Free access
Review
1 July 2006

Tuberculosis Chemotherapy: the Influence of Bacillary Stress and Damage Response Pathways on Drug Efficacy

SUMMARY

The global tuberculosis (TB) control effort is focused on interrupting transmission of the causative agent, Mycobacterium tuberculosis, through chemotherapeutic intervention in active infectious disease. The insufficiency of this approach is manifest in the inexorable annual increase in TB infection and mortality rates and the emergence of multidrug-resistant isolates. Critically, the limited efficacy of the current frontline anti-TB drug combination suggests that heterogeneity of host and bacillary physiologies might impair drug activity. This review explores the possibility that strategies enabling adaptation of M. tuberculosis to hostile in vivo conditions might contribute to the subversion of anti-TB chemotherapy. In particular, evidence that infecting bacilli are exposed to environmental and host immune-mediated DNA-damaging insults suggests a role for error-prone DNA repair synthesis in the generation of chromosomally encoded antibiotic resistance mutations. The failure of frontline anti-TB drugs to sterilize a population of susceptible bacilli is independent of genetic resistance, however, and instead implies the operation of alternative tolerance mechanisms. Specifically, it is proposed that the emergence of persister subpopulations might depend on the switch to an altered metabolic state mediated by the stringent response alarmone, (p)ppGpp, possibly involving some or all of the many toxin-antitoxin modules identified in the M. tuberculosis genome.

INTRODUCTION

Tuberculosis (TB) was declared a global health emergency by the World Health Organization in 1993 and currently claims approximately 1.7 million lives per annum, more than can be attributed to any other bacterial infection (252). It is estimated that one-third of the world's population is infected with the causative agent, the obligate human pathogen Mycobacterium tuberculosis (76), with around 9 to 10 million new cases of TB being reported each year (252). Approximately 90 to 95% of initial infections are controlled by the cell-mediated immune response. However, TB immunity is static (171), and a residual population of viable bacteria may be maintained in a poorly understood state of clinical latency for extended periods (141). Approximately 5 to 10% of cases overall are thought to result from spontaneous reactivation of latent TB infection (219). The massive reservoir of viable bacteria in the estimated 2 billion asymptomatically infected individuals worldwide is, therefore, of supreme importance for the epidemiology and control of TB. However, the potential for bacterial populations to occupy discrete lesions in a single host complicates targeting of the intractable pathogen (22). Furthermore, recent years have witnessed the lethal synergy between M. tuberculosis and the human immunodeficiency virus (HIV) (55) as well as an increasing emergence of multidrug-resistant (MDR) strains (251).
Infection with an M. tuberculosis strain that is resistant to the two most commonly used frontline anti-TB drugs, isoniazid (INH) and rifampin, is defined as MDR TB and often culminates in incurable disease (109). Treatment of MDR TB is resource intensive, and the therapeutic strategies recommended for high-prevalence areas (81, 103) comprise combinations of second-line drugs that are more expensive, more toxic, and less effective than the drugs used in standard therapy (114). MDR strains constitute 1 to 3% of global TB isolates, and although worldwide distribution is characterized by localized “hot zones” (82, 183), MDR TB has been identified as a significant problem in every region under World Health Organization surveillance (251). Approximately 300,000 new cases of MDR TB emerge worldwide each year, with the most common pathway to multidrug resistance initiating with monoresistance to either INH or streptomycin (251). Furthermore, the prevalence of strains resistant to these single drugs correlates with those areas exhibiting the highest levels of MDR. Significantly, “super strains” resistant to at least three of the four frontline TB drugs make up 79% of all MDR TB cases.

CURRENT ANTI-TB CHEMOTHERAPY

The current standard chemotherapeutic regimen for active TB—directly observed therapy, short course (DOTS)—requires the supervised administration of a multidrug combination for a minimum period of 6 months. The frontline drugs are biased towards interference in cell wall integrity, with the actively dividing M. tuberculosis population in the lung cavity, which is key to transmission and emergence of drug resistance (101, 122), being the principal target. M. tuberculosis replication in pulmonary cavities most closely resembles optimal aerobic growth in vitro, and the effectiveness of the frontline drugs in treating acute TB is manifest in rapid bacillary clearance within the first 2 months of chemotherapy. Differences in drug susceptibilities between replicating and nondividing bacterial cells can be significant (112, 254), however, and the dependence of the frontline anti-TB drugs on actively replicating cells for activity is probably the greatest limitation of current therapy (119). This weakness is further reflected in the profound disparity between in vitro and in vivo efficacies of the majority of the frontline drugs (155), an additional factor dictating the complexity and duration of the DOTS regimen.
Poor antibiotic penetration, heterogeneity of host environments, and altered bacterial physiology and metabolic activity within those environments have all been blamed for impaired drug efficacies in vivo. However, the influence of the in vivo environment on drug efficacy should not be viewed as inevitably negative: pyrazinamide (PZA), for example, is inactive in vitro under standard culture conditions but displays strong sterilizing activity in vivo (257, 259). Moreover, the activity of PZA in vivo correlates with enhanced in vitro activity under low pH (257) and limiting oxygen (236), suggesting the relevance of these factors to the in vivo environment (discussed below). The absolute dependence on environmental factors for PZA activity has profound implications for the discovery of new anti-TB drugs, since it would almost certainly render this drug unidentifiable according to standard drug screening criteria, thereby eliminating a mainstay of current TB chemotherapy. At the very least, the differential effect of the in vivo environment on PZA versus the other frontline drugs emphasizes the need to understand the in vivo lifestyle of M. tuberculosis so that the factors influencing drug efficacy can be determined and new drug targets relevant to both latent and active disease can be identified.

BACILLARY SURVIVAL TACTICS

In the presence of a functional adaptive immune system, TB bacilli are sequestered in granulomas comprising differentiated macrophages and other immune cells maintained in complex structures by extracellular matrix components (58). Granulomas are thought to be characterized by hypoxia, low pH, and nutrient deprivation and are likely awash in inhibitory organic acids (145, 176, 177). Furthermore, the apparently static balance established between the host immune system and the pathogen's resistance (166, 196) suggests that bacilli might endure continual exposure to immune effectors, including reactive oxygen intermediates (ROI) and reactive nitrogen intermediates (RNI) (38, 131, 147, 169, 174, 239). Unfortunately, there is currently a critical lack of a tractable and inexpensive animal model which replicates all aspects of human TB (22, 125). However, despite the limitations, wild-type and mutant M. tuberculosis strains have been widely employed as bioprobes in the various animal models and in vitro to refine our understanding of the host environment as well as the mechanisms enabling long-term survival of M. tuberculosis.
Broad concepts which have emerged include the notion that persisting bacilli are metabolically active (26, 113, 209) and might adapt sufficiently to long-term nonreplication to reinitiate cellular division (106). The stimuli for entry into a state of nonreplicating persistence have been explored by applying stresses such as nutrient starvation (14, 60), oxygen limitation (21, 244), exposure to low-dose NO (234), and adaptation to stationary phase (106, 235). Analysis of gene expression profiles in response to these stresses, as well as phenotypic and physiological characterization, has largely validated the relevance of in vitro stresses to in vivo conditions. In addition, selective in situ analysis of RNA expression levels in murine and human lung tissue has corroborated key findings of the in vitro models (83, 213, 223, 227). Critically, the differential expression of selected M. tuberculosis genes in lung tissue of human TB patients has shown that M. tuberculosis possesses the capacity to respond to specific and variable microenvironments in vivo, and it might adopt a variety of metabolic states depending on host immune status and stage of infection (83, 122, 227). Some of the possible metabolic adaptations are considered briefly below, with emphasis on the types of DNA damage likely to be incurred by the bacilli during their occupation of granulomas. Mechanisms that have evolved to minimize or repair that damage, as well as the potential physiological and mutagenic consequences, are considered in subsequent sections, with particular reference to the emergence of drug-resistant clinical isolates.

THE IN VIVO ENVIRONMENT ENCOUNTERED BY M. TUBERCULOSIS

Persistence of M. tuberculosis hinges on its ability to survive (and replicate) for extended periods in pathogen-friendly phagosomes. After phagocytic uptake, M. tuberculosis interrupts the normal maturation pathway (48), preventing phagosome-lysosome fusion (168) and ensuring limited phagosomal acidification and a phagosomal compartment deficient in mature lysosomal hydrolases (201). Arrest of phagosomal maturation is a complex process which depends on the continual modulation of host cell trafficking events (48). Even where maturation is stalled, however, bacilli are exposed to a slight decrease in phagosomal pH (69). Notably, the transcriptional response of M. tuberculosis to acid stress in vitro (84) includes genes identified as being necessary for survival in vivo (205). In addition, several genes potentially involved in fatty acid metabolism are common to both the intraphagosomal and acid response transcriptomes of M. tuberculosis (74, 84, 207). Evidence of the dependence of PZA on acidic conditions for activity in vitro further reinforces the idea that bacilli are confronted by low pH during active disease (257, 259).
The notion that M. tuberculosis encounters environments of limiting oxygen availability during the course of human infection (62) has motivated efforts to characterize the physiological and metabolic changes associated with adaptation to hypoxia (242, 244). In particular, recent studies have identified a prominent role for the regulon controlled by the DosR-DosS/DosT two-component regulator system (179, 207, 211, 234). The DosR regulon includes genes associated with anaerobic metabolism and stabilization of cellular components and is induced in response to hypoxia and nontoxic NO concentrations (126, 179, 185, 199, 211, 234). Unfortunately, the induction of DosR in response to both hypoxia and NO exposure has complicated any interpretation of the specific contribution of each to the host environment. The identification of key features differentiating the respective transcriptomes (25) does, however, suggest that disruptions in respiratory pathways might induce specific regulons (22). Evidence of the induction of DosR regulon genes in vitro in response to starvation (106) and other stresses (126), for example, is consistent with the presence of independent signaling pathways and raises the intriguing possibility that dormancy genes might be induced by other stimuli in vivo. Whatever the stimulus, the upregulation of DosR in various models of TB infection (123, 179, 185, 199, 207, 211, 213, 234) is, nevertheless, suggestive of its relevance to the adaptation of bacilli to persistence in humans.
M. tuberculosis is also thought to be deprived of nutrients and essential elements during occupation of the phagosome and in granulomas. Recent in vitro models of starvation have identified broad metabolic adjustments that are likely to apply during persistence in vivo (14, 60, 106) and which reinforce the suggestion that glucose deficiency and an abundance of fatty acids characterize the phagosome of activated macrophages (208). Genes involved in β-oxidation, the glyoxylate shunt, gluconeogenesis, amino acid/amine degradation, RNA synthesis and modification, and transcription are induced in response to nutrient stress, while carbon-degradative pathways, de novo ATP generation, and purine/pyrimidine synthesis are downregulated. In addition, evidence suggests that sulfur (106) and iron (72, 97, 207, 227) are likely to become limiting, both of which might be required for the maintenance of redox balance. Finally, several genes associated with sodium dodecyl sulfate treatment and cell wall maintenance are upregulated in resting macrophages (207), suggesting that M. tuberculosis might also sustain damage to surface structures.

