首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 78 毫秒
1.
2.
3.
KatG (catalase-peroxidase) in Mycobacterium tuberculosis is responsible for activation of isoniazid (INH), a pro-drug used to treat tuberculosis infections. Resistance to INH is a global health problem most often associated with mutations in the katG gene. The origin of INH resistance caused by the KatG[S315G] mutant enzyme is examined here. Overexpressed KatG[S315G] was characterized by optical, EPR, and resonance Raman spectroscopy and by studies of the INH activation mechanism in vitro. Catalase activity and peroxidase activity with artificial substrates were moderately reduced (50 and 35%, respectively), whereas the rates of formation of oxyferryl heme:porphyrin π-cation radical and the decay of heme intermediates were ∼2-fold faster in KatG[S315G] compared with WT enzyme. The INH binding affinity for the resting enzyme was unchanged, whereas INH activation, measured by the rate of formation of an acyl-nicotinamide adenine dinucleotide adduct considered to be a bactericidal molecule, was reduced by 30% compared with WT KatG. INH resistance is suggested to arise from a redirection of catalytic intermediates into nonproductive reactions that interfere with oxidation of INH. In the resting mutant enzyme, a rapid evolution of 5-c heme to 6-c species occurred in contrast with the behavior of WT KatG and KatG[S315T] and consistent with greater flexibility at the heme edge in the absence of the hydroxyl of residue 315. Insights into the effects of mutations at residue 315 on enzyme structure, peroxidation kinetics, and specific interactions with INH are presented.Tuberculosis infection kills nearly 2 million people a year and is the leading cause of death due to infectious diseases in adults and in AIDS patients (1). The infection is usually treatable, and isoniazid (isonicotinic acid hydrazide (INH))4 has been a first line antibiotic against Mycobacterium tuberculosis since 1952 (2). The management of the disease is complicated by the fact that bacterial strains have been steadily acquiring and accumulating mutations that confer resistance to INH and other drugs (36). Recently, the appearance of multidrug-resistant tuberculosis, resistant to at least two first line antibiotics, and extensively drug-resistant bacteria (defined as multidrug-resistant tuberculosis plus resistance to at least one fluoroquinolone and at least one of the injectable second line drugs) has made the disease virtually incurable in a growing number of cases (7, 8). Despite the widespread emergence of antibiotic-resistant strains, the molecular mechanisms by which enzyme targets or pro-drug activating enzymes confer resistance are poorly understood.The pro-drug INH requires activation by M. tuberculosis catalase-peroxidase KatG, a heme enzyme classified in the Class I superfamily of fungal, plant, and bacterial peroxidases (9). KatG is important for the virulence of M. tuberculosis due to its role in oxidative stress management (10). This enzyme exhibits both high catalase activity and a broad spectrum peroxidase activity (9, 11) for which a physiologically relevant substrate has not been identified. In vitro, INH is oxidized by KatG (1215) to an acylating species, most likely an acyl radical, that forms an adduct (IN-NAD) when it reacts with NAD+ (16). This modified cofactor then acts as a potent inhibitor of the M. tuberculosis enoyl-acyl carrier protein reductase, InhA, and interferes with cell wall biosynthesis (17, 18). The most common INH resistance mutations in M. tuberculosis clinical isolates occur in katG (19), although mutations in other genes, including inhA, and the promoter for this enzyme (mabA-inhA operon) may cause resistance (2022). Dihydrofolate reductase has also been recently proposed as a target of isoniazid that can be inhibited by an IN-NADP adduct (23, 24). Issues remain to be resolved about INH action as well as resistance in a large set of clinical isolates.Replacements at residue Ser315 are the most commonly encountered in the mutated katG gene of INH-resistant strains (19, 22, 2528). Among these, S315T, which confers high level drug resistance (up to a 200-fold increase in minimum inhibitory concentration (MIC) that kills 50% of bacteria (29)) is the most frequent and is found in more than 50% of INH-resistant isolates of M. tuberculosis. In vitro, this mutant enzyme exhibits a very poor rate of peroxidation/activation of the antibiotic, although the enzyme has close to normal catalase activity and peroxidase activity with substrates other than INH (3032). According to the crystal structure of KatG[S315T] (33), the replacement of serine by threonine leads to a structurally modified substrate access channel. This channel leads from the surface