity assay ................................................................ 50 2.14 In vitro PtpA secretion analysis ...................................................................... 50 2.15 FACS analysis of antibody specificity ............................................................ 51 2.16 In vitro kinase assay ......................................................................................... 52 \t\r \u00C2\u00A0 ix 2.17 Phospho-amino acid analysis ......................................................................... 52 CHAPTER 3: RESULTS ...................................................................................... 54 3.1 Mtb PtpA excludes host Vacuolar-H+-ATPase to inhibit phagosome acidification ............................................................................................................... 54 3.1.1 Introduction .................................................................................................... 54 3.1.2 Mtb PtpA binds subunit H of human V-ATPase ............................................. 54 3.1.3 PtpA binds to amino acid 220-402 of subunit H ............................................. 55 3.1.4 The C-terminal alpha helix of PtpA binds subunit H ...................................... 55 3.1.5 PtpA binds V-ATPase subunit H in vivo ......................................................... 58 3.1.6 PtpA binding to subunit H is required for Mtb intracellular survival ................ 59 3.1.7 PtpA inhibits phagosome acidification ........................................................... 60 3.1.8 The V-ATPase machinery recruits the Class C VPS complex ....................... 63 3.1.9 Mtb PtpA blocks the interaction between Class C VPS and V-ATPase complexes ............................................................................................................... 64 3.1.10 VPS33B remains phosphorylated in THP-1 infected with binding-defective PtpA ........................................................................................................................ 66 3.1.11 PtpA binding to subunit H participates in the exclusion of the V-ATPase from the Mtb phagosome ................................................................................................ 68 3.2 PtpA is a substrate for the novel protein-tyrosine kinase Mtb PtkA .............. 73 3.2.1 Introduction .................................................................................................... 73 3.2.2 PtpA interacts with the protein encoded by Rv2232 ...................................... 75 3.2.3 ORF Rv2232 encodes an autophosphorylated protein-tyrosine kinase ......... 76 3.2.4 Sequence-function analysis of PtkA .............................................................. 78 3.2.4.1 Mutation of lysine residues ................................................................................... 78 3.2.4.2 Mutation of tyrosine residues ................................................................................ 79 3.2.4.3 Mutation of D85 .................................................................................................... 80 3.2.5 PtpA is a substrate of PtkA ............................................................................ 80 CHAPTER 4: DISCUSSION ................................................................................ 85 4.1 PtpA exclusion of the V-ATPase from phagosomal membrane ..................... 85 4.2 PtpA is a substrate for the protein-tyrosine kinase PtkA ............................... 91 4.3 Inhibitors against PtpA as novel antituberculosis drug ................................. 95 4.4 Future directions ................................................................................................. 98 REFERENCES .................................................................................................. 102 \t\r \u00C2\u00A0 x APPENDIX A: SUPPLEMENTAL FIGURES FOR SECTION 3.1 ..................... 124 \t\r \u00C2\u00A0 xi LIST OF TABLES \t\r \u00C2\u00A0 Table 1. Oligonucleotides used for DNA cloning in this study ............................. 41 Table 2. Oligonucleotides used for SDM in this study ......................................... 42 Table 3. Phagosomal pH of Mtb strains in THP-1 ............................................... 62 Table 4. Autophosphorylation kinetic values of parental and mutated PtkA ........ 80 Table 5. Kd values of PtpA interacting with parental and mutated PtkA proteins . 83 \t\r \u00C2\u00A0 xii LIST OF FIGURES Figure 1. Estimated TB incidence rates (2010). .................................................... 2 Figure 2. Phylogeny of selected mycobacterial species. ....................................... 4 Figure 3. Mtb infection cycle.. ................................................................................ 6 Figure 4. Receptor and signaling interactions during phagocytosis of microbes . 10 Figure 5. Stages of phagosome maturation. ........................................................ 11 Figure 6. Vacuolar H+-ATPase.. .......................................................................... 17 Figure 7. Working model for Class C VPS function in yeast ................................ 19 Figure 8. Model of Mtb inhibition of mycobacterial phagosome maturation. ........ 22 Figure 9. Scheme of PtpA catalytic activity. ......................................................... 30 Figure 10. PtpA interacts with V-ATPase subunit H in vitro ................................. 56 Figure 11. Computer-generated model of PtpA complexed with the V-ATPase subunit H.. ............................................................................................................ 57 Figure 12. PtpA and V-ATPase subunit H interacts in vivo ................................. 59 Figure 13. Mtb PtpA inhibits phagosome acidification.. ....................................... 61 Figure 14. Co-immunoprecipitation of V-ATPase subunits with Class C VPS in THP-1 ................................................................................................................... 64 Figure 15. Mtb PtpA disrupts the interaction between the Class C VPS and V- ATPase complexes during infection .................................................................... 65 Figure 16. Western blot analysis of VPS33B phosphorylation in vivo. ................ 67 Figure 17. Confocal microscopy of THP-1 infected with E. coli or indicated Mtb strains.. ................................................................................................................ 70 Figure 18. Confocal microscopy of infected THP-1 macrophage doubly transfected with GFP-subunit H and DsRed2-VPS33B.. ..................................... 72 Figure 19. Rv2232 interacts with PtpA in vitro ..................................................... 75 Figure 20. Rv2232 autophosphorylation in vitro .................................................. 76 Figure 21. Phospho-amino acid analysis of PtkA. ............................................... 77 Figure 22. Mutational studies of PtkA catalytic mechanism ................................. 79 Figure 23. PtkA phosphorylates PtpA in vitro.. .................................................... 81 Figure 24. PtkA phosphorylates PtpA in dose- and time-dependent manner ...... 81 \t\r \u00C2\u00A0 xiii Figure 25. PtkA phosphorylates PtpA on tyrosine residues ................................. 84 Figure 26. A model for the specific exclusion of V-ATPase and the inhibition of mycobacterial phagosome acidification by PtpA. ................................................ 88 Figure 27. Mtb survival in infected THP-1 treated with chalcone inhibitors.. ....... 96 Figure 28. In vitro analysis of subunit H and PtpA interaction. .......................... 124 Figure 29. PtpAL146A is a phosphatase-active mutant defective in binding subunit H. ....................................................................................................................... 126 Figure 30. Western blot analysis of PtpA expression in complemented \u00CE\u0094ptpA strain and in vitro PtpA secretion using \u00CE\u00B1-PtpA antibodies ................................ 127 Figure 31. Calibration of THP-1 phagosomal pH during Mtb infection with FACS. ........................................................................................................................... 128 Figure 32. Measurement of phagosomal pH in transfected THP-1 infected with E. coli. ..................................................................................................................... 130 Figure 33. Analysis of antibodies specificity.. .................................................... 132 Figure 34. Immunostaining control experiment with Mtb-infected THP-1. ......... 134 Figure 35. Immunostaining control experiment with E. coli-infected THP-1. ..... 135 \t\r \u00C2\u00A0 xiv LIST OF ABBREVIATIONS Ace Acetamide ADP Adenosine diphosphate Arf ADP ribosylation factor Arg Arginine Asp Aspartate BCG Bacillus Calmette-Gu\u00C3\u00A9rin CFP Culture filtrate proteins CFU Colony forming units CORVET Class C core vacuole/endosome tethering complex CR3 Complement receptor 3 Cys Cysteine DMEM Dulbecco's modified Eagle medium DMSO Dimethyl sulfoxide DTT Dithiothreitol EDTA Ethylenediaminetetraacetic acid EEA1 Early endosome antigen 1 FACS Fluorescence activated cell sorting \t\r \u00C2\u00A0 xv FCS Fetal calf serum Fc\u00CE\u00B3Rs Fc\u00CE\u00B3 Receptors FITC Fluorescein isothiocyanate GFP Green fluorescent protein GST Glutathione s-transferase HBSS Hank's buffered salt solution HEPES 4-(2-hydroxyethyl)-1-piperazineethanesulphonic acid His Histidine HIV Human immunodeficiency virus HOPS Homotypic vacuole fusion and protein sorting IPTG Isopropyl-\u00CE\u00B2-D-thio-galactoside ITAM Immunoreceptor tyrosine-based activation motifs LAMP Lysosome-associated membrane proteins LPS Lipopolysaccharide MDR Multi drug resistant mAGP Mycolyl-arabinogalactan-peptidoglycan MES 2-(N-morpholino)ethanesulfonic acid \t\r \u00C2\u00A0 xvi MOI Multiplicity of infection Mtb Mycobacterium tuberculosis NOX2 NADPH oxidase NK Natural killer OADC Oleic acid albumin dextrose complex PBS Phosphate buffered saline Phox Phagocyte oxidase PFA Paraformaldehyde PI3P Phosphatidylinositol 3-phosphate PtkA Protein-tyrosine kinase A PMA Phorbol-12-myristate-13-acetate PMSF Phenylmethanesulfonyl fluoride PtpA Protein-tyrosine phosphatase A RILP Rab7 interacting lysosomal protein ROS Reactive oxygen species RNS Reactive nitrogen species SE Succinimidyl ester \t\r \u00C2\u00A0 xvii Ser Serine SCID Severe combined immunodeficient SDM Site-directed mutagenesis SNARE Soluble N-ethylmaleimide-sensitive factor (NSF) attachment protein receptor TB Tuberculosis Thr Threonine TNF Tumor necrosis factor Trp Tryptophan Tyr Tyrosine VPS Vacuolar protein sorting WT Wild-type\t\r \u00C2\u00A0\t\r \u00C2\u00A0 XDR Extreme drug resistant \t\r \u00C2\u00A0 xviii ACKNOWLEDGEMENTS The work of this thesis would not be possible without strong support from my supervisor Dr. Yossef Av-Gay. I started in Dr. Av-Gay\u00E2\u0080\u0099s lab as an undergraduate science co-op student. It was during this time when I was inspired by Dr. Av-Gay to undertake the challenge of graduate studies and Tuberculosis research. Throughout my graduate program, Dr. Av-Gay gave me numerous opportunities to experience and learn the various aspects of scientific research, preparing me to be an independent researcher in the future. I especially want to thank Dr. Av- Gay for his selfless guidance and patience during my first year as a PhD student, helping me to persevere through the hardships I encountered while I was adapting to my new role in the research laboratory. My committee members, Dr. Lisa Craig, Dr. Zakaria Hmama and Dr. Steven Pelech have provided me with constant support over the years. Thank you for all your helpful comments and discussions that greatly improved the quality of my work. I would also like to thank all the past and present members of the Av-Gay lab with special thanks to Mary Ko for all the technical assistance. I owe gratitude to Dr. Horacio Bach for mentoring me, sharing his knowledge in research and teaching me the basic research techniques when I first started in Dr. Av-Gay\u00E2\u0080\u0099s lab. I have to thank Joseph Chao, Jim Sun, and Xingji Zheng for their friendships over the years and for creating a lively and joyful environment in the lab. I would like to acknowledge financial support from Canadian Institute of Health Research (CIHR), Michael Smith Foundation For Health Research, and the University of British Columbia. I must thank the support of my family, especially my parents, for their continued support and sacrifice to give me the opportunity to pursue my dream. Finally, I would like to thank Eileen Zhou for her unconditional care, support and advice that help me become what I am today. \t\r \u00C2\u00A0 xix I am also thankful to all those who helped me during this entire work. \t\r \u00C2\u00A0 1 CHAPTER 1: INTRODUCTION 1.1 Tuberculosis 1.1.1 Disease epidemiology Tuberculosis (TB) remains the leading cause of morbidity and mortality due to a single infectious agent (Figure 1). Despite intensive efforts to reduce the global spread of TB disease, one-third of the world\u00E2\u0080\u0099s population is believed to be exposed to Mtb, which causes nearly two million deaths annually. WHO estimated that one billion new infections would occur by 2020 if no urgent and efficient measures to control TB disease are taken [1]. Indeed, the emergence of multi-drug resistant (MDR) and extensively-drug resistant (XDR) TB are impacting upon the efficiency of the current treatment. MDR-TB is defined as a form of TB that is resistant to at least isoniazid and rifampin, which are two of the four front line antibiotics. On the other hand, XDR-TB is defined as resistance to second-line drugs, including fluoroquinolone, and at least one of the three injectable TB drugs, i.e. amikacin, kanamycin, and capreomycin [2]. While treatment for MDR-TB requires up to two years of multi-drug regimen, XDR-TB is virtually untreatable. In fact, in a recent outbreak of XDR-TB in Kwazulu-Natal in South Africa, 52 of the 53 patients died within an average of 25 days after being diagnosed [3]. Furthermore, the association with the Human Immunodeficiency Virus (HIV) has led to the explosive spread of TB in developing countries [4]. Patients who are co-infected with both HIV and TB have an up to 800 times greater risk of developing active TB disease symptoms and becoming infectious compared to people who are not infected with HIV [4]. To complicate this \t\r \u00C2\u00A0 2 problem, despite promising trials for several novel anti-TB therapeutics, no new TB drug was introduced into the clinic in the last 40 years. Therefore, there is an urgent need to understand the physiology and the mechanism of TB pathogenesis in order to identify new targets for the development of novel therapeutics. World Health Organization. Global Tuberculosis Control 2011. Figure 1. Estimated TB incidence rates (2010). This illustration shows the worldwide distribution of TB with the highest incidence rate in sub-Saharan Africa at 350 cases per 100,000 population [5]. Reprinted with permission from World Health Organization. 1.1.2 Origin of human Tuberculosis TB has plagued humankind throughout its recorded history. The etiological agent of TB, Mycobacterium tuberculosis (Mtb), was discovered by Robert Koch in \t\r \u00C2\u00A0 3 1882. Recent genetic studies showed that Mtb might have evolved from an early progenitor, which might have infected early hominids in East Africa around 3 million years ago [6]. Comparison of the genome sequences of various species of the Mtb complex indicates that all members, including Mtb, Mycobacterium africanum, Mycobacterium canettii and Mycobacterium bovis, had a common African ancestor approximately 35,000 - 15,000 years ago [7, 8]. However, despite popular belief, Mtb did not arise from a bovine TB strain as evident from comparative genomics analysis and molecular phylogeny (Figure 2). In fact, M. bovis has undergone numerous deletions in its genome relative to Mtb, resulting in a smaller chromosome (~66 kbp truncation) as compared to Mtb [9]. As it is unlikely that the deletions could be repaired by recombination in clonal organisms, this observation rules out the possibility that Mtb might have evolved from M. bovis. Furthermore, the most recent common ancestor of both Mtb and M. bovis was more similar to Mtb than M. bovis in chromosomal content [10], and this common ancestor is believed to have already been a human pathogen. 1.1.3 Mycobacterium tuberculosis Mtb is a rod-shaped, slow growing, high G+C content (65.6%), aerobic facultative intracellular pathogen. It belongs to the order Actinomycetales and suborder Corynebacterineae together with other related microogranisms such as nocardiae, rhodococci and corynebacteria. Mtb is specially classified as an acido-alcohol resistant bacteria in the division of Actinobacteria due its cell envelope having a very different chemical nature, as compared to Gram-positive and Gram-negative bacteria. The cell envelope of Mtb is composed mainly of a \t\r \u00C2\u00A0 4 mycolyl-arabinogalactan-peptidoglycan (mAGP) complex [11-15], which forms the inner part of the cell wall. Other additional lipids such as phthiocerol dimycocerosates, glycopeptidolipids, menaquinones, and glycosylated phenolpthiocerols intercalate in the mycolic acid layer and form the outer region of the cell wall [16-18]. It is this thick lipid-rich coat and capsule that result in the distinctive property of acid fastness and make the Mtb cell wall 10 \u00E2\u0080\u0093 100 times more impermeable than that of the notably impermeable bacillus Pseudomonas aeruginosa [19]. The low permeability of the cell wall confers upon Mtb an intrinsic resistance to host antimicrobial mechanisms and antibiotics. Smith et al. Nature Reviews Microbiology. 2009. 7: 537-544 Figure 2. Phylogeny of selected mycobacterial species. This illustration shows the evolutionary relationship between Mtb complex species. The deletion of chromosomal region RD9 differentiates Mtb from other the animal specific \t\r \u00C2\u00A0 5 strains. Reprinted with permission from Macmillan Publishers Ltd, Nature Reviews Microbiology. \t\r \u00C2\u00A0 1.1.4 Infection cycle Infection by Mtb occurs by aerosol route, and the infectious dose is estimated to be 5 to 10 bacteria. Once inhaled, Mtb is phagocytosed by the alveolar macrophages, inducing a localized inflammatory response. The infected macrophages release Tumour Necrosis Factor (TNF)-alpha and other inflammatory chemokines that drive the recruitment and activation of neutrophils, natural killer (NK) T cells, CD4+ T cells and CD8+ T cells [20, 21]. The recruited immune cells, in turn, initiate the inflammatory cascade of cytokines, intensifying cellular recruitment and remodeling of the infection site. It is this host response and recruitment of immune cells as an attempt to contain the infection that leads to the formation of the granuloma, a pathologic feature characteristic of TB (summarized in Figure 3) [22] . The granuloma, with extensive vascularization, consists of the core of infected alveolar macrophage, surrounded by other macrophages, monocytes, neutrophils, giant cells, foamy macrophages, and epithelioid macrophages [23, 24]. The lymphocytes, which are recruited to mediate the adaptive immune response, localize to the periphery of the granuloma associated with a fibrous capsule of collagen and other extracellular matrix components [23]. At this stage, generally 2-3 weeks post infection, the infected human host does not show apparent signs of disease or transmit the infection to others, and the mycobacterial infection is considered to be contained [23]. However, the granuloma can reactivate when the host becomes immune- \t\r \u00C2\u00A0 6 compromised, which is usually the consequence of aging, malnutrition or co- infection with HIV. The disease progression in the granuloma is marked by diminished vascular structure and a more prominent fibrous sheath, creating a hypoxic and necrotic center [24]. Ultimately, the granuloma caseates and ruptures into a structureless mass of cell debris, releasing thousands of infectious bacilli into the lung airways [24]. This contributes to the development of a persistent cough, which further aid the aerosol spreading of TB. Figure 3 (next page). Mtb infection cycle. The illustration describes the life cycle of Mtb. Upon inhalation, Mtb is phagocytosed by alveolar macrophages. Host immune responses lead to granuloma formation as an attempt to control infection. However, Mtb reactivates when host immunity is compromised, and the granuloma caseates and ruptures, releasing infectious bacilli that contribute to a productive cough and spread of TB. Reprinted with permission from Nature Publishing Group; Nature Reviews Immunology. \t\r \u00C2\u00A0 7 Russell et al. Nat Immunol. 2009. 10: 943-948 \t\r \u00C2\u00A0 8 1.1.5 Mycobacteria resistance to host killing As an intracellular pathogen, Mtb is capable of adapting to various environmental changes that it encounters within the host macrophage throughout the course of infection. Once Mtb is phagocytosed by the alveolar macrophage, the host cell attacks the invading pathogen with a wide range of antimicrobial mechanisms, including nutrient deprivation, hypoxia, reactive oxygen and nitrogen species (ROS and RNS), acidity and hydrolytic enzymes. A major mycobacterial component that allows Mtb to survive and persist within such a hostile host environment is its thick, waxy, lipid-rich cell wall, which, as mentioned above, is composed of high mycolic acid content. The various complex lipids within the mycobacterial cell envelope not only scavenge harmful reactive radicals from the extracellular environment but also contribute to intrinsic resistance against acidic or alkali conditions and antibiotics [25]. In addition to this protective capsule, successful TB pathogenesis is dependent on several other Mtb intracellular survival strategies. Numerous studies in the past have shown that Mtb is capable of neutralizing ROS and RNS [26], disrupting host signaling pathways [27-29], inhibiting apoptosis to suppress the adaptive immune response [30, 31], and, most importantly, arresting phagosome fusion with lysosomes [32-34]. A summary of some of these mechanisms, with the focus on the arrest of phagosome maturation, will be further discussed in Section.1.3. 1.2 Phagocytosis 1.2.1 Phagosome formation Phagocytosis is a major component of the innate immune defense against \t\r \u00C2\u00A0 9 infectious agents such as bacteria and viruses. The interaction of the microorganism with the phagocyte can occur through recognition of pathogen- associated molecules, such as surface carbohydrates, peptidoglycans or lipoproteins, by pattern recognition receptors [35, 36]. Alternatively, foreign invading particles can be coated by opsonins such as immunoglobulin G (IgG) and components of the complement cascade that attach to the pathogen surface [37-39]. The opsonins are recognized by the phagocytes\u00E2\u0080\u0099 surface receptors such as Fc\u00CE\u00B3 receptors (Fc\u00CE\u00B3Rs) and complement receptor 3 (CR3) [37, 39]. Engagement of the host phagocyte receptors induces a complex network of signaling events (summarized in Figure 4). One such event is the phosphorylation of immunoreceptor tyrosine-based activation motifs (ITAMs) on Fc\u00CE\u00B3Rs by Src-family kinases [40]. ITAM phosphorylation recruits and activates the tyrosine kinase SYK, which in turn phosphorylates various substrates [41]. These signaling pathways trigger cytoskeletal rearrangement, mediated by Rho GTPases Rac1 and Cdc42, and alterations in membrane trafficking, regulated by Rab and ADP Ribosylation Factor (Arf) GTPases, allowing the phagocyte to extend pseudopod around the foreign particle [42, 43]. Phosphoinositides such as phosphatidylinositol-4,5-bisphosphate and phosphatidylinositol-3,4,5- trisphosphate also play a role in mediating the engulfment of the foreign particle [44]. Accumulation of phosphoinositides at the site of particle engagement initiates actin assembly and drives pseudopod formation, ultimately leading to the encapsulation of the foreign particle into a vacuole termed as phagosome. \t\r \u00C2\u00A0 10 \t\r \u00C2\u00A0 Underhill et al. Annu. Rev. Immunol. 2002. 20:825-52 Figure 4. Receptor and signaling interactions during phagocytosis of microbes. Phagocytosis is initiated through recognition of the invading foreign particles or microbes by multiple receptors. Receptor engagement elicits a network of signal transduction pathways with the phagocyte, leading to various responses. Reprinted with permission from Annual Reviews: Annual Reviews in Immunology. 