DNA Damage In Vivo

M. tuberculosis is expected to sustain a variety of potentially DNA-damaging assaults in vivo (24, 162), primarily from host-generated antimicrobial ROI and RNI (1, 3, 146, 171, 198). DNA is a biological target for RNI and ROI (33, 262), and interaction with toxic radicals is mutagenic (262). Furthermore, damage to cellular components required for the protection or propagation of DNA can indirectly affect chromosomal integrity, while detoxification reactions might themselves yield endogenous damaging adducts (171). Additional endogenous reactive intermediates are also likely to be generated by the switch between aerobic and anaerobic metabolism and from the partial reduction of terminal electron acceptors during respiration (22).
Exposure of M. tuberculosis to potent oxidative agents in vitro results in minimal differential gene expression (23, 72, 89, 133, 212), however, and has been attributed to the impaired ability of the organism to mount a coherent oxidative stress response as a result of a natural deficiency in soxRS and oxyR regulons (70, 73, 89, 212). It is possible, though, that loss of oxidative stress response regulation might confer a selective advantage by allowing constitutive expression of detoxification genes (104)—for example, the peroxiredoxin AhpC (30, 43, 207, 217), superoxide dismutase (107, 245), thioredoxin (194, 247), and the KatG catalase-peroxidase (150, 153, 173) are expressed in M. tuberculosis in the absence of the central regulators. An extension of this hypothesis holds that the selective inactivation of the canonical oxidative stress response in M. tuberculosis might facilitate immune avoidance by preventing the potentially detrimental expression of immunogens during certain stages of the infection process (256). There is also the likelihood that, as in other organisms, some of the mechanisms implicated in antioxidant and antinitrosative defense function primarily to enable oxidative metabolism or in homeostatic signaling networks but fulfill a vital role in tolerance of host immune effectors (170, 256).
That the in vivo environment is DNA damaging is supported by several lines of evidence (23, 205, 207). Notably, damage repair pathways are essential for the virulence and survival of other intracellular pathogens (31). However, inactivation of alkylation repair and reversal (Rv1317c-ogt) (75) or recombinational repair (recA) (203, 204) pathways does not attenuate in vivo survival in mice, perhaps indicating that nitrosative or oxidative stresses do not induce cytotoxic DNA damage in the murine model. Alternatively, other repair pathways may function preferentially in mycobacteria; for example, a Ku-ligase system for double-strand break repair by nonhomologous end joining was recently characterized in M. tuberculosis (99) and might provide a possible alternative to homologous recombination for repair of such lesions (163). In addition, M. tuberculosis contains homologues of a putative DNA repair system that is highly conserved in thermophilic archaea and predicted to function in translesion synthesis (149). Several base excision repair enzymes have also been identified that are required for growth in vivo (205) but not in vitro under optimal conditions (206), implying a role in virulence. Furthermore, mycobacteria possess several conserved Fpg/Nei family DNA glycosylases (162), although the in vivo role of these homologues in base excision repair remains uncertain. Resistance to nitrosative stress in Mycobacterium smegmatis is dependent on a functional uracil DNA glycosylase (231). In addition, mutations in the uvrB-encoded excinuclease subunit result in elevated RNI susceptibility in vitro and reduced capacity to resist ROI and RNI in vivo (63). Together, these observations are suggestive of the increased importance of excision (base and nucleotide) repair pathways in mycobacteria (163). Moreover, the upregulation of uvrB and several other DNA repair genes in response to only certain damaging agents indicates that the type of lesion might determine the repair mechanism invoked (23).
Other mechanisms implicated in resistance against oxidative and nitrosative stress include the DosR-regulated α-crystallin (30, 77, 90, 182, 200, 211), while some of the physical properties characteristic of mycobacteria have also been implicated in the subversion of toxic ROI (13, 39, 40, 41, 87). M. tuberculosis produces high levels of mycothiol (172), the principal mycobacterial thiol that has been associated with resistance to various toxic species, including oxidants and frontline anti-TB drugs (32). The observation that mycobacteria in a state of low-oxygen-induced nonreplicating persistence are highly resistant to mitomycin C (188) also suggests that some protection of chromosomal DNA is provided by the NO- or hypoxia-mediated induction of M. tuberculosis uspA-like genes as part of the DosR regulon (179, 185, 211, 234). ROI and RNI inactivate proteins by targeting key residues, and a role for the M. tuberculosis peptide methionine sulfoxide reductase in the reversal of oxidative damage methionine residues has been demonstrated (218). Finally, it is possible that small biological molecules such as cysteine, methionine, and tyrosine, which interact with various reactive oxygen and nitrogen species, might function as biological traps for reactive intermediates; however, their relevance to mycobacterial oxidative and nitrosative defense is unknown (171).

GENETIC DRUG RESISTANCE

All Resistance Determinants in M. tuberculosis Are Chromosomally Encoded

The preferential occupation of specific environments within the host restricts the opportunities for the acquisition of novel, transmissible genetic elements by M. tuberculosis. Furthermore, recent evidence indicates that M. tuberculosis infections are clonal (105, 240), and genome analyses have revealed only minor roles for horizontal gene transfer in the macroevolution of clinical and laboratory mycobacterial strains (53, 54, 85, 111, 124, 229). M. tuberculosis does not possess epigenetic information in the form of plasmids, and no evidence exists for its natural competence (68). Conjugal DNA transfer has been demonstrated in M. smegmatis (186) and is mediated by the specialized, RD1-encoded secretory apparatus (86). However, the relevance of conjugal DNA transfer in M. tuberculosis and its implications for genetic variability and drug resistance have yet to be established. Consistent with the genetic isolation of M. tuberculosis, all drug resistance determinants are chromosomally encoded (167, 195), arising exclusively through the acquisition and maintenance of spontaneous chromosomal mutations in target or complementary genes or, rarely, from the inactivation of a target gene by a mobile genetic element (61, 134).

Emergence and Costs of Drug Resistance

For other pathogens, extended antibiotic chemotherapy has been implicated in the evolution of multiple-antibiotic resistance (143), an association likely to be replicated during M. tuberculosis infections. The long duration of the DOTS regimen, together with the toxic side effects of the frontline drugs and the temptation to cease therapy as symptoms subside, often leads to patient noncompliance (161). Furthermore, even where treatment schedules are adhered to, limited efficacy of one or more drugs can compromise combination therapy (142); recent epidemiological evidence suggests that monotherapy, effective or actual, is common (251).
M. tuberculosis has a long generation time and can adapt to prevailing growth conditions through a regulated shift to an alternative metabolic state (60, 233, 241). These factors are considered key to the ability of the organism to establish latent asymptomatic infection, but they might also result in the selective activity of specific antibiotics against discrete subpopulations (Global Alliance for TB Drug Development [http://www.tballiance.org ]). In addition, there is compelling evidence that infecting populations occupy diverse microenvironments within the host (34, 35, 83, 120, 122), some of which might be recalcitrant to antibiotic penetration or refractory to activity (78, 115). Significantly, the locally effective antibiotic concentration has been implicated in the evolution of low-level resistant variants during infection with other pathogens (11) and has been identified as an important determinant of mutation rate. Furthermore, different mutation rates and genotypes are thought to arise at intervals along a spectrum of applied concentrations (128, 260), a possibility with profound implications for the generation of diverse microbial populations in a single host (10). The parallel evolution of a single founder population into heterogeneous, antibiotic-resistant subpopulations within isolated loci has been demonstrated in patients undergoing active treatment for TB (122), for example.
The attachment of a fitness cost to resistance mutations (6) has led to the assumption that removal of antibiotic selective pressure will favor reversion as a result of a competitive replicative disadvantage. However, there is evidence that evolution in the absence of the selective antibiotic preferentially results in the acquisition of compensatory mutations that ameliorate the cost of resistance (5, 135, 136), rather than reversion. That is, the fitness cost more likely determines the stability and potential reversibility of the associated resistance mutation, with the ability to compensate genetically dictating the frequency of resistant mutants within a population. While the most fit mutants will be selected in a large population, lower-fitness, compensated mutants might become fixed during bottlenecks if they are formed at a higher rate than fitter, susceptible revertants (136, 148), particularly where genetic linkage exists between selected and nonselected resistance markers (80).
Information on the relative fitness of MDR TB isolates is limited (65); however, there is evidence that compensatory mutations can restore reproductive potential in monoresistant M. tuberculosis strains (210, 215). In addition, while drug-resistant M. tuberculosis strains more frequently possess low- rather than high-cost mutations (202), studies investigating the effects of resistance on virulence (16, 159, 181, 192, 202) have failed to establish a direct correlation (50). Instead, relative fitness in vitro appears to depend not only on the particular resistance mutation but also on the specific assay (151). Of course, there is the possibility that a resistance mutation might affect the ability of the pathogen to interact with the host environment and so might remain undetected in vitro (5). Mutations impairing virulence, for example, such as deletion of katG (140), will not survive selection in areas of high transmission (51).
The effects of resistance mutations on the fitness of M. tuberculosis are crucial to epidemiological predictions of the spread of MDR isolates (50). This concept has been further refined by recent evidence from mathematical models which suggests that, provided the relative fitness of an MDR strain remains above a defined threshold, a subpopulation of the low-fitness MDR strain will outcompete both the drug-sensitive strains and other, less fit MDR strains when confronted by a functioning TB control program (19, 49). As a result, the distribution of fitness (49) among circulating M. tuberculosis strains might be considered a more accurate predictive measure of resistance emergence. This, in turn, has led to the proposal that DOTS regimens be supplemented with anti-MDR strategies to limit resistance amplification, as well as further transmission of MDR strains (19, 49).
Coincident MDR TB prevalences and HIV infection rates (252) add a further degree of complexity and are suggestive of a positive correlation between MDR and HIV seropositivity. Although the identification of HIV as an independent risk factor for MDR TB is contentious (251), characteristic features of HIV/TB-associated clinical disease might favor the emergence of resistance. It has been suggested, for example, that disease outcomes in the treated TB patient are determined by a combination of host defense mechanisms and antimicrobial activity (98). That is, a functional immune system might be required to potentiate drug activity, thus preventing the evolution of resistance. Although direct evidence is scant, the enhanced sterilizing activity of PZA in vivo, in contrast with its poor activity in vitro (257), is suggestive of a synergistic interplay between drug and host (236).
The large bacterial populations associated with M. tuberculosis-infected immunocompromised individuals might provide an expanded subset for selection and transmission of rare mutation events. Furthermore, it has been suggested that the absence of a functioning immune response in those individuals might exacerbate the conditions implicated in the exposure of bacteria to monotherapy in immunocompetent patients (93); for example, uncontrolled replication and dissemination could produce drug-inaccessible compartments, while drug absorption might be compromised by other HIV-associated chronic infections. It has also been proposed that drug-resistant strains of reduced fitness might undergo compensatory adaptation during passage through a population of immunocompromised individuals, ultimately restoring their capacity to infect immunocompetent hosts (93). In general, it seems likely that increased TB incidence rates associated with high HIV prevalence will facilitate the spread of both susceptible and MDR strains (55); while slower to emerge in immunocompetent individuals, MDR strains could result in huge burdens of disease in the future (93). However, there is evidence to suggest that the impact of HIV on TB transmission (and therefore prevalence) is more complex and might depend on factors such as the duration of infectious period and the presence of a functioning TB control program (56, 57).

Mutation Rates and the Role of Mutators

Based on in vitro measures of rates of mutation to single drug resistance in M. tuberculosis (64), the emergence of MDR TB appears to require a larger bacillary population than is usually present during infection. However, the assumption that the risk of acquiring multiple resistance equals the product of individual mutation rates is likely an oversimplification, considering the complex interplay of factors that might operate to increase mutation rates in vivo. Small subpopulations of mutators characterize commensal and pathogenic bacterial populations in vivo (18, 67, 154, 180, 221, 232), consistent with the idea that elevated mutation rates may promote adaptation to the fluctuating host environment. However, the deleterious consequences of a constitutive mutator phenotype (20) ensure that maintenance of the mutator allele depends on genetic linkage to the resultant beneficial mutation (20, 232). In general, the selective advantage of a high mutation rate is transient, and regaining the wild-type genotype is essential to the long-term survival of the population (94, 95).
Stable, acquisitive evolution is thought to depend on minimal disturbances to established bacterial pathways and host-pathogen interactions (253), a concept consistent with the suggested adaptation of separate M. tuberculosis lineages to particular host populations (111). Instead, the long-term selection or counterselection of small-effect mutators is thought likely to exert greater influence on bacterial evolution (221), perhaps explaining the failure to observe a mutator phenotype in M. tuberculosis, despite its natural deficiency in several pathways associated with hypermutability in other organisms (162). A possible exception is provided by the W-Beijing genotype, which is most frequently associated with emergence of MDR TB (96). A high proportion of W-Beijing isolates contain mutations in genes required for elimination of damaged nucleotides (mutT) and reversal of alkylation damage to DNA (ogt) (193). In vitro assays of mutation rates have so far failed to demonstrate increased mutagenesis in W-Beijing isolates (246), although it is possible that the strains tested did not carry the characteristic “mutator” mutations. Furthermore, the prevalence of the W-Beijing lineage among MDR strains makes it tempting to speculate that the mutator phenotype might be manifest only under (stressful) in vivo conditions.
Inducible (environment-dependent) mutators, in contrast, increase global mutation rates specifically in response to applied stress (158). Those cells that survive produce progeny cells with normal mutation rates, thereby reducing the risk of unchecked mutagenesis. Whereas the acquisition of a mutator phenotype is a random event, it has been proposed that inducible mutagenesis is an adaptive response that has evolved by second-order selection to modulate mutation rates while limiting the costs associated with a constitutive mutator phenotype (158, 225, 226). That inducible mutator mechanisms are subject to selection has been inferred from the negative correlation between stress-induced mutagenesis and constitutive mutators (17). This idea is further reinforced by the observation that where stresses are frequent or of long duration, inducible mutators are selected as efficiently as mutator alleles (17). The variation in the strength, frequency, and nature of inducible mutagenesis mechanisms is thought to be reflective of the dynamic response of different pathogens to specific local environments (225).
The role of stationary-phase or stress-induced mutagenesis in bacterial adaptation has been subject to considerable recent attention (225). In particular, an association between inducible mutation pathways and the emergence of drug-resistant isolates of pathogenic bacteria has been described (4, 18, 189, 197), which might be especially relevant to the generation of antibiotic and stress resistance mutations in M. tuberculosis, whose microevolution within the host environment (as stated above) is driven by genetic rearrangement and point mutations (105, 195). In most bacterial systems studied to date, adaptation to environmental stress is predicated on the activity of SOS-inducible, error-prone repair polymerases of the Y polymerase superfamily (157, 220, 224, 255). Members of the Y family of DNA polymerases likely evolved to promote mutation avoidance and damage tolerance through a specialized ability to replicate across a variety of DNA lesions; however, the flip side of this ability is that the very properties enabling translesion synthesis are implicated in mutagenesis (88). The M. tuberculosis genome encodes two putative Y family polymerases of the DinB subclass (178), but, unusually, neither is upregulated in response to DNA damage (23, 66). Instead, their predicted physiological roles are fulfilled in M. tuberculosis by a novel, damage-inducible C family polymerase, DnaE2, which is solely responsible for damage-induced base substitution mutagenesis (23). Significantly, deletion of dnaE2 results in damage hypersensitivity and eliminates damage-induced base mutagenesis in vitro and is associated with late-stage attenuation as well as reduced emergence of drug resistance mutations in a murine infection model (23).
Coupled with the induction of dnaE2 during stationary-phase infection in mice (23), these observations suggest that genetically encoded antibiotic resistance mutations may arise as the result of DnaE2-mediated repair synthesis during persistent infection. According to this hypothesis, a range of host immune effectors and other environmental damaging agents, as well as endogenous oxidative and nitrosative metabolic stresses or antibiotics, might induce damage lesions. Stalled replication at a lesion induces dnaE2 expression either as part of the mycobacterial SOS response or by unknown regulatory mechanisms analogous to novobiocin-mediated dnaE2 induction (25). Error-prone repair synthesis by DnaE2 might fix mutations in chromosomal DNA at the site of the damage, in some cases conferring antibiotic resistance which is then selected.