of the enzyme to the heme edge at the propionate of pyrrole IV. Residues Asp137 and Ser315 delimit the narrowest region of the channel, which is reduced in width from 6 to 4.7 Å. The methyl group of threonine effectively restricts accessibility to the heme pocket and apparently interferes with specific interactions required for binding and activation of the drug. Although a binding site for INH in KatG is not specifically defined by x-ray crystallography at this time, a recently reported CCP-INH structure (yeast CCP is a homologous Class I peroxidase) presents what should be an excellent model of drug binding in KatG (34). Hydrogen bonds between the backbone carbonyl of Ser185 (Ser315 in M. tuberculosis KatG), a water molecule, and the pyridine nitrogen of the drug are found in the CCP-INH complex. Thus, it is reasonable that mutations at residue 315 in KatG have an impact on drug binding and activation but little impact on catalase or peroxidase activity with substrates that may not require the same specific interactions as high affinity INH binding.Beyond these studies, there is a substantial gap in the knowledge of the relationship between INH resistance due to the numerous other mutations in the katG gene and the lost drug activation function of the mutant enzymes. The main goal of the present study was to examine KatG[S315G] in vitro. We report the generation, overexpression, purification, and characterization of this enzyme found in clinical isolates of M. tuberculosis having low level INH resistance with MIC values up to 40-fold higher than WT strains (8 μg/μl versus 0.05 μg/μl) (22, 25). An interesting aspect of the problem is that in KatG[S315T], a steric influence on INH binding strongly interferes with activation, whereas resistance is still present with the glycine replacement of serine 315, which would not be assumed to interfere with substrate access or binding at the same locus.The application of optical stopped-flow spectrophotometry, isothermal titration calorimetry (ITC), optical titration, EPR spectroscopy, and rapid freeze-quench EPR (RFQ-EPR) allowed us to probe the functional and structural consequences of the mutation on INH activation. Our results strongly suggest that resistance is due to catalytic changes rather than major changes in specific interactions between the enzyme and INH. Importantly, the results demonstrate the validity of an in vitro INH activation approach used here, since we find a correlation between our observations and the in vivo behavior of INH-resistant M. tuberculosis strains for both KatG[S315T] and KatG[S315G].  相似文献   

4.
Matrix metalloproteinases (MMPs) have been extensively studied because of their functional attributes in development and diseases. However, relatively few in vivo functional studies have been reported on the roles of MMPs in postembryonic organ development. Amphibian metamorphosis is a unique model for studying MMP function during vertebrate development because of its dependence on thyroid hormone (T3) and the ability to easily manipulate this process with exogenous T3. The MMP stromelysin-3 (ST3) is induced by T3, and its expression correlates with cell death during metamorphosis. We have previously shown that ST3 is both necessary and sufficient for larval epithelial cell death in the remodeling intestine. To investigate the roles of ST3 in other organs and especially on different cell types, we have analyzed the effect of transgenic overexpression of ST3 in the tail of premetamorphic tadpoles. We report for the first time that ST3 expression, in the absence of T3, caused significant muscle cell death in the tail of premetamorphic transgenic tadpoles. On the other hand, only relatively low levels of epidermal cell death were induced by precocious ST3 expression in the tail, contrasting what takes place during natural and T3-induced metamorphosis when ST3 expression is high. This cell type-specific apoptotic response to ST3 in the tail suggests distinct mechanisms regulating cell death in different tissues. Furthermore, our analyses of laminin receptor, an in vivo substrate of ST3 in the intestine, suggest that laminin receptor cleavage may be an underlying mechanism for the cell type-specific effects of ST3.The extracellular matrix (ECM),3 the dynamic milieu of the cell microenvironment, plays a critical role in dictating the fate of the cell. The cross-talk between the cell and ECM and the timely catabolism of the ECM are crucial for tissue remodeling during development (1). Matrix metalloproteinases (MMPs), extrinsic proteolytic regulators of the ECM, mediate this process to a large extent. MMPs are a large family of Zn2+-dependent endopeptidases potentially capable of cleaving the extracellular as well as nonextracellular proteins (29). The MMP superfamily includes collagenases, gelatinases, stromelysins, and membrane-type MMPs based on substrate specificity and domain organization (24). MMPs have been implicated to influence a wide range of physiological and pathological processes (1013). The roles of MMPs appear to be very complex. For example, MMPs have been suggested to play roles in both tumor promotion and suppression (1319). Unfortunately, relatively few functional studies have been carried out in vivo, especially in relation to the mechanisms involved during vertebrate development.Amphibian metamorphosis presents a fascinating experimental model to study MMP function during postembryonic development. A unique and salient feature of the metamorphic process is the absolute dependence on the signaling of thyroid hormone (2023). This makes it possible to prevent metamorphosis by simply inhibiting the synthesis of endogenous T3 or to induce precocious metamorphosis by merely adding physiological levels of T3 in the rearing water of premetamorphic tadpoles. Gene expression screens have identified the MMP stromelysin-3 (ST3) as a direct T3 response gene (2427). Expression studies have revealed a distinct spatial and temporal ST3 expression profile in correlation with metamorphic event, especially cell death (25, 2831). Organ culture studies on intestinal remodeling have directly substantiated an essential role of ST3 in larval epithelial cell death and ECM remodeling (32). Furthermore, precocious expression of ST3 alone in premetamorphic tadpoles through transgenesis is sufficient to induce ECM remodeling and larval epithelial apoptosis in the tadpole intestine (33). Thus, ST3 appears to be necessary and sufficient for intestinal epithelial cell death during metamorphosis.ST3 was first isolated as a breast cancer-associated gene (34), and unlike most other MMPs, ST3 is secreted as an active protease through a furin-dependent intracellular activation mechanism (35). Like many other MMPs, ST3 is expressed in a number of pathological processes, including most human carcinomas (11, 3640), as well as in many developmental processes in mammals (10, 34, 4143), although the physiological and pathological roles of ST3 in vivo are largely unknown in mammals. Interestingly, compared with other MMPs, ST3 has only weak activities toward ECM proteins in vitro but stronger activities against non-ECM proteins like α1 proteinase inhibitor and IGFBP-1 (4446). Although ST3 may cleave ECM proteins strongly in the in vivo environment, these findings suggest that the cleavage of non-ECM proteins is likely important for its biological roles. Consistently, we have recently identified a cell surface receptor, laminin receptor (LR) as an in vivo substrate of ST3 in the tadpole intestine during metamorphosis (4749). Analyses of LR expression and cleavage suggest that LR cleavage by ST3 is likely an important mechanism by which ST3 regulates the interaction between the larval epithelial cells and the ECM to induce cell death during intestinal remodeling (47, 48).Here, to investigate the role of ST3 in the apoptosis in other tissues during metamorphosis and whether LR cleavage serves as a mechanism for ST3 to regulate the fate of different cell types, we have analyzed the effects of precocious expression of ST3 in premetamorphic tadpole tail. The tail offers an opportunity to examine the effects of ST3 on different cell types. The epidermis, the fast and slow muscles, and the connective tissue underlying the epidermis in the myotendinous junctions and surrounding the notochord constitute the major tissue types in tail (50). Even though death is the destiny of all these cell types, it is not clear whether they all die through similar or different mechanisms. Microscopic and histochemical analyses have shown that at least the muscle and epidermal cells undergo T3-dependent apoptosis during metamorphosis (23, 29, 51, 52). To study whether ST3 regulates apoptosis of these two cell types, we have made use of the transgenic animals that express a transgenic ST3 under the control of a heat shock-inducible promoter (33). We show that whereas extensive apoptosis is present in both the epidermis and muscles during natural as well as T3-induced metamorphosis, transgenic expression of ST3 induces cell death predominantly in the muscles. Furthermore, we show that LR is expressed in the epidermis and connective tissue but not in muscles of the tadpole tail. More importantly, LR cleavage products are present in the tail during natural metamorphosis but not in transgenic tadpoles overexpressing ST3. These results suggest that ST3 has distinct effects on the epidermis and muscles in the tail, possibly because of the tissue-specific expression and function of LR.  相似文献   

5.