1.2.2 Phagosome maturation Phagosome maturation starts immediately after the nascent compartment pinches off from the surface plasma membrane. Through progressive interactions with components in the endocytic pathway, the phagosome remodels its membrane and luminal contents [45]. It is unclear whether these interactions involve complete fusions with the incoming membranes or \u00E2\u0080\u009Ckiss and run\u00E2\u0080\u009D \t\r \u00C2\u00A0 11 transient passage. It is also unclear whether the maturation process occurs in a linear manner where the phagosome interacts with the early endosomes, late endosomes and lysosomes sequentially or a more complex network of membrane transfers occurs simultaneously. However, there is evidence that upon formation of the nascent phagosome, tubular structures from lysosomes located in the perinuclear region extend along the microtubules to fuse with and deliver crucial contents to the phagosome at an early stage [46]. This indicates that phagosome maturation might occur through a more complex process involving multi-directional interactions with various membrane compartments. Figure 5. Stages of phagosome maturation. Upon phagocytosis, the nascent phagosome interacts sequentially with components of the endocytic pathway to acquire antimicrobial properties, including an acidic pH caused by the proton pumping V-ATPase and hydrolytic enzymes such as Cathepsins. Progressive interaction with early endosome, late endosome and lysosome ultimately leads to the formation of the phagolysosome where the invading microbe is eliminated, and antigens are presented on the surface of the phagocyte to initiate adaptive immune response. \t\r \u00C2\u00A0 12 Nevertheless, the phagosome maturation process can be characterized by several stages that culminate in the creation of a highly acidic, oxidative and degradative compartment (summarized in Figure 5). Upon detaching from the plasma membrane, the nascent phagosome retains properties similar to the early endosome, having a mildly acidic environment (pH 6.0) with poor hydrolytic activity characterized by the presence of the inactive Procathepsin D protease [47]. The small GTPase Rab5 plays a key role during this early phase of phagosome maturation [48]. Rab5 associates with phagosome immediately after phagocytosis and facilitates the recruitment of Rab5 effector proteins such as VPS34 [49]. VPS34 is a class III phosphatidylinositol-3-kinase that generates phosphatidylinositol-3-phosphate (PI3P) on the early phagosomal membrane. PI3P is responsible for recruiting and anchoring effector proteins, including the early endosome antigen 1 (EEA1), to the cytosolic face of the phagosome through FYVE and PX domains [50, 51]. EEA1, which also interacts directly with Rab5, participates in tethering the phagosome with early endosomes [52]. Furthermore, EEA1 binds to the SNARE (soluble N-ethylmaleimide-sensitive factor (NSF) attachment protein receptor) protein syntaxin13 to drive phagosome fusion with early endosomes [53]. PI3P is also required for the recruitment and assembly of the NADPH oxidase complex (NOX2), which interacts with membrane phosphoinositides via the PX domains of its p40phox and p47phox subunits, on the phagosomal membrane [54]. NOX2 generates superoxide in the phagosomal lumen through a redox chain by transferring electrons from cytosolic NADPH to phagosomal oxygen [54]. The superoxide is in turn converted into \t\r \u00C2\u00A0 13 several other powerful microbicidal ROS and RNS, such as hydrogen peroxide, hydroxyl radicals and peroxynitrite when combined with nitric oxide radicals [55]. Further enrichment of PI3P and association of Rab5 downstream effector proteins in the phagosomal membrane are followed by the recruitment of the Rab7 GTPase, a characteristic marker of the late phagosome [56, 57]. The exchange of Rab5 for Rab7 is mediated by the Class C VPS HOPS complex, a key regulator of membrane tethering and fusion in the endocytic pathway [58, 59]. The interaction with both Rab5 and Rab7 indicates that the Class C VPS HOPS complex is crucial for the transition from early to late phagosome [59]. The composition and function of the Class C VPS complex will be discussed further in Section 1.2.4. A key downstream effector molecule recruited by Rab7 is the Rab- interacting lysosomal protein (RILP), which drives the centripetal movement of late phagosomes to perinuclear lysosomes by linking the membrane to the dynein\u00E2\u0080\u0093dynactin motor complex [60]. Fusion of the late phagosome with lysosomes is then driven by the Class C VPS HOPS complex, which tethers the apposed membranes and mediates the assembly of SNARE complex, completing the membrane coalescence event [61, 62]. At this stage, the phagosome has an acidic luminal pH (5.5 \u00E2\u0080\u0093 6.0) due to the acquisition of the proton-pumping vacuolar H+-ATPase (V-ATPase), whose structure and function will be further discussed in Section 1.2.3. The late phagosome is also enriched in proteases and lysosomal-associated membrane proteins (LAMPs), which are either imported from the Golgi complex or acquired by fusion with late endosomes [45]. \t\r \u00C2\u00A0 14 The phagosome maturation process ultimately leads to the formation of the phagolysosome, a terminal microbicidal organelle loaded with an arsenal of antimicrobial effectors that completely degrade the invading microorganisms. Further accumulation of the V-ATPases creates a profoundly acidic lumen (pH 4.5 \u00E2\u0080\u0093 5.0), the hallmark of phagosome maturation. Small antimicrobial peptides such as defensins and cathelicidins [63, 64], which directly disrupt membrane integrity through the formation of multimeric ion channels on bacterial pathogens, are delivered to the phagolysosome. The phagolysosome also has an elevated content of acid-activated hydrolytic enzymes, including lysosomal carboxypeptidase, phospholipase A2, \u00CE\u00B1-hexosaminidase, \u00CE\u00B2-glucuronidase, lysozyme and cathepsins, that target protein, lipid and carbohydrate components of ingested particle [65, 66]. These acid-activated hydrolases mediate the complete destruction of the invading foreign particles, whose constituents are then presented as peptide antigens by MHC molecules on the phagocytes\u00E2\u0080\u0099 surface to activate adaptive immunity. 1.2.2.1 Vacuolar H+-ATPase The ubiquitous eukaryotic V-ATPase is composed of a membrane bound sector (V0) and a cytosolic stalk and catalytic head (V1) [67, 68]. It is structurally and mechanistically related to the ATP synthase (F-ATPase), which is also a multi- subunit protein complex that operates through a rotary mechanism to synthesize ATP [69]. However, unlike the ATP synthase, the V-ATPase hydrolyzes ATP to drive active transport of protons across membranes. The integral V0 domain, which has a molecular mass of 260 kDa, consists of five different subunits (a, d, \t\r \u00C2\u00A0 15 c, c\u00E2\u0080\u0099 and c\u00E2\u0080\u0099\u00E2\u0080\u0099) that are responsible for proton translocation while the cytosolic domain contains eight different subunits (A, B, C, D, E, F, G, H) that are mainly involved in harnessing the energy from the hydrolysis of ATP (Figure 6). The complex is present on the membranous organelles such as lysosomes, endoplasmic reticulum and the Golgi apparatus. ATP hydrolysis is catalyzed by subunit A and B, which form the heterohexameric structure of the catalytic head [70, 71]. The energy from ATP hydrolysis drives the rotation of the entire rotary assembly consisting of the cytoplasmic central stalk (subunit D and F), subunit d and the integral membrane proteolipid ring (c, c\u00E2\u0080\u0099, c\u00E2\u0080\u0099\u00E2\u0080\u0099) [72]. Proton transport is dependent on a crucial glutamate residue in each of the proteolipid subunits that undergoes reversible protonation during the rotation [73]. The glutamate residue is protonated as it engages the cytoplasmic hemi-channel in subunit a. Rotation of the proteolipid ring then exposes the protonated residue to the lumenal hemi- channel, where an arginine residue forces the deprotonation of the glutamate residue leading to proton translocation [74]. Subunit E, C, G and H form the peripheral stator that maintains the orientation of the V-ATPase subunits during the rotational catalysis [75]. Interestingly, subunit H is the only subunit of V- ATPase of which an analogous subunit is not found in F-ATPase, distinguishing the V-ATPase complex from the F-ATPase complex. A deletion mutant of the subunit H gene in yeast showed that it is essential for the activity but not the assembly of the enzyme complex, suggesting that subunit H is a regulator of the V-ATPase [76]. This finding was further supported by the observation that when V1 and V0 domains of the V-ATPase complex dissociate, subunit H remains \t\r \u00C2\u00A0 16 bound to cytosolic V1 domain and inhibits the unproductive ATPase activity of the dissembled V1 domain [77]. In addition to the elucidation of the proton transport mechanism, recent studies showed that the V-ATPase complex is capable of interacting with other proteins to regulate their activities [78-80]. In particular, the small GTPases ADP-ribosylation factor nucleotide site opener (ARNO) and Arf6, which are endocytosis regulators involved in carrier-coat vesicle formation, have been found to be recruited directly by V-ATPase to endosomes, in a pH dependent manner, to initiate vesicle formation. These observations reveal that V-ATPase is a multi-functional enzyme complex that modulates a network of cellular processes. Furthermore, studies of Drosophila neurons and mammalian renal medulla cells have shown interaction between subunits of V-ATPase and SNARE proteins, reinforcing the notion that V-ATPase might act as a scaffolding structure to modulate membrane trafficking and fusion [81, 82]. Accumulation of the V-ATPase complex and acidification of the phagosome lumen is therefore not only crucial for the inhibition of bacterial survival and activation of acid-activated hydrolytic enzymes, but also proper trafficking of the phagosome within the endocytic pathway. \t\r \u00C2\u00A0 17 Forgac et al. Nat. Rev. Mol. Cell. Biol. 2007. 8:917-929 Figure 6. Vacuolar H+-ATPase. This illustration shows a model of the overall structure of the S. cerevisiae V-ATPase complex, comprising of the V1 cytosolic domain and V0 membrane integral domain. The Arg-735 residue in subunit a is the key residue involved in proton translocation. Reprinted with permission from Nature Publishing Group: Nature Reviews Molecular Cell Biology. 1.2.2.2 Class C VPS The Class C Vacuolar Protein Sorting (VPS) complex comprises four subunits, VPS11, 16, 18 and 33, that were first identified in classical genetics screens in Saccharomyces cerevisiae [83, 84]. Mutations in each of the four genes resulted in a failure of vacuolar lysosome formation in mutant yeast. This observation, together with biochemical studies, led to the conclusion that the Class C VPS complexes play a key role in the endocytic trafficking pathway, especially at the endosome-to-lysosome stage of transport where it functions together with Rab \t\r \u00C2\u00A0 18 GTPases to tether membrane compartments and to assemble trans-SNARE complexes for membrane fusion [62, 85, 86]. In particular, Class C VPS complex was found to bind to the vacuolar SNARE syntaxin7 to regulate its activity [87]. More recent studies in S. cerevisiae further indicate that the Class C VPS complex acts as a core protein assembly that can interact with other VPS accessory subunits to participate at different stages of membrane transport. Specifically, the Class C VPS complex can interact with VPS8 and VPS3 to form the CORVET (class C core/vacuole endosome tethering) complex on Rab5- positive early endosomes to mediate fusion with late endosomes and with VPS41 and VPS39 to form the HOPS (homotypic fusion and protein sorting) complex that, together with Rab7, regulates lysosomal transport [88, 89] (Figure 7). Although the CORVET complex has not yet been identified in mammalian cells, it likely exists and has the same functions given the high similarity in the transport machinery between yeast and mammals. Peplowska et al. [88] showed that the HOPS and CORVET complexes could reversibly interconvert and also identified two intermediates depending on which subunits had been exchanged (Figure 7). Two models for the conversion have been postulated [88]. The first model assumes that the intermediates can assemble de novo and bind to different Rab GTPases on endocytic membranes. The second model links the observed order of intermediates to the Rab GTPase cycle, displacing and recruiting Rab GEFs (guanine exchange factors) and Rab effectors as necessary. In fact, VPS39, the GEF that interacts with and activates Rab7, can also bind to Rab5, indicating that the conversion of CORVET to HOPS is dependent on and coincides with the \t\r \u00C2\u00A0 19 exchange of Rab5 for Rab7 [59, 88]. The phagosome maturation process is therefore largely dependent on the dynamics of the interaction between Rab5 and Rab7 GTPases and the Class C VPS complex, and the recruitment of their respective downstream effector proteins to direct the membrane trafficking and fusion along the endocytic pathway. Nickerson et al. Curr Opin Cell Biol. 2009. 21: 543-551 Figure 7. Working model for Class C VPS function in yeast. Class C VPS complex serves as the core for the CORVET and HOPS complexes through reversible association with CORVET-specific subunits (VPS3 and VPS8) and HOPS-specific subunits (VPS39 and VPS41). CORVET and HOPS regulate membrane fusion events at late endosomes and lysosomes, respectively, through interaction with Rab GTPases. Hybrid intermediate complexes containing both HOPS and CORVET subunits are known to exist. Reprinted with permission from Elsevier Inc.: Current Opinion in Cell Biology. 1.3 Mtb arrest of phagosome maturation 1.3.1 Overview The ability of Mtb to persist and replicate within the host macrophage after phagocytosis is central to TB pathogenesis [90]. As mentioned above, intracellular survival of Mtb results from a combination of factors including a unique cell wall structure, which physically shields the bacterium from \t\r \u00C2\u00A0 20 bactericidal and hydrolytic enzymes [91], and secretion of enzymes to combat host ROS and RNS [92, 93]. Although these factors contribute to Mtb persistence within the macrophage, one recurring and highly important feature of this pathogen is inhibition of the normal phagosome maturation process, thereby abrogating physical fusion of phagosome with lysosomes and ultimately protecting the bacterium from a bactericidal environment (summarized in Figure 8). Mtb arrests phagosome maturation at an early stage, resulting in a mycobacterial phagosome that possesses similar properties of the early endosome, characterized by membrane acquisition of the early endosome marker Rab5 [94]. However, the key Rab5 effectors, EEA1 and VPS34, are excluded from the mycobacterial phagosome [27]. This phenotype is due to Mtb\u00E2\u0080\u0099s capability of interfering with the macrophage Ca2+ fluxes during phagosome maturation [95]. The block in Ca2+ fluxes, mediated by Mtb cell wall glycolipid ManLAM (mannose-capped lipoarabinomannan), results in the inhibition of the Ca2+/Calmodulin protein kinase CaMKII signaling cascade, which is required for the recruitment of VPS34 to the phagosomal membrane [28]. The absence of VPS34 impairs the generation of PI3P in the phagosomal membrane, resulting in the failure to recruit EEA1 and the inhibition of the phagosome maturation process. Another major characteristic of the mycobacterial phagosome is the lack of acidification, restricting the compartment to a pH of 6.4 where Mtb could survive and persist. It is a common theme among intracellular pathogens that \t\r \u00C2\u00A0 21 enter the host cells through the phagocytic or endocytic pathway to possess mechanisms to counter the acidic environment within the phagosome. For instance, Salmonella enterica maintains the pH homeostasis by upregulation of its acid tolerance response genes [96]. Yersinia pseudotuberculosis blocks phagosome acidification by direct inhibition of V-ATPase activity [97]. Legionella pneumophila secretes SidK into the host phagocyte to inhibit vacuole acidification through interaction with V-ATPase subunit A [98]. In the case of Mtb, it is long accepted that the lack of phagosome acidification is mainly due to exclusion of the V-ATPase from the phagosomal membrane [99]. Recent studies have also demonstrated that Mtb might be capable of disrupting the Rab7 GTPase mediated membrane trafficking pathway. Sun et al. [100, 101] showed that nucleoside diphosphate kinase (Ndk) is secreted by M. bovis BCG into the cytosol of infected cells to act as a GTPase-activating protein (GAP) that catalyzes the conversion of the active Rab7-GTP to the inactive Rab7-GDP form. The lack of active Rab7 on the phagosomal membrane results in the failure to recruit RILP, thus halting phagosome fusion with lysosomes. \t\r \u00C2\u00A0 22 Philips. Cell Microbiol. 2008. 10:2408-15 Figure 8. Model of Mtb inhibition of mycobacterial phagosome maturation. Mtb proteins or lipids are colored red, and host proteins or lipids are illustrated in blue. SapM, a PI3P phosphatase, and the cell wall glycolipid ManLAM together interfere with PI3P production on the phagosomal membrane, thereby leading to the exclusion of EEA1 from the mycobacterial phagosome. PtpA shuts down the membrane fusion machinery through dephosphorylation of VPS33B in the Class C VPS complex. A Rab7 GTPase-activating protein (GAP) also disrupts Rab7 activation, impairing phagosome trafficking to the lysosome. Reprinted with permission from Blackwell Publishing Ltd: Cellular Microbiology. 1.3.2 Mtb phosphatase interference with host signaling pathways In addition to the aforementioned mycobacterial proteins and lipids, Mtb utilizes a wide repertoire of signal transduction systems to sense and respond to the hostile host environment and modulate the host signaling pathways. These \t\r \u00C2\u00A0 23 systems include eleven \u00E2\u0080\u009Ctwo-component\u00E2\u0080\u009D systems, eleven eukaryotic-like protein- serine/threonine kinases (STPKs) (PknA-PknL), two protein-tyrosine phosphatases (PTPs) (PtpA and PtpB), and the newly identified protein-tyrosine kinase (PTK) (PtkA) [32, 102-105]. The role of mycobacterial kinases in Mtb pathophysiology, particularly the regulation of cell wall biogenesis and dormancy, has been widely studied, spurring research efforts in the identification of compounds targeting these signal transduction components. In recent years, as a better understanding of the cross-talk between host and mycobacterial signaling pathways is gained, there was an immense interest in targeting this host-pathogen interface for novel antituberculosis drugs. Among the proteins secreted by Mtb into the host cytosol, three phosphatases, PtpA, PtpB and SapM, are essential for Mtb survival in human macrophages and in animal models of infection [32, 34, 106, 107]. While SapM was found to be a lipid phosphatase that inhibits the proper generation of phosphatidylinositol 3-phosphate (PI3P) that is essential for phagosome recruitment of EEA1 [34], PtpA was shown by us to target the human VPS33B in order to inhibit phagosome-lysosome fusion [32]. Although the host cognate substrate of PtpB remains to be elusive, this phosphatase was shown to contribute to the establishment of latent infections. Thus, identifying potential inhibitors that specifically target these Mtb phosphatases is the main focus of many laboratories with the ultimate goal to generate new TB drugs [106, 108-113]. \t\r \u00C2\u00A0 24 1.3.2.1 PtpB PtpB, along with PtpA, was first identified from the genome sequence of Mtb H37Rv through its homology to known eukaryotic PTPs. Specifically, PtpB possesses the typical C(X)5R(S/T) signature motif in the active site phosphate binding loop (P-loop) of eukaryotic PTPs. Koul et al. [104] showed that Mtb PtpB is secreted into culture media and its phosphatase activity is specific for phosphotyrosine substrates. Additionally, crystallographic analyses revealed that PtpB possesses the distinct features of dual phosphotyrosine binding sites and a two-helix lid structure that covers and protects the active site of the enzyme in an oxidative environment [114]. From these observations, it has been postulated that Mtb secretes PtpB into either the phagosome or host cytosol during infection to interact with host proteins and disrupt macrophage signaling pathways. However, a recent study, which re-examined the sequence homology and enzymatic activity of PtpB, indicates that PtpB, within its C(X)5R(S/T) signature motif, contains the conserved AGK and DRT motifs (CFAGKDRT) found in phosphatase and tensin homolog (PTEN) and myotubularin (MTM) phosphoinositide lipid phosphatase, respectively [115]. With an active site region that also resembles the eukaryotic dual-specificity phosphatase that can dephosphorylate both phosphotyrosine and phosphoserine/threonine, PtpB was found to possess the unique property of triple-specificity for phosphoinositides, phosphotyrosine and phosphoserine/phosphothreonine [115]. Based upon these results, it is reasonable to suggest that PtpB might also disrupt host phosphoinositide metabolism and its associated signaling pathways, which are \t\r \u00C2\u00A0 25 known to have a key role in phagosome maturation. However, thus far, the cognate substrates of PtpB remain elusive. Despite the lack of information on PtpB targets within the host macrophage, there is compelling evidence that PtpB is essential for TB pathogenesis [106, 107]. Indeed, ptpB deletion in Mtb H37Rv strain led to decreased bacterial survival in IFN-\u00CE\u00B3 activated murine J774A.1 macrophage-like cells and a significant reduction of Mtb persistence in the organs of infected guinea pigs 6 weeks post-infection as compared to infection with the parental strain [107]. This late replication defect of the \u00CE\u0094ptpB mutant coincides with the development of efficient T lymphocyte-mediated responses, which supports a role for PtpB in Mtb\u00E2\u0080\u0099s evasion of adaptive immune response. A more recent study examining the effect of PtpB on IFN-\u00CE\u00B3-mediated signaling pathways showed that expression of PtpB in activated RAW264.7 murine macrophage-like cells resulted in reduced phosphorylation of the ERK1/2 and p38 mitogen-activated protein kinases (MAPK) [116]. Suppression of ERK1/2 and p38 activities led to decreased production of IL-6 [116], which has been implicated in the induction of innate and adaptive immune response during Mtb infection. Furthermore, PtpB expression resulted in the inhibition of macrophage apoptosis in response to IFN- \u00CE\u00B3 activation as evident from increased Akt phosphorylation and decreased caspase-3 activation [116]. In fact, it has been previously observed that mycobacterial infection can induce the activation of ERK1/2 and p38 MAPK during entry into macrophages; however, this activation of MAPK signaling cascades is only observed with attenuated or non-pathogenic strains whereas \t\r \u00C2\u00A0 26 pathogenic species quickly diminish the host kinase activities upon infection [117]. Although there is no evidence suggesting that ERK1/2 and p38 are direct substrates of PtpB, these observations, taken together, indicate that PtpB participates in disrupting host signal transductions to subvert host immune response to Mtb infection and to promote Mtb intracellular survival. It is interesting to note that Mtb antigens such as ESAT-6, Heparin binding hemagglutinin (HBHA) and various cell wall glycolipids and glycoproteins are known to elicit the MAPK and PI3K/Akt signaling pathways [118-121]. It is therefore likely that PtpB plays a key role as a damper balancing host immune detection with evasion during Mtb infection. The relevance of PtpB to Mtb intracellular survival was further confirmed by a recent study showing that specific inhibition of PtpB impairs intracellular survival of Mtb within murine macrophages [106]. Multiple studies have already identified compounds that specifically target and inhibit PtpB; however, PtpB\u00E2\u0080\u0099s exact mechanism of action within the host macrophage remains to be elucidated. 1.3.2.2 SapM SapM (secreted acid phosphatase of Mtb) was first identified from fractionated culture filtrate of the Mtb H37Rv strain, which was known to contain phosphatase activity, as a non-specific acid phosphatase [122]. Interestingly though, SapM does not exhibit significant homology with any prokaryotic acid phosphatase. In addition, SapM lacks any conserved sequence motifs of protein-serine/threonine phosphatases, protein-tyrosine phosphatases, metallophosphoesterases, and histidine phosphatases. The absence of any identifiable motifs posed questions \t\r \u00C2\u00A0 27 on SapM function. However, SapM does possess two histidine residues that are highly conserved in fungal acid phosphatases. In fact, three homologous sequences were identified in SapM, belonging to known fungal acid phosphatases from Penicillium chrysogenum, Aspergillus nigerficuum and Kluyveromyces lactis, could be found [122-125]. Furthermore, the presence of a typical N-terminal signal peptide indicates that SapM is secreted by Mtb through the general secretion system (Sec) pathway [122]. The secretory nature of SapM, therefore, indicates that it might function to interact with host molecules during infection. When SapM was tested on various phospho-substrates, it was initially found that it does not have activity on phospholipids such as phosphatidylcholine, phosphatidylethanolamine, and phosphatidic acid [122]. Rather, it exhibits a notable activity on phosphoenolpyruvate, glycerophosphate, GTP, NADPH, phosphotyrosine, and trehalose-6-phosphate [122]. Subsequent studies by Vergne et al. [34] showed that PI3P, a membrane trafficking regulatory lipid essential for phagosomal acquisition of lysosomal constituents, is retained on phagosomes harboring dead mycobacteria but is continuously eliminated from phagosomes with live bacilli. While host macrophage PI3P phosphatases such as MTMs were not responsible for the absence of PI3P from the mycobacterial phagosome, SapM was found to possess PI3P phosphatase activity and to be responsible for the removal of PI3P [34]. This indicated that SapM, together with the cell wall glycolipid ManLAM, mediates the arrest of phagosome maturation during Mtb infection through disrupting the recruitment of PI3P effector proteins \t\r \u00C2\u00A0 28 such as EEA1. Indeed, addition of SapM to an in vitro assay inhibited phagosome fusion with late endosomes [34]. A more recent study of SapM in M. bovis BCG demonstrated that deletion of sapM results in a better vaccine strain that improved survival of mice challenged with Mtb [126]. The \u00CE\u0094sapM BCG vaccine was more effective than the parental vaccine in inducing recruitment and activation of CD11c+ MHC-IIint CD40int dendritic cells (DCs) to the draining lymph nodes. Reduced activation of DCs in animal vaccinated with wild type BCG could be the result of SapM inhibition of phagosome maturation, thereby preventing the subsequent antigen presentation and the initiation of adaptive immunity. 