INTRINSIC DRUG RESISTANCE

M. tuberculosis is intrinsically resistant to many of the antibiotics and chemotherapeutics in current medical use (117). This intrinsic resistance is, at least in part, because of the relative impermeability of the mycolic acid-rich mycobacterial cell envelope (27). The contribution of each of the various classes of mycolic acids to cell wall composition has been shown to vary depending on growth phase and oxygen tension and differs for in vitro versus intraphagosomal survival (13). In addition, adaptation to the stationary phase is characterized by a thickening of the mycobacterial cell wall (59), which might be especially relevant to antimicrobial tolerance in vivo. However, limited permeability accounts to only a certain degree for the inherent antibiotic resistance of mycobacteria (117), and mycobacteria possess several general mechanisms which might limit the intracellular accumulation, or interfere with the activity, of those antimycobacterial compounds that are able to penetrate the cell wall. For example, a recently identified mycobacterial protein, MfpA, has been implicated in low-level fluoroquinolone resistance (108). Although the precise physiological role of MfpA is not known, its structure resembles that of double-stranded DNA, thus enabling interference in fluoroquinolone drug action by binding to the cellular target, DNA gyrase.
The concentration achieved by antibiotics inside bacterial cells is a function of the number of efflux systems capable of extruding toxic compounds (175). Although hydrophilic diffusion in M. tuberculosis is constrained by the limited number and restrictive structure of mycobacterial porins (79, 228), mycobacteria contain a large number of putative drug transporters (138). Significantly, recent evidence suggests a role for active efflux in diminished intracellular drug concentrations (190). Expression of the efpA-encoded drug transport homologue is induced in INH-treated M. tuberculosis (248), for example, while exposure of M. tuberculosis to either INH or ethambutol results in upregulation of iniA, encoding a transmembrane protein implicated in tolerance to both frontline anti-TB drugs through an MDR pump-like mechanism (52). Overexpression of certain transport genes has similarly been shown to confer low-level resistance to specific antimicrobials (2, 44, 71, 144, 214, 222). Conversely, targeted disruption of specific efflux pump genes is associated with increased antibiotic susceptibility (139). Finally, the demonstration that disruptions in efflux gene regulation markedly increase resistance to a variety of drugs is suggestive of a potential evolutionary mechanism for the development of multidrug resistance in pathogenic mycobacteria (139). Indirect support for this contention might be provided by the recent identification of the WhiB7-regulated multidrug resistance system in M. tuberculosis (164).

PHENOTYPIC DRUG TOLERANCE

Specialized Persister Cells

The ability of bacterial populations to resist killing by bactericidal factors has been the subject of renewed interest recently, particularly because of the potential implications for drug tolerance (9, 127, 160). Most in vitro antibiotic kill curves are biphasic, comprising early exponential decay followed by bacteriostasis and reduced or delayed bactericidal activity (36, 102, 250). The surviving subpopulation comprises phenotypically tolerant bacteria, or persisters, that can be enriched by prolonged antibiotic exposure (250). In contrast to the case for heritable resistance, however, persisters retain genetic susceptibility to the drug.
Although primarily associated with biofilm or stationary-phase populations (28, 137, 216), persisters were originally observed in populations of rapidly growing planktonic bacteria (15). In fact, separate persister classes have been classified (9) according to the requirement for stationary-phase passage, with type I persisters emerging slowly during stationary-phase exit and type II persisters arising spontaneously in growing populations as a result of a reversible metabolic switch. Of course, because it is not associated with a long-term fitness cost, reversible antibiotic tolerance confers a selective advantage; that is, phenotypic heterogeneity offers a genetically homogeneous population an “insurance policy” (130) against elimination by antibiotics or other stresses.
Mathematical modeling suggests that the optimal switching rate between normal and type II persister cells is a function of the frequency of environmental changes, implying that bacterial persistence mechanisms constitute an adaptation to the distribution of environmental change (130). There is some evidence, however, that the switch to a persistent state might be preferentially induced by certain classes of antibiotics (152, 160), in some cases involving the SOS response (160). Furthermore, the switch to a persistent phenotype might be favored where bactericidal antibiotics are ineffective against slowly dividing or nondividing cells (9, 160), a characteristic of current frontline antituberculosis drugs.

The Stringent Response, Toxin-Antitoxin Loci, and Applicability to Phenotypic Drug Tolerance in M. tuberculosis

The mechanisms governing the switch to a persistent phenotype remain to be identified; however, accumulating evidence implicates the activity of the prokaryotic stringent response regulator, RelA (100, 230). RelA catalyzes the hyperphosphorylation of GTP to (p)ppGpp during amino acid and carbon source starvation (37). Binding of the (p)ppGpp alarmone to the RNA polymerase β subunit inhibits transcription of translation machinery components, stimulates amino acid biosynthesis and transport operons, and decreases transcription rates (8, 12, 37, 42, 237). In addition, (p)ppGpp affects the global transcriptional response to changing environmental conditions by mediating association of alternative σ factors with RNA polymerase (121, 132). Given the similarities between persistent and stringent physiology, the requirement for RelA is not surprising. However, the identification of a high proportion of toxin components of toxin-antitoxin (TA) modules among those genes induced in a persister population (127) suggests that RelA, through the activity of (p)ppGpp, might not constitute the sole mediator of persistence.
TA modules were originally characterized as factors ensuring episomal stability by postsegregational killing of cured segregants (91, 116). However, subsequent analysis has revealed the widespread distribution of chromosomally encoded TA modules in prokaryotic genomes (184). The stable toxins inhibit transcription and translation by various mechanisms, including inhibition of DNA gyrase activity (118) and cleavage of mRNA (45-47, 187, 258) and require neutralization by labile antitoxins. Compelling evidence of the potential role of TA-like modules in persistence is provided by the recent characterization of the Escherichia coli hipA7 allele (129), originally identified in a screen for high-persistence mutants (165). Significantly, abrogation of (p)ppGpp synthetase activity in the hipA7 mutant strain eliminates the high-persistence phenotype (129), implying the requirement for a functional stringent response in the persister state. These observations have been incorporated in a general model of stress physiology (92) in which the (p)ppGpp-mediated stringent response, in combination with TA activity, effects a regulated switch to a persistent phenotype.
The failure of frontline drugs to sterilize M. tuberculosis infections in vivo (156) is reminiscent of incomplete killing in vitro (110) and suggests that a switch to persistent physiology might hinder anti-TB drug efficacy. Significantly, resistance to frontline antimicrobial compounds emerges during long-term in vitro adaptation of M. tuberculosis (110, 238) and mimics resistance in bacilli recovered from tuberculous lesions in vivo (241, 243). Similarly, starvation of M. tuberculosis in vitro has recently been shown to increase tolerance of a wide range of antimycobacterial compounds (254). The stringent response in M. tuberculosis mediated by the M. tuberculosis Rel protein has been the subject of detailed investigation and has been shown to be required for survival of M. tuberculosis during long-term starvation in vitro (191) and for persistence in mice (60, 123). M. tuberculosis also contains an unusually large number of putative TA loci in its genome (7, 184). Although most of these TA loci have yet to be investigated, sequence-specific mRNA interferase activity has recently been demonstrated in two M. tuberculosis MazF homologues (261). That study further showed that four of the putative M. tuberculosis MazF proteins cause cell growth arrest when overexpressed in E. coli (261). Coupled with their abundance in M. tuberculosis, as well as the functionality and physiological importance of the stringent response regulator, these observations are strongly suggestive of possible involvement of the TA genes in phenotypic drug tolerance in M. tuberculosis.

CONCLUDING REMARKS

The efficacy of TB chemotherapy is a function of drug activity against potentially heterogeneous populations of M. tuberculosis in disparate anatomical loci within the host, as well as the ability of the organism to subvert that activity. The current frontline drugs are a modern intervention, however, and M. tuberculosis has evolved mechanisms over thousands of years (29) to enable survival in the variable and hostile in vivo environment. In this review, it has been proposed that the same mechanisms might affect drug efficacy (Fig. 1). Specifically, inherent characteristics such as cell wall physiology and active efflux of toxic metabolites, persistence mechanisms such as (p)ppGpp- and TA-mediated translational inhibition, and adaptive mechanisms such as induced mutagenesis are potentially implicated in impaired antibiotic activity or the emergence of chromosomally encoded resistance mutations. Furthermore, although this review has categorized separate tolerance or resistance mechanisms, it is likely that a complex interaction governs survival in vivo; for example, phenotypic tolerance likely enables the emergence of antibiotic-resistant mutants by ensuring that a subpopulation survives the extended course of chemotherapy. Finally, key assumptions require further investigation. For example, there is limited information on the ability of the different anti-TB drugs to penetrate lesions in vivo, while the contribution of M. tuberculosis Rel and the TA modules to drug tolerance has emerged as an important area for future study.
FIG. 1.
FIG. 1. Mechanisms and pathways limiting drug efficacy in M. tuberculosis. Solid lines represent known interactions; broken lines represent those assumed or postulated.

Acknowledgments

D.F.W. was supported by Collaborative Research Initiative grant no. 065578 from the Wellcome Trust (to Neil G. Stoker and V.M.), and V.M. was supported by an International Research Scholar's grant from the Howard Hughes Medical Institute and grants from the Medical Research Council of South Africa, National Research Foundation, and University of the Witwatersrand.
We thank Helena Boshoff, Megan Murray, Harvey Rubin, Ken Duncan, Neil Stoker, and members of the Mizrahi lab for critically reviewing the manuscript and John McKinney for stimulating discussions.