Mycobacterium leprae, which has undergone reductive evolution leaving behind a minimal set of essential genes, has retained intervening sequences in four of its genes implicating a vital role for them in the survival of the leprosy bacillus. A single in-frame intervening sequence has been found embedded within its recA gene. Comparison of the M. leprae recA intervening sequence with the known intervening sequences indicated that it has the consensus amino acid sequence necessary for being a LAGLIDADG-type homing endonuclease. In light of massive gene decay and function loss in the leprosy bacillus, we sought to investigate whether its recA intervening sequence encodes a catalytically active homing endonuclease. Here we show that the purified M. leprae RecA intein (PI-MleI) binds to cognate DNA and displays endonuclease activity in the presence of alternative divalent cations, Mg2+ or Mn2+. A combination of approaches, including four complementary footprinting assays such as DNase I, copper-phenanthroline, methylation protection, and KMnO4, enhancement of 2-aminopurine fluorescence, and mapping of the cleavage site revealed that PI-MleI binds to cognate DNA flanking its insertion site, induces helical distortion at the cleavage site, and generates two staggered double strand breaks. Taken together, these results implicate that PI-MleI possesses a modular structure with separate domains for DNA target recognition and cleavage, each with distinct sequence preferences. From a biological standpoint, it is tempting to speculate that our findings have implications for understanding the evolution of the LAGLIDADG family of homing endonucleases.Mycobacterium leprae, a Gram-positive rod-shaped bacillus, mostly found in warm tropical countries, is the bacterium that causes leprosy in humans (1). The lack of understanding of the basic biology of M. leprae is believed to be the key factor for the failure of leprosy research to advance. The genome sequence of M. leprae contains 3.27 Mb and has an average G + C content of 57.8%, values much lower than the corresponding values for Mycobacterium tuberculosis, which are ∼4.41 Mb and 65.6% G + C, respectively (2). There are some 1500 genes that are common to both M. leprae and M. tuberculosis. The comparative genome analysis suggests that both species of mycobacteria are derived from a common ancestor and, at one stage, had gene pools of similar size. The downsizing of the M. tuberculosis genome from ∼4.41 to 3.27 Mb of M. leprae would account for the loss of some 1200 protein-coding sequences (1, 3). There is evidence that many of the genes that were present in the genome of M. leprae have truly been lost (1, 3). Comparative genomics of M. leprae with that of M. tuberculosis indicate that the former has undergone substantial downsizing, losing more than 2000 genes, thus suggesting an extreme case of reductive evolution in a microbial pathogen (1). With the availability of the M. leprae genome sequence, using functional genomics approaches, it is possible to identify the gene products, elucidate the mechanism of their action, and identify novel drug targets for rational design of new therapeutic regimens and drugs to treat leprosy.Eubacterial RecA proteins catalyze a set of biochemical reactions that are essential for homologous recombination, DNA repair, restoration of stalled replication forks, and SOS response (47). RecA protein and the process of homologous recombination, which is the main mechanism of genetic exchange, are evolutionarily conserved among a range of organisms (4, 7). Perhaps the most striking development in the field of RecA protein biology was the discovery of an in-frame insertion of an intein-coding sequence in the recA genes of M. tuberculosis and M. leprae (8, 9). In these organisms, RecA is synthesized as a large precursor, which undergoes protein splicing to excise the intein, and the two flanking domains called exteins are ligated together to generate a functionally active RecA protein (9, 10). The milieu in which RecA precursor undergoes splicing differs substantially between M. tuberculosis and M. leprae. M. leprae RecA precursor (79 kDa) undergoes splicing only in mycobacterial species, whereas M. tuberculosis RecA precursor (85 kDa) is spliced efficiently in Escherichia coli as well (911). Intriguingly, M. tuberculosis and M. leprae RecA inteins differ greatly in their size, primary sequence, and location within the recA gene, thereby suggesting two independent origins during evolution (9). The occurrence of inteins in the obligate mycobacterial pathogens, M. tuberculosis, M. leprae, and Mycobacterium microti, suggested that RecA inteins might play a role in mycobacterial functions related to pathogenesis or virulence (9). Previously, we have shown that M. tuberculosis RecA intein (PI-MtuI),2 which contains Walker A motif, displays dual target specificity in the presence of alternative cofactors in an ATP-dependent manner (12, 13).Since their discovery in Saccharomyces cerevisiae (14, 15), a large number of putative homing endonucleases have been found in a diverse range of proteins in all the three domains of life (1619). The majority of inteins possess the protein splicing and homing endonuclease activities (18, 19). Homing endonucleases