1.3.2.3 PtpA As mentioned before, PtpA was identified from the genome sequence of Mtb H37Rv through its homology to known eukaryotic PTPs. The gene encoding PtpA is located within an apparent operon together with ptkA and Rv2235 upstream and downstream, respectively. PtpA is classified as a low molecular weight PTP (LMW-PTP) based upon its sequence and structural similarity with the family of 18-kDa enzymes involved in cell growth regulation in eukaryotes [127]. It contains the conserved C(X)5R(S/T) signature motif typical of functional PTPs, where the Cys-11 residue serves as the nucleophile, attacking the phosphoryl group on the Tyr residue of the protein substrate. The active site Arg- 17 residue forms bidentate hydrogen bonds with the phosphoryl group in the substrate through its guanidinium group, stabilizing the transition state during catalysis. The invariant Asp-126 residue is critical for release of substrate and regeneration of active enzyme (Figure 9). Despite overall topology similarity with \t\r \u00C2\u00A0 29 the eukaroytic LMW-PTPs, significant deviations regarding the amino acids involved can be found in the crevice leading to the active site region. Different physiological function may have arisen from the different substrate specificities [127]. Interestingly, the presence of an additional cysteine (Cys-16) residue in the active site may form an intramolecular disulfide bridge with Cys-11. The formation of the disulfide bond was proposed to protect the nucleophilic Cys-11 from excessive oxidation, conferring rapid sensing and responding to the redox status of the environment [128]. On the other hand, a more recent study showed that PtpA, when treated with reactive nitrogen species (RNS) in vitro, is nitrosylated on the solvent-exposed Cys-53 residue, leading to a 50% reduction in phosphatase activity. On the other hand, both Cys-11 and Cys-16, which have lower solvent-accessibility, are unaffected by the treatment [129]. While the existence of the active site disulfide bridge in PtpA has not been observed directly, Cys-53 nitrosylation indicates that PtpA could be susceptible to the oxidative environment in the host macrophage. This would then raise questions on how PtpA could maintain its activity in the macrophage to disrupt host signaling pathways. It is tempting to speculate that another reduction system to protect PtpA from oxidative or nitrosative stress might exist. The exact mechanism of redox regulation of PtpA and its role in Mtb pathogenesis remains to be elucidated. \t\r \u00C2\u00A0 30 Bach et al. Cell Host Microbe. 2008. 3: 316-22 Figure 9. Scheme of PtpA catalytic activity. PtpA cysteine at position 11 attacks nucleophilically the phosphorylated tyrosine substrate. PtpA aspartate at position 126 protonates the phenolic oxygen and a cysteinyl-phosphate intermediate is formed facilitating the cleavage of the P-O bond. The transient intermediate is hydrolyzed by an activated water molecule. Reprinted with permission from Elsevier Inc.: Cell Host & Microbe. The fact that PtpA is secreted from Mtb and the absence of a gene encoding a typical protein tyrosine kinase (PTK) in the Mtb genome sequence indicates that PtpA is involved in dephosphorylation of host tyrosine phosphorylated protein and thus may be involved in the modulation of host signaling pathways during infection. Indeed, initial studies in our laboratory demonstrated that PtpA is released by Mtb H37Rv into culture media and that PtpA expression is up regulated upon Mtb entry into host macrophages [104, 105]. Furthermore, PtpA export into the host macrophage cytosol can be detected directly with immuno-electron microscopy and western blot analysis of the cytosolic fractions of infected macrophages. More importantly the phenotype observed after expression of neutralizing single-chain antibodies against PtpA within macrophages clearly demonstrate the presence of PtpA within the cytosol of the infected macrophage [32]. Yet, the exact mechanism by which PtpA translocates into the host macrophage cytosol remains to be identified. The lack of an amino-terminal signal sequence rules out the SecA1 and twin-arginine \t\r \u00C2\u00A0 31 translocation (Tat) pathway for the secretion of PtpA. While the SecA2 or ESX/Type VII export systems are possible candidates responsible for PtpA secretion, there is also evidence indicating that bacterial proteins less than 70 kDa in molecular size can cross the phagosomal membrane [130]. This observation might be explained with the recent discovery that the ESX-1 secretion system with its substrate ESAT-6 can perturb host cell membrane to facilitate the translocation of mycobacterial proteins into the macrophage cytosol [131]. Further study is required to investigate whether PtpA is a direct substrate for the ESX/Type VII secretion system. Within Mtb-infected macrophage, PtpA was found to disrupt key components of the endocytic pathway in order to arrest phagosome maturation [32]. Initial studies showed that expression of PtpA in RAW 264.7 murine macrophage-like cells results in a decrease of phagocytic activity and an increase of F-actin nucleation on phagosomes in vivo and in vitro. The authors of this study postulated that F-actin polymerization might physically block direct contact between donor and acceptor membranes, thereby inhibiting phagosome\u00E2\u0080\u0093 lysosome fusion. However, in a separate study, PtpA was shown to be dispensable for Mtb growth in a mouse infection model [132]. Nevertheless, using substrate trapping and radioactive \u00CE\u00B332P-ATP\t\r \u00C2\u00A0 labeling approaches, the human Vacuolar Protein Sorting 33B (VPS33B), a subunit of the Class C VPS complex, was identified as the cognate substrate of PtpA. As the Class C VPS is a key regulator of membrane trafficking, tethering and the assembly of trans- SNAREs complexes in the endocytic pathway [133], PtpA dephosphorylation of \t\r \u00C2\u00A0 32 VPS33B inactivates the membrane fusion mechanism, leading to inhibition of phagosome-lysosome fusion [32]. In support of this, deletion of ptpA in Mtb H37Rv impaired the intracellular survival of the bacteria with host human THP-1 macrophage-like cells, and phagosomes harboring the \u00CE\u0094ptpA strain showed increased fusion with the lysosomes and accumulation of lysosomal markers as compared to the parental strain. PtpA is, therefore, essential for the pathogenesis of TB. 1.4 Aims of the study Although the cognate substrate of PtpA within human macrophages has been identified, the complete mechanism of action of PtpA in relation to phagosome maturation arrest has not been characterized. In particular, from the substrate trapping experiment performed to identify VPS33B, we observed that recombinant wild type PtpA protein, which should not have substrate-trapping capability, was able to pull down another human protein with an apparent MW of 50 KDa from THP-1 macrophage-like cell lysate. MALDI-TOF mass spectrometry identified this protein as the subunit H of the V-ATPase complex, indicating that PtpA might be capable of directly disrupting phagosome acidification, an essential arm of the phagosome maturation process. The significance of this discovery stems from the lack of a mechanistic explanation for the long established paradigm of Mtb pathogenesis: the absence of acidification in the mycobacterial phagosome. The overall objective of this thesis is therefore to characterize the complete mechanism by which PtpA contributes to Mtb persistence within the macrophage. The working hypothesis for this project is \t\r \u00C2\u00A0 33 that Mtb PtpA interferes with multiple host macrophage pathways that are essential for phagosome maturation in order to promote Mtb intracellular survival. To investigate this hypothesis, our initial goal was to investigate whether the subunit H of V-ATPase is a bona fide catalytic or binding substrate of PtpA. When the interaction between PtpA and V-ATPase subunit H was established to be genuine, our aim was to identify the downstream effects of this host-pathogen protein-protein interaction on the various aspects of the phagosome maturation process. In the second part of the project, we performed biochemical studies on PtkA, encoded by an open reading frame (Rv2232) upstream of ptpA within the PtpA operon, and investigated its functional relationship with PtpA. Overall, the study has provided significant insight into Mtb evasion of host macrophage killing. Specifically, we discovered the long-sought mechanism behind the lack of acidification in mycobacterial phagosome and identified a novel pathway required for successful phagosome maturation. The complete elucidation of human signaling pathways disrupted by PtpA will perhaps lead to new therapeutics to combat the global spread of TB. \t\r \u00C2\u00A0 34 CHAPTER 2: MATERIALS AND METHODS 2.1 Reagents and chemicals 2.1.1 Commercial reagents DMEM, Fetal calf serum (FCS) and HBSS were purchased from Gibco Laboratories (Burlington, Ontario, Canada). Other culture reagents were purchased from Sigma-Aldrich (St. Louis, MO). Protease inhibitor mixture, PMSF, trypsin-EDTA, and glutathione-agarose beads were purchased from Sigma- Aldrich. Protein A-agarose beads were from Bio-Rad laboratories (Hercules, CA). Nickel-Nitriloacetic acid (Ni-NTA) polyhistidine-Tag purification resin was purchased from Qiagen (Mississauga, ON, Canada). pHrodo Succinimidyl Ester (SE), Alexa-Fluor 350 Carobxylic Acid SE and Alexa-Fluor 488 Carobxylic Acid SE were purchased from Invitrogen (Carlsbad, CA). Aldehyde/sulfate latex beads (diameter, 4 \u00C2\u00B5m) were obtained from Interfacial Dynamics (Portland, OR). [\u00CE\u00B3-32P]-adenosine 5`-triphosphate (ATP) and [\u00CE\u00B3-32P]-guanosine 5`-triphosphate (GTP) were purchased from Perkin Elmer (Boston, MA). Restriction enzymes, Pfu DNA polymerase, and T4 DNA ligase were purchased from Fermentas (Burlington, ON, Canada) or New England Biolabs (Ipswich, MA). Luria-Bertani (LB) broth and LB agar were from Fisher Scientific (Pittsburgh, PA). Oleic acid, dextrose and catalase complex (OADC) and 7H9 and 7H10 agar culture media were from Difco Laboratories, (Detroit, MI). Kanamycin was purchased from Gibco Laboratories. Hygromycin was purchased from Roche. MinElute PCR purification and plasmid purification Mini and Midi Kits were from Qiagen. \t\r \u00C2\u00A0 35 2.1.2 Antibodies Rabbit polyclonal anti-VPS33B and rabbit polyclonal anti-subunit E of V-ATPase (ATP6V1E1) IgG were purchased from Proteintech Group (Chicago, IL). Rabbit polyclonal anti-subunit H of V-ATPase (ATP6V1H) IgG and rabbit polyclonal anti- subunit B of V-ATPase (ATP6V1B) IgG were purchased from Santa Cruz Biotechnology (Santa Cruz, CA). Mouse monoclonal Anti-His-tag antibody and rabbit polyclonal anti-VPS18 IgG were purchased from Applied Biological Materials (Vancouver, British Columbia, Canada). Mouse polyclonal anti-subunit H of V-ATPase IgG was purchased from Sigma-Aldrich, 4G10 Anti- phoshotyrosine antibody was purchased from Millipore (Billerica, MA). Rabbit anti-PtpA was described previously [134]. Secondary horseradish peroxidase (HRP) conjugated goat anti-rabbit antibodies was purchased from Sigma-Aldrich. Secondary horseradish peroxidase (HRP) conjugated goat anti-mouse antibodies was purchased from Cedarlane Labs. Texas Red conjugated goat anti-rabbit IgG, Alexa Fluor 488-conjugated goat anti-rabbit IgG and Alexa Fluor 488-conjugated goat anti-mouse IgG were purchased from Invitrogen. FITC-conjugated anti- rabbit and anti-mouse IgG were purchased from Sigma Aldrich. Specificity of the commercial antibodies was tested with FACS analysis, Western blot analysis and immunostaining control experiments. 2.2 Bacteria 2.2.1 Strain maintenance and generation Mtb H37Rv and derivative strains were grown in Middlebrook 7H9 broth (BD Diagnostic Systems, Mississauga, ON, Canada) supplemented with 10% (v/v) \t\r \u00C2\u00A0 36 OADC (oleic acid, albumin and dextrose solution; BD Diagnostic Systems) and 0.05% (v/v) Tween-80 (Sigma-Aldrich) at 37\u00C2\u00B0C standing or rolling in incubator. Red fluorescent Mtb H37Rv was generated by transforming the wild type Mtb H37Rv strain with pSMT3 vector encoding the DsRed protein under the control of the Hsp60 promoter (pSMT3-DsRed). For complementation of \u00CE\u0094ptpA, DNA sequences encoding wild type PtpA or mutant PtpA under the control of the native promoter, or the hps60 promoter, were cloned into the pKP201 integrative vector [135]. The \u00CE\u0094ptpA strain was co-transformed with the resulting plasmids and pBSint, a non-replicating plasmid that provides integrase in trans but is subsequently lost from the cells, thereby reducing the chances of integrase- mediated excision of the complementing DNA [136]. Transformants were selected on 7H11 medium with 50 \u00C2\u00B5g/mL hygromycin. 2.2.2 Bacterial preparation for infection Bacteria in mid-log phase were harvested by centrifugation at 12,000 rpm. In indicated experiments, bacteria were stained by pHrodo SE (Invitrogen), Alexa- Fluor 488 carboxylic acid SE (Invitrogen) or Alexa-Fluor 350 carboxylic acid SE (Invitrogen) at a final concentration of 10 \u00C2\u00B5g/mL at 37\u00C2\u00B0C for 1 hour. After staining, the bacteria were washed three times with 7H9 with 0.05% Tween-80. Prior to infection of differentiated THP-1 macrophage-like cells, the bacteria are also opsonized with human serum for 30 minutes at 37oC, washed three times with 7H9 with 0.05% Tween-80 and passed through 25 gauge needles several times to prevent bacteria clumping. \t\r \u00C2\u00A0 37 2.2.3 Mycobacterial lysis Bacterial cell pellets were resuspended in 500 \u00C2\u00B5L of 50 mM Tris pH 7.4, 100 mM NaCl, 5 mM EDTA, 1% Triton X-100 and 1 mM PMSF, then mixed with 80 mg of 0.1 mm silica beads (Biospec) and lysed by bead-beating at 30 seconds intervals for 3 times. Subsequently, soluble lysates were obtained by centrifugation at 12,000 rpm for 15 minutes. 2.2.4 Transformation of mycobacteria Competent mycobacteria were prepared by harvesting mid-log phase bacteria and washing three times with 10% glycerol. 200 \u00C2\u00B5L of competent cells were then mixed with plasmid DNA in a 0.2 cm electroporation cuvette (Bio-Rad, Hercules, CA). Cells were electroporated with a 2.5V pulse, and allowed to recover in 7H9 supplemented with 10% OADC and 0.05% Tween-80 in the absence of antibiotics for 2 hours (M. smegmatis) or overnight (Mtb) at 37oC. Transformants were selected on 7H10-OADC agar supplemented with appropriate antibiotics. 2.3 Cell culture 2.3.1 Maintenance of tissue culture and differentiation The human monocytic leukemia THP-1 cell line (TIB-202, American Type Culture Collection, Manassas, VA) was maintained in 125 cm2 tissue culture flasks (Corning Inc., Corning, NY) in RPMI 1640 medium supplemented with 10% FBS, and 1% L-Glutamine. THP-1 cells were seeded onto 12-well (1.0 x 106 /well) or 24-well (5.0 x 105 /well) tissue culture plates and differentiated into macrophage- like cells with 40 ng/mL 12-phorbol 13-myristate acetate (PMA) in complete RPMI media. \t\r \u00C2\u00A0 38 2.3.2 Single transfection of differentiated THP-1 THP-1 cells were transfected with the mammalian pEGFP vector harboring the GFP-tagged PtpA constructs (pEGFP-N1-GFP) using the Nucleofector Cell Line Nucleofector Kit V (Lonza) according to manufacturer\u00E2\u0080\u0099s protocol. Transfected macrophages were seeded in 12 well or 96 well plates and differentiated with PMA as described above. Expression of GFP-tagged proteins was measured 12 hours after transfection by spectrofluorometry (Emission 535 nm). 2.3.3 Double transfection of differentiated THP-1 THP-1 cells were seeded onto coverslips in 24 well plates at a density of 1 million cells per well, and differentiated with PMA at 40 ng/mL overnight. Transfection was carried out by magnetofection using PolyMAG reagent (Chemicell, Berlin, Germany). In brief, 1.5 \u00C2\u00B5g of each DNA construct (pEGFP-N1-H and pDsRed2- N1-VPS33B) was mixed with 1 \u00C2\u00B5L of PolyMAG reagent in a volume of 200 \u00C2\u00B5L serum free and supplement free RPMI 1640 (Sigma). This mixture was incubated for 15 minutes at room temperature and then added to each well in 300 \u00C2\u00B5L of complete RPMI media. Thereafter, plates were placed on a magnetic field and incubated for 30 minutes at 37oC. The magnet was then removed and cells were incubated overnight for expression of the fluorescent proteins. 2.4 Vectors and DNA manipulation 2.4.1 RNA isolation and cDNA synthesis Total RNA was prepared from lysates of THP-1-derived macrophages (1 x 107 cells) using Qiagen RNeasy Mini kit according to the manufacturer's instructions. Synthesis of cDNA was performed using Revert Aid First strand cDNA synthesis \t\r \u00C2\u00A0 39 kit (Fermentas). For each cDNA synthesis 5 \u00C2\u00B5g of total RNA and 0.5 \u00C2\u00B5g oligo(dT) primers were used. 2.4.2 Cloning The gene encoding Mtb PtpA was PCR-amplified from Mtb H37Rv genomic DNA using the PtpA-F1 primer with an EcoRI site, and PtpA-R1 primer with an XhoI site. The amplified gene was inserted into pGEX-4T3 vector (GE Healthcare) to generate the pGEX-4T3-PtpA plasmid for GST-tagged PtpA fusion protein expression. For cloning into the pALACE vector to create pALACE-PtpA for His- tagged fusion protein expression construct, the gene encoding Mtb PtpA was amplified from Mtb H37Rv genomic DNA using the PtpA-F2 primer, containing an AflII site, and PtpA-R2 primer, containing a ClaI site. For constructing the \u00CE\u0094ptpA complementation plasmid, the ptpA gene was amplified from pGEX-4T3-PtpA using PtpA-F3 primer, containing an EcoRI site, and PtpA-R2 primer, containing a ClaI site. The amplicon was inserted downstream of the Mtb hsp60 promoter in the pMV261 mycobacterial vector, generating a pMV261-PtpA plasmid. The entire promoter-gene construct was then excised using XbaI and ClaI restriction enzymes and inserted into the pKP201 mycobacterial complementation vector, generating pKP201-hPtpA plasmid. To construct the PtpA-GFP mammalian transfection construct, ptpA was amplified from pGEX-4T3-PtpA using PtpA-F4 primer, containing an XhoI site, and PtpA-R4 primer, containing an EcoRI site. The amplified gene was inserted into pEGFP-N1 vector (Clontech), generating pEGFP-N1-PtpA plasmid. The gene encoding subunit B (ATP6V1B1) of the human V-ATPase was amplified from THP-1 cDNA using B-F1, containing a XhoI \t\r \u00C2\u00A0 40 site, and B-R1, containing a HindIII site. The amplicon was inserted into pRSET- b his-tagged fusion protein expression vector, generating pRSET-b-ATP6V1B1 plasmid. The genes encoding subunit H (ATP6V1H) of the human V-ATPase was amplified from cDNA prepared from THP-1 using H-F1, containing a XhoI site, and H-R1, containing a HindIII site. The amplicon was inserted into pRSET-b, a his-tagged fusion protein expression vector (Invitrogen), generating pRSET-b- ATP6V1H plasmid. Human VPS33B his-tagged fusion protein expression plasmid (pBO1-VPS33B) was purchased from Genecopoeia Inc. (Rockville, MD). To construct VPS33B-DsRed2 mammalian transfection construct, VPS33B gene was amplified from pBO1-VPS33B using 33-F1 primer, containing a HindIII site and 33-R1 primer, containing a EcoRI site. The amplicon was inserted into pDsRed2-N1 (Clontech), generating the pDsRed2-N1-VPS33B plasmid. To construct Subunit H-GFP mammalian transfection construct, the gene encoding subunit H (ATP6V1H) was amplified from pRSET-b-ATP6V1H using H-F2 primer, containing a XhoI site, and H-R2 primer, containing an EcoRI site. The amplified gene was inserted into pEGFP-N1 vector, generating pEGFP-N1-H plasmid. Mtb ptkA (Rv2232) was amplified from its genomic DNA using the PtkA-F1 primer, containing an AflII site, and PtkA-R1 primer, containing a ClaI site. The amplified gene was inserted into pET151DTOPO to generate pET151DTOPO- PtkA His-tagged PtkA fusion protein expression construct. To clone ptkA into the pALACE vector, the ptkA gene was excised from pET151DTOPO-PtkA using the AflII and ClaI restriction sites, and the insert was ligated with the pALACE vector digested with the same enzymes, generating pALACE-PtkA. All plasmid \t\r \u00C2\u00A0 41 constructs were subsequently verified by sequencing at Eurofins MWG Operon (Ebersberg, Germany). The oligonucleotides used for DNA cloning in this study are described in Table 1. Table 1. Oligonucleotides used for DNA cloning in this study Oligonucleotides Sequence (5\u00E2\u0080\u0099 \u00C3\u00A8\u00EF\u0083\u00A8 3\u00E2\u0080\u0099)a PtpA-F1 ATATATGAATTCCGTGTCTGATCCGCTG PtpA-R1 ATATATCTCGAGTCAACTCGGTCCGTTC PtpA-F2 ATATATCTTAAGGTGTCTGATCCGCTG PtpA-R2 ATATATATCGATTCAACTCGGTCCGTTC PtpA-F3 ATATATGAATTCGTGTCTGATCCGCTG PtpA-F4 ATATCGCTCGAGATGTCTGATCCGCTG PtpA-R4 ATATATGAATTCGACTCGGTCCGTTCCGCGC B-F1 ATATATCTCGAGAATGGCCATGGAG B-R1 ATATATAAGCTTCTAGAGCGCAGTGTC H-F1 ATATATCTCGAGAATGACCAAAATGG H-R1 ATATATAAGCTTTTAGCTTCGGGCGGC 33-F1 ATAGTGAAGCTTATGGCTTTTCCCCATC 33-R1 ATATATGAATTCGGGCTTTCACCTCACTC H-F2 ATATCGCTCGAGATGACCAAAATGGATATC H-R2 ATATATGAATTCGGCTTCGGGCGGCAGCGGTC PtkA-F1 CACCCTTAAGGTGTCTTCGCCTCGTGAAC PtkA-R1 ATATATATCGATTCAGACCACCTAGCGCCT aThe recognition sequence for the restriction site is underlined. 2.4.3 Site-directed mutagenesis Site-directed mutagenesis (SDM) was carried out according to the QuikChange II-E site-directed mutagenesis kit (Stratagene, La Jolla, CA) protocol. Pfu DNA polymerase (Fermentas) was used in the SDM reactions. Oligonucleotides were designed with 15 bases flanking both sides of the mutated codons (Table 2). . The products obtained after PCR amplification were digested with DpnI restriction enzymes and chemically transformed into E. coli strain DH5\u00CE\u00B1. The positive clones were grown in LB media containing appropriate antibiotics, and the plasmids were extracted using Qiagen Mini Prep Plasmid Extraction Kit. Point \t\r \u00C2\u00A0 42 mutations were verified by sequencing at Eurofins MWG Operon (Ebersberg, Germany). Table 2. Oligonucleotides used for SDM in this study Oligonucleotides Sequence (5\u00E2\u0080\u0099 \u00C3\u00A8\u00EF\u0083\u00A8 3\u00E2\u0080\u0099)a PtkA-D85A-F CAGCTGGTGATCTTCGCTCTGGACGGCACGCTG PtkA-D85A-R CAGCGTGCCGTCCAGAGCGAAGATCACCAGCTG PtkA-K184M-F GCCGTCGCCACCTCCATGGCAGAGCCGACCGCA PtkA-K184M-R TGCGGTCGGCTCTGCCATGGAGGTGGCGACGGC PtkA-K217M-F GGCTCGCGAGGCAGCATGGTCGACGTGCTGGCC PtkA-K217M-R GGCCAGCACGTCGACCATGCTGCCTCGCGAGCC PtkA-K270M-F GGCATCTTTATCGACATGACCTCCACCACCGTC PtkA-K270M-R GACGGTGGTGGAGGTCATGTCGATAAAGATGCC PtkA-Y146A-F GAGGCGATCGTAGCCGCCCGGGCCGACTACAGC PtkA-Y146A-R GCTGTAGTCGGCCCGGGCGGCTACGATCGCCTC PtkA-Y150A-F GCCTACCGGGCCGACGCCAGCGCCCGCGGTTGG PtkA-Y150A-R CCAACCGCGGGCGCTGGCGTCGGCCCGGTAGGC PtkA-Y262A-F GTGGTCGGCTGGGGCGCCGGGCGCGCCGACTTT PtpA-Y262A-R AAAGTCGGCGCGCCCGGCGCCCCAGCCGACCAC PtpA-Y67A-F TTGCGAGCCCACGGCGCCCCTACCGACCACCGG PtpA-Y67A-R CCGGTGGTCGGTAGGGGCGCCGTGGGCTCGCAA PtpA-Y128A-F GATGTCGAGGATCCCGCCTATGGCGATCACTCC PtpA-Y128A-R GGAGTGATCGCCATAGGCGGGATCCTCGACATC PtpA-Y129A-F GTCGAGGATCCCTACGCCGGCGATCACTCCGAC PtpA-Y129A-R GTCGGAGTGATCGCCGGCGATGGGATCCTCGAC PtpA-Y128-129A-F GATGTCGAGGATCCCGCCGCCGGCGATCACTCCGAC PtpA-Y128-129A-R GTCGGAGTGATCGCCGGCGGCGGGATCCTCGACATC PtpA-L146A-F CATCGAATCCGCCGCGCCCGGCCTGCACGAC PtpA-L146A-R GTCGTGCAGGCCGGGCGCGGCGGATTCGATG aThe substituted nucleotide is highlighted in bold font. 2.5 Protein expression and purification 2.5.1 Expression of recombinant proteins in E. coli GST-tagged PtpA was expressed in E. coli strain BL21 transformed with pGEX- 4T3-PtpA plasmid. An overnight starter culture of the transformed BL21 was grown and inoculated into 1 L of LB media with 100 \u00C2\u00B5g/mL ampicillin. The culture was then grown to an OD600 of 0.6 at 37\u00C2\u00B0C and expression was induced with 0.4 mM IPTG at 19\u00C2\u00B0C overnight. The bacteria were harvested by centrifugation at \t\r \u00C2\u00A0 43 5,000 rpm, resuspended in PBS containing 1 mM DTT, 0.1 mM PMSF and lysed by sonication. The soluble fraction of the bacterial lysates was obtained by centrifugation at 13,000 rpm and GST-tagged proteins were purified from the soluble fraction by affinity chromatography on glutathione-agarose resin (Sigma- Alrich). To express and purify histidine-tagged recombinant protein, pRSET-b- ATP6V1H, pBO1-VPS33B, pRSET-b-ATP6V1B1 and pET151D-TOPO-PtkA plasmid were transformed into E. coli strain BL21. The transformed BL-21 was used to grow overnight starter cultures, which were then inoculated into 1 L of LB media supplemented with appropriate antibiotics. The cultures were then grown to an OD600 of 0.6 at 37\u00C2\u00B0C, and expression was induced with 0.4 mM IPTG at 19\u00C2\u00B0C or room temperature overnight. The bacteria were harvested by centrifugation at 5,000 rpm, resuspended in lysis buffer (50 mM Na2HPO4/NaH2PO4 pH = 7.4, 500 mM NaCl, 10 mM imidazole and 1 mM PMSF) and lysed by sonication. The soluble fraction of the bacterial lysates was obtained by centrifugation at 13,000 rpm and GST-tagged proteins were purified from the soluble fraction by affinity chromatography on Ni-NTA polyhistidine-tag purification resin (Qiagen). 2.5.2 Expression of recombinant proteins in M. smegmatis Competent M. smegmatis cells were transformed with pALACE-PtpA or pALACE-PtkA plasmid by electroporation and plated at 37\u00C2\u00B0C on 7H10 plates containing 1% dextrose and 50 \u00C2\u00B5g/mL hygromycin. Starter cultures of the transformed bacteria were grown over two nights in 7H9 media containing 0.05% \t\r \u00C2\u00A0 44 Tween-80, 50 \u00C2\u00B5g/mL hygromycin and supplemented with 1% dextrose. The starter cultures were inoculated into 1 L of 7H9 media containing 0.05% Tween- 80, 50 \u00C2\u00B5g/mL hygromycin and supplemented with 1% dextrose. The cultures were then grown at 37\u00C2\u00B0C to an OD600 of 1.0, and the cells were harvested by centrifugation at 5000 rpm for 30 minutes. The cell pellets were resuspended in 7H9 media supplemented with 0.05% Tween-80, 50 \u00C2\u00B5g/mL hygromycin. Acetamide was added to a final concentration of 0.2% to induce protein expression at room temperature for 24 hours. The bacteria were harvested by centrifugation at 5000 rpm for 30 minutes, resuspended in lysis buffer (50 mM Na2HPO4/NaH2PO4 pH = 7.4, 500 mM NaCl, 10 mM imidazole and 1 mM PMSF) and lysed by sonication. The soluble fraction of the bacterial lysates was obtained by centrifugation at 13,000 rpm and histidine-tagged proteins were purified by affinity chromatography on Ni-NTA polyhistidine-tag purification resin (Qiagen). 