REFERENCES

1.
Adams, L. B., M. C. Dinauer, D. E. Morgenstern, and J. L. Krahenbuhl. 1997. Comparison of the roles of reactive oxygen and nitrogen intermediates in the host response to Mycobacterium tuberculosis using transgenic mice. Tuber. Lung Dis.78:237-246.
2.
Ainsa, J. A., M. C. Blokpoel, I. Otal, D. B. Young, K. A. De Smet, and C. Martin. 1998. Molecular cloning and characterization of Tap, a putative multidrug efflux pump present in Mycobacterium fortuitum and Mycobacterium tuberculosis. J. Bacteriol.180:5836-5843.
3.
Akaki, T., H. Tomioka, T. Shimizu, S. Dekio, and K. Sato. 2000. Comparative roles of free fatty acids with reactive nitrogen intermediates and reactive oxygen intermediates in expression of the anti-microbial activity of macrophages against Mycobacterium tuberculosis. Clin. Exp. Immunol.121:302-310.
4.
Alonso, A., E. Campanario, and J. L. Martínez. 1999. Emergence of multidrug-resistant mutants is increased under antibiotic selective pressure in Pseudomonas aeruginosa. Microbiology145:2857-2862.
5.
Andersson, D. I. 2003. Persistence of antibiotic resistant bacteria. Curr. Opin. Microbiol.6:452-456.
6.
Andersson, D. I., and B. R. Levin. 1999. The biological cost of antibiotic resistance. Curr. Opin. Microbiol.2:489-493.
7.
Arcus, V. L., P. B. Rainey, and S. J. Turner. 2005. The PIN-domain toxin-antitoxin array in mycobacteria. Trends Microbiol.13:360-365.
8.
Artsimovitch, I., V. Patian, S. Sekine, M. N. Vassylyeva, T. Hosaka, K. Ochi, S. Yokoyama, and D. G. Vassylyev. 2004. Structural basis for transcription regulation by alarmone ppGpp. Cell117:299-310.
9.
Balaban, N. Q., J. Merrin, R. Chait, L. Kowalik, and S. Leibler. 2004. Bacterial persistence as a phenotypic switch. Science305:1622-1625.
10.
Baquero, F., and M. C. Negri. 1997. Selective compartments for resistant microorganisms in antibiotic gradients. Bioessays19:731-736.
11.
Baquero, F., M. C. Negri, M. I. Morosini, and J. Blazquez. 1998. Antibiotic-selective environments. Genetics27:S5-S11.
12.
Barker, M. M., T. Gaal, C. A. Josaitis, and R. L. Gourse. 2001. Mechanisms of regulation of transcription initiation by ppGpp. I. Effects of ppGpp on transcription initiation in vivo and in vitro. J. Mol. Biol.305:673-688.
13.
Barry, C. E., III, and K. Mdluli. 1996. Drug sensitivity and environmental adaptation of mycobacterial cell wall components. Trends Microbiol.4:275-281.
14.
Betts, J. C., P. T. Lukey, L. C. Robb, R. A. McAdam, and K. Duncan. 2002. Evaluation of a nutrient starvation model of Mycobacterium tuberculosis persistence by gene and protein expression profiling. Mol. Microbiol.43:717-731.
15.
Bigger, J. W. 1944. Treatment of staphylococcal infections with penicillin. Lancetii:497-500.
16.
Billington, O. J., T. D. McHugh, and S. H. Gillespie. 1999. Physiological cost of rifampin resistance induced in vitro in Mycobacterium tuberculosis. Antimicrob. Agents Chemother.43:1866-1869.
17.
Bjedov, I., O. Tenaillon, B. Gerard, V. Souza, E. Denamur, M. Radman, F. Taddei, and I. Matic. 2003. Stress-induced mutagenesis in bacteria. Science300:1404-1409.
18.
Björkholm, B., M. Sjolund, P. G. Falk, O. G. Berg, L. Engstrand, and D. I. Andersson. 2001. Mutation frequency and biological cost of antibiotic resistance in Helicobacter pylori. Proc. Natl. Acad. Sci. USA98:14607-14612.
19.
Blower, S. M., and T. Chou. 2004. Modeling the emergence of the ‘hot zones’: tuberculosis and the amplification dynamics of drug resistance. Nat. Med.10:1111-1116.
20.
Boe, L., M. Danielsen, S. Knudsen, J. B. Petersen, J. Maymann, and P. R. Jensen. 2000. The frequency of mutators in populations of E. coli. Mutat. Res.448:47-55.
21.
Boon, C., and T. Dick. 2002. Mycobacterium bovis BCG response regulator essential for hypoxic dormancy. J. Bacteriol.184:6760-6767.
22.
Boshoff, H. I. M., and C. E. Barry III. 2005. Tuberculosis: metabolism and respiration in the absence of growth. Nat. Rev. Microbiol.3:70-80.
23.
Boshoff, H. I. M., M. B. Reed, C. E. Barry III, and V. Mizrahi. 2003. DnaE2 polymerase contributes to in vivo survival and the emergence of drug resistance in Mycobacterium tuberculosis. Cell113:183-193.
24.
Boshoff, H. I. M., S. I. Durbach, and V. Mizrahi. 2001. DNA metabolism in Mycobacterium tuberculosis: implications for drug resistance and strain variability. Scand. J. Infect. Dis.33:101-105.
25.
Boshoff, H. I. M., T. G. Myers, B. R. Copp, M. R. McNeil, M. A. Wilson, and C. E. Barry III. 2004. The transcriptional responses of M. tuberculosis to inhibitors of metabolism: novel insights into drug mechanisms of action. J. Biol. Chem.279:40174-40184.
26.
Bouley, D. M., N. Ghori, K. L. Mercer, S. Falkow, and L. Ramakrishnan. 2001. Dynamic nature of host-pathogen interactions in Mycobacterium marinum granulomas. Infect. Immun.69:7820-7831.
27.
Brennan, P. J., and H. Nikaido. 1995. The envelope of mycobacteria. Annu. Rev. Biochem.64:29-63.
28.
Brooun, A., S. Liu, and K. Lewis. 2000. A dose-response study of antibiotic resistance in Pseudomonas aeruginosa biofilms. Antimicrob. Agents Chemother.44:640-646.
29.
Brosch, R., S. V. Gordon, M. Marmiesse, P. Brodin, C. Buchrieser, K. Eiglmeier, T. Garnier, C. Gutierrez, G. Hewinson, K. Kremer, L. M. Parsons, A. S. Pym, S. Samper, D. van Soolingen, and S. T. Cole. 2002. A new evolutionary scenario for the Mycobacterium tuberculosis complex. Proc. Natl. Acad. Sci. USA99:3684-3689.
30.
Bryk, R., P. Griffin, and C. Nathan. 2000. Peroxynitrite reductase activity of bacterial peroxiredoxins. Nature407:211-215.
31.
Buchmeier, N. A., C. J. Lipps, M. Y. So, and F. Heffron. 1993. Recombination-deficient mutants of Salmonella typhimurium are avirulent and sensitive to the oxidative burst of macrophages. Mol. Microbiol.7:933-936.
32.
Buchmeier, N. A., G. L. Newton, T. Koledin, and R. C. Fahey. 2003. Association of mycothiol with protection of Mycobacterium tuberculosis from toxic oxidants and antibiotics. Mol. Microbiol.47:1723-1732.
33.
Burney, S., J. L. Caulfield, J. C. Niles, J. S. Wishnok, and S. R. Tannenbaum. 1999. The chemistry of DNA damage from nitric oxide and peroxynitrite. Mutat. Res.424:37-49.
34.
Canetti, G. 1955. The tubercle bacillus in the pulmonary lesion of man. Springer Publishing Co., Inc., New York, N.Y.
35.
Canetti, G., M. Le Lirzin, G. Porven, N. Rist, and F. Grumbach. 1968. Some comparative aspects of rifampicin and isoniazid. Tubercle49:367-376.
36.
Carret, G., J. P. Flandrois, and J. R. Lobry. 1991. Biphasic kinetics of bacterial killing by quinolones. J. Antimicrob. Chemother.27:319-327.
37.
Cashel, M., D. R. Gentry, V. H. Hernandez, and D. Vinella. 1996. The stringent response, p. 1458-1496. In F. C. Neidhardt, R. Curtiss III, J. L. Ingraham, E. C. C. Lin, K. Low, Jr., B. Magasanik, W. S. Reznikoff, M. Riley, M. Schaechter, and H. E. Umbarger (ed.), Escherichia coli and Salmonella: cellular and molecular biology. ASM Press, Washington, D.C.
38.
Chan, J., and J. L. Flynn. 1999. Nitric oxide in Mycobacterium tuberculosis infection, p. 281-310. In F. Fang (ed.), Nitric oxide and infection. Plenum, New York, N.Y.
39.
Chan, J., and S. H. E. Kaufmann. 1994. Immune mechanisms of protection, p. 389-415. In B. R. Bloom (ed.), Tuberculosis: pathogenesis, protection and control. ASM Press, Washington, D.C.
40.
Chan, J., T. Fujiwara, P. Brennan, M. McNeil, S. J. Turco, J. C. Sibille, M. Snapper, P. Aisen, and B. R. Bloom. 1989. Microbial glycolipids: possible virulence factors that scavenge oxygen radicals. Proc. Natl. Acad. Sci. USA86:2453-2457.
41.
Chan, J., X. D. Fan, S. W. Hunter, P. J. Brennan, and B. R. Bloom. 1991. Lipoarabinomannan, a possible virulence factor involved in persistence of Mycobacterium tuberculosis within macrophages. Infect. Immun.59:1755-1761.
42.
Chang, D. E., D. J. Smalley, and T. Conway. 2002. Gene expression profiling of Escherichia coli growth transitions: an expanded stringent response model. Mol. Microbiol.45:289-306.
43.
Chen, L., Q. W. Xie, and C. Nathan. 1998. Alkyl hydroperoxide reductase subunit C (AhpC) protects bacterial and human cells against reactive nitrogen intermediates. Mol. Cell1:795-805.
44.
Choudhuri, B. S., S. Bhakta, R. Barik, J. Basu, M. Kundu, and P. Chakrabarti. 2002. Overexpression and functional characterization of an ABC (ATP-binding cassette) transporter encoded by the genes drrA and drrB of Mycobacterium tuberculosis. Biochem. J.367:279-285.
45.
Christensen, S. K., and K. Gerdes. 2003. RelE toxins from bacteria and archaea cleave mRNAs on translating ribosomes, which are rescued by tmRNA. Mol. Microbiol.48:1389-1400.
46.
Christensen, S. K., K. Pedersen, F. G. Hansen, and K. Gerdes. 2003. Toxin-antitoxin loci as stress-response elements: ChpAK/MazF and ChpBK cleave translated RNAs and are counteracted by tmRNA. J. Mol. Biol.332:809-819.
47.
Christensen, S. K., M. Mikkelsen, K. Pedersen, and K. Gerdes. 2001. RelE, a global inhibitor of translation, is activated during nutritional stress. Proc. Natl. Acad. Sci. USA98:14328-14333.
48.
Chua, J., I. Vergne, S. Master, and V. Deretic. 2004. A tale of two lipids: Mycobacterium tuberculosis phagosome maturation arrest. Curr. Opin. Microbiol.7:71-77.
49.
Cohen, T., and M. Murray. 2004. Modeling epidemics of mutidrug-resistant M. tuberculosis of heterogeneous fitness. Nat. Med.10:1117-1121.
50.
Cohen, T., B. Sommers, and M. Murray. 2003. The effect of drug resistance on the fitness of Mycobacterium tuberculosis. Lancet Infect. Dis.3:13-21.
51.
Cohen, T., M. C. Becerra, and M. B. Murray. 2004. Isoniazid resistance and the future of drug-resistant tuberculosis. Microb. Drug Resist.10:280-285.
52.
Colangeli, R., D. Helb, S. Sridharan, J. Sun, M. Varma-Basil, M. H. Hazbon, R. Harbacheuski, N. J. Megjugorac, W. R. Jacobs, Jr., A. Holzenburg, J. C. Sacchettini, and D. Alland. 2005. The Mycobacterium tuberculosis iniA gene is essential for activity of an efflux pump that confers drug tolerance to both isoniazid and ethambutol. Mol. Microbiol.55:1829-1840.
53.
Cole, S. T., K. Eiglmeier, J. Parkhill, K. D. James, N. R. Thomson, P. R. Wheeler, N. Honoré, T. Garnier, C. Churcher, D. Harris, K. Mungall, D. Basham, D. Brown, T. Chillingworth, R. Connor, R. M. Davies, K. Devlin, S. Duthoy, T. Feltwell, A. Fraser, N. Hamlin, S. Holroyd, T. Hornsby, K. Jagels, C. Lacroix, J. Maclean, S. Moule, L. Murphy, K. Oliver, M. A. Quail, M. A. Rajandream, K. M. Rutherford, S. Rutter, K. Seeger, S. Simon, M. Simmonds, J. Skelton, R. Squares, S. Squares, S. Stevens, K. Taylor, S. Whitehead, J. R. Woodward, and B. G. Barrell. 2001. Massive gene decay in the leprosy bacillus. Nature409:1007-1011.
54.
Cole, S. T., R. Brosch, J. Parkhill, T. Garnier, C. Churcher, D. Harris, S. V. Gordon, K. Eiglmeier, S. Gas, C. E. Barry III, F. Tekaia, K. Badcock, D. Basham, D. Brown, T. Chillingworth, R. Connor, R. Davies, K. Devlin, T. Feltwell, S. Gentles, N. Hamlin, S. Holroyd, T. Hornsby, K. Jagels, A. Krogh, J. McLean, S. Moule, L. Murphy, K. Oliver, J. Osborne, M. A. Quail, M. A. Rajandream, J. Rogers, S. Rutter, K. Seeger, J. Skelton, R. Squares, S. Squares, J. E. Sulston, J. Taylor, S. Whitehead, and B. G. Barrell. 1998. Deciphering the biology of Mycobacterium tuberculosis from the complete genome sequence. Nature393:537-544.
55.
Corbett, E. L., C. J. Watt, N. Walker, D. Maher, B. G. Williams, M. C. Raviglione, and C. Dye. 2003. The growing burden of tuberculosis: global trends and interactions with the HIV epidemic. Arch. Intern. Med.163:1009-1021.
56.
Corbett, E. L., S. Charalambous, K. Fielding, T. Clayton, R. J. Hayes, K. M. De Cock, and G. J. Churchyard. 2003. Stable incidence rates of tuberculosis (TB) among human immunodeficiency virus (HIV)-negative South African gold miners during a decade of epidemic HIV-associated TB. J. Infect. Dis.188:1156-1163.
57.
Corbett, E. L., S. Charalambous, V. M. Moloi, K. Fielding, A. D. Grant, C. Dye, K. M. De Cock, R. J. Hayes, B. G. Williams, and G. J. Churchyard. 2004. Human immunodeficiency virus and the prevalence of undiagnosed tuberculosis in African gold miners. Am. J. Respir. Crit. Care Med.170:673-679.
58.
Cosma, C. L., D. R. Sherman, and L. Ramakrishnan. 2003. The secret lives of pathogenic mycobacteria. Annu. Rev. Microbiol.57:641-676.
59.
Cunningham, A. F., and C. L. Spreadbury. 1998. Mycobacterial stationary phase induced by low oxygen tension: cell wall thickening and localization of the 16-kilodalton alpha-crystallin homolog. J. Bacteriol.180:801-808.
60.
Dahl, J. L., C. N. Kraus, H. I. Boshoff, B. Doan, K. Foley, D. Avarbock, G. Kaplan, V. Mizrahi, H. Rubin, and C. E. Barry III. 2003. The role of RelMtb-mediated adaptation to stationary phase in long-term persistence of Mycobacterium tuberculosis in mice. Proc. Natl. Acad. Sci. USA100:10026-10031.
61.
Dale, J. W. 1995. Mobile genetic elements in mycobacteria. Eur. Respir. J. Suppl.20:633s-648s.
62.
Dannenberg, A. M., Jr. 1993. Immunopathogenesis of pulmonary tuberculosis. Hosp. Prac.28:51-58.
63.
Darwin, K. H., and C. F. Nathan. 2005. Role for nucleotide excision repair in virulence of Mycobacterium tuberculosis. Infect. Immun.73:4581-4587.
64.
David, H. L. 1970. Probability distribution of drug-resistant mutants in unselected populations of Mycobacterium tuberculosis. Appl. Microbiol.20:810-814.
65.
Davies, A. P., O. J. Billington, B. A. Bannister, W. R. Weir, T. D. McHugh, and S. H. Gillespie. 2000. Comparison of fitness of two isolates of Mycobacterium tuberculosis, one of which had developed multi-drug resistance during the course of treatment. J. Infect.41:184-187.
66.