are a class of diverse rare-cutting enzymes that promote site-specific transposition of their encoding genetic elements by inflicting double-stranded DNA breaks via different cleavage mechanisms in alleles lacking these elements (1823). In addition, these are characterized by their ability to bind long DNA target sites (14–40 bp), and their tolerance of minor sequence changes in their binding region. These have been divided into highly divergent subfamilies on the basis of conserved sequence and structural motifs as follows: LAGLIDADG, GIY-YIG, HNH, His-Cys box, and the more recently identified PD(D/E)XK families (1824). LAGLIDADG homing enzymes, which include the largest family, contain one or two copies of the conserved dodecapeptide motif and utilize an extended protein-DNA interface covering up to 40 bp to acquire their necessary specificity (1822). The LAGLIDADG sequence is a part of the conserved 10- or 12-residue sequence motif defining the family of LAGLIDADG-type homing endonucleases; therefore, it is designated as deca- or dodecapeptide motif (19).Comparison of the M. leprae recA intervening sequence with known intervening sequences indicated that it has the consensus amino acid sequence necessary for being a LAGLIDADG-type homing endonuclease (25, 26). In light of massive gene decay and function loss in the leprosy bacillus, and dissimilarities in size and primary structures among mycobacterial inteins, we sought to investigate whether M. leprae recA intervening sequence encodes a catalytically active homing endonuclease. In this study, we show that the purified M. leprae RecA intein (PI-MleI) binds to cognate DNA and displays endonuclease activity in the presence of alternative divalent cations Mg2+ or Mn2+. Furthermore, using a variety of approaches, we have mapped the positions of PI-MleI binding as well as cleavage in the cognate DNA, thus providing the most comprehensive analysis of PI-MleI. Taken together, these results suggest that PI-MleI possesses a modular structure with functionally separable domains for DNA target recognition and cleavage, each with distinct sequence preferences. These results provide insights into understanding the function and evolution of the family of LAGLIDADG homing endonucleases.  相似文献   

6.
The thrombin mutant W215A/E217A (WE) is a potent anticoagulant both in vitro and in vivo. Previous x-ray structural studies have shown that WE assumes a partially collapsed conformation that is similar to the inactive E* form, which explains its drastically reduced activity toward substrate. Whether this collapsed conformation is genuine, rather than the result of crystal packing or the mutation introduced in the critical 215–217 β-strand, and whether binding of thrombomodulin to exosite I can allosterically shift the E* form to the active E form to restore activity toward protein C are issues of considerable mechanistic importance to improve the design of an anticoagulant thrombin mutant for therapeutic applications. Here we present four crystal structures of WE in the human and murine forms that confirm the collapsed conformation reported previously under different experimental conditions and crystal packing. We also present structures of human and murine WE bound to exosite I with a fragment of the platelet receptor PAR1, which is unable to shift WE to the E form. These structural findings, along with kinetic and calorimetry data, indicate that WE is strongly stabilized in the E* form and explain why binding of ligands to exosite I has only a modest effect on the E*-E equilibrium for this mutant. The E* → E transition requires the combined binding of thrombomodulin and protein C and restores activity of the mutant WE in the anticoagulant pathway.Thrombin is the pivotal protease of blood coagulation and is endowed with both procoagulant and anticoagulant roles in vivo (1). Thrombin acts as a procoagulant when it converts fibrinogen into an insoluble fibrin clot, activates clotting factors V, VIII, XI, and XIII, and cleaves PAR12 and PAR4 on the surface of human platelets thereby promoting platelet aggregation (2). Upon binding to thrombomodulin, a receptor present on the membrane of endothelial cells, thrombin becomes unable to interact with fibrinogen and PAR1 but increases >1,000-fold its activity toward the zymogen protein C (3). Activated protein C generated from the thrombin-thrombomodulin complex down-regulates both the amplification and progression of the coagulation cascade (3) and acts as a potent cytoprotective agent upon engagement of EPCR and PAR1 (4).The dual nature of thrombin has long motivated interest in dissociating its procoagulant and anticoagulant activities (512). Thrombin mutants with anticoagulant activity help rationalize the bleeding phenotypes of several naturally occurring mutations and could eventually provide new tools for pharmacological intervention (13) by exploiting the natural protein C pathway (3, 14, 15). Previous mutagenesis studies have led to the identification of the E217A and E217K mutations that significantly shift thrombin specificity from fibrinogen toward protein C relative to the wild type (1012). Both constructs were found to display anticoagulant activity in vivo (10, 12). The subsequent discovery of the role of