2.6 In vitro pull down assay In vitro pull down assays were carried out by incubation of THP-1 lysate with recombinant wild-type his-tagged PtpA proteins at 4oC overnight. The mixture was purified using Ni-NTA resin (Qiagen) and resolved in SDS-PAGE. Captured proteins were identified with MALDI-TOF mass spectrometry and analyzed with western blotting to confirm identity. 2.7 ALPHAScreen protein-protein interaction assay ALPHAScreen assays performed using the Histidine (Nickel Chelate) Detection Kit (Perkin Elmer) according to the manufacturer\u00E2\u0080\u0099s protocol. Specifically, purified \t\r \u00C2\u00A0 45 GST-tagged recombinant proteins were first biotinylated using the EZ-link Biotinylation Kit (Pierce). The biotinylated recombinant proteins were serial diluted in white opaque 384-well microplates (PerkinElmer), and purified his- tagged recombinant proteins were added to same wells. Nickel-chelating acceptor beads were added to the proteins and incubated for 30 minutes at 25oC. Streptavidin donor beads were added to the wells and the reactions were incubated for 30 minutes at 25oC. The reaction kinetics were then monitored on the Fusion-\u00CE\u00B1-HT Multimode Microplate Reader (PerkinElmer) for luminescence signal generated from protein-protein interaction (counts per second (cps)). 2.8 Mycobacterial \u00E2\u0080\u009CSplit-Trp\u00E2\u0080\u009D protein-protein interaction assay In vivo interaction was detected using a modified \u00E2\u0080\u009CSplit-Trp\u00E2\u0080\u009D assay first developed by O\u00E2\u0080\u0099Hare et al. [137, 138]. Interaction between two proteins of interest, which are individually fused to the split protein fragments (Ntrp and Ctrp), would reconstitute the activity of the N-(5\u00E2\u0080\u0099-phosphoribosyl)-anthranilate isomerase enzyme in tryptophan synthesis; thereby, enabling the auxtrophic \u00CE\u0094hisA M. smegmatis to grow on tryptophan deficient media. Briefly, the DNA encoding PtpA and subunit H were cloned into pJC10 and pJC11 vectors, generating translational fusion constructs with the Ntrp and Ctrp fragments of the N-(5\u00E2\u0080\u0099- phosphoribosyl)-anthranilate isomerase under the control of the acetamidase promoter respectively [138]. The resulting plasmids were transformed into the \u00CE\u0094hisA M. smegmatis strain. Positive transformants were grown in LB broth supplemented with 0.05% Tween-80, 50 \u00C2\u00B5g/mL hygromycin and 30 \u00C2\u00B5g/mL \t\r \u00C2\u00A0 46 apramycin to OD600 of 1.0 and spotted onto 7H9 agar medium with 1% glucose, 60 \u00C2\u00B5g/mL histidine and 0.02% acetamide to induce protein expression. 2.9 Infection 2.9.1 Infection of THP-1 Infection of THP-1 macrophage-like cells was performed using human serum- opsonized Mtb or E. coli DH5\u00CE\u00B1 at multiplicity of infection (MOI) of 10:1 or 1:1. For Mtb intracellular survival studies, infected macrophages were harvested at defined time points, lysed with 0.025% SDS, serially diluted, and plated on 7H10 agar medium supplemented with OADC and appropriate antibiotics. The plates were incubated at 37oC for 2 weeks to allow for growth of visible colonies. 2.9.2 Phagocytosis assay THP-1 macrophage-like cells were transfected and seeded as mentioned above (Section 2.3.2 and 2.3.3). E. coli was labeled with 50 ug/mL Alexa Fluor 350 carboxylic acid SE (Invitrogen) at 37oC for 1 hour. The bacteria were then washed with PBS and opsonized with human serum. The transfected cells were infected with labeled E. coli at a MOI of 10:1. Phagocytosis was synchronized as described previously [139]. Non-internalized bacteria were washed away after incubation at 37oC for 15 minutes, and phagocytosis was measured by spectrofluorometry. 2.10 Immunoprecipitation THP-1 cells were infected and lysed as above (Section 2.9.1). Lysates were centrifuged and soluble fractions were used in co-immunoprecipitation assays. \t\r \u00C2\u00A0 47 Briefly, rabbit anti-VPS33B or rabbit anti-subunit H IgG were added to 2 mg or 4 mg of soluble lysate (final IgG dilution, 1:100) and the mixture was incubated at room temperature for one hour. Affi-Gel Protein-A Agarose Resin (Bio-Rad) was blocked with 0.5% BSA, washed with PBS, then added to the mixture and incubated on a shaker at room temperature for one hour. The resin was then washed with PBS, and SDS sample buffer was added to release IgG and bound material. The resulting samples were resolved on SDS-PAGE and transferred onto a nitrocellulose membrane for Western blot analysis with the indicated primary antibodies. Protein A-HRP was used as the secondary detection reagent. For Western blot analysis with anti-phosphotyrosine 4G10, recombinant VPS33B protein was phosphorylated in kinase buffer as previously described [32] with 0.5 \u00C2\u00B5M ATP and assayed as a control along with the immunoprecipitated samples. 2.11 Measurement of phagosomal pH 2.11.1 Overview Assessment of the phagosomal pH of E. coli and Mtb phagosomes was performed according to a previously published procedure with modifications [140, 141]. Before infection, bacteria were labeled with the pH-sensitive pHrodo probe (Invitrogen), which emits red fluorescence in acidic environment, and pH- insensitive probes Alexa Fluor 350 carboxylic acid SE (Blue) or Alexa Fluor 488 carboxylic acid SE (Green) (Invitrogen). E. coli Phagosomal pH was measured with spectrofluorometry in the Fusion-Alpha HT microplate reader (Perkin Elmer) and Mtb phagosomal pH was measured with a FACSCalibur flow cytometer (BD Bioscience). The fluorescence ratio values of the pH-sensitive and pH-insensitive \t\r \u00C2\u00A0 48 probes were used to determine phagosomal pH according to a calibration curve. pH calibration was performed by incubating Mtb- or E. coli-infected THP-1 macrophage-like cells in 10 mM phosphate-citrate buffer of pre-determined pH (5.0 - 8.0) containing 145 mM KCl, 1 mM MgCl2, 0.5 mM CaCl2 and 10 mM glucose, together with 20 \u00C2\u00B5M nigericin and 4 \u00C2\u00B5M monensin, which were added as ionophores in order to equilibrate intracellular and intraphagosomal pH with extracelluar pH. As a positive control, concanamycin (50 nM) was used to inhibit macrophage V-ATPase proton transport. 2.11.2 Phagosomal pH of Mtb phagosomes Mtb strains were first labeled with 20 \u00C2\u00B5M pHrodo SE (Invitrogen) at 37oC for one hour. The bacteria were washed with 7H9 supplemented with 0.05% Tween-80 three times before being labeled with 25 \u00C2\u00B5g/mL Alexa Fluor 488 carboxylic acid SE (Invitrogen). The bacteria were washed and opsonized with human serum then used to infect THP-1-derived macrophages at a MOI of 10:1 for 2 hours at 37oC. Non-internalized bacteria were washed away, and cells were reincubated at 37oC for another 2 hours. Cells were then washed, scraped off the plate and fixed with 2.5% paraformaldehyde. pHrodo is stable after fixation with paraformaldehyde [142]. Fixed cells were then analyzed by flow cytometry, and the data was processed with the FlowJo 8.7 software. Color compensation was performed to prevent signal overlapping. Mean fluorescence intensities of pHrodo and Alexa Fluor 488 were used to calculate pHagosomal pH. \t\r \u00C2\u00A0 49 2.11.3 Phagosomal pH of E. coli phagosomes Transfection of THP-1, E. coli labeling with pHrodo SE and Alexa Fluor 350 carboxylic acid SE and THP-1 infection were performed as described before. For pH calibration, the infected cells were incubated in 10 mM phosphate-citrate buffer with pre-determined pH (5.0 - 8.0) after non-internalized bacteria were washed away. pHrodo fluorescence (Emission 620 nm) and Alexa Fluor 350 fluorescence (Emission 460 nm) were measured simultaneously in the Fusion-\u00CE\u00B1 HT microplate reader (Perkin Elmer). 2.12 Digital confocal microscopy 2.12.1 Intracellular staining Immunofluorescence staining was performed on cells adhered to glass coverslips. Cell fixation was performed using 2.5% paraformaldehyde in PBS for 20 minutes at 37\u00C2\u00B0C. Subsequently, the cells were washed with PBS and incubated in PBS containing 0.2% saponin and 1% normal goat serum (permeabilization buffer) for 15 minutes at room temperature. Specific primary antibody (10 \u00C2\u00B5g/mL) in permeabilization buffer was added to the coverslips and incubated for 45 minutes at room temperature. Then, the cells were washed three times with PBS and incubated with suitable secondary FITC-, Alexa-Fluor 488- or Texas Red- conjugated antibody for 45 minutes. Cells were then washed extensively with PBS. Double immunostaining was performed sequentially with rabbit anti- VPS33B at 10 \u00C2\u00B5g/mL and mouse anti-subunit H at 10 \u00C2\u00B5g/mL in permeabilization buffer. Texas Red-conjugated goat anti-rabbit IgG (Invitrogen) and Alexa-Fluor \t\r \u00C2\u00A0 50 488-conjugated goat anti-mouse IgG (Invitrogen), both of which are highly cross- adsorbed to minimize cross-reactivity, were used at a dilution of 1:1000. 2.12.2 Microscopy Fixed adherent cells on cover slips were mounted on microscope slides in FluorSaveTM (Calbiochem-Novabiochem, La Jolla, CA) to prevent photobleaching. Digital confocal microscopy using Axioplan II epifluorescence microscope (Carl Zeiss Inc., Thornwood, NY) equipped with 63x/1.4 Plan-Apochromatobjective (Carl Zeiss Inc) was used to analyze the slides. Images were captured using Retiga EX CCD digital camera (QImaging, BC, Canada) operated by Northern Eclipse software (Empix Imaging Inc. Ontario, Canada). 2.13 In vitro phosphatase activity assay para-nitrophenyl phosphate (pNPP) was used as an artificial chromogenic substrate to assay for PtpA phosphatase activity in vitro. The phosphatase reaction contained 50 mM Tris-HCl pH 7.4, 5 mM MgCl2, 2 mM MnCl2, 1 mM DTT and 50 mM pNPP together with 20 \u00C2\u00B5g of purified recombinant PtpA proteins. The reaction was incubated at 37oC, and the absorbance A450 was measured at 10 minutes interval on Bio-rad Model 680 microplate reader. 2.14 In vitro PtpA secretion analysis The DNA encoding His-tagged PtpA under the control of the hsp60 promoter was cloned into the pMV261 vector. The resulting plasmid was introduced into the wild type H37Rv Mtb strain, and the transformed strain was grown to an OD600 = 0.8 in 50 mL of protein-free Sauton\u00E2\u0080\u0099s medium (containing (per L) 4 g of L- \t\r \u00C2\u00A0 51 asparagine, 0.5 g K2HPO4, 0.5 g MgSO4, 50 mg ferric ammonium citrate, 2 g citric acid, 1 mg ZnSO4 and 60 mL glycerol) supplemented with 0.05% Tween-80. To test whether covalent labeling with fluorescent dyes affect secretion of proteins by Mtb, pHrodo and Alexa-Fluor 488 labeling was performed according to the manufacturer\u00E2\u0080\u0099s protocol. The unlabeled and labeled bacteria were then re- inoculated into 50 mL of fresh Sauton\u00E2\u0080\u0099s medium supplemented with 0.05% Tween-80 and grown for 48 hours at 37oC. The culture filtrate proteins were then harvested and precipitated with 10% TCA and 0.1% SDS as described before [143]. The samples were resolved on 10% SDS-PAGE and PtpA secretion into the culture supernatant was analyzed with Western blotting using rabbit anti-PtpA and mouse anti-His-tag antibodies. 2.15 FACS analysis of antibody specificity H37Rv Mtb harboring a plasmid construct for the expression of DsRed (pSMT- DsRed) were grown to O.D.600 = 0.8 in 7H9 medium supplemented with 10% OADC and 0.05% Tween-80. The bacteria were washed three times with 7H9 supplemented with 0.05% Tween-80 (7H9T), and 1.0 x 108 was resuspended in 100 \u00C2\u00B5L 7H9T. The indicated primary antibodies were added to the bacteria at a final concentration of 5 \u00C2\u00B5g/mL and incubated at 37oC with shaking for one hour. The bacteria were then washed three times with 7H9T and incubated with either Alexa-Fluor conjugated goat anti-rabbit IgG (Invitrogen) or Alexa-Fluor 488 goat anti-mouse IgG (Invitrogen) secondary antibodies at a dilution of 1:1000 for one hour at 37oC with shaking. The bacteria were then washed with 7H9T and fixed with 2.5% paraformaldehyde for 30 minutes at 37oC. The fixed bacteria were \t\r \u00C2\u00A0 52 subjected to analysis with FACSCalibur for antibody binding. Expression of DsRed in Mtb allowed specific gating to distinguish the bacterial population from background noise signal. Color compensation was performed to prevent signal overlapping. 2.16 In vitro kinase assay The in vitro kinase reactions contained 25 mM Tris\u00E2\u0080\u0093HCl, pH 7.2, 5 mM MgCl2, 2 mM MnCl2, 1 mM DTT, 10\u00E2\u0080\u0093300 ng of recombinant PtkA kinase, and 0\u00E2\u0080\u0093500nM of substrates. The reactions were started by addition of 10 \u00C2\u00B5Ci of \u00CE\u00B3-[32P]ATP (Perkin-Elmer) or 10 \u00C2\u00B5Ci of \u00CE\u00B3-[32P]GTP followed by 5\u00E2\u0080\u009330 minutes incubation at room temperature (25\u00C2\u00B0C). At the end of the incubation period, reactions were stopped by the addition of SDS-sample loading buffer and heated at 95 \u00C2\u00B0C for 5 minutes. Samples were resolved by SDS\u00E2\u0080\u0093PAGE on 8% polyacrylamide gels and the gels were silver-stained, and dried. The 32P-radioactively labeled protein bands were detected using a PhosphorImager SI (Molecular Dynamics). Bands corresponding to the phosphorylated proteins were cut out and analyzed by scintillation counting (Beckman Coulter LS 6500). Kinase inhibitors were added to the kinase reaction buffer at the following concentrations: staurosporine (100 nM), tartaric acid (10 mM), wortmannin (1 nM), methyl piperazine (100 nM), and phenyl piperazine (100 nM). 2.17 Phospho-amino acid analysis Recombinant PtkA was incubated with \u00CE\u00B3-[32P]-ATP in the kinase buffer described above (Section 2.16). The sample was hydrolyzed with 6 N HCl at 110oC for 1 hour and separated on a cellulose TLC plate by ascending chromatography [144]. \t\r \u00C2\u00A0 53 For PtpA phosphorylation by PtkA, the reaction was stopped with sample buffer, loaded into SDS-PAGE 12% containing 8 M urea, and subjected to electrophoresis. The gel was electroblotted onto an Immobilon PVDF membrane [145]. Phosphorylated proteins bound to the membrane were detected by autoradiography using a PhosphorImager apparatus. The 32P-labelled protein band corresponding to the migration of PtpA was excised from the membrane and hydrolyzed following the same procedure as described above. After hydrolysis, samples were separated on ascending TLC chromatography [144]. After migration, radioactive amino acids were detected by autoradiography. Phosphoserine, phosphothreonine and phosphotyrosine standards were separated on cellulose TLC plate in parallel and visualized by staining with ninhydrin. \t\r \u00C2\u00A0 54 CHAPTER 3: RESULTS 3.1 Mtb PtpA excludes host Vacuolar-H+-ATPase to inhibit phagosome acidification 3.1.1 Introduction Mtb pathogenicity partly depends on its ability to inhibit phagosome acidification and maturation processes after engulfment by macrophages. We showed that secreted PtpA, a key protein tyrosine phosphatase needed for Mtb pathogenicity within host macrophages, binds to subunit H of the host V-ATPase machinery, a multi-subunit protein complex in the phagosome membrane that drives luminal acidification. Furthermore, we showed for the first time that macrophage Class C VPS complex, a key regulator of endosomal membrane fusion, associates with V-ATPase during phagosome maturation, indicating a novel role for V-ATPase in coordinating membrane trafficking and phagosome-lysosome fusion. PtpA interaction with V-ATPase is required for the previously reported dephosphorylation of VPS33B and subsequent exclusion of V-ATPase from the phagosome during Mtb infection. These findings showed that inhibition of phagosome acidification in the mycobacterial phagosome is directly attributed to PtpA, revealing the mechanism behind the long-established paradigm of acidification block by Mtb. 3.1.2 Mtb PtpA binds subunit H of human V-ATPase Earlier, we used a substrate trapping assay with a mutant of PtpA to pull down the catalytic substrate of PtpA, VPS33B, from THP-1 cell lysate [32]. Interestingly, when we used wild type recombinant PtpA as bait, we were able to pull down \t\r \u00C2\u00A0 55 another 50 kDa macrophage protein [32]. We identified this protein by MALDI- TOF mass spectrometry to be subunit H of human V-ATPase (Figure 28A, Appendix A) and verified its identity by Western analysis (Figure 10A). In vitro kinase assay and phosphoamino acid analysis showed that subunit H is phosphorylated on threonine and therefore could not serve as a catalytic substrate for PtpA, which is a tyrosine phosphatase (Figure 28B and 28C, Appendix A). In vitro protein-protein interaction analysis of PtpA and subunit H demonstrated direct binding of mycobacterial PtpA to host V-ATPase subunit H (Figure 10B and 10C). A hyperbolic curve fitting the 1:1 Langmuir Binding Model and a dissociation constant (Kd) of 1.1 x 10-7 M indicate a relatively high binding affinity. 3.1.3 PtpA binds to amino acid 220-402 of subunit H We determined the PtpA binding site on subunit H by constructing overlapping polypeptide fragments of subunit H. As shown in Figure 10D, only fragments that included the linker region (amino acid 220-402) that connects the N-terminal and C-terminal domains of subunit H were able to bind to PtpA with high affinity. This is consistent with previous predictions that the linker region of yeast subunit H is involved in protein-protein interactions [146]. 3.1.4 The C-terminal alpha helix of PtpA binds subunit H Subunit H was reported to interact with di-leucine based motifs on the Human Immunodeficiency Virus (HIV) Nef protein [78]. Therefore, we performed site- directed mutagenesis to investigate whether di-leucine motifs or other leucine \t\r \u00C2\u00A0 56 residues on PtpA participate in binding subunit H (Figure 29A, Appendix A). One mutation in the C-terminal alpha helix (PtpAL146A) abolished PtpA binding to subunit H (Figure 10E). These findings, together with the results presented above, are consistent with the predicted interaction of PtpA bound with subunit H using the protein-docking algorithm, 3D-Garden (Figure 11A and 11B) [147]. Interestingly, the PtpAL146A mutant retained its phosphatase activity with kinetics similar to the wild type protein (Figure 29B, Appendix A). This result is in line with our observation that subunit H is not a catalytic substrate but rather a binding partner of PtpA. Figure 10. PtpA interacts with V-ATPase subunit H in vitro. (A) In vitro pull- down using recombinant His-tagged Mtb PtpA captured human V-ATPase \t\r \u00C2\u00A0 57 subunit H from THP-1 macrophage lysate. The eluates were analyzed on Western blots using rabbit anti-subunit H and anti-PtpA antibodies. (B) Recombinant PtpA was subjected to ALPHAScreen Assay with increasing concentration of V-ATPase subunit H. GST served as negative control. Curve fitting yielded a dissociation constant (Kd) of 1.1 x 10-7 M. (C) The reciprocal experiment with increasing PtpA concentration yielded similar results. His-tagged Rv0323c protein was used as negative control. (D) Mapping of the PtpA binding site on V-ATPase subunit H. Overlapping truncated recombinant subunit H protein fragments were incubated with recombinant PtpA proteins in ALPHAScreen assay. Amino acid 220-402 of subunit H is the minimal region required for binding PtpA. (E) Site-directed mutagenesis generated a recombinant mutant PtpA protein (PtpAL146A) that failed to interact with subunit H. \t\r \u00C2\u00A0\t\r \u00C2\u00A0 Figure 11. Computer-generated model of PtpA complexed with the V- ATPase subunit H. Based upon 3D-Garden macromolecular docking analysis [147], the crystal structure of Mtb PtpA (PDB accession number 1U2P) and S. cerevisiae V-ATPase subunit H (PDB accession number 1HO8) were oriented into the predicted binding position. (A) Surface representation of PtpA and subunit H as separated molecules. The C-terminal alpha helix of PtpA is colored in magenta. L146A mutation on PtpA likely alters the C-terminal alpha helix, leading to the defective binding. (B) Computer prediction of PtpA binding on subunit H with the C-terminal alpha helix of PtpA docking on to the cleft region between the N-terminal and C-terminal domain of subunit H. \t\r \u00C2\u00A0 \t\r \u00C2\u00A0 58 3.1.5 PtpA binds V-ATPase subunit H in vivo We applied immunoprecipitation with specific antibodies to investigate the in vivo interaction between PtpA and subunit H during macrophage infection. Antibodies against subunit H were able to pull down PtpA from lysates of THP-1 cells infected with H37Rv Mtb expressing wild type PtpA, confirming the physiological relevance of PtpA-subunit H interaction during Mtb infection (Figure 12A). To further monitor PtpA interaction with subunit H, we used a modified \u00E2\u0080\u009CSplit Trp\u00E2\u0080\u009D mycobacterial assay for detecting protein-protein interactions [137, 138]. In this system, a tryptophan auxotrophic strain of M. smegmatis grows in the absence of tryptophan only if the tested proteins interact with each other. As shown in Figure 12B, co-expression of subunit H and PtpA fusion proteins restored mycobacterial growth on 7H9 agar plates, confirming protein-protein interaction within a surrogate host. Co-expression of subunit H with the binding- defective PtpAL146A did not lead to mycobacterial growth, indicating that this phosphatase-active PtpA mutant lost its ability to bind subunit H. The phosphatase-defective PtpAD126A retained its ability to bind to subunit H, further indicating the interaction is independent of the catalytic activity of PtpA (Figure 12B). \t\r \u00C2\u00A0 59 Figure 12. PtpA and V-ATPase subunit H interacts in vivo. (A) PtpA co- immunoprecipitated with subunit H from THP-1 infected with Mtb expressing PtpA. Immunoprecipitation (IP) was performed with either rabbit anti-subunit H (\u00CE\u00B1-H) antibodies (Ab) or rabbit IgG as negative control. (B) PtpA and subunit H interacted in the \u00E2\u0080\u009CSplit-Trp\u00E2\u0080\u009D protein fragment complementation assay to facilitate Ntrp and Ctrp reassembly required for Trp biosynthesis, thus enabling growth of the M. smegmatis Trp- strain coexpressing Ntrp-PtpA and subunit H-Ctrp under acetamide (ACE) induction (Middle panel). Ntrp-PtpAL146A failed to interact with subunit H-Ctrp to restore M. smegmatis growth. Ntrp-ESAT6 and CFP10-Ctrp were used as a positive control. The negative control consisted of Ntrp and Ctrp fragments alone. Top panel shows that all transformed strains are capable of growing on Trp-supplemented media. No growth could be observed in the absence of acetamide induction and exogenous Trp (Bottom panel) (C) Loss of PtpA binding to V-ATPase subunit H impairs Mtb survival within host macrophages. V-ATPase binding-defective PtpAL146A expressing Mtb strain showed a 2-log reduction in bacterial load compared to the wild type strain at 6 days post-infection. 3.1.6 PtpA binding to subunit H is required for Mtb intracellular survival To test whether PtpA binding to subunit H is required for Mtb pathogenicity, we investigated the ability of the binding-defective PtpAL146A Mtb strain to survive in \t\r \u00C2\u00A0 60 THP-1 macrophage-like cells. As shown in Figure 12C, a \u00CE\u0094ptpA strain complemented with a construct encoding PtpAL146A was attenuated within the macrophage in a manner similar to that of the \u00CE\u0094ptpA knockout strain, while the parental and the complement strains were able to establish a stable infection after 3 days. Expression and stability of the wild type and mutant PtpA proteins in these strains were confirmed with Western blot analysis (Figure 30A, Appendix A). At 6 days post-infection, the binding defective strain showed a 2-log reduction in CFU compared to the wild type strain. The PtpAD126A strain also exhibited a similar intracellular survival phenotype as the \u00CE\u0094ptpA strain. While the parental and complemented strains, entered into growth phase within the macrophage, the \u00CE\u0094ptpA and the binding-defective strains were continually cleared, establishing the importance of PtpA interaction with the V-ATPase machinery for Mtb survival within macrophages. 