Davis, E. O., E. M. Dullaghan, and L. Rand. 2002. Definition of the mycobacterial SOS box and use to identify LexA-regulated genes in Mycobacterium tuberculosis. J. Bacteriol.184:3287-3295.
67.
Denamur, E., S. Bonacorsi, A. Giraud, P. Duriez, F. Hilali, C. Amorin, E. Bingen, A. Andremont, B. Picard, F. Taddei, and I. Matic. 2002. High frequency of mutator strains among human uropathogenic Escherichia coli isolates. J. Bacteriol.184:605-609.
68.
Derbyshire, K. M., and S. Bardarov. 2000. DNA transfer in mycobacteria: conjugation and transduction, p. 93-107. In G. F. Hatfull and W. R. Jacobs, Jr. (ed.), Molecular genetics of mycobacteria. ASM Press, Washington, D.C.
69.
Deretic, V., and R. A. Fratti. 1999. Mycobacterium tuberculosis phagosome. Mol. Microbiol.31:1603-1609.
70.
Deretic, V., W. Phillip, S. Dhandayuthapani, M. H. Mudd, R. Curcic, T. Garbe, B. Heym, L. E. Via, and S. T. Cole. 1995. Mycobacterium tuberculosis is a natural mutant with an inactivated oxidative-stress regulatory gene: implications for sensitivity to isoniazid. Mol. Microbiol.17:889-900.
71.
De Rossi, E., M. Branzoni, R. Cantoni, A. Milano, G. Riccardi, and O. Ciferri. 1998. mmr, a Mycobacterium tuberculosis gene conferring resistance to small cationic dyes and inhibitors. J. Bacteriol.180:6068-6071.
72.
De Voss, J. J., K. Rutter, B. G. Schroeder, H. Su, Y. Zhu, and C. E. Barry, III. 2000. The salicylate-derived mycobactin siderophores of Mycobacterium tuberculosis are essential for growth in macrophages. Proc. Natl. Acad. Sci. USA97:1252-1257.
73.
Dhandayuthapani, S., Y. Zhang, M. H. Mudd, and V. Deretic. 1996. Oxidative stress response and its role in sensitivity to isoniazid in mycobacteria: characterization and inducibility of AhpC by peroxides in Mycobacterium smegmatis and lack of expression in M. aurum and M. tuberculosis. J. Bacteriol.178:3641-3649.
74.
Dubnau, E., P. Fontan, R. Manganelli, S. Soares-Appel, and I. Smith. 2002. Mycobacterium tuberculosis genes induced during infection of human macrophages. Infect. Immun.70:2787-2795.
75.
Durbach, S. I., B. Springer, E. E. Machowski, R. J. North, K. G. Papavinasasundaram, M. J. Colston, E. C. Böttger, and V. Mizrahi. 2003. DNA alkylation damage as a sensor of nitrosative stress in Mycobacterium tuberculosis. Infect. Immun.71:997-1000.
76.
Dye, C., C. J. Watt, D. M. Bleed, S. M. Hosseini, and M. C. Raviglione. 2005. Evolution of tuberculosis control and prospects for reducing tuberculosis incidence, prevalence, and deaths globally. JAMA293:2767-2775.
77.
Ehrt, S., M. U. Shiloh, J. Ruan, M. Choi, S. Gunzburg, C. Nathan, Q. Xie, and L. W. Riley. 1997. A novel antioxidant gene from Mycobacterium tuberculosis. J. Exp. Med.186:1885-1896.
78.
Elliott, A. M., S. E. Berning, M. D. Iseman, and C. A. Peloquin. 1995. Failure of drug penetration and acquisition of drug resistance in chronic tuberculous empyema. Tuber. Lung. Dis.76:463-467.
79.
Engelhardt. H., C. Heinz, and M. Niederweis. 2002. A tetrameric porin limits the cell wall permeability of Mycobacterium smegmatis. J. Biol. Chem.277:37567-37572.
80.
Enne, V. I., D. M. Livermore, P. Stephens, and L. M. Hall. 2001. Persisence of sulphonamide resistance in Escherichia coli in the UK despite national prescribing restriction. Lancet357:1325-1328.
81.
Farmer, P., and J. Y. Kim. 1998. Community based approaches to the control of multidrug resistant tuberculosis: introducing “DOTS-plus.” BMJ317:671-674.
82.
Farmer, P., J. Furin, J. Bayona, M. Becerra, C. Henry, H. Hiatt, J. Y. Kim, C. Mitnick, E. Nardell, and S. Shin. 1999. Management of MDR-TB in resource-poor countries. Int. J. Tuberc. Lung Dis.3:643-645.
83.
Fenhalls, G., L. Stevens, L. Moses, J. Bezuidenhout, J. C. Betts, P. van Helden, P. T. Lukey, and K. Duncan. 2002. In situ detection of Mycobacterium tuberculosis transcripts in human lung granulomas reveals differential gene expression in necrotic lesions. Infect. Immun.70:6330-6338.
84.
Fisher, M. A., B. B. Plikaytis, and T. M. Shinnick. 2002. Microarray analysis of the Mycobacterium tuberculosis transcriptional response to the acidic conditions found in phagosomes. J. Bacteriol.184:4025-4032.
85.
Fleischmann, R. D., D. Alland, J. A. Eisen, L. Carpenter, O. White, J. Peterson, R. DeBoy, R. Dodson, M. Gwinn, D. Haft, E. Hickey, J. F. Kolonay, W. C. Nelson, L. A. Umayam, M. Ermolaeva, S. L. Salzberg, A. Delcher, T. Utterback, J. Weidman, H. Khouri, J. Gill, A. Mikula, W. Bishai, W. R. Jacobs, Jr., J. C. Venter, and C. M. Fraser. 2002. Whole-genome comparison of Mycobacterium tuberculosis clinical and laboratory strains. J. Bacteriol.184:5479-5490.
86.
Flint, J. L., J. C. Kowalski, P. K. Karnati, and K. M. Derbyshire. 2004. The RD1 virulence locus of Mycobacterium tuberculosis regulates DNA transfer in Mycobacterium smegmatis. Proc. Natl. Acad. Sci. USA101:12598-12603.
87.
Flynn, J. L., and J. Chan. 2001. Immunology of tuberculosis. Annu. Rev. Immunol.19:93-129.
88.
Friedberg, E. C., P. L. Fischhaber, and C. Kisker. 2001. Error-prone DNA polymerases: novel structures and the benefits of infidelity. Cell107:9-12.
89.
Garbe, T. R., N. S. Hibler, and V. Deretic. 1996. Response of Mycobacterium tuberculosis to reactive oxygen and nitrogen intermediates. Mol. Med.2:134-142.
90.
Garbe, T. R., N. S. Hibler, and V. Deretic. 1999. Response to reactive nitrogen intermediates in Mycobacterium tuberculosis: induction of the 16-kilodalton alpha-crystallin homolog by exposure to nitric oxide donors. Infect. Immun.67:460-465.
91.
Gerdes, K., P. B. Rasmussen, and S. Molin. 1986. Unique type of plasmid maintenance function: postsegregational killing of plasmid-free cells. Proc. Natl. Acad. Sci. USA83:3116-3120.
92.
Gerdes, K., S. K. Christensen, and A. Løbner-Olesen. 2005. Prokaryotic toxin-antitoxin stress response loci. Nat. Rev. Microbiol.3:371-382.
93.
Gillespie, S. H. 2002. Evolution of drug resistance in Mycobacterium tuberculosis: clinical and molecular perspective. Antimicrob. Agents Chemother.46:267-274.
94.
Giraud, A., I. Matic, O. Tenaillon, A. Clara, M. Radman, M. Fons, and F. Taddei. 2001. Costs and benefits of high mutation rates: adaptive evolution of bacteria in the mouse gut. Science291:2606-2608.
95.
Giraud, A., M. Radman, I. Matic, and F. Taddei. 2001. The rise and fall of mutator bacteria. Curr. Opin. Microbiol.4:582-585.
96.
Glynn, J. R., J. Whiteley, P. J. Bifani, K. Kremer, and D. van Soolingen. 2002. Worldwide occurrence of Beijing/W strains of Mycobacterium tuberculosis: a systematic review. Emerg. Infect. Dis.8:843-849.
97.
Gold, B., G. M. Rodriguez, S. A. Marras, M. Pentecost, and I. Smith. 2001. The Mycobacterium tuberculosis IdeR is a dual function regulator that controls transcription of genes involved in iron acquisition, iron storage and survival in macrophages. Mol. Microbiol.42:851-865.
98.
Gomez, J. E., and J. D. McKinney. 2004. M. tuberculosis persistence, latency, and drug tolerance. Tuberculosis (Edinburgh)84:29-44.
99.
Gong, C., P. Bongiomo, A. Martins, N. C. Stephanou, H. Zhu, S. Shuman, and M. S. Glickman. 2005. Mechanism of nonhomologous end-joining in mycobacteria: a low-fidelity repair system driven by Ku, ligase D and ligase C. Nat. Struct. Mol. Biol.12:304-312.
100.
Greenway, D. L., and R. R. England. 1999. The intrinsic resistance of Escherichia coli to various antimicrobial agents requires ppGpp and sigma s. Lett. Appl. Microbiol.29:323-326.
101.
Grosset, J. 2003. Mycobacterium tuberculosis in the extracellular compartment: an underestimated adversary. Antimicrob. Agents Chemother.47:833-836.
102.
Guerillot, F., G. Carret, and J. P. Flandrois. 1993. Mathematical model for comparison of time-killing curves. Antimicrob. Agents Chemother.37:1685-1689.
103.
Gupta, R., M. C. Raviglione, and M. A. Espinal. 2001. Should tuberculosis programmes invest in second-line treatments for multidrug-resistant tuberculosis (MDR-TB)? Int. J. Tuberc. Lung Dis.5:1078-1079.
104.
Gupta, S., and D. Chatterji. 2005. Stress responses in mycobacteria. IUBMB Life57:149-159.
105.
Gutacker, M. M., J. C. Smoot, C. A. Lux Migliaccio, S. M. Ricklefs, S. Hua, D. V. Cousins, E. A. Graviss, E. Shashkina, B. N. Kreiswirth, and J. M. Musser. 2002. Genome-wide analysis of synonymous single nucleotide polymorphisms in Mycobacterium tuberculosis complex organisms: resolution of genetic relationships among closely related microbial strains. Genetics162:1533-1543.
106.
Hampshire, T., S. Soneji, J. Bacon, B. W. James, J. Hinds, K. Laing, R. A. Stabler, P. D. Marsh, and P. D. Butcher. 2004. Stationary phase gene expression of Mycobacterium tuberculosis following progressive nutrient depletion: a model for persistent organisms. Tuberculosis84:228-238.
107.
Harth, G., and M. A. Horwitz. 1999. Export of recombinant Mycobacterium tuberculosis superoxide dismutase is dependent upon both information in the protein and mycobacterial export machinery. A model for studying export of leaderless proteins by pathogenic mycobacteria. J. Biol. Chem.274:4281-4292.
108.
Hegde, S. S., M. W. Vetting, S. L. Roderick, L. A. Mitchenall, A. Maxwell, H. E. Takiff, and J. S. Blanchard. 2005. A fluoroquinolone resistance protein from Mycobacterium tuberculosis that mimics DNA. Science308:1480-1483.
109.
Heifets, L. B., and G. A. Cangelosi. 1999. Drug susceptibility testing of Mycobacterium tuberculosis: a neglected problem at the turn of the century. Int. J. Tuberc. Lung Dis.3:564-581.
110.
Herbert, D., C. N. Paramasivan, P. Venkatesan, G. Kubendiran, R. Prabhakar, and D. A. Mitchison. 1996. Bactericidal action of ofloxacin, sulbactam-ampicillin, rifampin, and isoniazid on logarithmic- and stationary-phase cultures of Mycobacterium tuberculosis. Antimicrob. Agents Chemother.40:2296-2299.
111.
Hirsh, A. E., A. G. Tsolaki, K. DeRiemer, M. W. Feldman, and P. M. Small. 2004. Stable association between strains of Mycobacterium tuberculosis and their human host populations. Proc. Natl. Acad. Sci. USA101:4871-4876.
112.
Hobby, G. L., and T. F. Lenert. 1957. The in vitro action of antituberculous agents against multiplying and non-multiplying microbial cells. Am. Rev. Tuberc.76:1031-1048.
113.
Hu, Y. M., J. A. Mangan, J. Dhillon, K. M. Sole, D. A. Mitchison, P. D. Butcher, and A. R. Coates. 2000. Detection of mRNA transcripts and active transcription in persistent Mycobacterium tuberculosis induced by exposure to rifampin or pyrazinamide. J. Bacteriol.182:6358-6365.
114.
Iseman, M. D. 1993. Treatment of multidrug-resistant tuberculosis. N. Engl. J. Med.329:784-791.
115.
Iseman, M. D., and L. A. Madsen. 1991. Chronic tuberculous empyema with bronchopleural fistula resulting in treatment failure and progressive drug resistance. Chest100:124-127.
116.
Jaffe, A., T. Ogura, and S. Hiraga. 1985. Effects of the ccd function of the F plasmid on bacterial growth. J. Bacteriol.163:841-849.
117.
Jarlier, V., and H. Nikaido. 1994. Mycobacterial cell wall: structure and role in natural resistance to antibiotics. FEMS Microbiol. Lett.123:11-18.
118.
Jiang, Y., J. Pogliano, D. R. Helinski, and I. Konieczny. 2002. ParE toxin encoded by the broad-host-range plasmid RK2 is an inhibitor of Escherichia coli gyrase. Mol. Microbiol.44:971-979.
119.
Jindani, A., C. J. Doré, and D. A. Mitchison. 2003. The bactericidal and sterilising activities of antituberculosis drugs during the first 14 days. Am. J. Respir. Crit. Care Med.167:1348-1354.
120.
Jindani, A., V. R. Aber, E. A. Edwards, and D. A. Mitchison. 1980. The early bactericidal activity of drugs in patients with pulmonary tuberculosis. Am. Rev. Respir. Dis.121:939-949.
121.
Jishage, M., K. Kvint, V. Shingler, and T. Nyström. 2002. Regulation of sigma factor competition by the alarmone ppGpp. Genes Dev.16:1260-1270.
122.
Kaplan, G., F. A. Post, D. L. Moreira, H. Wainwright, B. N. Kreiswirth, M. Tanverdi, B. Mathema, S. V. Ramaswamy, G. Walther, L. M. Steyn, C. E. Barry III, and L. G. Bekker. 2003. Mycobacterium tuberculosis growth at the cavity surface: a microenvironment with failed immunity. Infect. Immun.71:7099-7108.
123.
Karakousis, P. C., T. Yoshimatsu, G. Lamichhane, S. C. Woolwine, E. L. Nuermberger, J. Grosset, and W. R. Bishai. 2004. Dormancy phenotype displayed by extracellular Mycobacterium tuberculosis within artificial granulomas in mice. J. Exp. Med.200:647-657.
124.
Kato-Maeda, M., J. T. Rhee, T. R. Gingeras, H. Salamon, J. Drenkow, N. Smittipat, and P. M. Small. 2001. Comparing genomes within the species Mycobacterium tuberculosis. Genome Res.11:547-554.
125.
Kaufmann, S. H. E., S. T. Cole, V. Mizrahi, E. Rubin, and C. Nathan. 2005. Mycobacterium tuberculosis and the host response. J. Exp. Med.201:1693-1697.
126.
Kendall, S. L., F. Movahedzadeh, S. C. G. Rison, L. Wernisch, T. Parish, K. Duncan, J. C. Betts, and N. G. Stoker. 2004. The Mycobacterium tuberculosis dosRS two-component system is induced by multiple stresses. Tuberculosis84:247-255.
127.
Keren, I., D. Shah, A. Spoering, N. Kaldalu, and K. Lewis. 2004. Specialized persister cells and the mechanism of multidrug tolerance in Escherichia coli. J. Bacteriol.186:8172-8180.
128.
Kohler, T., M. Michea-Hamzehpour, P. Plesiat, A. L. Kahr, and J. C. Pechere. 1997. Differential selection of multidrug efflux systems by quinolones in Pseudomonas aeruginosa. Antimicrob. Agents Chemother.41:2540-2543.
129.
Korch, S. B., T. A. Henderson, and T. M. Hill. 2003. Characterization of the hipA7 allele of Escherichia coli and evidence that high persistence is governed by (p)ppGpp synthesis. Mol. Microbiol.50:1199-1213.
130.
Kussell, E., R. Kishony, N. Q. Balaban, and S. Leibler. 2005. Bacterial persistence: a model of survival in changing environments. Genetics169:1807-1814.