Trp-215 in controlling the balance between pro- and anti-coagulant activities of thrombin (16) made it possible to construct the double mutant W215A/E217A (WE) featuring >19,000-fold reduced activity toward fibrinogen but only 7-fold loss of activity toward protein C (7). These properties make WE the most potent anticoagulant thrombin mutant engineered to date and a prototype for a new class of anticoagulants (13). In vivo studies have revealed an extraordinary potency, efficacy, and safety profile of WE when compared with direct administration of activated protein C or heparin (1719). Importantly, WE elicits cytoprotective effects (20) and acts as an antithrombotic by antagonizing the platelet receptor GpIb in its interaction with von Willebrand factor (21).What is the molecular mechanism underscoring the remarkable functional properties of WE? The mutant features very low activity toward synthetic and physiological substrates, including protein C. However, in the presence of thrombomodulin, protein C is activated efficiently (7). A possible explanation is that WE assumes an inactive conformation when free but is converted into an active form in the presence of thrombomodulin. The ability of WE to switch from inactive to active forms is consistent with recent kinetic (22) and structural (23, 24) evidence of the significant plasticity of the trypsin fold. The active form of the protease, E, coexists with an inactive form, E*, that is distinct from the zymogen conformation (25). Biological activity of the protease depends on the equilibrium distribution of E* and E, which is obviously different for different proteases depending on their physiological role and environmental conditions (25). The E* form features a collapse of the 215–217 β-strand into the active site and a flip of the peptide bond between residues Glu-192 and Gly-193, which disrupts the oxyanion hole. These changes have been documented crystallographically in thrombin and other trypsin-like proteases such as αI-tryptase (26), the high temperature requirement-like protease (27), complement factor D (28), granzyme K (29), hepatocyte growth factor activator (30), prostate kallikrein (31), prostasin (32, 33), complement factor B (34), and the arterivirus protease nsp4 (35). Hence, the questions that arise about the molecular mechanism of WE function are whether the mutant is indeed stabilized in the inactive E* form and whether it can be converted to the active E form upon thrombomodulin binding.Structural studies of the anticoagulant mutants E217K (36) and WE (37) show a partial collapse of the 215–217 β-strand into the active site that abrogates substrate binding. The collapse is similar to, but less pronounced than, that observed in the structure of the inactive E* form of thrombin where Trp-215 relinquishes its hydrophobic interaction with Phe-227 to engage the catalytic His-57 and residues of the 60-loop after a 10 Å shift in its position (24). These more substantial changes have been observed recently in the structure of the anticoagulant mutant Δ146–149e (38), which has proved that stabilization of E* is indeed a molecular mechanism capable of switching thrombin into an anticoagulant. It would be simple to assume that both E217K and WE, like Δ146–149e, are stabilized in the E* form. However, unlike Δ146–149e, both E217K and WE carry substitutions in the critical 215–217 β-strand that could result into additional functional effects overlapping with or mimicking a perturbation of the E*-E equilibrium. A significant concern is that both structures suffer from crystal packing interactions that may have biased the conformation of side chains and loops near the active site (24). The collapsed structures of E217K and WE may be artifactual unless validated by additional structural studies where crystal packing is substantially different.To address the second question, kinetic measurements of chromogenic substrate hydrolysis by WE in the presence of saturating amounts of thrombomodulin have been carried out (37), but these show only a modest improvement of the kcat/Km as opposed to >57,000-fold increase observed when protein C is used as a substrate (7, 37). The modest effect of thrombomodulin on the hydrolysis of chromogenic substrates is practically identical to that seen upon binding of hirugen to exosite I (37) and echoes the results obtained with the wild type (39) and other anticoagulant thrombin mutants (7, 9, 10, 12, 38). That argues against the ability of thrombomodulin alone to significantly shift the E*-E equilibrium in favor of the E form. Binding of a fragment of the platelet receptor PAR1 to exosite I in the D102N mutant stabilized in the E* form (24) does trigger the transition to the E form (23), but evidence that a similar long-range effect exists for the E217K or WE mutants has not been presented.In this study we have addressed the two unresolved questions about the mechanism of action of the anticoagulant thrombin mutant WE. Here we present new structures of the mutant in its human and murine versions, free and bound to a fragment of the thrombin receptor PAR1 at exosite I. The structures are complemented by direct energetic assessment of the binding of ligands to exosite I and its effect on the E*-E equilibrium.  相似文献   

7.