3.1.7 PtpA inhibits phagosome acidification The in vivo PtpA-subunit H interaction and the impaired intracellular survival of the binding-defective PtpAL146A strain indicated that PtpA interferes with the phagosome acidification process. To examine this hypothesis, we analyzed the pH of Mtb-containing phagosomes by FACS. Parental and mutant strains were dually labeled with the pH-sensitive pHrodo fluorescent dye and the pH- insensitive Alexa Fluor 488 prior to infection of THP-1. Covalent labeling of the bacteria with fluorescent dyes does not affect PtpA secretion (Figure 30B, Appendix A). Overlaid FACS histograms showed a clear increase in the mean fluorescence intensity of pHrodo in phagosomes harboring \u00CE\u0094ptpA, \u00CE\u0094ptpA \t\r \u00C2\u00A0 61 complemented with the phosphatase-inactive ptpAD126A, and \u00CE\u0094ptpA complemented with the binding-defective ptpAL146A (corresponding to pH 5.8, 5.6, and 5.55 respectively), as compared to phagosomes containing parental Mtb and \u00CE\u0094ptpA complemented with wild type ptpA (corresponding to pH 6.7 and 6.4, respectively) (Figure 13A and 13B, Figure 31, Appendix A and Table 3). These results indicated a direct functional role for PtpA in the inhibition of Mtb phagosome acidification, which is dependent on both the binding ability to subunit H and phosphatase activity of PtpA. Figure 13. Mtb PtpA inhibits phagosome acidification. (A) The indicated Mtb strains were labeled with the pH-sensitive fluorescent dye (pHrodo) and used to infect THP-1 macrophage-like cells. Phagosomal pH of the infected macrophages was measured with FACS. Wild type Mtb maintained a \t\r \u00C2\u00A0 62 phagosomal pH of around 6.7 while phagosomes of both \u00CE\u0094ptpA and PtpAL146A strains were acidified to around pH 5.5. * = p < 0.05; ** = p < 0.01, significant difference compared with H37Rv by Student\u00E2\u0080\u0099s t test. (B) Overlaid FACS histograms of the pHrodo fluorescence intensity of the wild type H37Rv and \u00CE\u0094ptpA phagosomes show a shift in phagosomal pH in the presence of PtpA. (C) THP-1 macrophage-like cells were transfected with GFP-tagged constructs of wild type and mutant PtpA proteins and infected with pHrodo-labeled E. coli. Phagosomal pH was measured with spectrofluorometry during the course of infection. Wild type PtpA inhibits E. coli phagosome acidification (pH 6.5) while the binding-defective PtpAL146A failed to block acidification (pH 5.6). Concanamycin (CMA) was added as a positive control for inhibition of phagosome acidification. Untransfected THP-1 and THP-1 transfected with pEGFP vector served as negative control. (D) Digital confocal microscopy of the infected macrophages in (C) confirms the inhibition of phagosome acidification by Mtb PtpA. Green: Expression of GFP or GFP-tagged PtpA constructs. Red: pHrodo fluorescence in acidified phagosomes. Blue: Alexa Fluor 350 labeled E. coli. Table 3. Phagosomal pH of Mtb strains in THP-1 \t\r \u00C2\u00A0 AF488: Alexa Fluor 488; MFI: Mean Fluorescence Intensity; N/A: Not Applicable; NS: Not Significant. To examine PtpA\u00E2\u0080\u0099s ability to modulate phagosome acidification independently of other Mtb proteins, we transfected THP-1 macrophage-like cells with plasmids expressing GFP-fused parental or mutated forms of PtpA. All constructs were expressed at similar levels, and the survival and phagocytic ability of the macrophages were not impaired by the transfection (Figure 32A and pHrodo MFI AF 488 MFI Fluorescence Ratio (pHrodo/AF488) pH Standard Error p value WT 42.23 15.40 2.74 6.73 0.38 NA \u00CE\u0094ptpA 47.53 14.17 3.36 5.78 0.04 0.012 Complement 44.20 15.15 2.92 6.42 0.01 NS D126A 50.87 14.47 3.52 5.58 0.09 0.0067 L146A 49.17 13.90 3.54 5.55 0.09 0.0063 \t\r \u00C2\u00A0 63 32B, Appendix A). Transfected macrophages were infected with E. coli labeled with pHrodo and Alexa Fluor 350, and the phagosomal pH was assessed by spectrofluorometry (Figure 32C, Appendix A). Expression of wild type PtpA within THP-1 macrophage-like cells inhibited acidification of E. coli containing phagosomes, whereas both catalytically inactive PtpAD126A and binding mutant PtpAL146A were unable to block phagosome acidification, showing a marked decrease in phagosomal pH similar to untransfected and empty vector control macrophages (Figure 13C). Digital confocal microscopy of the infected macrophages confirmed these findings (Figure 13D and Figure 32D, Appendix A). These results established that PtpA alone is able to prevent effectively phagosome acidification in the macrophage. This inhibition depends on both PtpA phosphatase activity and its ability to bind to host V-ATPase machinery. 3.1.8 The V-ATPase machinery recruits the Class C VPS complex The interaction of PtpA with both subunit H and the previously described VPS33B [32] indicated that V-ATPase and Class C VPS complexes might be in close proximity or even interacting directly during phagosome maturation and endosome-lysosome fusion. To investigate this hypothesis, we conducted immunoprecipitation experiments in THP-1 macrophage-like cells using antibodies against VPS33B as bait and probed for subunits H, B and E of the V- ATPase complex and VPS18 of the Class C VPS complex by Western blot. In uninfected THP-1, faint bands corresponding to subunits of the V-ATPase complex were captured with VPS33B and VPS18 (Figure 14). This indicates that V-ATPase interacts transiently with the Class C VPS to regulate normal \t\r \u00C2\u00A0 64 endosome-lysosome fusion in resting cells. Furthermore, this interaction is up- regulated upon phagocytosis of E. coli as demonstrated by the prominent bands of V-ATPase subunits captured 2 hours post-infection (Figure 14), further supporting that the Class C VPS complex interaction with V-ATPase plays a key role in phagosome maturation and fusion with lysosomes. Figure 14. Co-immunoprecipitation of V-ATPase subunits with Class C VPS in THP-1. V-ATPase subunits (H, B and E) can be captured with VPS33B and VPS18 from uninfected THP-1 macrophage-like cells when immunoprecipitation was performed using 4 mg of soluble lysate. The interaction was not detected when 2 mg of lysate were used. Prominent bands corresponding to V-ATPase subunits can be pulled down from 2 mg of lysate from E. coli-infected THP-1 macrophage-like cells. 3.1.9 Mtb PtpA blocks the interaction between Class C VPS and V-ATPase complexes When THP-1 macrophage-like cells were infected with Mtb H37Rv, V-ATPase subunit H, B or E were not captured using antibodies against VPS33B (Figure 15A). Mtb inhibition of V-ATPase and Class C VPS interaction was maintained 24 hours post-infection (Figure 15B). However, the V-ATPase subunits co- immunoprecipitated with VPS33B and VPS18 in THP-1 macrophage-like cells \t\r \u00C2\u00A0 65 infected with \u00CE\u0094ptpA Mtb. Complementation of \u00CE\u0094ptpA mutant with wild type ptpA restored the ability of Mtb to inhibit Class C VPS interaction with the V-ATPase complex. Furthermore, complementation with ptpAD126A but not ptpAL146A restored Mtb inhibition of Class C VPS and V-ATPase interaction (Figure 15A). These results indicated that Mtb infection disrupted the V-ATPase and Class C VPS association, that PtpA is responsible for blocking this interaction, and that PtpA binding to subunit H is specifically necessary for this disruption. Figure 15 (next page). Mtb PtpA disrupts the interaction between the Class C VPS and V-ATPase complexes during infection. Immunoprecipitation (IP) was performed with anti-VPS33B (\u00CE\u00B1-VPS33B) using 2 mg of soluble lysate from THP-1 macrophage-like cells infected with the indicated strain for 2 hours and followed by Western analysis with the indicated antibody. Rabbit IgG was used as negative control. (A) The Class C VPS (VPS33B and VPS18) interacts with the V-ATPase complexes (Subunit B, H, and E) from THP-1 infected with E. coli, \u00CE\u0094ptpA and PtpAL146A. The parental H37Rv, PtpAD126A and the complemented strain disrupted the host proteins interaction. (B) Mtb disruption of V-ATPase interaction with Class C VPS can be observed 24 hours post-infection. Immunoprecipitation (IP) was performed with either rabbit IgG (IgG) or anti- VPS33B (\u00CE\u00B1-VPS33B) using 2 mg of soluble lysate from THP-1 infected for 24 hours. For the \u00CE\u0094ptpA and \u00CE\u0094ptpA complemented with PtpAL146A strains, V-ATPase interaction with Class C VPS remains upregulated 24 hours post-infection, suggesting continued delivery of lysosomal contents and killing of the bacteria in the phagosomes. \t\r \u00C2\u00A0 66 3.1.10 VPS33B remains phosphorylated in THP-1 infected with binding- defective PtpA VPS33B phosphorylation is required for phagosome-lysosome fusion and its dephosphorylation by Mtb PtpA blocks this process [32]. Since the Class C VPS \t\r \u00C2\u00A0 67 complex is recruited to the V-ATPase, we asked whether PtpA binding to V- ATPase subunit H affects the phosphorylation status of VPS33B during Mtb infection. To address this question, we immunoprecipitated VPS33B and analyzed its phosphorylation with the anti-phosphotyrosine antibody 4G10. When E. coli was used to infect macrophages, VPS33B was phosphorylated (Figure 16). However, as expected, infection with Mtb inhibited VPS33B phosphorylation. Consistent with these findings, \u00CE\u0094ptpA strain failed to reduce VPS33B phosphorylation, and this was reversed by complementation with wild type ptpA (COMP). Interestingly, VPS33B remained phosphorylated in macrophages infected with \u00CE\u0094ptpA complemented with either the phosphatase-defective ptpAD126A, or the binding-defective ptpAL146A, despite being catalytically active (Figure 16). This indicates that the binding of PtpA to subunit H of V-ATPase is a prerequisite for VPS33B dephosphorylation. This is consistent with our findings that the loss of PtpA interaction with subunit H abolished its functions within the host cell. \t\r \u00C2\u00A0 Figure 16. Western blot analysis of VPS33B phosphorylation in vivo. VPS33B remained phosphorylated in macrophages infected with the PtpAL146A expressing strain. Top panel was probed with anti-phosphotyrosine antibody 4G10, and bottom panel is probed with anti-VPS33B to ensure equal loading. \t\r \u00C2\u00A0 68 3.1.11 PtpA binding to subunit H participates in the exclusion of the V- ATPase from the Mtb phagosome Disruption of the interaction between Class C VPS and V-ATPase by PtpA suggests that PtpA directly impairs V-ATPase trafficking to the Mtb phagosome. To verify this hypothesis, we used digital confocal microscopy to monitor Class C VPS and V-ATPase localization within infected macrophages, represented by VPS33B and subunit H, respectively. In control experiments, where macrophages were infected with E. coli, both VPS33B and subunit H co-localized to the phagosome (Figure 17A). In contrast, phagosomes containing Mtb were decorated with VPS33B but excluded subunit H, indicating a lack of V-ATPase recruitment. To directly assign a role to PtpA, we examined cells that were infected with the \u00CE\u0094ptpA strain and found co-localization of VPS33B and subunit H on the mycobacterial phagosomal membrane. These findings were confirmed with the observation that macrophages infected with \u00CE\u0094ptpA complemented with wild type ptpA led to a phenotype similar to that observed with wild type strain (Figure 17B). Additionally, co-recruitment of subunit H and VPS33B to bacilli- containing phagosomes was observed for \u00CE\u0094ptpA strain complemented with either the catalytically inactive ptpAD126A or the binding-defective ptpAL146A. Therefore, PtpA binding to V-ATPase subunit H is required for blocking V-ATPase trafficking to the phagosome, which ultimately leads to the inhibition of Mtb phagosome acidification. Although we have shown that VPS33B and subunit H antibodies are highly specific (Figure 33, Figure 34 and Figure 35, Appendix A), to overcome \t\r \u00C2\u00A0 69 any potential cross-reactivity, we also performed double transfection of THP-1 with GFP and DsRed2 fused constructs of subunit H and VPS33B respectively, allowing for direct visualization of these proteins and their localization within the macrophage. Co-localization of GFP-subunit H and DsRed2-VPS33B in uninfected macrophages provides further support that V-ATPase and Class C VPS functions in the same pathway to regulate normal endosome-lysosome fusion. When transfected macrophages were infected with the same Mtb strains, similar patterns of subunit H and VPS33B localization was observed as compared to the immunofluorescent stained macrophages (Figure 18). This result further supports that PtpA binding to subunit H is needed for the exclusion of V-ATPase from the phagosome. \t\r \u00C2\u00A0 70 Figure 17. Confocal microscopy of THP-1 infected with E. coli or indicated Mtb strains. (A) Localization of VPS33B and V-ATPase subunit H was detected with immunofluorescence staining. V-ATPase was excluded from the Mtb phagosome in macrophages infected with the wild type strain while the \u00CE\u0094ptpA phagosome acquired V-ATPase. The phagosomes of the binding-defective \t\r \u00C2\u00A0 71 PtpAL146A expressing strain also failed to exclude host V-ATPase. (B) Quantification of the confocal data shown in (A). Values are the mean \u00C2\u00B1 SD of phagosome colocalization with V-ATPase subunit H in 50-80 cells from three independent experiments. *** = p < 0.001, significant difference compared with H37Rv by Student\u00E2\u0080\u0099s t test. \t\r \u00C2\u00A0 72 \t\r \u00C2\u00A0 Figure 18. Confocal microscopy of infected THP-1 macrophage doubly transfected with GFP-subunit H and DsRed2-VPS33B. (A) The transfected macrophages were infected with E. coli or indicated Mtb strains, and the localization of subunit H and VPS33B was directly visualized. V-ATPase was \t\r \u00C2\u00A0 73 excluded from the Mtb phagosome in macrophages infected with the wild type strain while the \u00CE\u0094ptpA phagosome acquired V-ATPase. The phagosomes of the binding-defective PtpAL146A expressing strain also failed to exclude host V- ATPase. (B) Quantification of the confocal data shown in (A). *** = p < 0.001, significant difference compared with H37Rv by Student\u00E2\u0080\u0099s t test. \t\r \u00C2\u00A0 3.2 PtpA is a substrate for the novel protein-tyrosine kinase Mtb PtkA 3.2.1 Introduction Protein tyrosine phosphorylation has long been recognized to play a key role in the regulation of numerous fundamental cellular processes in eukaryotes [148] and various bacterial species [149-151]. Thus far, bacterial phosphotyrosine signaling has been shown to be involved in cell division, antibiotic production, capsule synthesis, and host infection [150-155]. The first indication of bacterial protein-tyrosine kinase activity was demonstrated in Escherichia coli, in which the presence of phosphotyrosine in partial hydrolysates of proteins was shown [156]. Since then, a large number of protein-tyrosine kinases and phosphatases were identified in various bacterial species. These include two protein kinases from Streptococcus pneumoniae that autophosphorylate on tyrosine residues and are involved in the regulation of capsular polysaccharide production [150, 152]. Pathogenic Yersiniae contain an extrachromosomal virulence plasmid that encodes a PTP named YopH, which is injected into the host through a Type III secretion system encoded by the same plasmid [157]. Once inside the host cells, YopH modulates host-signaling pathways by dephosphorylating p130Cas to inhibit phagocytosis of the bacteria [157]. \t\r \u00C2\u00A0 74 In Mtb, the presence of tyrosine phosphorylation activity has been predicted since the identification of a 55 kDa protein that was recognized by the 4G10 anti-phosphotyrosine antibody in cell extracts from virulent but not avirulent Mtb strains [158]. Furthermore, Mtb possesses two protein-tyrosine phosphatases, PtpA and PtpB, further implying the existence of phosphotyrosine signaling activity in Mtb [104, 105]. Previous studies have shown that genes encoding kinases are often located in the proximity or in the same operon with the genes encoding their substrates [159]. Studies on bacterial PTKs and PTPs also showed that their corresponding genes are frequently clustered in an operon such that the expression of the PTKs and PTPs can be coordinated and regulated in concert [e.g. 160]. Within the Mtb genome, ptpA is located in an apparent operon with Rv2232 and Rv2235 upstream and downstream, respectively. The Rv2232 gene encodes a protein annotated as a member of the Haloacid Dehalogenase like- hydrolases (HAD) super-family, which catalyze the hydrolysis of various molecules. Interestingly though, with more than 3000 sequenced proteins, this super-family comprises enzymes, such as phosphatases, ATPases, phosphonatase (P-C bond hydrolysis), and sugar phosphomutases [161, 162] that are specialized in phosphoryl transfer. In this study, we show that the Rv2232 encoded protein is a novel protein-tyrosine kinase that phosphorylates PtpA. \t\r \u00C2\u00A0 75 3.2.2 PtpA interacts with the protein encoded by Rv2232 As described earlier, genes in the proximity of Mtb kinases and phosphatases are proposed to act as substrates or be under the control of these regulatory proteins. We used AlphaScreen technology to determine and measure the interaction between PtpA and Rv2232. As shown in Figure 19, the protein encoded by Rv2232 binds to PtpA (Kd = 3.0 \u00C2\u00B5M). Interestingly, in the presence of ATP, a stronger affinity was measured as reflected by a Kd of 1.3 \u00C2\u00B5M. GTP was also able to serve as a phosphate donor with a Kd of 1.8 \u00C2\u00B5M. These results suggest that: (i) the presence of the phosphate donor increases the interaction affinity of the protein encoded by Rv2232 for PtpA, (ii) GTP serves as an alternative phosphate donor, and (iii) ATP is the preferred substrate based on the lower Kd. Figure 19. Rv2232 interacts with PtpA in vitro. Protein\u00E2\u0080\u0093protein interaction between Rv2232 and PtpA was measured using AlphaScreen technology. The \t\r \u00C2\u00A0 76 dissociation constant (Kd) was calculated using biotinylated Rv2232 and His6- tagged PtpA in the presence or absence of 50 \u00C2\u00B5M ATP (A) or GTP (B). GST was used as negative control. In the absence of ATP or GTP, Rv2232 interacts with PtpA in vitro with a Kd of 3.0 \u00C2\u00B5M. Addition of either ATP (A) or GTP (B) results in a stronger interaction between Rv2232 and PtpA with a Kd of 1.3 \u00C2\u00B5M (ATP) and 1.8 \u00C2\u00B5M (GTP). GST negative control shows a Kd of 2.9 x 10 7 \u00C2\u00B5M, indicating absence of significant protein-protein interaction. 3.2.3 ORF Rv2232 encodes an autophosphorylated protein-tyrosine kinase Since ptpA is clustered with Rv2232, and rationalizing that as a result of such proximity, the Rv2232-encoded protein might serve as a substrate for the tyrosine phosphatase PtpA, the phosphorylation status of Rv2232 was investigated. Indeed, in vitro phosphorylation assays using purified recombinant protein expressed from the Rv2232 gene revealed that this protein possesses ATP concentration- and time-dependent autophosphorylation activity (Figure 20). Figure 20. Rv2232 autophosphorylation in vitro. Recombinant PtkA is autophosphorylated in an ATP dose- (A) and time- (B) dependent manner. The \t\r \u00C2\u00A0 77 upper panels in (A) and (B) show phosphorylation visualized by autoradiography, and the lower panels represent the silver-stained gel. Molecular size markers (MM) were shown. To determine the identity of the phosphorylated residues, phospho-amino acid analysis of the [32P]-phosphorylated protein was performed under acidic conditions. As shown in Figure 21, this protein was found to be phosphorylated on tyrosine residues. These data provided support for the protein encoded by Rv2232 to be the first protein tyrosine auto-kinase identified in Mtb, and thus we named it as PtkA. Figure 21. Phospho-amino acid analysis of PtkA. PtkA is phosphorylated on tyrosine residues as revealed by phospho-amino acid analysis. The migration of hydrolysed protein was detected by TLC against phospho-amino acids standards \t\r \u00C2\u00A0 78 (Retention Factor (Rf): phosphor-serine (p-Ser)=0.22; phosphor-threonine (p- Thr)=0.25; phosphor-tyrosine (p-Tyr)=0.30; PtkA=0.29).. 3.2.4 Sequence-function analysis of PtkA Bacterial protein-tyrosine kinase signatures, termed Walker A and B motifs [163], and the glycine-rich loop GXGXXGXV motif [164], are not present in PtkA. As we were unable to find any conserved kinase signature from in silico analysis of PtkA, we constructed a series of mutants in order to identify key functional residues in the protein. We mutated all three lysine residues with alanine residues in Rv2232 rationalizing that they might be required for ATP binding. Furthermore, all three tyrosine residues present in Rv2232 were mutated to determine potential phosphorylation sites. Lastly, the first aspartate (D85) in the DXD motif, which is conserved and essential for the catalytic activity in the HAD super-family members, was mutated to determine its role in the catalytic mechanism of Rv2232 [165]. All of the mutated proteins were individually tested for their ability to undergo autophosphorylation by means of radiolabeled \u00CE\u00B3-[32P]- ATP incorporation in an in vitro kinase assay (Table 4). 3.2.4.1 Mutation of lysine residues Binding of ATP in an enzymatic autophosphorylation reaction depends on lysine residues [166]. Thus, we constructed PtkA mutants K184M, K217M, and K270M and found all of them unable to bind ATP in a dose-dependent manner (Figure 22). Moreover, a concomitant increase of 2 to 14-fold in the KM values revealed a decrease in the affinity of these mutants for ATP as detailed in Table 1. \t\r \u00C2\u00A0 79 3.2.4.2 Mutation of tyrosine residues We have shown earlier that PtkA is phosphorylated on tyrosine residues (Figure 21). To precisely map the tyrosine phosphorylation site, Y146A, Y150A and Y262A mutants were constructed. Only the Y262A mutant failed to undergo autophosphorylation, while the Y146A and Y150A mutants incorporated 32P in a dose-dependent manner (Figure 22). These results indicate that Y262 is the autophosphorylated tyrosine in the PtkA backbone. Interestingly, the affinity of these proteins for ATP binding and the enzymatic efficiency of 32P incorporation were not affected as reflected by similar values measured for KM and Kcat/KM respectively (Table 4). Figure 22. Mutational studies of PtkA catalytic mechanism. Autophosphorylation activity of a series of purified recombinant PtkA point mutant \t\r \u00C2\u00A0 80 proteins was assessed by incubation with increasing concentrations of \u00CE\u00B3-[32P] ATP. The reactions were resolved on silver-stained SDS-PAGE, and phosphorylation was visualized with autoradiography. Molecular size markers (MM) were shown. 3.2.4.3 Mutation of D85 The mutation D85A, which represents the first aspartate in the conserved DXD motif within the HAD super-family was found to be essential for PtkA autophosphorylation activity (Figure 22). This finding is supported by a considerable decrease in the enzymatic efficiency according to the measured Kcat/KM for this mutant, as well as by the significant decrease in the ATP-binding affinity (Table 4). Table 4. Autophosphorylation kinetic values of parental and mutated PtkA PtkA Vmax (pmol\u00C2\u009F\u00EF\u0082\u009F min-1\u00C2\u009F\u00EF\u0082\u009F mg-1) Km (nM) Kcat (s-1) Kcat/Km (M-1\u00C2\u009F\u00EF\u0082\u009F s-1) Wild type 1331 27 9.73 x 10-4 3.60 x 104 Y146A 1877 54 1.40 x 10-4 2.59 x 104 Y150A 1967 68 1.47 x 10-3 2.16 x 104 Y262A 855 59 6.38 x 10-4 1.08 x 104 D85A 533 118 3.98 x 10-4 3.40 x 103 K184M 9 x 10-5 40 6.79 x 10-11 1.68 x 10-3 K217M 117 89 8.71 x 10-5 9.76 x 102 K270M 302 379 2.26 x 10-4 5.97 x 102 3.2.5 PtpA is a substrate of PtkA To test whether PtkA can serve as a substrate for the tyrosine phosphatase PtpA, PtpA was supplemented in the autophosphorylation reaction. Interestingly, we found the opposite of the hypothesized relationship. PtkA autophosphorylation did not decrease upon addition of PtpA. Moreover, PtpA acted as a substrate of PtkA. As shown in Figure 23, PtkA phosphorylates PtpA (lane 1) and the generic \t\r \u00C2\u00A0 81 tyrosine kinase substrate myelin basic protein (MBP) (lane 2). Moreover, PtkA phosphorylated PtpA in an ATP dose- and time-dependent manner (Figure 24A and 24B), demonstrating that the proitein-tyrosine phosphatase PtpA is a cognate substrate of PtkA. Figure 23. PtkA phosphorylates PtpA in vitro. PtkA was incubated with PtpA in an in vitro kinase reaction including \u00CE\u00B3-[32P]-ATP. Myelin basic protein (MBP) was used as an artificial substrate. The results show that PtpA is a cognate substrate for PtkA. Molecular size markers (MM) were shown. Figure 24. PtkA phosphorylates PtpA in dose- and time-dependent manner. (A) Increasing concentrations of PtpA were incubated with PtkA in an in vitro \t\r \u00C2\u00A0 82 kinase reaction with \u00CE\u00B3-[32P]-ATP, subjected to silver-stained SDS-PAGE and visualized with autoradiography. An increase in PtpA phosphorylation signal can be observed as PtpA concentration increases, indicating a genuine PtkA kinase activity on PtpA. (B) PtkA and PtpA were incubated in an in vitro kinase reaction over the indicated time periods. Increase in PtpA phosphorylation over time can be observed. The lower panel in (B) represents the silver-stained SDS-PAGE, showing equal loading of PtpA. Molecular size markers (MM) were shown. To determine the amino acid residues affecting the interaction between PtkA and PtpA, we performed protein-protein interaction studies with a series of PtkA mutants in an AlphaScreen assay. The results listed in Table 5 show that K184M, Y146A, and Y150A PtkA mutants interacted with PtpA as reflected with Kd values ranging between 0.85 - 1.5 \u00C2\u00B5M. This range is similar to the Kd of 1.27 \u00C2\u00B5M obtained for the wild-type protein. However, K217M, K270M, Y262A and D85A mutants showed Kd values between 3- to 6-fold higher than that obtained with native PtkA (Kd = 1.27 \u00C2\u00B5M) (Table 5). These results support the finding that Y262A is the phosphorylated residue (Kd = 6.6 \u00C2\u00B5M), as a decrease of 5-fold in the interaction was measured with respect to the wild type protein (Kd = 1.27 \u00C2\u00B5M). In addition, K217M and K270M, with Kd of 4.98 and 7.08 \u00C2\u00B5M, respectively, appear to be the two residues involved in ATP binding as shown in Table 4. The measured Kd of D85A mutant (3.32 \u00C2\u00B5M) also indicated that this residue is involved in the reaction as an interaction with PtpA of about 3-fold weaker than the wild type protein (Kd = 1.27 \u00C2\u00B5M) was measured. We cannot accurately determine if K184 is an essential amino acid for the catalytic activity of the enzyme, as its autophosphorylation was not observed (Figure 22). Nevertheless, an interaction similar to the wild-type protein was measured (Kd = 1.47 \u00C2\u00B5M). \t\r \u00C2\u00A0 83 Table 5. Kd values of PtpA interacting with parental and mutated PtkA proteins PtkA PtpA Kd (\u00C2\u00B5M) Wild type Wild type 1.27 Y146A Wild type 0.97 Y150A Wild type 0.85 Y262A Wild type 6.60 D85A Wild type 3.32 K184M Wild type 1.47 K217M Wild type 4.98 K270M Wild type 7.08 Wild type Y128A/Y129A 1.24 3.2.6 PtkA can phosphorylate both Y128 and Y129 residues of PtpA Phospho-amino acid analysis of PtkA-phosphorylated PtpA showed that PtpA, like PtkA, is phosphorylated on tyrosine residues (Figure 25A). To determine the specific tyrosine phosphorylated residue, we replaced the three tyrosine residues in PtpA with alanine, generating Y67A, Y128A, and Y129A variant forms of the PtpA protein. All three PtpA tyrosine variants were able to incorporate 32P (Figure 25B, lanes 2-5). The proximity of both Y128 and Y129 residues indicated that upon mutation of one of these residues, the phosphorylation would undergo on the adjacent residue. Therefore, we constructed a double mutant, where both Y128 and Y129 were replaced with alanine. PtkA failed to phosphorylate this double mutant (Figure 25B, lane 1), indicating that both residues exist in the phosphorylated