131.
Lau, Y. L., G. C. Chan, S. Y. Ha, Y. F. Hui, and K. Y. Yuen. 1998. The role of the phagocytic respiratory burst in host defense against Mycobacterium tuberculosis. Clin. Infect. Dis.26:226-227.
132.
Laurie, A. D., L. M. Bernardo, C. C. Sze, E. Skarfstad, A. Szalewska-Palasz, T. Nyström, and V. Shingler. 2003. The role of the alarmone (p)ppGpp in sigma N competition for core RNA polymerase. J. Biol. Chem.278:1494-1503.
133.
Lee, B. Y., and M. A. Horwitz. 1995. Identification of macrophage and stress-induced proteins of Mycobacterium tuberculosis. J. Clin. Investig.96:245-249.
134.
Lemaitre, N., W. Sougakoff, C. Truffot-Pernot, and V. Jarlier. 1999. Characterization of new mutations in pyrazinamide-resistant strains of Mycobacterium tuberculosis and identification of conserved regions important for the catalytic activity of pyrazinamidase PncA. Antimicrob. Agents Chemother.43:1761-1763.
135.
Levin, B. R. 2002. Models for the spread of resistant pathogens. Neth. J. Med.60:58-64.
136.
Levin, B. R., V. Perrot, and N. Walker. 2000. Compensatory mutations, antibiotic resistance and the population genetics of adaptive evolution in bacteria. Genetics54:985-997.
137.
Lewis, K. 2001. Riddle of biofilm resistance. Antimicrob. Agents Chemother.45:999-1007.
138.
Li, X. Z., and H. Nikaido. 2004. Efflux-mediated drug resistance in bacteria. Drugs64:159-204.
139.
Li, X. Z., L. Zhang, and H. Nikaido. 2004. Efflux pump-mediated intrinsic drug resistance in Mycobacterium smegmatis. Antimicrob. Agents Chemother.48:2415-2423.
140.
Li, Z., C. Kelley, F. Collins, D. Rouse, and S. Morris. 1998. Expression of katG in Mycobacterium tuberculosis is associated with its growth and persistence in mice and guinea pigs. J. Infect. Dis.177:1030-1035.
141.
Lillebaek, T., A. Dirksen, I. Baess, B. Strunge, V. O. Thomsen, and A. B. Andersen. 2002. Molecular evidence of endogenous reactivation of Mycobacterium tuberculosis after 33 years of latent infection. J. Infect. Dis.185:401-404.
142.
Lipsitch, M., and B. R. Levin. 1998. Population dynamics of tuberculosis treatment: mathematical models of the roles of non-compliance and bacterial heterogeneity in the evolution of drug resistance. Int. J. Tuberc. Lung Dis.2:187-199.
143.
Lipsitch, M., C. T. Bergstrom, and B. R. Levin. 2000. The epidemiology of antibiotic resistance in hospitals: paradoxes and prescriptions. Proc. Natl. Acad. Sci. USA97:1938-1943.
144.
Liu, J., H. E. Takiff, and H. Nikaido. 1996. Active efflux of fluoroquinolones in Mycobacterium smegmatis mediated by LfrA, a multidrug efflux pump. J. Bacteriol.178:3791-3795.
145.
Loebel, R. O., E. Shorr, and H. B. Richardson. 1933. The influence of adverse conditions upon the respiratory metabolism and growth of human tubercle bacilli. J. Bacteriol.26:167-200.
146.
MacMicking, J. D., R. J. North, R. LaCourse, J. S. Mudgett, S. K. Shah, and C. F. Nathan. 1997. Identification of nitric oxide synthase as a protective locus against tuberculosis. Proc. Natl. Acad. Sci. USA94:5243-5248.
147.
MacMicking, J., Q. W. Xie, and C. Nathan. 1997. Nitric oxide and macrophage function. Annu. Rev. Immunol.15:323-350.
148.
Maisnier-Patin, S., O. G. Berg, L. Liljas, and D. I. Andersson. 2002. Compensatory adaptation to the deleterious effect of antibiotic resistance in Salmonella typhimurium. Mol. Microbiol.46:355-366.
149.
Makarova, K. S., L. Aravind, N. V. Grishin, I. B. Rogozin, and E. V. Koonin. 2002. A DNA repair system specific for thermophilic Archaea and bacteria predicted by genomic context analysis. Nucleic Acids Res.30:482-496.
150.
Manca, C., S. Paul, C. E. Barry III, V. H. Freedman, and G. Kaplan. 1999. Mycobacterium tuberculosis catalase and peroxidase activities and resistance to oxidative killing in human monocytes in vitro. Infect. Immun.67:74-79.
151.
Mariam, D. H., Y. Mengistu, S. E. Hoffner, and D. I. Andersson. 2004. Effect of rpoB mutations conferring rifampin resistance on fitness of Mycobacterium tuberculosis. Antimicrob. Agents Chemother.48:1289-1294.
152.
Massey, R. C., A. Buckling, and S. J. Peacock. 2001. Phenotypic switching of antibiotic resistance circumvents permanent costs in Staphylococcus aureus. Curr. Biol.11:1810-1814.
153.
Master, S., T. C. Zahrt, J. Song, and V. Deretic. 2001. Mapping of Mycobacterium tuberculosis katG promoters and their differential expression in infected macrophages. J. Bacteriol.183:4033-4039.
154.
Matic, I., M. Radman, F. Taddei, B. Picard, C. Doit, E. Bingen, E. Denamur, and J. Elion. 1997. Highly variable mutation rates in commensal and pathogenic E. coli. Science277:1833-1834.
155.
McCune, R. M., Jr., W. McDermott, and R. Tompsett. 1956. The fate of Mycobacterium tuberculosis in mouse tissues as determined by the microbial enumeration technique. II. The conversion of tuberculous infection to the latent state by the administration of pyrazinamide and a companion drug. J. Exp. Med.104:763-802.
156.
McCune, R. M., Jr., and R. Tompsett. 1956. Fate of Mycobacterium tuberculosis in mouse tissues as determined by the microbial enumeration technique. I. The persistence of drug-susceptible tubercle bacilli in the tissues despite prolonged antimicrobial therapy. J. Exp. Med.104:737-762.
157.
McKenzie, G. J., P. L. Lee, M. J. Lombardo, P. J. Hastings, and S. M. Rosenberg. 2001. SOS mutator DNA polymerase IV functions in adaptive mutation and not adaptive amplification. Mol. Cell7:571-579.
158.
Metzgar, D., and C. Wills. 2000. Evidence for the adaptive evolution of mutation rates. Cell101:581-584.
159.
Middlebrook, G., and M. L. Cohn. 1953. Some observations on the pathogenicity of isoniazid resistant variants of tubercle bacilli. Science118:297-299.
160.
Miller, C., L. E. Thomsen, C. Gaggero, R. Mosseri, H. Ingmer, and S. N. Cohen. 2004. SOS response induction by beta-lactams and bacterial defense against antibiotic lethality. Science305:1629-1631.
161.
Mitchison, D. A. 1992. The Garrod Lecture. Understanding the chemotherapy of tuberculosis—current problems. J. Antimicrob. Chemother.29:477-493.
162.
Mizrahi, V., and S. J. Andersen. 1998. DNA repair in Mycobacterium tuberculosis. What have we learnt from the genome sequence? Mol. Microbiol.29:1331-1339.
163.
Mizrahi, V., M. Buckstein, and H. Rubin. 2005. Nucleic acid metabolism, p. 369-378. In S. T. Cole, K. D. Eisenach, D. N. McMurray, and W. R. Jacobs, Jr. (ed.), Tuberculosis and the tubercle bacillus. ASM Press, Washington, D.C.
164.
Morris, R. P., L. Nguyen, J. Gatfield, K. Visconti, K. Nguyen, D. Schnappinger, S. Ehrt, Y. Liu, L. Heifets, J. Pieters, G. Schoolnik, and C. J. Thompson. 2005. Ancestral antibiotic resistance in Mycobacterium tuberculosis. Proc. Natl. Acad. Sci. USA102:12200-12205.
165.
Moyed, H. S., and K. P. Bertrand. 1983. hipA, a newly recognized gene of Escherichia coli K-12 that affects frequency of persistence after inhibition of murein synthesis. J. Bacteriol.155:768-775.
166.
Muñoz-Elías, E. J., J. Timm, T. Botha, W. T. Chan, J. E. Gomez, and J. D. McKinney. 2005. Replication dynamics of Mycobacterium tuberculosis in chronically infected mice. Infect. Immun.73:546-551.
167.
Musser, J. M. 1995. Antimicrobial agent resistance in mycobacteria: molecular genetic insights. Clin. Microbiol. Rev.8:496-514.
168.
Mwandumba, H. C., D. G. Russell, M. H. Nyirenda, J. Anderson, S. A. White, M. E. Molyneux, and S. B. Squire. 2004. Mycobacterium tuberculosis resides in nonacidified vacuoles in endocytically competent alveolar macrophages from patients with tuberculosis and HIV infection. J. Immunol.172:4592-4598.
169.
Nathan, C. 1992. Nitric oxide a secretory product of mammalian cells. FASEB J.6:3051-3064.
170.
Nathan, C. 2003. Specificity of a third kind: reactive oxygen and nitrogen intermediates in cell signaling. J. Clin. Investig.111:769-778.
171.
Nathan, C., and M. U. Shiloh. 2000. Reactive oxygen and nitrogen intermediates in the relationship between mammalian hosts and microbial pathogens. Proc. Natl. Acad. Sci. USA97:8841-8848.
172.
Newton, G. L., K. Arnold, M. S. Price, C. Sherrill, S. B. del Cardayré, Y. Aharonowitz, G. Cohen, J. Davies, R. C. Fahey, and C. Davis. 1996. Distribution of thiols in microorganisms: mycothiol is a major thiol in most actinomycetes. J. Bacteriol.178:1990-1995.
173.
Ng, V. H., J. S. Cox, A. O. Sousa, J. D. MacMicking, and J. D. McKinney. 2004. Role of KatG catalase-peroxidase in mycobacterial pathogenesis: countering the phagocyte oxidase burst. Mol. Microbiol.52:1291-1302.
174.
Nicholson, S., M. Bonecini-Almeida, J. R. L. Silva, C. Nathan, Q. W. Xie, R. Mumford, J. R. Weidner, J. Calaycay, J. Geng, N. Boechat, C. Linhares, W. Rom, and J. L. Ho. 1996. Inducible nitric oxide synthase in pulmonary alveolar macrophages from patients with tuberculosis. J. Exp. Med.184:2293-2302.
175.
Nikaido, H. 1994. Prevention of drug access to bacterial targets: permeability barriers and active efflux. Science264:382-388.
176.
Nyka, W. 1967. Method for staining both acid-fast and chromophobic tubercle bacilli with carbolfuschin. J. Bacteriol.93:1458-1460.
177.
Nyka, W. 1974. Studies on the effects of starvation on mycobacteria. Infect. Immun.9:843-850.
178.
Ohmori, H., E. C. Friedberg, R. P. Fuchs, M. F. Goodman, F. Hanaoka, D. Hinkle, T. A. Kunkel, C. W. Lawrence, Z. Livneh, T. Nohmi, L. Prakash, S. Prakash, T. Todo, G. C. Walker, Z. Wang, and R. Woodgate. 2001. The Y family of DNA polymerases. Mol. Cell8:7-8.
179.
Ohno, H., G. Zhu, V. P. Mohan, D. Chu, S. Kohno, W. R. Jacobs, Jr., and J. Chan. 2003. The effects of reactive nitrogen intermediates on gene expression in Mycobacterium tuberculosis. Cell. Microbiol.5:637-648.
180.
Oliver, A. R., P. Canton, F. Campo, F. Baquero, and J. Blazquez. 2000. High frequency of hypermutable Pseudomonas aeruginosa in cystic fibrosis lung infection. Science288:1251-1253.
181.
Ordway, D. J., M. G. Sonnenberg, S. A. Donahue, J. T. Belisle, and I. M. Orme. 1995. Drug-resistant strains of Mycobacterium tuberculosis exhibit a range of virulence for mice. Infect. Immun.63:741-743.
182.
Ouellet, H., Y. Ouellet, C. Richard, M. Labarre, B. Wittenberg, J. Wittenberg, and M. Guertin. 2002. Truncated hemoglobin HbN protects Mycobacterium bovis from nitric oxide. Proc. Natl. Acad. Sci. USA99:5902-5907.
183.
Pablos-Mendez, A. 2000. Working alliance for TB drug development, Cape Town, South Africa, February 8th, 2000. Int. J. Tuberc. Lung Dis.4:489-490.
184.
Pandey, D. P., and K. Gerdes. 2005. Toxin-antitoxin loci are highly abundant in free-living but lost from host-associated prokaryotes. Nucleic Acids Res.33:966-976.
185.
Park, H. D., K. M. Guinn, M. I. Harrell, R. Liao, M. I. Voskuil, M. Tompa, G. K. Schoolnik, and D. R. Sherman. 2003. Rv3133c/dosR is a transcription factor that mediates the hypoxic response of Mycobacterium tuberculosis. Mol. Microbiol.48:833-843.
186.
Parsons, L. M., C. S. Jankowski, and K. M. Derbyshire. 1998. Conjugal transfer of chromosomal DNA in Mycobacterium smegmatis. Mol. Microbiol.28:571-582.
187.
Pedersen, K., A. V. Zavialov, M. Y. Pavlov, J. Elf, K. Gerdes, and M. Ehrenberg. 2003. The bacterial toxin RelE displays codon-specific cleavage of mRNAs in the ribosomal A site. Cell112:131-140.
188.
Peh, H. L., A. Toh, B. Murugasu-Oei, and T. Dick. 2001. In vitro activities of mitomycin C against growing and hypoxic dormant bacilli. Antimicrob. Agents Chemother.45:2403-2404.
189.
Phillips, I., E. Culebras, F. Moreno, and F. Baquero. 1987. Induction of the SOS response by new 4-quinolones. J. Antimicrob. Chemother.20:631-638.
190.
Piddock, L. J., K. J. Williams, and V. Ricci. 2000. Accumulation of rifampicin by Mycobacterium aurum,Mycobacterium smegmatis and Mycobacterium tuberculosis. J. Antimcrob. Chemother.45:159-165.
191.
Primm, T., S. Andersen, V. Mizrahi, D. Avarbock, H. Rubin, and C. E. Barry III. 2000. The stringent response of Mycobacterium tuberculosis is required for long-term survival. J. Bacteriol.182:4889-4898.
192.
Pym, A. S., B. Saint-Joanis, and S. T. Cole. 2002. Effect of katG mutations on the virulence of Mycobacterium tuberculosis and the implication for transmission in humans. Infect. Immun.70:4955-4960.
193.
Rad, M. E., P. Bifani, C. Martin, K. Kremer, S. Samper, J. Rauzier, B. Kreiswirth, J. Blazquez, J. Jouan, D. van Soolingen, and B. Gicquel. 2003. Mutations in putative mutator genes of Mycobacterium tuberculosis strains of the W-Beijing family. Emerg. Infect. Dis.9:838-845.
194.
Raman, S., R. Hazra, C. C. Dascher, and R. N. Husson. 2004. Transcription regulation by the Mycobacterium tuberculosis alternative sigma factor SigD and its role in virulence. J. Bacteriol.186:6605-6616.
195.
Ramaswamy, S., and J. M. Musser. 1998. Molecular genetic basis of antimicrobial agent resistance in Mycobacterium tuberculosis: 1998 update. Tuber. Lung Dis.79:111-118.
196.
Rees, R. J. W., and P. D. Hart. 1961. Analysis of the host-parasite equilibrium in chronic murine tuberculosis by total and viable bacillary counts. Br. J. Exp. Pathol.42:83-88.
197.
Ren, L., M. S. Rahman, and M. Z. Humayun. 1999. Escherichia coli cells exposed to streptomycin display a mutator phenotype. J. Bacteriol.181:1043-1044.
198.
Rich, E. A., M. Torres, E. Sada, K. Finegan, B. D. Hamilton, and Z. Toossi. 1997. Mycobacterium tuberculosis (MTB)-stimulated production of nitric oxide by human alveolar macrophages and relationship of nitric oxide production to growth inhibition of MTB. Tuber. Lung Dis.78:247-255.
199.
Roberts, D. M., R. P. Liao, G. Wisedchaisri, W. G. J. Hol, and D. R. Sherman. 2004. Two sensor kinases contribute to the hypoxic response of Mycobacterium tuberculosis. J. Biol. Chem.279:23082-23087.