The general stress response of Bacillus subtilis can be activated by stimuli such as the addition of salt or ethanol and with blue light. In the latter response, YtvA activates σB through a cascade of Rsb proteins, organized in stressosomes. YtvA is composed of an N-terminal LOV (light, oxygen, and voltage) domain and a C-terminal STAS (sulfate transporter and anti-sigma factor) domain and shows light-modulated GTP binding in vitro. Here, we examine the mechanism of YtvA-mediated activation of σB in vivo with site-directed mutagenesis. Constitutive off and constitutive on mutations have been identified. Disruption of GTP binding in the STAS domain eliminates light activation of σB. In contrast, modification of sites relevant for phosphorylation of STAS domains does not affect the stress response significantly. The data obtained are integrated into a model for the structure of full-length YtvA, which presumably functions as a dimer.LOV2 domains (1), members of the superfamily of PAS domains (2, 3), are abundant in all domains of life and were first identified in plant phototropins (4). These photoreceptors regulate stomatal opening, phototropism, etc. and contain two N-terminal LOV domains that confer light regulation on the C-terminal Ser/Thr kinase domain (4). They also occur in bacteria, in which YtvA from Bacillus subtilis has been best characterized (for a review, see e.g. Ref. 5). Its N-terminal LOV domain binds FMN and shows the typical LOV photochemistry (6, 7): covalent adduct formation between a cysteine and the FMN chromophore. A linker helix, denoted Jα (7), connects the LOV domain to a STAS domain. The latter domain is present in many regulators of the general stress response of B. subtilis (8, 9). Stress via the addition of salt or ethanol (for a review, see Ref. 10) and blue light (11, 12) activates the general stress response via the environmental pathway, which integrates various signals via a large multiprotein complex, called the stressosome (13, 14). YtvA, which mediates light activation of σB (11, 12, 15), co-purifies with other STAS domain proteins in the stressosomes (16).When cells are stressed, STAS domains of several stressosome proteins (e.g. RsbS and RsbR) are phosphorylated by another intrinsic stressosome component, the serine/threonine kinase RsbT (9, 14, 17, 18). Next, RsbT is released from the complex to trigger RsbU, a protein phosphatase, thus (indirectly) activating σB (19). Phosphorylation of YtvA, however, has never been detected. Rather, it has been demonstrated in vitro that YtvA shows light-dependent GTP binding, presumably at its NTP-binding site in the STAS domain (20).Little is known about the mechanism of signal transmission in and by YtvA, except that in the C62A mutant, photochemistry in vitro (12) and light activation of σB in vivo (12, 15) are abolished. More detailed information is available for LOV domains of phototropins. A conserved glutamine, which is in hydrogen-bonding contact with the isoalloxazine ring of FMN, rotates its side chain by 180° upon covalent adduct formation (21). Replacement of this residue by leucine in the LOV2 domain of Phy3 from Adiantum results in a considerable reduction of the light-induced structural change (22). The corresponding mutation in phototropin 1 from Arabidopsis impairs autophosphorylation activity (23). The signal generated in the LOV2 domain is transmitted to the downstream kinase domain of phototropin 1 of Avena sativa through disruption of the interaction between its central β-sheet and the C-terminal linker region, the Jα-helix (24).Here, we study the mechanism of activation of YtvA in vivo, i.e. light-induced activation of the σB response, with site-directed mutagenesis. We focus on three regions of the protein, the flavin-binding pocket, the β-sheet of the LOV domain, and the GTP-binding site, and on potential phosphorylation sites of the STAS domain. We demonstrate that light-activated GTP binding is crucial for functional YtvA. A computational approach was used to model the structure of full-length YtvA. The model suggests that light modulates accessibility of the GTP-binding site of the STAS domain of YtvA.  相似文献   

8.
Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.  相似文献   

9.