form. The binding affinity between PtkA and the PtpA Y128-129A double mutant was not affected by the residue replacement, as a Kd of 1.24 \u00C2\u00B5M is very similar to the Kd of 1.27 \u00C2\u00B5M obtained between PtkA and the parental PtpA. \t\r \u00C2\u00A0 84 Figure 25. PtkA phosphorylates PtpA on tyrosine residues. (A) Phospho- amino acid analysis of PtpA phosphorylated by PtkA in an in vitro kinase reaction with \u00CE\u00B3-[32P]-ATP. Retention factor (Rf): phosphoserine = 0.49; phosphothreonine = 0.55; phosphotyrosine = 0.59; PtpA = 0.61). (B) Site directed mutagenesis of PtpA was performed to mutate candidate tyrosine phosphorylation sites to alanine. The recombinant mutant proteins were purified and subjected to in vitro kinase reaction with PtkA in the presence of \u00CE\u00B3-[32P]-ATP. The reactions were resolved on silver-stained SDS-PAGE and visualized with autoradiography. PtkA failed to phosphorylate PtpAY128A-Y129A, pointing to Y128 and Y129 as the residues phosphorylated by PtkA. \t\r \u00C2\u00A0 85 CHAPTER 4: DISCUSSION 4.1 PtpA exclusion of the V-ATPase from phagosomal membrane More than a decade ago, Sturgill-Koszycki et al. demonstrated that macrophages fail to acidify phagosomes containing Mycobacteria because these phagosomes did not accumulate the V-ATPase responsible for phagosomal acidification [99]. This suggested that Mtb actively inhibits the fusion of its phagosome with vesicles harboring the V-ATPase complex [99]. Despite the lack of a mechanistic explanation, the exclusion of the V-ATPase complex during Mtb infection was a long-established paradigm. We showed that the lack of phagosome acidification is directly attributed to Mtb secreted protein PtpA, which specifically inhibits V- ATPase trafficking to the mycobacterial phagosome during phagosome maturation. During Mtb infection, PtpA positions itself within a cleft linker region between the N-terminal and C-terminal domains of the V-ATPase subunit H, which was predicted to serve as a platform for the interaction with other proteins [146]. In fact, the V-ATPase subunit H and the HIV Nef protein were previously shown to interact through a specific di-leucine motif [78]. We were able to identify a catalytically active Leucine mutant of PtpA (Leu146Ala), which lost its ability to bind to subunit H but retained its phosphatase activity. Due to its location within the core region of PtpA, this mutation most likely altered the conformation of the entire C-terminal alpha helix, leading to the defective binding. These and our binding sites mapping results together with a computer-generated model based upon a protein-docking algorithm (3D-Garden) confirmed the cleft region of \t\r \u00C2\u00A0 86 subunit H as the binding site for the C-terminal alpha helix of PtpA (Figure 11A and 11B). The observed binding to the V-ATPase subunit H provided the first evidence that PtpA has a major role in interfering with the Mtb phagosome acidification process. Indeed, we showed that the loss of PtpA interaction with the host V-ATPase machinery resulted in the failure to block phagosome acidification and impaired Mtb survival within the host macrophage. The interaction with V-ATPase subunit H and dephosphorylation of VPS33B are both required for PtpA inhibition of macrophage phagosome-lysosome fusion and phagosome acidification. In fact, the loss of either its phosphatase activity or binding to subunit H renders PtpA non-functional within the host cell. We have shown earlier that Mtb PtpA dephosphorylates the host macrophage protein VPS33B, a key regulator of membrane fusion, leading to inhibition of phagosome-lysosome fusion [32]. The identification of V-ATPase subunit H as an additional key partner to the process of PtpA-mediated phagosome maturation arrest suggested a possible link between Class C VPS and V-ATPase. During phagosome maturation, lysosomal V-ATPase is directly recruited to the phagosomal membrane for luminal acidification [46]. The Class C VPS complex has been implicated in this process as a membrane tethering factor. Although studies of Drosophila neurons and mammalian renal medulla cells have shown interaction between subunits of V-ATPase and SNARE proteins [81, 82], our studies of PtpA in Mtb demonstrate for the first time a direct interaction between the Class C VPS and V-ATPase. We further showed that this interaction could be detected, albeit at a weaker level, in non-infected \t\r \u00C2\u00A0 87 macrophages, indicating that association of Class C VPS and V-ATPase is not limited to phagosome maturation only, but occurs transiently in resting cell to regulate normal endosome-lysosome fusion. However, this association is likely upregulated during phagocytosis in order to deliver lysosomal contents to the phagosome. This provides compelling evidence that V-ATPase is a key player in the membrane fusion machinery, particularly in cooperating with the Class C VPS complex to tether the fusing phagosomal or endosomal membrane with the lysosome (Figure 6). Disruption by Mtb PtpA further exemplifies the importance of this interaction in host defense mechanism. The exact mechanism by which PtpA is secreted across the phagosomal membrane remains unclear. However, using electron microscopy, neutralizing antibodies and Western blot analysis we have previously shown that PtpA is present in the host cytosol milieu [32, 167]. There is evidence suggesting that bacterial proteins less than 70 kDa in molecular size can cross the phagosomal membrane [130]. This observation might be related to the more recent discovery that the ESX-1 secretion system with its substrate ESAT-6 can activate the inflammasome and perturb host cell membrane to facilitate the translocation of mycobacterial proteins into the macrophage cytosol [131, 168]. Interestingly, a recent study showed that Mtb deletion mutant in secA2, which encodes for a secondary general secretion pathway that does not require the typical N-terminal signal peptide, failed to inhibit phagosome acidification [169]. This similar phenotype to the \u00CE\u0094ptpA strain suggests that the SecA2 protein secretion system \t\r \u00C2\u00A0 88 might be responsible for the secretion of PtpA. Further investigation is needed to elucidate the mechanistic details of PtpA secretion. Figure 26. A model for the specific exclusion of V-ATPase and the inhibition of mycobacterial phagosome acidification by PtpA. During infection, the V-ATPase recruits the Class C VPS complex, possibly associated with VPS39 and VPS41 as the HOPS complex, aiding the tethering of the fusing phagosomal and lysosomal membranes. PtpA, secreted into the host cytosol by Mtb, binds to subunit H of the V-ATPase complex, disrupting the interaction between the two protein complexes and localizing itself near its catalytic substrate VPS33B. PtpA then dephosphorylates and inactivates VPS33B, thereby shutting down the membrane fusion machinery. Binding to subunit H therefore allows PtpA to specifically inhibit V-ATPase trafficking to the mycobacterial phagosome. With these results, in our current model, we suggest a two-step process for PtpA exclusion of V-ATPase and inactivation of VPS33B. During Mtb infection, PtpA binding to subunit H may first disrupt initial membrane tethering and localize PtpA to the proximity of its catalytic substrate VPS33B (Figure 26). Subsequent dephosphorylation of VPS33B would then inactivate the entire membrane fusion machinery and its downstream effectors, preventing delivery of V-ATPase to the \t\r \u00C2\u00A0 89 mycobacterial phagosome. This would explain the failure of the phosphatase- defective mutant PtpAD126A, which retains the capability of binding to V-ATPase and disrupting V-ATPase and Class C VPS complexes association, to prevent V- ATPase trafficking to the Mtb phagosome. Our observation that VPS33B remains phosphorylated in both the phosphatase-defective and binding-defective strain supports such a two-step process, indicating that PtpA binding to subunit H is a prerequisite for VPS33B dephosphorylation. This is also supported by our confocal data where both the phosphatase-defective PtpAD126A and binding- defective PtpAL146A strains failed to inhibit V-ATPase trafficking to Mtb phagosomes. Yet, the mycobacterial phagosome is not an isolated compartment; rather, it interacts extensively with the early endosomes to acquire nutrients, such as iron-bound transferrin, to allow Mtb survival within the host cell [170]. In fact, the Mtb cell wall glycolipid, phosphatidylinositol mannoside (PIM), is capable of stimulating homotypic fusion of early endosomes in an ATP-, cytosol-, and N- ethylmaleimide sensitive factor-dependent manner [171]. Recent advances in the study of membrane trafficking in S. cerevisiae show that vesicle fusion in the endocytic pathway depends on two membrane tethering protein complexes, CORVET (class C core vacuole/endosome tethering) and HOPS (homotypic fusion and protein sorting) [88, 172]. Class C VPS complex serves as the core of both CORVET and HOPS complexes through reversible association with CORVET-specific (VPS3 and VPS8) and HOPS-specific (VPS39 and VPS41) accessory subunits, which mediate early to late endosome fusion events and \t\r \u00C2\u00A0 90 fusion with lysosomes, respectively [reviewed in 173]. Although the CORVET complex has not yet been identified in mammalian cells, it likely exists and has the same functions given the high similarity in the transport machinery between yeast and mammals. On the other hand, the mammalian HOPS complex has been identified and is known to play the same role as the S. cerevisiae homologs in mediating the conversion of Rab5 to Rab7 and tethering the fusing membranes. Our results indicate that V-ATPase may specifically interact with the HOPS complex during phagosome-lysosome fusion (Figure 26). Therefore, by binding to the V-ATPase, PtpA could specifically localize to the phagosome- lysosome interface while CORVET-mediated early phagosome-endosome fusion process will remain intact. This specific localization mechanism would then allow PtpA to distinguish HOPS from CORVET, thereby specifically excluding V- ATPase to inhibit mycobacterial phagosome acidification. In fact, a link between the V-ATPase and HOPS complexes was suggested by a previous study demonstrating that HOPS specific subunit VPS41 failed to function in yeast which had mutations in the V-ATPase complex [174]. This indicates that in macrophages, the HOPS-specific subunits, VPS39 and VPS41, might be direct effectors of the V-ATPase complex during phagosome-lysosome fusion (Figure 26), and that the CORVET complex might be recruited to the phagocytic pathway through a different mechanism. As the HOPS-specific subunits are known to interact with activated Rab7 during endosome maturation [58], it is likely that Rab7 is the upstream activator of V-ATPase association with the Class C VPS complex. Further experiments will be needed to fully characterize these host \t\r \u00C2\u00A0 91 cellular pathways in the context of mycobacterial infection. Nonetheless, our study clearly demonstrated that the absence of V-ATPase and the lack of mycobacterial phagosome acidification are directly attributed to the Mtb protein tyrosine phosphatase PtpA. 4.2 PtpA is a substrate for the protein-tyrosine kinase PtkA As protein phosphorylation plays a fundamental role in a wide range of cellular processes, it is not surprising that a large number of structurally distinct protein kinases have evolved [175, 176]. The HAD enzyme super-family comprises enzymes with hydrolytic activities, but kinase activities were not reported as of yet. We provided evidence that while Mtb PtkA was annotated as a HAD super- family member it exhibited an authentic protein-tyrosine kinase activity and its substrate was PtpA. Although earlier evidence indicated that Mtb possesses protein-tyrosine phosphorylation activity [158], post-genomic bioinformatics analysis failed to identify a corresponding protein-tyrosine kinase. Extensive in silico analysis revealed that PtkA does not possess any signature or pattern related to tyrosine kinases. For instance, the Walker A and B motifs [163], which are present in bacterial autophosphorylating tyrosine kinases, are absent in PtkA. Our findings are in line with recent data describing novel protein-tyrosine kinases and exemplified by the MasK protein from Myxococcus xanthus [177] and WaaP from Pseudomonas aeruginosa [178], which do not contain conserved pattern homologous to classical tyrosine kinases [163] but have been reported as self- phosphorylated tyrosine kinases. Therefore, we conclude that protein tyrosine \t\r \u00C2\u00A0 92 kinase activity does not necessarily require the canonical protein-tyrosine kinase motifs. We carried out several independent experiments to demonstrate that PtkA possesses tyrosine kinase activity. These include the incorporation of 32P on tyrosine residues, detection of the radioactive phosphotyrosine residues in a TLC autoradiogram, and mutational analysis of PtkA. These experiments showed that the amino acid Y262 is the autophosphorylated tyrosine residue on PtkA. This finding indicates that Y146 and Y150 are not involved in the catalytic activity of PtkA. In an attempt to determine other residues involved in the catalytic activity of PtkA, we carried out mutational analyses on all three lysine residues in this protein. Mutations of these residues resulted in no incorporation of 32P, indicating that these residues could play a role in the enzymatic mechanism. Since individual mutations of each of the three lysines reduced the enzymatic activity to nearly undetectable levels, we, at this stage are not able to determine which one of them is providing the hydrogen bonds to both the phosphate and the nucleophilic [179]. Interestingly, we found that although the mutant K184M was not phosphorylated, it binds PtpA similarly as the wild-type protein. The inability of all three lysine mutants to incorporate 32P is in agreement with previous observations that two lysines are required for ATP binding in tyrosine kinase Ptk of Acinetobacter johnsonii [163]. The core mechanism of HAD proteins is the transfer of a phosphoryl group from a specific phosphate ester to an active site aspartate, and then to a water \t\r \u00C2\u00A0 93 molecule. The nucleophilic mechanism in HAD super-family members is driven by the first aspartate in the DXD motif. To demonstrate that this residue is involved in the catalytic activity of PtkA, D85 residue was replaced by alanine. The resulting mutant protein did not show any kinase activity indicating that D85 residue is indeed involved in the catalytic mechanism of the enzyme. These findings further supports the essentiality of the DXD motif for the catalytic activity of the HAD super-family members [161] and the inclusion of PtkA in this superfamily. Our findings that PtpA is the cognate substrate of PtkA are in agreement with the observations that PtpA is secreted and acts on host proteins. Therefore, it seems unlikely that its substrate would be present within the bacilli. Although preliminary reports have shown that PtkA is upregulated in the SCID mouse model of Mtb infection [180] and in murine macrophages [181], the role of PtpA phosphorylation by PtkA still needs to be defined. Mutational analyses of PtpA showed that PtkA could phosphorylate two adjacent tyrosine residues of PtpA. This result is in concordance with the findings that in human low-molecular- weight protein tyrosine phosphatase B, two adjacent tyrosine residues are phosphorylated by a protein-tyrosine kinase [182]. Thus, we hypothesize that phosphorylation of PtpA is needed to retain its activity within the host macrophages. Alternatively, phosphorylation might be needed for the PtpA secretion process from Mtb, through the phagosome membrane to the host cytosol. Nevertheless, due to the discrepancy between the molecular mass of PtpA, (about 18 kDa) and the reported tyrosine phosphorylated 55 kDa protein in \t\r \u00C2\u00A0 94 Mtb [158], we do not rule out broader activity for PtkA depending on differential environmental signals. It is very likely that PtkA has multiple substrates within Mtb and controls a larger network of signal cascades. In fact, a recent study on Mtb STPKs phosphorylation motifs found that PtkA is phosphorylated on Ser/Thr residues in vivo when Mtb is exposed to peroxide stress, indicating another level of regulation for PtkA tyrosine kinase activity [183]. Since peroxide stress is a condition that Mtb encounters within the hostile host environment, PtkA activity might be controlled by a peroxide-sensing STPK during entry into the host macrophage. Therefore, it is tempting to speculate that this hypothetical signaling mechanism might allow Mtb to sense the phagosomal environment within the macrophage, thereby, activating or upregulating either PtpA\u00E2\u0080\u0099s activity or secretion to halt the phagosome maturation process. The exact relationship between Mtb STPK control of PtkA and PtkA phosphorylation of PtpA will require further investigation. Nevertheless, we have already demonstrated that PtpA and its secretion are essential to progression of Mtb infection [32]. This together with the large industrial know-how, and small molecule libraries of kinase inhibitors, raised the interesting notion that PtkA might yet be another attractive target for drug development against this notorious disease. To conclude, our data provide evidence that: (a) PtkA is a new tyrosine kinase belonging to the HAD super-family, (b) PtkA is able to use either ATP or GTP as phosphate donors, (c) PtkA neither possesses kinase signatures nor resembles a protein-tyrosine kinase, but is autophosphorylated in tyrosine residues, and (d) PtkA is able to autophosphorylate and transfer the incorporated \t\r \u00C2\u00A0 95 phosphate group to the tyrosine phosphatase PtpA, although the role of PtpA phosphorylation in Mtb physiology remains to be determined. 4.3 Inhibitors against PtpA as novel antituberculosis drug The essentiality of PtpA in the intracellular survival and pathogenesis of Mtb in various infection models has spurred great interest in exploiting PtpA as a novel antitubercular drug target [111]. A major challenge for the development of Mtb PtpA inhibitors is to demonstrate efficacy in vivo due to the fact that PtpA is not essential for bacterial growth in in vitro. Therefore, the traditional high throughput screening approach for in vitro growth inhibition would not be effective for the discovery of PtpA inhibitors. Nevertheless, several programs have already developed selective inhibitors against PtpA. Platforms adopted from cancer research for the design of PTP inhibitors have been established, and compounds which can specifically inhibit PtpA enzymatic activity and reduce mycobacterial survival within host macrophages, have been successfully identified and characterized [106, 109, 110, 112, 113, 184]. Pioneering in the identification of PtpA inhibitory compounds, Manger et al. tested natural products such as stevastelins, roseophilins and prodigiosins, obtaining IC50 values between 8.8 \u00C2\u00B5M and 28.7 \u00C2\u00B5M [110]. However, as PtpA has a high sequence and structural similarity with human PTP, specificity is an issue with the identified compounds as they show inhibition against human PTPs including PTP1B and the dual-specificity phosphatase Cdc25A. More recently, using fragment-based discovery approach, where small chemical entities are \t\r \u00C2\u00A0 96 screened for affinity and subsequently linked to generate specific inhibitors, a group of aryl difluoromethylphosphonic acid (DFMP) compounds were identified as potent inhibitors of PtpA with Ki value of 1.4 \u00C2\u00B1 0.3 \u00C2\u00B5M [113]. This group of inhibitors, due to stringent design, was reported to have greater than 70-fold specificity for Mtb PtpA as compared to highly homologous human phosphatases. In another study, Chiaradia et al. [109] have recently characterized a family of synthetic chalcones, essential intermediate compounds in flavonoid biosynthesis in plants, as inhibitors of PtpA with IC50 values between 8.4 \u00C2\u00B5M and 53.7 \u00C2\u00B5M. When these compounds were analyzed for their efficacy in the THP-1 infection model, one inhibitor (named 4d), with minimal toxicity against the THP-1 macrophage-like cells, was able to reduce the bacillary load by 50% at 48 hours post infection and 77% at 96 hours post infection [111] (Figure 27). This was the first demonstration of the effectiveness of PtpA inhibitors as a potential antitubercular drug. Figure 27. Mtb survival in infected THP-1 treated with chalcone inhibitors. Differentiated THP-1 macrophage-like cells were infected with the H37Rv Mtb \t\r \u00C2\u00A0 97 strain and treated with 10 \u00C2\u00B5M of the indicated inhibitor at 0 hr. A second and third dose were added at 24 hr and 48 hr, respectively. Survival of intracellular Mtb (CFU) was determined by plating on 7H10 agar. DMSO is used as a negative control (-). Data are expressed as mean CFU for triplicate wells with standard error. Our findings in this dissertation that PtpA binds to the host macrophage V- ATPase subunit H represents another avenue for targeting and inhibiting PtpA activity within the host cell. Since PtpA binding to the V-ATPase machinery precedes the dephosphorylation of VPS33B, targeting PtpA\u00E2\u0080\u0099s binding capability might prevent its localization to the phagosome-lysosome fusion interface and completely inactivate PtpA functions. The identification of the binding site for subunit H on PtpA might further aid the development of compounds that can specifically block the interaction with the host protein complex. It is interesting to note that the V-ATPase subunit H is also a binding target for HIV Nef, an accessory protein that is required for viral infectivity and pathogenicity [78]. HIV Nef plays a key role in downregulating the surface expression of CD4 and MHC-I on T cells by advancing the endocytosis and degradation of these cell surface proteins [185, 186]. Nef-dependent reduction of surface MHC-I protects HIV-infected primary T cells from recognition and killing by HIV-specific cytotoxic T lymphocytes whereas downregulation of CD4 prevents superinfection and allows optimal production of viral particles [187, 188]. It has been proposed that Nef, which binds to the cytoplasmic tail of CD4, drives internalization and degradation of CD4 by interacting with V-ATPase subunit H, which, in turn, binds to and recruits medium chain (\u00C2\u00B52) of the adaptor protein complex 2 (AP-2) [78, 189]. AP-2 is responsible for the recruitment of clathrin \t\r \u00C2\u00A0 98 proteins, driving the formation of clathrin-coated pit and the subsequent endocytosis of surface CD4. Therefore, the importance of the V-ATPase subunit H in both Mtb and HIV infections indicates that a single inhibitory compound that can prevent both PtpA and Nef recognition of subunit H might be capable of limiting both Mtb and HIV proliferation within the co-infected human host. However, toxicity could be an issue as in order for such an inhibitory compound to impede both PtpA and Nef binding, it would most likely have to target the host V-ATPase complex, which is essential for numerous human physiological processes. Nevertheless, continued advances in drug discovery approaches might allow the development of such an inhibitory compound with low toxicity profile. This would then be a potential novel therapeutic against the current twin epidemic of Mtb and HIV co-infection, \u00E2\u0080\u009Ckilling two birds with one stone\u00E2\u0080\u009D. 4.4 Future directions The work presented in this thesis described the role of PtpA in the arrest of phagosome maturation during Mtb entry into the host macrophage. PtpA- mediated specific exclusion of the V-ATPase proton pump from the mycobacterial phagosome explains the mechanism behind the inhibition of phagosome acidification during macrophage infection with Mtb. PtpA is, therefore, is essential for the pathogenesis of TB, allowing Mtb to subvert macrophage antimicrobial mechanisms and persist within the human host. Furthermore, within Mtb, PtpA is a substrate for the protein tyrosine kinase PtkA, indicating a novel signal transduction pathway for the regulation of PtpA function. Based on the work thus far, the PtpA project has the potential to branch into several research \t\r \u00C2\u00A0 99 directions, such as the global host macrophage phosphoproteome response to PtpA and the physiological relevance of PtkA phosphorylation of PtpA. To investigate the global phosphoproteome response of the host macrophages to PtpA during Mtb infection, a quantitative proteomic analysis approach based upon the iTRAQ (isobaric tag for relative and absolute quantitation) methodology can be performed to fully reveal downstream host signaling cascades affected by PtpA. From our previous study, we found that VPS33B possesses protein-tyrosine kinase activity, indicating that it might regulate multiple downstream signaling pathways that are involved in membrane trafficking and fusion [32]. Therefore, PtpA dephosphorylation of VPS33B suggests that PtpA can disrupt multiple host proteins in order to perturb phagosome maturation and to promote Mtb intracellular survival. Human macrophages will be infected with the parental H37Rv wild type (WT) and the \u00CE\u0094ptpA Mtb strains, and the phosphorylation response of the infected macrophages to the different strains will be compared. The differential phosphorylation profiles of the host cells should then reveal the complete network of host signaling pathways that are modulated by PtpA, thereby, providing a complete overview of the physiological functions of PtpA in relation to Mtb pathogenicity. It will also be interesting to examine the effect of PtpA on the recruitment of proteins to the phagosome using the same iTRAQ approach. However, a hurdle of this analysis would be to isolate and obtain highly pure samples of Mtb-containing phagosomes from infected macrophages. \t\r \u00C2\u00A0 100 To investigate the role of PtkA and its phosphorylation of PtpA in Mtb pathophysiology, we have constructed