200.
Ruan, J., G. St. John, S. Ehrt, L. Riley, and C. Nathan. 1999. noxR3, a novel gene from Mycobacterium tuberculosis, protects Salmonella typhimurium from nitrosative and oxidative stress. Infect. Immun.67:3276-3283.
201.
Russell, D. G., H. C. Mwandumba, and E. E. Rhoades. 2002. Mycobacterium and the coat of many lipids. J. Cell Biol.158:421-426.
202.
Sander, P., B. Springer, T. Prammananan, A. Sturmfels, M. Kappler, M. Pletschette, and E. C. Böttger. 2002. Fitness cost of chromosomal drug resistance-conferring mutations. Antimicrob. Agents Chemother.46:1204-1211.
203.
Sander, P., E. C. Böttger, B. Springer, B. Steinmann, M. Rezwan, E. Stavropolous, and M. J. Colston. 2003. A recA deletion mutant of Mycobacterium bovis BCG confers protection equivalent to that of wild type BCG but shows increased genetic stability. Vaccine21:4124-4127.
204.
Sander, P., K. G. Papavinasasundaram, T. Dick, E. Stavropolous, K. Ellrott, B. Springer, M. J. Colston, and E. C. Böttger. 2001. Mycobacterium bovis BCG recA deletion mutant shows increased susceptibility to DNA-damaging agents but wild-type survival in a mouse infection model. Infect. Immun.69:3562-3568.
205.
Sassetti, C. M., and E. J. Rubin. 2003. Genetic requirements for mycobacterial survival during infection. Proc. Natl. Acad. Sci. USA100:12989-12994.
206.
Sassetti, C. M., D. H. Boyd, and E. J. Rubin. 2003. Genes required for mycobacterial growth defined by high-density mutagenesis. Mol. Microbiol.48:77-84.
207.
Schnappinger, D., S. Ehrt, M. I. Voskuil, Y. Liu, J. A. Mangan, I. M. Monahan, G. Dolganov, B. Efron, P. D. Butcher, C. Nathan, and G. K. Schoolnik. 2003. Transcriptional adaptation of Mycobacterium tuberculosis within macrophages: insights into the phagosomal environment. J. Exp. Med.198:693-704.
208.
Segal, W., and H. Bloch. 1956. Biochemical differentiation of Mycobacterium tuberculosis grown in vivo and in vitro. J. Bacteriol.72:132-141.
209.
Sever, J. L., and G. P. Youmans. 1957. Enumeration of viable tubercle bacilli from the organs of nonimmunized and immunized mice. Am. Rev. Tuberc. Pulm. Dis.76:16-635.
210.
Sherman, D. R., K. Mdluli, M. J. Hickey, T. M. Arain, S. L. Morris, C. E. Barry III, and C. K. Stover. 1996. Compensatory ahpC gene expression in isoniazid-resistant Mycobacterium tuberculosis. Science272:1641-1643.
211.
Sherman, D. R., M. Voskuil, D. Schnappinger, R. Liao, M. I. Harrell, and G. K. Schoolnik. 2001. Regulation of the Mycobacterium tuberculosis hypoxic response gene encoding α-crystallin. Proc. Natl. Acad. Sci. USA98:7534-7539.
212.
Sherman, D. R., P. J. Sabo, M. J. Hickey, T. M. Arain, G. G. Mahairas, Y. Yuan, C. E. Barry III, and C. K. Stover. 1995. Disparate responses to oxidative stress in saprophytic and pathogenic mycobacteria. Proc. Natl. Acad. Sci. USA92:6625-6629.
213.
Shi, L., Y. J. Jung, S. Tyagi, M. L. Gennaro, and R. J. North. 2003. Expression of Th1-mediated immunity in mouse lungs induces a Mycobacterium tuberculosis transcription pattern characteristic of nonreplicating persistence. Proc. Natl. Acad. Sci. USA100:241-246.
214.
Silva, P. E., F. Bigi, M. de la Paz Santangelo, M. I. Romano, C. Martin, A. Cataldi, and J. A. Ainsa. 2001. Characterization of P55, a multidrug efflux pump in Mycobacterium bovis and Mycobacterium tuberculosis. Antimicrob. Agents Chemother.45:800-804.
215.
Slayden, R. A., and C. E. Barry III. 2000. The genetics and biochemistry of isoniazid resistance in Mycobacterium tuberculosis. Microbes Infect.2:659-669.
216.
Spoering, A. L., and K. Lewis. 2001. Biofilms and planktonic cells of Pseudomonas aeruginosa have similar resistance to killing by antimicrobials. J. Bacteriol.183:6746-6751.
217.
Springer, B., S. Master, P. Sander, T. Zahrt, M. McFalone, J. Song, K. G. Papavinasasundaram, M. J. Colston, E. Böttger, and V. Deretic. 2001. Silencing of oxidative stress response in Mycobacterium tuberculosis: expression patterns of ahpC in virulent and avirulent strains and effect of ahpC inactivation. Infect. Immun.69:5967-5973.
218.
St. John, G., N. Brot, J. Ruan, H. Erdjument-Bromage, P. Tempst, H. Weissbach, and C. Nathan. 2001. Peptide methionine sulfoxide reductase from Escherichia coli and Mycobacterium tuberculosis protects bacteria against oxidative damage from reactive nitrogen intermediates. Proc. Natl. Acad. Sci. USA98:9901-9906.
219.
Styblo, K. 1991. Epidemiology of tuberculosis. InSelected papers of the Royal Netherlands Tuberculosis Association,vol. 24. KNCV Tuberculosis Foundation Press, The Hague, The Netherlands.
220.
Sung, H. M., G. Yeamans, C. A. Ross, and R. E. Yasbin. 2003. Roles of YqjH and YqjW, homologs of the Escherichia coli UmuC/DinB or Y superfamily of DNA polymerases, in stationary-phase mutagenesis and UV-induced mutagenesis of Bacillus subtilis. J. Bacteriol.185:2153-2160.
221.
Taddei, F., M. Radman, J. Maynard-Smith, B. Toupance, P. H. Gouyon, and B. Godelle. 1997. Role of mutator alleles in adaptive evolution. Nature387:700-702.
222.
Takiff, H. E., M. Cimino, M. C. Musso, T. Weisbrod, R. Martinez, M. B. Delgado, L. Salazar, B. R. Bloom, and W. R. Jacobs, Jr. 1996. Efflux pump of the proton antiporter family confers low-level fluoroquinolone resistance in Mycobacterium smegmatis. Proc. Natl. Acad. Sci. USA93:362-366.
223.
Talaat, A. M., R. Lyons, S. T. Howard, and S. A. Johnston. 2004. The temporal expression profile of Mycobacterium tuberculosis infection in mice. Proc. Natl. Acad. Sci. USA101:4602-4607.
224.
Tegova, R., A. Tover, K. Tarassova, M. Tark, and M. Kivisaar. 2004. Involvement of error-prone DNA polymerase IV in stationary-phase mutagenesis in Pseudomonas putida. J. Bacteriol.186:2735-2744.
225.
Tenaillon, O., E. Denamur, and I. Matic. 2004. Evolutionary significance of stress-induced mutagenesis in bacteria. Trends Microbiol.12:264-270.
226.
Tenaillon, O., F. Taddei, M. Radman, and I. Matic. 2001. Second-order selection in bacterial evolution: selection acting on mutation and recombination rates in the course of adaptation. Res. Microbiol.152:11-16.
227.
Timm, J., F. A. Post, L. G. Bekker, G. B. Walther, H. C. Wainwright, R. Manganelli, W. T. Chan, L. Tsenova, B. Gold, I. Smith, G. Kaplan, and J. D. McKinney. 2003. Differential expression of iron-, carbon-, and oxygen-responsive mycobacterial genes in the lungs of chronically infected mice and tuberculosis patients. Proc. Natl. Acad. Sci. USA100:14321-14326.
228.
Trias, J., and R. Benz. 1994. Permeability of the cell wall of Mycobacterium smegmatis. Mol. Microbiol.14:283-290.
229.
Tsolaki, A. G., A. E. Hirsh, K. DeRiemer, J. A. Enciso, M. Z. Wong, M. Hannan, Y. O. Goguet de la Salmoniere, K. Aman, M. Kato-Maeda, and P. M. Small. 2004. Functional and evolutionary genomics of Mycobacterium tuberculosis: insights from genomic deletions in 100 strains. Proc. Natl. Acad. Sci. USA101:4865-4870.
230.
Tuomanen, E., D. T. Durack, and A. Tomasz. 1986. Antibiotic tolerance among clinical isolates of bacteria. Antimicrob. Agents Chemother.30:521-527.
231.
Venkatesh, R., J. P. Kumar, P. S. Krishna, R. Majunath, and U. Varshney. 2003. Importance of uracil DNA glycosylase in Pseudomonas aeruginosa and Mycobacterium smegmatis, G+C-rich bacteria, in mutation prevention, tolerance to acidified nitrite, and endurance in mouse macrophages. J. Biol. Chem.278:24350-24358.
232.
Visser, J. A. 2002. The fate of microbial mutators. Microbiology148:1247-1252.
233.
Voskuil, M. I. 2004. Mycobacterium tuberculosis gene expression during environmental conditions associated with latency. Tuberculosis84:138-143.
234.
Voskuil, M. I., D. Schnappinger, K. C. Visconti, M. I. Harrell, G. M. Dolganov, D. R. Sherman, and G. K. Schoolnik. 2003. Inhibition of respiration by nitric oxide induces a Mycobacterium tuberculosis dormancy program. J. Exp. Med.198:705-713.
235.
Voskuil, M. I., K. C. Visconti, and G. K. Schoolnik. 2004. Mycobacterium tuberculosis gene expression during adaptation to stationary phase and low-oxygen dormancy. Tuberculosis84:218-227.
236.
Wade, M. M., and Y. Zhang. 2004. Anaerobic incubation conditions enhance pyrazinamide activity against Mycobacterium tuberculosis. J. Med. Microbiol.53:769-773.
237.
Wagner, R. 2002. Regulation of ribosomal RNA synthesis in E. coli: effects of the global regulator guanosine tetraphosphate (ppGpp). J. Mol. Microbiol. Biotechnol.4:331-340.
238.
Wallis, R. S., S. Patil, S. H. Cheon, K. Edmonds, M. Phillips, M. D. Perkins, M. Joloba, A. Namale, J. L. Johnson, L. Teixeira, R. Dietze, S. Siddiqi, R. D. Mugerwa, K. Eisenach, and J. J. Ellner. 1999. Drug tolerance in Mycobacterium tuberculosis. Antimicrob. Agents Chemother.43:2600-2606.
239.
Wang, C. H., C. Y. Liu, H. C. Lin, C. T. Yu, K. F. Chung, and H. P. Kuo. 1998. Increased exhaled nitric oxide in active pulmonary tuberculosis due to inducible NO synthase upregulation in alveolar macrophages. Eur. Respir. J.11:809-815.
240.
Warren, R. M., T. C. Victor, E. M. Streicher, M. Richardson, N. Beyers, N. C. Gey van Pittius, and P. D. van Helden. 2004. Patients with active tuberculosis often have different strains in the same sputum specimen. Am. J. Respir. Care Med.169:610-614.
241.
Wayne, L. G. 1994. Dormancy of Mycobacterium tuberculosis and latency of disease. Eur. J. Clin. Microbiol. Infect. Dis.13:908-914.
242.
Wayne, L. G., and C. D. Sohaskey. 2001. Nonreplicating persistence of Mycobacterium tuberculosis. Annu. Rev. Microbiol.55:139-163.
243.
Wayne, L. G., and H. A. Sramek. 1994. Metronidazole is bactericidal to dormant cells of Mycobacterium tuberculosis. Antimicrob. Agents Chemother.38:2054-2058.
244.
Wayne, L. G., and L. G. Hayes. 1996. An in vitro model for sequential shiftdown of Mycobacterium tuberculosis through two stages of nonreplicating persistence. Infect. Immun.64:2062-2069.
245.
Weldingh, K., I. Rosenkrands, S. Jacobsen, P. B. Rasmussen, M. J. Elhay, and P. Andersen. 1998. Two-dimensional electrophoresis for analysis of Mycobacterium tuberculosis culture filtrate and purification and characterization of six novel proteins. Infect. Immun.66:3492-3500.
246.
Werngren, J., and S. E. Hoffner. 2003. Drug-susceptible Mycobacterium tuberculosis Beijing genotype does not develop mutation-conferred resistance to rifampin at an elevated rate. J. Clin. Microbiol.41:1520-1524.
247.
Wieles, B., S. Nagai, H. G. Wiker, M. Harboe, and T. H. Ottenhoff. 1995. Identification and functional characterization of thioredoxin of Mycobacterium tuberculosis. Infect. Immun.63:4946-4948.
248.
Wilson, M., J. DeRisi, H. H. Kristensen, P. Imboden, S. Rane, P. O. Brown, and G. K. Schoolnik. 1999. Exploring drug-induced alterations in gene expression in Mycobacterium tuberculosis by microarray hybridisation. Proc. Natl. Acad. Sci. USA96:12833-12838.
249.
Wink, D. A., K. S. Kasprzak, C. M. Maragos, R. K. Elespuru, M. Misra, T. M. Dunams, T. A. Cebula, W. H. Koch, A. W. Andrews, J. S. Allen, and L. K. Keefer. 1991. DNA deaminating ability and genotoxicity of nitric oxide and its progenitors. Science254:1001-1003.
250.
Wiuff, C., R. M. Zappala, R. R. Regoes, K. N. Garner, F. Baquero, and B. R. Levin. 2005. Phenotypic tolerance: antibiotic enrichment of noninherited resistance in bacterial populations. Antimicrob. Agents Chemother.49:1483-1494.
251.
World Health Organization. 2004.The WHO/IUATLD Global Project on Anti-Tuberculosis Drug Resistance Surveillance 1999-2002. World Health Organization, Geneva, Switzerland.
252.
World Health Organization. 2005. Global tuberculosis control: surveillance, planning, financing: WHO report 2005. WHO/HTM/TB/2005.349. World Health Organization, Geneva, Switzerland.
253.
Wright, B. E. 2004. Stress-directed adaptive mutations and evolution. Mol. Microbiol.52:643-650.
254.
Xie, Z., N. Siddiqi, and E. J. Rubin. 2005. Differential antibiotic susceptibilities of starved Mycobacterium tuberculosis isolates. Antimicrob. Agents Chemother.49:4778-4780.
255.
Yeiser, B., E. D. Pepper, M. F. Goodman, and S. E. Finkel. 2002. SOS-induced DNA polymerases enhance long-term survival and evolutionary fitness. Proc. Natl. Acad. Sci. USA99:8737-8741.
256.
Zahrt, T. C., and V. Deretic. 2002. Reactive nitrogen and oxygen intermediates and bacterial defenses: unusual adaptations in Mycobacterium tuberculosis. Antioxid. Redox Signal.4:141-159.
257.
Zhang, Y., and D. A. Mitchison. 2003. The curious characteristics of pyrazinamide. Int. J. Tuberc. Lung Dis.7:6-21.
258.
Zhang, Y., J. Zhang, K. P. Hoeflich, M. Ikura, G. Qing, and M. Inouye. 2003. MazF cleaves cellular mRNAs specifically at ACA to block protein synthesis in Escherichia coli. Mol. Cell12:913-923.
259.
Zhang, Y., M. M. Wade, A. Scorpio, H. Zhang, and Z. Sun. 2003. Mode of action of pyrazinamide: disruption of Mycobacterium tuberculosis membrane transport and energetics by pyrazinoic acid. J. Antimicrob. Chemother.52:790-795.
260.
Zhou, J., Y. Dong, X. Zhao, S. Lee, A. Amin, S. Ramaswamy, J. Domagala, J. M. Musser, and K. Drlica. 2000. Selection of antibiotic-resistant bacterial mutants: allelic diversity among fluoroquinolone-resistant mutations. J. Infect. Dis.182:517-525.
261.
Zhu, L., Y. Zhang, J.-S. Teh, J. Zhang, N. Connell, H. Rubin, and M. Inouye. 2006. Characterization of mRNA interferases from Mycobacterium tuberculosis. J. Biol. Chem. [Online.]
262.
Zhuang, J. C., T. L. Wright, T. DeRojas-Walker, S. R. Tannenbaum, and G. N. Wogan. 2000. Nitric oxide-induced mutations in the HPRT gene of human lymphoblastoid TK6 cells and in Salmonella typhimurium. Environ. Mol. Mutagen.35:39-47.