a ptkA deletion mutant in the H37Rv background. The in vitro growth characteristics and in vivo survival in various infection models, including THP-1 macrophage-like cells, can be examined. If PtkA is involved in regulating PtpA activity or secretion, we should observe phenotypes similar to those seen with the \u00CE\u0094ptpA mutant strain. PtpA has also been demonstrated to be secreted into the culture media in vitro [105]. Therefore, to investigate whether PtkA controls secretion of PtpA, we can probe for the presence of PtpA in the culture filtrate proteins of the \u00CE\u0094ptkA mutant strain as an initial study. Furthermore, as mentioned above, a previous study has shown that PtkA might be phosphorylated by Mtb STPK under peroxide stress conditions in vitro [183], indicating an upstream regulatory control of PtkA activity. It would be interesting to investigate whether PtkA can indeed be a substrate for one of the eleven STPKs in Mtb. Since the biological functions of Mtb STPKs, such as regulation of cell wall biosynthesis, dormancy and cell division, have been extensively studied [138, 190-196], identification of such a signal transduction pathway might provide further insight into the physiological role of PtkA within Mtb. It is possible that PtkA might, in addition to PtpA, regulate a larger network of Mtb proteins involved in host-pathogen interaction and adaptation to the hostile host environment. Lastly, within the ptpA operon, Rv2235 encodes a transmembrane protein annotated as conserved hypothetical protein. An effort to comprehensively identify all genes required for Mtb growth using the transposon site hybridization \t\r \u00C2\u00A0 101 (TRaSH) technique predicted that Rv2235 is essential for Mtb growth [197]. However, the genomic location and transmembrane feature of Rv2235 indicates that this protein might also play a role in the regulation of PtpA secretion. It will be interesting to elucidate and characterize the function of the protein encoded by Rv2235. \t\r \u00C2\u00A0 102 REFERENCES 1. Dye, C. and Williams, B.G. (2010) The population dynamics and control of tuberculosis. Science 328, 856-861 2. Goldman, R.C., Plumley, K.V., and Laughon, B.E. (2007) The evolution of extensively drug resistant tuberculosis (XDR-TB): history, status and issues for global control. Infect Disord Drug Targets 7, 73-91 3. Gandhi, N.R., Moll, A., Sturm, A.W., Pawinski, R., Govender, T., Lalloo, U., Zeller, K., Andrews, J., and 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, 1575-1580 4. Vitoria, M., Granich, R., Gilks, C.F., Gunneberg, C., Hosseini, M., Were, W., Raviglione, M., and De Cock, K.M. (2009) The global fight against HIV/AIDS, tuberculosis, and malaria: current status and future perspectives. Am J Clin Pathol 131, 844-848 5. WHO (2011) Estimates of the incidence of TB. In WHO Report 2011: Global Tuberculosis Control 2011, pp. 10 - 14 6. Gutierrez, M.C., Brisse, S., Brosch, R., Fabre, M., Omais, B., Marmiesse, M., Supply, P., and Vincent, V. (2005) Ancient origin and gene mosaicism of the progenitor of Mycobacterium tuberculosis. PLoS Pathog 1, e5 7. Brosch, R., Gordon, S.V., Marmiesse, M., Brodin, P., Buchrieser, C., Eiglmeier, K., Garnier, T., Gutierrez, C., Hewinson, G., Kremer, K., Parsons, L.M., Pym, A.S., Samper, S., van Soolingen, D., and Cole, S.T. (2002) A new evolutionary scenario for the Mycobacterium tuberculosis complex. Proc Natl Acad Sci U S A 99, 3684-3689 8. Sreevatsan, S., Pan, X., Stockbauer, K.E., Connell, N.D., Kreiswirth, B.N., Whittam, T.S., and Musser, J.M. (1997) Restricted structural gene polymorphism in the Mycobacterium tuberculosis complex indicates evolutionarily recent global dissemination. Proc Natl Acad Sci U S A 94, 9869-9874 9. Garnier, T., Eiglmeier, K., Camus, J.C., Medina, N., Mansoor, H., Pryor, M., Duthoy, S., Grondin, S., Lacroix, C., Monsempe, C., Simon, S., Harris, B., Atkin, R., Doggett, J., Mayes, R., Keating, L., Wheeler, P.R., Parkhill, J., Barrell, B.G., Cole, S.T., Gordon, S.V., and Hewinson, R.G. (2003) The complete \t\r \u00C2\u00A0 103 genome sequence of Mycobacterium bovis. Proc Natl Acad Sci U S A 100, 7877- 7882 10. Gutacker, M.M., Smoot, J.C., Migliaccio, C.A., Ricklefs, S.M., Hua, S., Cousins, D.V., Graviss, E.A., Shashkina, E., Kreiswirth, B.N., and Musser, J.M. (2002) Genome-wide analysis of synonymous single nucleotide polymorphisms in Mycobacterium tuberculosis complex organisms: resolution of genetic relationships among closely related microbial strains. Genetics 162, 1533-1543 11. Besra, G.S., Khoo, K.H., McNeil, M.R., Dell, A., Morris, H.R., and Brennan, P.J. (1995) A new interpretation of the structure of the mycolyl-arabinogalactan complex of Mycobacterium tuberculosis as revealed through characterization of oligoglycosylalditol fragments by fast-atom bombardment mass spectrometry and 1H nuclear magnetic resonance spectroscopy. Biochemistry 34, 4257-4266 12. Daffe, M., Brennan, P.J., and McNeil, M. (1990) Predominant structural features of the cell wall arabinogalactan of Mycobacterium tuberculosis as revealed through characterization of oligoglycosyl alditol fragments by gas chromatography/mass spectrometry and by 1H and 13C NMR analyses. J Biol Chem 265, 6734-6743 13. Dover, L.G., Cerdeno-Tarraga, A.M., Pallen, M.J., Parkhill, J., and Besra, G.S. (2004) Comparative cell wall core biosynthesis in the mycolated pathogens, Mycobacterium tuberculosis and Corynebacterium diphtheriae. FEMS Microbiol Rev 28, 225-250 14. McNeil, M., Daffe, M., and Brennan, P.J. (1991) Location of the mycolyl ester substituents in the cell walls of mycobacteria. J Biol Chem 266, 13217- 13223 15. McNeil, M., Daffe, M., and Brennan, P.J. (1990) Evidence for the nature of the link between the arabinogalactan and peptidoglycan of mycobacterial cell walls. J Biol Chem 265, 18200-18206 16. Bhowruth, V., Alderwick, L.J., Brown, A.K., Bhatt, A., and Besra, G.S. (2008) Tuberculosis: a balanced diet of lipids and carbohydrates. Biochem Soc Trans 36, 555-565 17. Brennan, P.J. and Crick, D.C. (2007) The cell-wall core of Mycobacterium tuberculosis in the context of drug discovery. Curr Top Med Chem 7, 475-488 \t\r \u00C2\u00A0 104 18. Brennan, P.J. and Nikaido, H. (1995) The envelope of mycobacteria. Annu Rev Biochem 64, 29-63 19. Jarlier, V. and Nikaido, H. (1990) Permeability barrier to hydrophilic solutes in Mycobacterium chelonei. J Bacteriol 172, 1418-1423 20. Algood, H.M., Chan, J., and Flynn, J.L. (2003) Chemokines and tuberculosis. Cytokine Growth Factor Rev 14, 467-477 21. Flynn, J.L. and Chan, J. (2005) What's good for the host is good for the bug. Trends Microbiol 13, 98-102 22. Russell, D.G. (2007) Who puts the tubercle in tuberculosis? Nat Rev Microbiol 5, 39-47 23. Dheda, K., Booth, H., Huggett, J.F., Johnson, M.A., Zumla, A., and Rook, G.A. (2005) Lung remodeling in pulmonary tuberculosis. J Infect Dis 192, 1201- 1209 24. Kaplan, G., Post, F.A., Moreira, A.L., Wainwright, H., Kreiswirth, B.N., Tanverdi, M., Mathema, B., Ramaswamy, S.V., Walther, G., Steyn, L.M., Barry, C.E., 3rd, and Bekker, L.G. (2003) Mycobacterium tuberculosis growth at the cavity surface: a microenvironment with failed immunity. Infect Immun 71, 7099- 7108 25. Barry, C.E., 3rd, Lee, R.E., Mdluli, K., Sampson, A.E., Schroeder, B.G., Slayden, R.A., and Yuan, Y. (1998) Mycolic acids: structure, biosynthesis and physiological functions. Prog Lipid Res 37, 143-179 26. Ehrt, S. and Schnappinger, D. (2009) Mycobacterial survival strategies in the phagosome: defence against host stresses. Cell Microbiol 11, 1170-1178 27. Fratti, R.A., Chua, J., Vergne, I., and Deretic, V. (2003) Mycobacterium tuberculosis glycosylated phosphatidylinositol causes phagosome maturation arrest. Proc Natl Acad Sci U S A 100, 5437-5442 \t\r \u00C2\u00A0 105 28. Vergne, I., Chua, J., and Deretic, V. (2003) Tuberculosis toxin blocking phagosome maturation inhibits a novel Ca2+/calmodulin-PI3K hVPS34 cascade. J Exp Med 198, 653-659 29. Kang, P.B., Azad, A.K., Torrelles, J.B., Kaufman, T.M., Beharka, A., Tibesar, E., DesJardin, L.E., and Schlesinger, L.S. (2005) The human macrophage mannose receptor directs Mycobacterium tuberculosis lipoarabinomannan-mediated phagosome biogenesis. J Exp Med 202, 987-999 30. Briken, V. and Miller, J.L. (2008) Living on the edge: inhibition of host cell apoptosis by Mycobacterium tuberculosis. Future Microbiol 3, 415-422 31. Lee, J., Repasy, T., Papavinasasundaram, K., Sassetti, C., and Kornfeld, H. (2011) Mycobacterium tuberculosis induces an atypical cell death mode to escape from infected macrophages. PLoS One 6, e18367 32. Bach, H., Papavinasasundaram, K.G., Wong, D., Hmama, Z., and Av-Gay, Y. (2008) Mycobacterium tuberculosis virulence is mediated by PtpA dephosphorylation of human vacuolar protein sorting 33B. Cell Host Microbe 3, 316-322 33. Wong, D., Bach, H., Sun, J., Hmama, Z., and Av-Gay, Y. (2011) Mycobacterium tuberculosis protein tyrosine phosphatase (PtpA) excludes host vacuolar-H+-ATPase to inhibit phagosome acidification. Proc Natl Acad Sci U S A 108, 19371-19376 34. Vergne, I., Chua, J., Lee, H.H., Lucas, M., Belisle, J., and Deretic, V. (2005) Mechanism of phagolysosome biogenesis block by viable Mycobacterium tuberculosis. Proc Natl Acad Sci U S A 102, 4033-4038 35. Taylor, M.E. and Drickamer, K. (1993) Structural requirements for high affinity binding of complex ligands by the macrophage mannose receptor. J Biol Chem 268, 399-404 36. Ezekowitz, R.A., Sastry, K., Bailly, P., and Warner, A. (1990) Molecular characterization of the human macrophage mannose receptor: demonstration of multiple carbohydrate recognition-like domains and phagocytosis of yeasts in Cos-1 cells. J Exp Med 172, 1785-1794 \t\r \u00C2\u00A0 106 37. Bharadwaj, D., Mold, C., Markham, E., and Du Clos, T.W. (2001) Serum amyloid P component binds to Fc gamma receptors and opsonizes particles for phagocytosis. J Immunol 166, 6735-6741 38. Bharadwaj, D., Stein, M.P., Volzer, M., Mold, C., and Du Clos, T.W. (1999) The major receptor for C-reactive protein on leukocytes is fcgamma receptor II. J Exp Med 190, 585-590 39. Ehlers, M.R. (2000) CR3: a general purpose adhesion-recognition receptor essential for innate immunity. Microbes Infect 2, 289-294 40. Daeron, M. (1997) Fc receptor biology. Annu Rev Immunol 15, 203-234 41. Ghazizadeh, S., Bolen, J.B., and Fleit, H.B. (1994) Physical and functional association of Src-related protein tyrosine kinases with Fc gamma RII in monocytic THP-1 cells. J Biol Chem 269, 8878-8884 42. Caron, E. and Hall, A. (1998) Identification of two distinct mechanisms of phagocytosis controlled by different Rho GTPases. Science 282, 1717-1721 43. Niedergang, F., Colucci-Guyon, E., Dubois, T., Raposo, G., and Chavrier, P. (2003) ADP ribosylation factor 6 is activated and controls membrane delivery during phagocytosis in macrophages. J Cell Biol 161, 1143-1150 44. Araki, N., Johnson, M.T., and Swanson, J.A. (1996) A role for phosphoinositide 3-kinase in the completion of macropinocytosis and phagocytosis by macrophages. J Cell Biol 135, 1249-1260 45. Desjardins, M., Huber, L.A., Parton, R.G., and Griffiths, G. (1994) Biogenesis of phagolysosomes proceeds through a sequential series of interactions with the endocytic apparatus. J Cell Biol 124, 677-688 46. Sun-Wada, G., Tabata, H., Kawamura, N., Aoyama, M., and Wada, Y. (2009) Direct recruitment of H+-ATPase from lysosomes for phagosomal acidification. J Cell Sci 122, 2504-2513 47. Beatty, W.L., Rhoades, E.R., Ullrich, H.J., Chatterjee, D., Heuser, J.E., and Russell, D.G. (2000) Trafficking and release of mycobacterial lipids from infected macrophages. Traffic 1, 235-247 \t\r \u00C2\u00A0 107 48. Bucci, C., Parton, R.G., Mather, I.H., Stunnenberg, H., Simons, K., Hoflack, B., and Zerial, M. (1992) The small GTPase rab5 functions as a regulatory factor in the early endocytic pathway. Cell 70, 715-728 49. Vieira, O.V., Botelho, R.J., Rameh, L., Brachmann, S.M., Matsuo, T., Davidson, H.W., Schreiber, A., Backer, J.M., Cantley, L.C., and Grinstein, S. (2001) Distinct roles of class I and class III phosphatidylinositol 3-kinases in phagosome formation and maturation. J Cell Biol 155, 19-25 50. Gaullier, J.M., Simonsen, A., D'Arrigo, A., Bremnes, B., Stenmark, H., and Aasland, R. (1998) FYVE fingers bind PtdIns(3)P. Nature 394, 432-433 51. Kanai, F., Liu, H., Field, S.J., Akbary, H., Matsuo, T., Brown, G.E., Cantley, L.C., and Yaffe, M.B. (2001) The PX domains of p47phox and p40phox bind to lipid products of PI(3)K. Nat Cell Biol 3, 675-678 52. Callaghan, J., Nixon, S., Bucci, C., Toh, B.H., and Stenmark, H. (1999) Direct interaction of EEA1 with Rab5b. Eur J Biochem 265, 361-366 53. McBride, H.M., Rybin, V., Murphy, C., Giner, A., Teasdale, R., and Zerial, M. (1999) Oligomeric complexes link Rab5 effectors with NSF and drive membrane fusion via interactions between EEA1 and syntaxin 13. Cell 98, 377- 386 54. Babior, B.M. (2004) NADPH oxidase. Curr Opin Immunol 16, 42-47 55. Bedard, K. and Krause, K.H. (2007) The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev 87, 245-313 56. Bucci, C., Thomsen, P., Nicoziani, P., McCarthy, J., and van Deurs, B. (2000) Rab7: a key to lysosome biogenesis. Mol Biol Cell 11, 467-480 57. Harrison, R.E., Bucci, C., Vieira, O.V., Schroer, T.A., and Grinstein, S. (2003) Phagosomes fuse with late endosomes and/or lysosomes by extension of membrane protrusions along microtubules: role of Rab7 and RILP. Mol Cell Biol 23, 6494-6506 \t\r \u00C2\u00A0 108 58. Plemel, R.L., Lobingier, B.T., Brett, C.L., Angers, C.G., Nickerson, D.P., Paulsel, A., Sprague, D., and Merz, A.J. (2011) Subunit organization and Rab interactions of Vps-C protein complexes that control endolysosomal membrane traffic. Mol Biol Cell 22, 1353-1363 59. Rink, J., Ghigo, E., Kalaidzidis, Y., and Zerial, M. (2005) Rab conversion as a mechanism of progression from early to late endosomes. Cell 122, 735-749 60. Jordens, I., Fernandez-Borja, M., Marsman, M., Dusseljee, S., Janssen, L., Calafat, J., Janssen, H., Wubbolts, R., and Neefjes, J. (2001) The Rab7 effector protein RILP controls lysosomal transport by inducing the recruitment of dynein- dynactin motors. Curr Biol 11, 1680-1685 61. Antonin, W., Holroyd, C., Fasshauer, D., Pabst, S., Von Mollard, G.F., and Jahn, R. (2000) A SNARE complex mediating fusion of late endosomes defines conserved properties of SNARE structure and function. The EMBO journal 19, 6453-6464 62. Sato, T.K., Rehling, P., Peterson, M.R., and Emr, S.D. (2000) Class C Vps protein complex regulates vacuolar SNARE pairing and is required for vesicle docking/fusion. Mol Cell 6, 661-671 63. Lehrer, R.I., Lichtenstein, A.K., and Ganz, T. (1993) Defensins: antimicrobial and cytotoxic peptides of mammalian cells. Annu Rev Immunol 11, 105-128 64. Zanetti, M. (2005) The role of cathelicidins in the innate host defenses of mammals. Curr Issues Mol Biol 7, 179-196 65. Pillay, C.S., Elliott, E., and Dennison, C. (2002) Endolysosomal proteolysis and its regulation. Biochem J 363, 417-429 66. Claus, V., Jahraus, A., Tjelle, T., Berg, T., Kirschke, H., Faulstich, H., and Griffiths, G. (1998) Lysosomal enzyme trafficking between phagosomes, endosomes, and lysosomes in J774 macrophages. Enrichment of cathepsin H in early endosomes. J Biol Chem 273, 9842-9851 67. Forgac, M. (2007) Vacuolar ATPases: rotary proton pumps in physiology and pathophysiology. Nat Rev Mol Cell Biol 8, 917-929 \t\r \u00C2\u00A0 109 68. Jefferies, K.C., Cipriano, D.J., and Forgac, M. (2008) Function, structure and regulation of the vacuolar (H+)-ATPases. Arch Biochem Biophys 476, 33-42 69. Yoshida, M., Muneyuki, E., and Hisabori, T. (2001) ATP synthase--a marvellous rotary engine of the cell. Nat Rev Mol Cell Biol 2, 669-677 70. Liu, Q., Kane, P.M., Newman, P.R., and Forgac, M. (1996) Site-directed mutagenesis of the yeast V-ATPase B subunit (Vma2p). J Biol Chem 271, 2018- 2022 71. Shao, E., Nishi, T., Kawasaki-Nishi, S., and Forgac, M. (2003) Mutational analysis of the non-homologous region of subunit A of the yeast V-ATPase. J Biol Chem 278, 12985-12991 72. Yokoyama, K., Nakano, M., Imamura, H., Yoshida, M., and Tamakoshi, M. (2003) Rotation of the proteolipid ring in the V-ATPase. J Biol Chem 278, 24255- 24258 73. Fillingame, R.H., Angevine, C.M., and Dmitriev, O.Y. (2002) Coupling proton movements to c-ring rotation in F(1)F(o) ATP synthase: aqueous access channels and helix rotations at the a-c interface. Biochim Biophys Acta 1555, 29- 36 74. Kawasaki-Nishi, S., Nishi, T., and Forgac, M. (2001) Yeast V-ATPase complexes containing different isoforms of the 100-kDa a-subunit differ in coupling efficiency and in vivo dissociation. J Biol Chem 276, 17941-17948 75. Muench, S.P., Huss, M., Song, C.F., Phillips, C., Wieczorek, H., Trinick, J., and Harrison, M.A. (2009) Cryo-electron microscopy of the vacuolar ATPase motor reveals its mechanical and regulatory complexity. J Mol Biol 386, 989-999 76. Ho, M.N., Hirata, R., Umemoto, N., Ohya, Y., Takatsuki, A., Stevens, T.H., and Anraku, Y. (1993) VMA13 encodes a 54-kDa vacuolar H(+)-ATPase subunit required for activity but not assembly of the enzyme complex in Saccharomyces cerevisiae. J Biol Chem 268, 18286-18292 77. Parra, K.J., Keenan, K.L., and Kane, P.M. (2000) The H subunit (Vma13p) of the yeast V-ATPase inhibits the ATPase activity of cytosolic V1 complexes. J Biol Chem 275, 21761-21767 \t\r \u00C2\u00A0 110 78. Geyer, M., Yu, H., Mandic, R., Linnemann, T., Zheng, Y.H., Fackler, O.T., and Peterlin, B.M. (2002) Subunit H of the V-ATPase binds to the medium chain of adaptor protein complex 2 and connects Nef to the endocytic machinery. J Biol Chem 277, 28521-28529 79. Hurtado-Lorenzo, A., Skinner, M., El Annan, J., Futai, M., Sun-Wada, G.H., Bourgoin, S., Casanova, J., Wildeman, A., Bechoua, S., Ausiello, D.A., Brown, D., and Marshansky, V. (2006) V-ATPase interacts with ARNO and Arf6 in early endosomes and regulates the protein degradative pathway. Nat Cell Biol 8, 124- 136 80. Lu, M., Ammar, D., Ives, H., Albrecht, F., and Gluck, S.L. (2007) Physical interaction between aldolase and vacuolar H+-ATPase is essential for the assembly and activity of the proton pump. J Biol Chem 282, 24495-24503 81. Hiesinger, P.R., Fayyazuddin, A., Mehta, S.Q., Rosenmund, T., Schulze, K.L., Zhai, R.G., Verstreken, P., Cao, Y., Zhou, Y., Kunz, J., and Bellen, H.J. (2005) The v-ATPase V0 subunit a1 is required for a late step in synaptic vesicle exocytosis in Drosophila. Cell 121, 607-620 82. Li, G., Alexander, E.A., and Schwartz, J.H. (2003) Syntaxin isoform specificity in the regulation of renal H+-ATPase exocytosis. J Biol Chem 278, 19791-19797 83. Duli\u00C4\u0087, V. and Riezman, H. (1990) Saccharomyces cerevisiae mutants lacking a functional vacuole are defective for aspects of the pheromone response. J Cell Sci 97 ( Pt 3), 517-525 84. Wada, Y., Ohsumi, Y., and Anraku, Y. (1992) Genes for directing vacuolar morphogenesis in Saccharomyces cerevisiae. I. Isolation and characterization of two classes of vam mutants. J Biol Chem 267, 18665-18670 85. Akbar, M.A., Ray, S., and Kr\u00C3\u00A4mer, H. (2009) The SM protein Car/Vps33A regulates SNARE-mediated trafficking to lysosomes and lysosome-related organelles. Mol Biol Cell 20, 1705-1714 86. Seals, D.F., Eitzen, G., Margolis, N., Wickner, W.T., and Price, A. (2000) A Ypt/Rab effector complex containing the Sec1 homolog Vps33p is required for homotypic vacuole fusion. Proc Natl Acad Sci U S A 97, 9402-9407 \t\r \u00C2\u00A0 111 87. Kim, B.Y., Kr\u00C3\u00A4mer, H., Yamamoto, A., Kominami, E., Kohsaka, S., and Akazawa, C. (2001) Molecular characterization of mammalian homologues of class C Vps proteins that interact with syntaxin-7. J Biol Chem 276, 29393-29402 88. Peplowska, K., Markgraf, D.F., Ostrowicz, C.W., Bange, G., and Ungermann, C. (2007) The CORVET tethering complex interacts with the yeast Rab5 homolog Vps21 and is involved in endo-lysosomal biogenesis. Dev Cell 12, 739-750 89. Peterson, M.R. and Emr, S.D. (2001) The class C Vps complex functions at multiple stages of the vacuolar transport pathway. Traffic 2, 476-486 90. Cooper, A.M. (2009) Cell-mediated immune responses in tuberculosis. Annu Rev Immunol 27, 393-422 91. Jozefowski, S., Sobota, A., and Kwiatkowska, K. (2008) How Mycobacterium tuberculosis subverts host immune responses. Bioessays 30, 943-954 92. Lee, W.L., Gold, B., Darby, C., Brot, N., Jiang, X., de Carvalho, L.P., Wellner, D., St John, G., Jacobs, W.R., Jr., and Nathan, C. (2009) Mycobacterium tuberculosis expresses methionine sulphoxide reductases A and B that protect from killing by nitrite and hypochlorite. Mol Microbiol 71, 583-593 93. Colangeli, R., Haq, A., Arcus, V.L., Summers, E., Magliozzo, R.S., McBride, A., Mitra, A.K., Radjainia, M., Khajo, A., Jacobs, W.R., Jr., Salgame, P., and Alland, D. (2009) The multifunctional histone-like protein Lsr2 protects mycobacteria against reactive oxygen intermediates. Proc Natl Acad Sci U S A 106, 4414-4418 94. Via, L.E., Deretic, D., Ulmer, R.J., Hibler, N.S., Huber, L.A., and Deretic, V. (1997) Arrest of mycobacterial phagosome maturation is caused by a block in vesicle fusion between stages controlled by rab5 and rab7. J Biol Chem 272, 13326-13331 95. Malik, Z.A., Thompson, C.R., Hashimi, S., Porter, B., Iyer, S.S., and Kusner, D.J. (2003) Cutting edge: Mycobacterium tuberculosis blocks Ca2+ signaling and phagosome maturation in human macrophages via specific inhibition of sphingosine kinase. J Immunol 170, 2811-2815 \t\r \u00C2\u00A0 112 96. Prost, L.R., Daley, M.E., Le Sage, V., Bader, M.W., Le Moual, H., Klevit, R.E., and Miller, S.I. (2007) Activation of the bacterial sensor kinase PhoQ by acidic pH. Mol Cell 26, 165-174 97. Tsukano, H., Kura, F., Inoue, S., Sato, S., Izumiya, H., Yasuda, T., and Watanabe, H. (1999) Yersinia pseudotuberculosis blocks the phagosomal acidification of B10.A mouse macrophages through the inhibition of vacuolar H(+)-ATPase activity. Microb Pathog 27, 253-263 98. Xu, L., Shen, X., Bryan, A., Banga, S., Swanson, M.S., and Luo, Z.Q. (2010) Inhibition of host vacuolar H+-ATPase activity by a Legionella pneumophila effector. PLoS Pathog 6, e1000822 99. Sturgill-Koszycki, S., Schlesinger, P.H., Chakraborty, P., Haddix, P.L., Collins, H.L., Fok, A.K., Allen, R.D., Gluck, S.L., Heuser, J., and Russell, D.G. (1994) Lack of acidification in Mycobacterium phagosomes produced by exclusion of the vesicular proton-ATPase. Science 263, 678-681 100. Sun, J., Deghmane, A.E., Soualhine, H., Hong, T., Bucci, C., Solodkin, A., and Hmama, Z. (2007) Mycobacterium bovis BCG disrupts the interaction of Rab7 with RILP contributing to inhibition of phagosome maturation. J Leukoc Biol 82, 1437-1445 101. Sun, J., Wang, X., Lau, A., Liao, T.Y., Bucci, C., and Hmama, Z. (2010) Mycobacterial nucleoside diphosphate kinase blocks phagosome maturation in murine RAW 264.7 macrophages. PLoS One 5, e8769 102. Av-Gay, Y. and Everett, M. (2000) The eukaryotic-like Ser/Thr protein kinases of Mycobacterium tuberculosis. Trends Microbiol 8, 238-244 103. Bach, H., Wong, D., and Av-Gay, Y. (2009) Mycobacterium tuberculosis PtkA is a novel protein tyrosine kinase whose substrate is PtpA. Biochem J 420, 155-160 104. Koul, A., Choidas, A., Treder, M., Tyagi, A.K., Drlica, K., Singh, Y., and Ullrich, A. (2000) Cloning and characterization of secretory tyrosine phosphatases of Mycobacterium tuberculosis. J Bacteriol 182, 5425-5432 \t\r \u00C2\u00A0 113 105. Cowley, S.C., Babakaiff, R., and Av-Gay, Y. (2002) Expression and localization of the Mycobacterium tuberculosis protein tyrosine phosphatase PtpA. Res Microbiol 153, 233-241 106. Beresford, N.J., Mulhearn, D., Szczepankiewicz, B., Liu, G., Johnson, M.E., Fordham-Skelton, A., Abad-Zapatero, C., Cavet, J.S., and Tabernero, L. (2009) Inhibition of MptpB phosphatase from Mycobacterium tuberculosis impairs mycobacterial survival in macrophages. J Antimicrob Chemother 63, 928-936 107. Singh, R., Rao, V., Shakila, H., Gupta, R., Khera, A., Dhar, N., Singh, A., Koul, A., Singh, Y., Naseema, M., Narayanan, P.R., Paramasivan, C.N., Ramanathan, V.D., and Tyagi, A.K. (2003) Disruption of mptpB impairs the ability of Mycobacterium tuberculosis to survive in guinea pigs. Mol Microbiol 50, 751- 762 108. Chiaradia, L.D., Martins, P.G., Cordeiro, M.N., Guido, R.V., Ecco, G., Andricopulo, A.D., Yunes, R.A., Vernal, J., Nunes, R.J., and Terenzi, H. (2011) Synthesis, Biological Evaluation, And Molecular Modeling of Chalcone Derivatives As Potent Inhibitors of Mycobacterium tuberculosis Protein Tyrosine Phosphatases (PtpA and PtpB). J Med Chem 109. Chiaradia, L.D., Mascarello, A., Purificacao, M., Vernal, J., Cordeiro, M.N., Zenteno, M.E., Villarino, A., Nunes, R.J., Yunes, R.A., and Terenzi, H. (2008) Synthetic chalcones as efficient inhibitors of Mycobacterium tuberculosis protein tyrosine phosphatase PtpA. Bioorg Med Chem Lett 18, 6227-6230 110. Manger, M., Scheck, M., Prinz, H., von Kries, J.P., Langer, T., Saxena, K., Schwalbe, H., Furstner, A., Rademann, J., and Waldmann, H. (2005) Discovery of Mycobacterium tuberculosis protein tyrosine phosphatase A (MptpA) inhibitors based on natural products and a fragment-based approach. Chembiochem 6, 1749-1753 111. Mascarello, A., Chiaradia, L.D., Vernal, J., Villarino, A., Guido, R.V., Perizzolo, P., Poirier, V., Wong, D., Martins, P.G., Nunes, R.J., Yunes, R.A., Andricopulo, A.D., Av-Gay, Y., and Terenzi, H. (2010) Inhibition of Mycobacterium tuberculosis tyrosine phosphatase PtpA by synthetic chalcones: kinetics, molecular modeling, toxicity and effect on growth. Bioorg Med Chem 18, 3783-3789 112. Grundner, C., Perrin, D., Hooft van Huijsduijnen, R., Swinnen, D., Gonzalez, J., Gee, C.L., Wells, T.N., and Alber, T. (2007) Structural basis for \t\r \u00C2\u00A0 114 selective inhibition of Mycobacterium tuberculosis protein tyrosine phosphatase PtpB. Structure 15, 499-509 113. Soellner, M.B., Rawls, K.A., Grundner, C., Alber, T., and Ellman, J.A. (2007) Fragment-based substrate activity screening method for the identification of potent inhibitors of the Mycobacterium tuberculosis phosphatase PtpB. J Am Chem Soc 129, 9613-9615 114. Grundner, Ng, and Alber (2005) Mycobacterium tuberculosis protein tyrosine phosphatase PtpB structure reveals a diverged fold and a buried active site. Structure 13, 1625-1634 115. Beresford, N., Patel, S., Armstrong, J., Szoor, B., Fordham-Skelton, A.P., and Tabernero, L. (2007) MptpB, a virulence factor from Mycobacterium tuberculosis, exhibits triple-specificity phosphatase activity. Biochem J 406, 13- 18 116. Zhou, B., He, Y., Zhang, X., Xu, J., Luo, Y., Wang, Y., Franzblau, S.G., Yang, Z., Chan, R.J., Liu, Y., Zheng, J., and Zhang, Z.Y. (2010) Targeting mycobacterium protein tyrosine phosphatase B for antituberculosis agents. Proceedings of the National Academy of Sciences of the United States of America 107, 4573-4578 117. Roach, S.K. and Schorey, J.S. (2002) Differential regulation of the mitogen-activated protein kinases by pathogenic and nonpathogenic mycobacteria. Infection and immunity 70, 3040-3052 118. A, S.K., Bansal, K., Holla, S., Verma-Kumar, S., Sharma, P., and Balaji, K.N. (2012) ESAT-6 induced COX-2 expression involves coordinated interplay between PI3K and MAPK signaling. Mol Immunol 49, 655-663 119. Ganguly, N., Giang, P.H., Basu, S.K., Mir, F.A., Siddiqui, I., and Sharma, P. (2007) Mycobacterium tuberculosis 6-kDa early secreted antigenic target (ESAT-6) protein downregulates lipopolysaccharide induced c-myc expression by modulating the extracellular signal regulated kinases 1/2. BMC Immunol 8, 24 120. Jung, S.B., Yang, C.S., Lee, J.S., Shin, A.R., Jung, S.S., Son, J.W., Harding, C.V., Kim, H.J., Park, J.K., Paik, T.H., Song, C.H., and Jo, E.K. (2006) The mycobacterial 38-kilodalton glycolipoprotein antigen activates the mitogen- activated protein kinase pathway and release of proinflammatory cytokines \t\r \u00C2\u00A0 115 through Toll-like receptors 2 and 4 in human monocytes. Infection and immunity 74, 2686-2696 121. Kim, K.H., Yang, C.S., Shin, A.R., Jeon, S.R., Park, J.K., Kim, H.J., and Jo, E.K. (2011) Mycobacterial Heparin-binding Hemagglutinin Antigen Activates Inflammatory Responses through PI3-K/Akt, NF-kappaB, and MAPK Pathways. Immune Netw 11, 123-133 122. Saleh, M.T. and Belisle, J.T. (2000) Secretion of an acid phosphatase (SapM) by Mycobacterium tuberculosis that is similar to eukaryotic acid phosphatases. J Bacteriol 182, 6850-6853 123. Ehrlich, K.C., Montalbano, B.G., Mullaney, E.J., Dischinger, H.C., Jr., and Ullah, A.H. (1994) An acid phosphatase from Aspergillus ficuum has homology to Penicillium chrysogenum PhoA. Biochem Biophys Res Commun 204, 63-68 124. Ferminan, E. and Dominguez, A. (1997) The KIPHO5 gene encoding a repressible acid phosphatase in the yeast Kluyveromyces lactis: cloning, sequencing and transcriptional analysis of the gene, and purification and properties of the enzyme. Microbiology 143 ( Pt 8), 2615-2625 125. Haas, H., Redl, B., Friedlin, E., and Stoffler, G. (1992) Isolation and analysis of the Penicillium chrysogenum phoA gene encoding a secreted phosphate-repressible acid phosphatase. Gene 113, 129-133 126. Festjens, N., Bogaert, P., Batni, A., Houthuys, E., Plets, E., Vanderschaeghe, D., Laukens, B., Asselbergh, B., Parthoens, E., De Rycke, R., Willart, M.A., Jacques, P., Elewaut, D., Brouckaert, P., Lambrecht, B.N., Huygen, K., and Callewaert, N. (2011) Disruption of the SapM locus in Mycobacterium bovis BCG improves its protective efficacy as a vaccine against M. tuberculosis. EMBO Mol Med 3, 222-234 127. Madhurantakam, C., Rajakumara, E., Mazumdar, P.A., Saha, B., Mitra, D., Wiker, H.G., Sankaranarayanan, R., and Das, A.K. (2005) Crystal structure of low-molecular-weight protein tyrosine phosphatase from Mycobacterium tuberculosis at 1.9-A resolution. J Bacteriol 187, 2175-2181 128. Raugei, G., Ramponi, G., and Chiarugi, P. (2002) Low molecular weight protein tyrosine phosphatases: small, but smart. Cell Mol Life Sci 59, 941-949 \t\r \u00C2\u00A0 116 129. Ecco, G., Vernal, J., Razzera, G., Martins, P.A., Matiollo, C., and Terenzi, H. (2010) Mycobacterium tuberculosis tyrosine phosphatase A (PtpA) activity is modulated by S-nitrosylation. Chem Commun (Camb) 46, 7501-7503 130. Teitelbaum, R., Cammer, M., Maitland, M.L., Freitag, N.E., Condeelis, J., and Bloom, B.R. (1999) Mycobacterial infection of macrophages results in membrane-permeable phagosomes. Proc Natl Acad Sci U S A 96, 15190-15195 131. Carlsson, F., Kim, J., Dumitru, C., Barck, K.H., Carano, R.A., Sun, M., Diehl, L., and Brown, E.J. (2010) Host-detrimental role of Esx-1-mediated inflammasome activation in mycobacterial infection. PLoS Pathog 6, e1000895 132. Grundner, Cox, and Alber (2008) Protein tyrosine phosphatase PtpA is not required for Mycobacterium tuberculosis growth in mice. FEMS Microbiol Lett 287, 181-184 133. Rieder, S.E. and Emr, S.D. (1997) A novel RING finger protein complex essential for a late step in protein transport to the yeast vacuole. Mol Biol Cell 8, 2307-2327 134. 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., and 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. Mol Microbiol 52, 1691-1702 135. Frota, C.C., Papavinasasundaram, K.G., Davis, E.O., and Colston, M.J. (2004) The AraC family transcriptional regulator Rv1931c plays a role in the virulence of Mycobacterium tuberculosis. Infect Immun 72, 5483-5486 136. Springer, B., Sander, P., Sedlacek, L., Ellrott, K., and Bottger, E.C. (2001) Instability and site-specific excision of integration-proficient mycobacteriophage L5 plasmids: development of stably maintained integrative vectors. Int J Med Microbiol 290, 669-675 137. O'Hare, H., Juillerat, A., Dianiskova, P., and Johnsson, K. (2008) A split- protein sensor for studying protein-protein interaction in mycobacteria. J Microbiol Methods 73, 79-84 \t\r \u00C2\u00A0 117 138. Chao, J.D., Papavinasasundaram, K.G., Zheng, X., Chavez-Steenbock, A., Wang, X., Lee, G.Q., and Av-Gay, Y. (2010) Convergence of Ser/Thr and two- component signaling to coordinate expression of the dormancy regulon in Mycobacterium tuberculosis. J Biol Chem 285, 29239-29246 139. Iyer, S.S., Barton, J.A., Bourgoin, S., and Kusner, D.J. (2004) Phospholipases D1 and D2 coordinately regulate macrophage phagocytosis. J Immunol 173, 2615-2623 140. Vergne, I., Constant, P., and Laneelle, G. (1998) Phagosomal pH determination by dual fluorescence flow cytometry. Anal Biochem 255, 127-132 141. Bernardo, J., Long, H.J., and Simons, E.R. (2010) Initial cytoplasmic and phagosomal consequences of human neutrophil exposure to Staphylococcus epidermidis. Cytometry A 77, 243-252 142. Miksa, M., Komura, H., Wu, R., Shah, K.G., and Wang, P. (2009) A novel method to determine the engulfment of apoptotic cells by macrophages using pHrodo succinimidyl ester. J Immunol Methods 342, 71-77 143. Joshi, B., Kumar, G., Shankar, H., Bisht, D., Sharma, P., Singhal, N., and Katoch, V.M. (2010) A Simple and Rapid Method of Sample Preparation from Culture Filtrate of M. Tuberculosis for Two-Dimensional Gel Electrophoresis. Braz J Microbiol 41, 295-299 144. Av-Gay, Y., Jamil, S., and Drews, S.J. (1999) Expression and characterization of the Mycobacterium tuberculosis serine/threonine protein kinase PknB. Infect Immun 67, 5676-5682 145. Hildebrandt, E. and Fried, V.A. (1989) Phosphoamino acid analysis of protein immobilized on polyvinylidene difluoride membrane. Anal Biochem 177, 407-412 146. Sagermann, M., Stevens, T.H., and Matthews, B.W. (2001) Crystal structure of the regulatory subunit H of the V-type ATPase of Saccharomyces cerevisiae. Proc Natl Acad Sci USA 98, 7134-7139 147. Lesk, V.I. and Sternberg, M.J. (2008) 3D-Garden: a system for modelling protein-protein complexes based on conformational refinement of ensembles generated with the marching cubes algorithm. Bioinformatics 24, 1137-1144 \t\r \u00C2\u00A0 118 148. Hunter, T. (2009) Tyrosine phosphorylation: thirty years and counting. Curr Opin Cell Biol 21, 140-146 149. Zhao, X. and Lam, J.S. (2002) WaaP of Pseudomonas aeruginosa is a novel eukaryotic type protein-tyrosine kinase as well as a sugar kinase essential for the biosynthesis of core lipopolysaccharide. J Biol Chem 277, 4722-4730 150. Morona, J.K., Morona, R., Miller, D.C., and Paton, J.C. (2002) Streptococcus pneumoniae capsule biosynthesis protein CpsB is a novel manganese-dependent phosphotyrosine-protein phosphatase. J Bacteriol 184, 577-583 151. Matsumoto, A., Hong, S.K., Ishizuka, H., Horinouchi, S., and Beppu, T. (1994) Phosphorylation of the AfsR protein involved in secondary metabolism in Streptomyces species by a eukaryotic-type protein kinase. Gene 146, 47-56 152. Rubens, C.E., Heggen, L.M., Haft, R.F., and Wessels, M.R. (1993) Identification of cpsD, a gene essential for type III capsule expression in group B streptococci. Mol Microbiol 8, 843-855 153. Wu, J., Ohta, N., Zhao, J.L., and Newton, A. (1999) A novel bacterial tyrosine kinase essential for cell division and differentiation. Proc Natl Acad Sci U S A 96, 13068-13073 154. Sau, S., Bhasin, N., Wann, E.R., Lee, J.C., Foster, T.J., and Lee, C.Y. (1997) The Staphylococcus aureus allelic genetic loci for serotype 5 and 8 capsule expression contain the type-specific genes flanked by common genes. Microbiology 143 ( Pt 7), 2395-2405 155. Sau, S., Sun, J., and Lee, C.Y. (1997) Molecular characterization and transcriptional analysis of type 8 capsule genes in Staphylococcus aureus. J Bacteriol 179, 1614-1621 156. Manai, M. and Cozzone, A.J. (1982) Endogenous protein phosphorylation in Escherichia coli extracts. Biochem Biophys Res Commun 107, 981-988 157. Black, D.S. and Bliska, J.B. (1997) Identification of p130Cas as a substrate of Yersinia YopH (Yop51), a bacterial protein tyrosine phosphatase that \t\r \u00C2\u00A0 119 translocates into mammalian cells and targets focal adhesions. EMBO J 16, 2730-2744 158. Chow, K., Ng, D., Stokes, R., and Johnson, P. (1994) Protein tyrosine phosphorylation in Mycobacterium tuberculosis. FEMS Microbiol Lett 124, 203- 207 159. Av-Gay, Y. and Everett, M. (2000) The eukaryotic-like Ser/Thr protein kinases of Mycobacterium tuberculosis. Trends Microbiol 8, 238-244 160. Vincent, C., Doublet, P., Grangeasse, C., Vaganay, E., Cozzone, A.J., and Duclos, B. (1999) Cells of Escherichia coli contain a protein-tyrosine kinase, Wzc, and a phosphotyrosine-protein phosphatase, Wzb. J Bacteriol 181, 3472-3477 161. Burroughs, A.M., Allen, K.N., Dunaway-Mariano, D., and Aravind, L. (2006) Evolutionary genomics of the HAD superfamily: understanding the structural adaptations and catalytic diversity in a superfamily of phosphoesterases and allied enzymes. J Mol Biol 361, 1003-1034 162. Allen, K.N. and Dunaway-Mariano, D. (2004) Phosphoryl group transfer: evolution of a catalytic scaffold. Trends Biochem Sci 29, 495-503 163. Doublet, P., Vincent, C., Grangeasse, C., Cozzone, A.J., and Duclos, B. (1999) On the binding of ATP to the autophosphorylating protein, Ptk, of the bacterium Acinetobacter johnsonii. FEBS Lett 445, 137-143 164. Leonard, C.J., Aravind, L., and Koonin, E.V. (1998) Novel families of putative protein kinases in bacteria and archaea: evolution of the \"eukaryotic\" protein kinase superfamily. Genome Res 8, 1038-1047 165. Collet, J.F., Stroobant, V., Pirard, M., Delpierre, G., and Van Schaftingen, E. (1998) A new class of phosphotransferases phosphorylated on an aspartate residue in an amino-terminal DXDX(T/V) motif. J Biol Chem 273, 14107-14112 166. Saraste, M., Sibbald, P.R., and Wittinghofer, A. (1990) The P-loop-a common motif in ATP- and GTP-binding proteins. Trends Biochem Sci 15, 430- 434 \t\r \u00C2\u00A0 120 167. Bach, H., Sun, J., Hmama, Z., and Av-Gay, Y. (2006) Mycobacterium avium subsp. paratuberculosis PtpA is an endogenous tyrosine phosphatase secreted during infection. Infect Immun 74, 6540-6546 168. Mishra, B.B., Moura-Alves, P., Sonawane, A., Hacohen, N., Griffiths, G., Moita, L.F., and Anes, E. (2010) Mycobacterium tuberculosis protein ESAT-6 is a potent activator of the NLRP3/ASC inflammasome. Cell Microbiol 12, 1046-1063 169. Sullivan, J.T., Young, E.F., McCann, J.R., and Braunstein, M. (2012) The Mycobacterium tuberculosis SecA2 System Subverts Phagosome Maturation to Promote Growth in Macrophages. Infect Immun 170. Clemens, D.L. and Horwitz, M.A. (1996) The Mycobacterium tuberculosis phagosome interacts with early endosomes and is accessible to exogenously administered transferrin. J Exp Med 184, 1349-1355 171. Vergne, I., Fratti, R.A., Hill, P.J., Chua, J., Belisle, J., and Deretic, V. (2004) Mycobacterium tuberculosis phagosome maturation arrest: mycobacterial phosphatidylinositol analog phosphatidylinositol mannoside stimulates early endosomal fusion. Mol Biol Cell 15, 751-760 172. Wurmser, A.E., Sato, T.K., and Emr, S.D. (2000) New component of the vacuolar class C-Vps complex couples nucleotide exchange on the Ypt7 GTPase to SNARE-dependent docking and fusion. J Cell Biol 151, 551-562 173. Nickerson, D.P., Brett, C.L., and Merz, A.J. (2009) Vps-C complexes: gatekeepers of endolysosomal traffic. Curr Opin Cell Biol 21, 543-551 174. Takeda, K., Cabrera, M., Rohde, J., Bausch, D., Jensen, O.N., and Ungermann, C. (2008) The vacuolar V1/V0-ATPase is involved in the release of the HOPS subunit Vps41 from vacuoles, vacuole fragmentation and fusion. FEBS Lett 582, 1558-1563 175. Hanks, S.K., Quinn, A., and Hunter, T. (1988) The protein kinase family: conserved features and deduced phylogeny of the catalytic domains. Science 241, 42-52 176. Hunter, T. (1987) A thousand and one protein kinases. Cell 50, 823-829 \t\r \u00C2\u00A0 121 177. Thomasson, B., Link, J., Stassinopoulos, A.G., Burke, N., Plamann, L., and Hartzell, P.L. (2002) The GTPase, MglA, interacts with a tyrosine kinase to control type-IV pili-mediated motility of Myxococcus xanthus. Mol Microbiol 46, 1399-1413 178. Zhao, X. and Lam, J.S. (2002) WaaP pf Pseudomonas aeruginosa is a novel eukaryotic type protein-tyrosine kinase as well as a sugar kinase essential for the biosynthesis of core lipopolysaccharide. J Biol Chem 277, 4722-4730 179. Morais, M.C., Zhang, W., Baker, A.S., Zhang, G., Dunaway-Mariano, D., and Allen, K.N. (2000) The crystal structure of bacillus cereus phosphonoacetaldehyde hydrolase: insight into catalysis of phosphorus bond cleavage and catalytic diversification within the HAD enzyme superfamily. Biochemistry 39, 10385-10396 180. Talaat, A.M., Lyons, R., Howard, S.T., and Johnston, S.A. (2004) The temporal expression profile of Mycobacterium tuberculosis infection in mice. Proc Natl Acad Sci U S A 101, 4602-4607 181. Srivastava, V., Rouanet, C., Srivastava, R., Ramalingam, B., Locht, C., and Srivastava, B.S. (2007) Macrophage-specific Mycobacterium tuberculosis genes: identification by green fluorescent protein and kanamycin resistance selection. Microbiology 153, 659-666 182. Tailor, P., Gilman, J., Williams, S., Couture, C., and Mustelin, T. (1997) Regulation of the low molecular weight phosphotyrosine phosphatase by phosphorylation at tyrosines 131 and 132. J Biol Chem 272, 5371-5374 183. Prisic, S., Dankwa, S., Schwartz, D., Chou, M.F., Locasale, J.W., Kang, C.M., Bemis, G., Church, G.M., Steen, H., and Husson, R.N. (2010) Extensive phosphorylation with overlapping specificity by Mycobacterium tuberculosis serine/threonine protein kinases. Proc Natl Acad Sci U S A 107, 7521-7526 184. Noren-Muller, A., Wilk, W., Saxena, K., Schwalbe, H., Kaiser, M., and Waldmann, H. (2008) Discovery of a new class of inhibitors of Mycobacterium tuberculosis protein tyrosine phosphatase B by biology-oriented synthesis. Angew Chem Int Ed Engl 47, 5973-5977 185. Benson, R.E., Sanfridson, A., Ottinger, J.S., Doyle, C., and Cullen, B.R. (1993) Downregulation of cell-surface CD4 expression by simian \t\r \u00C2\u00A0 122 immunodeficiency virus Nef prevents viral super infection. J Exp Med 177, 1561- 1566 186. Schwartz, O., Marechal, V., Le Gall, S., Lemonnier, F., and Heard, J.M. (1996) Endocytosis of major histocompatibility complex class I molecules is induced by the HIV-1 Nef protein. Nat Med 2, 338-342 187. Lama, J., Mangasarian, A., and Trono, D. (1999) Cell-surface expression of CD4 reduces HIV-1 infectivity by blocking Env incorporation in a Nef- and Vpu- inhibitable manner. Curr Biol 9, 622-631 188. Ross, T.M., Oran, A.E., and Cullen, B.R. (1999) Inhibition of HIV-1 progeny virion release by cell-surface CD4 is relieved by expression of the viral Nef protein. Curr Biol 9, 613-621 189. Geyer, M., Fackler, O.T., and Peterlin, B.M. (2002) Subunit H of the V- ATPase involved in endocytosis shows homology to beta-adaptins. Mol Biol Cell 13, 2045-2056 190. Dasgupta, A., Datta, P., Kundu, M., and Basu, J. (2006) The serine/threonine kinase PknB of Mycobacterium tuberculosis phosphorylates PBPA, a penicillin-binding protein required for cell division. Microbiology 152, 493-504 191. Molle, Kremer, Girard-Blanc, Besra, Cozzone, and Prost (2003) An FHA phosphoprotein recognition domain mediates protein EmbR phosphorylation by PknH, a Ser/Thr protein kinase from Mycobacterium tuberculosis. Biochemistry 42, 15300-15309 192. Molle, Soulat, Jault, Grangeasse, Cozzone, and Prost (2004) Two FHA domains on an ABC transporter, Rv1747, mediate its phosphorylation by PknF, a Ser/Thr protein kinase from Mycobacterium tuberculosis. FEMS Microbiol Lett 234, 215-223 193. Perez, Garcia, Bach, Waard, d., Jacobs, Av-Gay, Bubis, and Takiff (2006) Mycobacterium tuberculosis transporter MmpL7 is a potential substrate for kinase PknD. Biochem Biophys Res Commun 348, 6-12 194. Sharma, K., Gupta, M., Krupa, A., Srinivasan, N., and Singh, Y. (2006) EmbR, a regulatory protein with ATPase activity, is a substrate of multiple \t\r \u00C2\u00A0 123 serine/threonine kinases and phosphatase in Mycobacterium tuberculosis. FEBS J 273, 2711-2721 195. Thakur, M. and Chakraborti, P.K. (2006) GTPase activity of mycobacterial FtsZ is impaired due to its transphosphorylation by the eukaryotic-type Ser/Thr kinase, PknA. J Biol Chem 281, 40107-40113 196. Zheng, Papavinasasundaram, and Av-Gay (2007) Novel substrates of Mycobacterium tuberculosis PknH Ser/Thr kinase. Biochem Biophys Res Commun 355, 162-168 197. Sassetti, C.M., Boyd, D.H., and Rubin, E.J. (2003) Genes required for mycobacterial growth defined by high density mutagenesis. Mol Microbiol 48, 77- 84 \t\r \u00C2\u00A0 124 APPENDIX A: SUPPLEMENTAL FIGURES FOR SECTION 3.1 \t\r \u00C2\u00A0 Figure 28. In vitro analysis of subunit H and PtpA interaction. (A) Recombinant PtpA pulled down V-ATPase subunit H from THP-1 cell lysate. Recombinant His-tagged PtpA proteins were used as a bait to pull down target proteins from THP-1 human macrophages lysates. Lane 1: THP-1 cell lysate alone. Lane 2: THP-1 cell lysate incubated with His-tagged wild type PtpA protein. PtpA pulled down V-ATPase subunit H (blue circle). MALDI-TOF mass spectrometry was used to analyze the captured proteins (1). (B) Phosphorylated subunit H is not a catalytic substrate of PtpA. (Top Panel) PhosphorImage of subunit H recombinant proteins phosphorylated with \u00CE\u00B3[32P]ATP and incubated with increasing concentration of PtpA recombinant proteins. Lane 1. Subunit H \t\r \u00C2\u00A0 125 alone. Lane 2. PtpA alone. Lane 3. Subunit H and PtpA (1:1 molar ratio). Lane 4. Subunit H and PtpA (1:2 molar ratio). Lane 5. Subunit H and PtpA (1:3 molar ratio). (Bottom Panel) Silver stained SDS-PAGE of the corresponding gel. (C) Phosphoamino acid analysis shows subunit H is Thr-phosphorylated. Phosphoamino acid analysis was performed as described previously (1). Retention factor (Rf): Phospho-serine (P-Ser), 0.44; Phospho-threonine (P-Thr), 0.49; Phospho-tyrosine (P-Tyr), 0.55; Subunit H, 0.50. \t\r \u00C2\u00A0 126 Figure 29. PtpAL146A is a phosphatase-active mutant defective in binding subunit H. (A) Site-directed mutagenesis of di-leucine motifs and leucine residues in PtpA. Highlighted amino acids are the residues mutated and tested. L146A mutation creates a catalytically active mutant PtpA that does not interact with subunit H of the V-ATPase. (B) PtpAL146A remains catalytically active with similar kinetics as the wild type enzyme despite point mutation on its C-terminal alpha helix. p-nitrophenyl phosphate (pNPP) is used as a chromogenic substrate to detect phosphatase activity of PtpA. The reactions were analyzed in a spectrophotometer for absorbance at 450 nm. \t\r \u00C2\u00A0 127 Figure 30. Western blot analysis of PtpA expression in complemented \u00CE\u0094ptpA strain and in vitro PtpA secretion using \u00CE\u00B1-PtpA antibodies. (A) A weak band corresponding to PtpA expression is observed for the parental H37Rv as expected. \u00CE\u0094ptpA strains complemented with wild type ptpA (COMP), ptpAD126A or ptpAL146A shows clear expression of PtpA proteins. Recombinant PtpA was used as a positive control. No expression was observed for the \u00CE\u0094ptpA strain. (B) Covalent labeling of the bacteria with pHrodo and Alexa-Fluor 488 does not affect PtpA secretion as shown by the prominent bands corresponding to His- tagged PtpA in the culture filtrate. Soluble lysates from the bacteria were analyzed to ensure expression of His-tagged PtpA. Recombinant PtpA was used as a positive marker control. \t\r \u00C2\u00A0 128 Figure 31. Calibration of THP-1 phagosomal pH during Mtb infection with FACS. (A) Calibration curve of phagosomal pH in THP-1 macrophage-like cells infected with Mtb dual-labeled with pHrodo and Alexa Fluor 488. The fluorescence intensity ratios of the fluorescence probes were calculated and plotted for each pH. (B) Overlaid FACS histograms of the pHrodo fluorescence intensities of Mtb phagosomes at pre-calibrated pH (5.0 - 8.0). (C-F) Overlaid FACS histograms of pHrodo fluorescence intensities of THP-1 infected with the \t\r \u00C2\u00A0 129 indicated Mtb strains. (C) Parental H37Rv. (D) \u00CE\u0094ptpA complemented with the wild type ptpA gene. (E) \u00CE\u0094ptpA complemented with the mutant ptpAD126A gene. (F) \u00CE\u0094ptpA complemented with the mutant ptpAL146A gene. \t\r \u00C2\u00A0 130 Figure 32. Measurement of phagosomal pH in transfected THP-1 infected with E. coli. (A) Expression of GFP-tagged wild type and mutant PtpA proteins were assessed by measurement of GFP fluorescence (Emission 535 nm) in transfected THP-1 macrophage-like cells. Similar levels of expression were found for the wild type and mutant proteins. Untransfected THP-1 serves as a negative control, and THP-1 transfected with the pEGFP vector was used as a positive control. (B) Phagocytosis of E. coli labeled with Alexa Fluor 350 by THP-1 was not affected by transfection and expression of GFP-tagged wild type and mutant PtpA proteins. There was no difference in the level of phagocytosis of THP-1 transfected cells with untransfected cells as shown by the similar levels of Alexa Fluor 350 fluorescence (Emission 460 nm). (C) Calibration of THP-1 phagosomal pH during E. coli infection with spectrofluorometry. The fluorescence ratios of pHrodo (Emission 620 nm) and Alexa Fluor 350 (Emission 460 nm) were plotted for each pH to generate the calibration curve of phagosomal pH in THP-1 \t\r \u00C2\u00A0 131 macrophage-like cells infected with E. coli dual-labeled with the pH-sensitive pHrodo and pH-insensitive Alexa Fluor 350 dyes. (D) Separate digital confocal microscopy fluorescent images of transfected THP-1 macrophage-like cells infected with dual-labeled E. coli. Green: Expression of GFP or GFP-tagged PtpA constructs. Red: pHrodo fluorescence in acidified phagosomes. Blue: Alexa Fluor 350 labeled E. coli. Localization of GFP-PtpA to the phagosomes can be observed. \t\r \u00C2\u00A0 132 \t\r \u00C2\u00A0 Figure 33. Analysis of antibodies specificity. (A) FACS analysis of the ability of indicated antibodies to non-specifically recognize extracellular components of Mtb. None of the tested antibodies showed significant binding to the bacteria. Rabbit anti-LqpH (19 kDa lipoprotein that is known to be present on the Mtb cell \t\r \u00C2\u00A0 133 wall) was used as a positive control. (B) Western blot analysis of antibodies specificity with soluble lysates from E. coli DH5\u00CE\u00B1, Mtb H37Rv, THP-1, THP-1 infected with E. coli DH5\u00CE\u00B1 (THP-1\u00CE\u00B1) and THP-1 infected with Mtb H37Rv (THP- 1Rv). 25 \u00C2\u00B5g of each lysate was resolved on SDS-PAGE and probed with the indicated antibodies. The tested antibodies showed specific detection of the target proteins with minimal non-specific background signal. \t\r \u00C2\u00A0 134 Figure 34. Immunostaining control experiment with Mtb-infected THP-1. Immunostaining control experiment was performed to examine potential cross- reactivity among the secondary antibodies. Rabbit anti-VPS33B (R\u00CE\u00B1-VPS33B) and Mouse anti-subunit H (M\u00CE\u00B1-H) were used as primary antibodies. Texas Red- conjugated goat-anti rabbit IgG and Alexa-Fluor 488-conjugated goat-anti mouse IgG were used as secondary antibodes. Fluorescent signal was only detected when the correct pair of primary and secondary antibodies was used. The secondary antibodies do not react with primary antibodies from another species, demonstrating the specificity of the secondary antibodies for double immunofluorescence staining. \t\r \u00C2\u00A0 135 Figure 35. Immunostaining control experiment with E. coli-infected THP-1. Immunostaining control experiment was performed to examine potential cross- reactivity among the secondary antibodies. Rabbit anti-VPS33B (R\u00CE\u00B1-VPS33B) and Mouse anti-subunit H (M\u00CE\u00B1-H) were used as primary antibodies. Texas Red- conjugated goat-anti rabbit IgG and Alexa-Fluor 488-conjugated goat-anti mouse IgG were used as secondary antibodes. Fluorescent signal was only detected when the correct pair of primary and secondary antibodies was used. The secondary antibodies do not react with primary antibodies from another species, demonstrating the specificity of the secondary antibodies for double immunofluorescence staining. "@en . "Thesis/Dissertation"@en . "2012-11"@en . "10.14288/1.0072795"@en . "eng"@en . "Experimental Medicine"@en . "Vancouver : University of British Columbia Library"@en . "University of British Columbia"@en . "Attribution-NonCommercial-NoDerivs 3.0 Unported"@en . "http://creativecommons.org/licenses/by-nc-nd/3.0/"@en . "Graduate"@en . "Role of Mycobacterium tuberculosis protein tyrosine phosphatase A in the pathogenesis of tuberculosis"@en . "Text"@en . "http://hdl.handle.net/2429/42331"@en .