Information & Contributors

Information

Published In

cover image Clinical Microbiology Reviews
Clinical Microbiology Reviews
Volume 19Number 3July 2006
Pages: 558 - 570
PubMed: 16847086

History

Published online: 1 July 2006

Permissions

Request permissions for this article.

Contributors

Authors

Digby F. Warner [email protected]
MRC/NHLS/WITS Molecular Mycobacteriology Research Unit, DST/NRF Centre of Excellence for Biomedical TB Research, School of Pathology, University of the Witwatersrand and National Health Laboratory Service, Johannesburg 2000, South Africa
Valerie Mizrahi [email protected]
MRC/NHLS/WITS Molecular Mycobacteriology Research Unit, DST/NRF Centre of Excellence for Biomedical TB Research, School of Pathology, University of the Witwatersrand and National Health Laboratory Service, Johannesburg 2000, South Africa

Metrics & Citations

Metrics

Note:

  • For recently published articles, the TOTAL download count will appear as zero until a new month starts.
  • There is a 3- to 4-day delay in article usage, so article usage will not appear immediately after publication.
  • Citation counts come from the Crossref Cited by service.

Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. For an editable text file, please select Medlars format which will download as a .txt file. Simply select your manager software from the list below and click Download.

View Options

Figures and Media

Figures

Media

Tables

Share

Share

Share the article link

Share with email

Email a colleague

Share on social media

American Society for Microbiology ("ASM") is committed to maintaining your confidence and trust with respect to the information we collect from you on websites owned and operated by ASM ("ASM Web Sites") and other sources. This Privacy Policy sets forth the information we collect about you, how we use this information and the choices you have about how we use such information.
FIND OUT MORE about the privacy policy