Paneth cells are a secretory epithelial lineage that release dense core granules rich in host defense peptides and proteins from the base of small intestinal crypts. Enteric α-defensins, termed cryptdins (Crps) in mice, are highly abundant in Paneth cell secretions and inherently resistant to proteolysis. Accordingly, we tested the hypothesis that enteric α-defensins of Paneth cell origin persist in a functional state in the mouse large bowel lumen. To test this idea, putative Crps purified from mouse distal colonic lumen were characterized biochemically and assayed in vitro for bactericidal peptide activities. The peptides comigrated with cryptdin control peptides in acid-urea-PAGE and SDS-PAGE, providing identification as putative Crps. Matrix-assisted laser desorption ionization time-of-flight mass spectrometry experiments showed that the molecular masses of the putative α-defensins matched those of the six most abundant known Crps, as well as N-terminally truncated forms of each, and that the peptides contain six Cys residues, consistent with identities as α-defensins. N-terminal sequencing definitively revealed peptides with N termini corresponding to full-length, (des-Leu)-truncated, and (des-Leu-Arg)-truncated N termini of Crps 1–4 and 6. Crps from mouse large bowel lumen were bactericidal in the low micromolar range. Thus, Paneth cell α-defensins secreted into the small intestinal lumen persist as intact and functional forms throughout the intestinal tract, suggesting that the peptides may mediate enteric innate immunity in the colonic lumen, far from their upstream point of secretion in small intestinal crypts.Antimicrobial peptides (AMPs)2 are released by epithelial cells onto mucosal surfaces as effectors of innate immunity (15). In mammals, most AMPs derive from two major families, the cathelicidins and defensins (6). The defensins comprise the α-, β-, and θ-defensin subfamilies, which are defined by the presence of six cysteine residues paired in characteristic tridisulfide arrays (7). α-Defensins are highly abundant in two primary cell lineages: phagocytic leukocytes, primarily neutrophils, of myeloid origin and Paneth cells, which are secretory epithelial cells located at the base of the crypts of Lieberkühn in the small intestine (810). Neutrophil α-defensins are stored in azurophilic granules and contribute to non-oxidative microbial cell killing in phagolysosomes (11, 12), except in mice whose neutrophils lack defensins (13). In the small bowel, α-defensins and other host defense proteins (1418) are released apically as components of Paneth cell secretory granules in response to cholinergic stimulation and after exposure to bacterial antigens (19). Therefore, the release of Paneth cell products into the crypt lumen is inferred to protect mitotically active crypt cells from colonization by potential pathogens and confer protection against enteric infection (7, 20, 21).Under normal, homeostatic conditions, Paneth cells are not found outside the small bowel, although they may appear ectopically in response to local inflammation throughout the gastrointestinal tract (22, 23). Paneth cell numbers increase progressively throughout the small intestine, occurring at highest numbers in the distal ileum (24). Mouse Paneth cells express numerous α-defensin isoforms, termed cryptdins (Crps) (25), that have broad spectrum antimicrobial activities (6, 26). Collectively, α-defensins constitute approximately seventy percent of the bactericidal peptide activity in mouse Paneth cell secretions (19), selectively killing bacteria by membrane-disruptive mechanisms (2730). The role of Paneth cell α-defensins in gastrointestinal mucosal immunity is evident from studies of mice transgenic for human enteric α-defensin-5, HD-5, which are immune to infection by orally administered Salmonella enterica sv. typhimurium (S. typhimurium) (31).The biosynthesis of mature, bactericidal α-defensins from their inactive precursors requires activation by lineage-specific proteolytic convertases. In mouse Paneth cells, inactive ∼8.4-kDa Crp precursors are processed intracellularly into microbicidal ∼4-kDa Crps by specific cleavage events mediated by matrix metalloproteinase-7 (MMP-7) (32, 33). MMP-7 null mice exhibit increased susceptibility to systemic S. typhimurium infection and decreased clearance of orally administered non-invasive Escherichia coli (19, 32). Although the α-defensin proregions are sensitive to proteolysis, the mature, disulfide-stabilized peptides resist digestion by their converting enzymes in vitro, whether the convertase is MMP-7 (32), trypsin (34), or neutrophil serine proteinases (35). Because α-defensins resist proteolysis in vitro, we hypothesized that Paneth cell α-defensins resist degradation and remain in a functional state in the large bowel, a complex, hostile environment containing varied proteases of both host and microbial origin.Here, we report on the isolation and characterization of a population of enteric α-defensins from the mouse colonic lumen. Full-length and N-terminally truncated Paneth cell α-defensins were identified and are abundant in the distal large bowel lumen.  相似文献   

10.
11.
12.
13.
14.
15.
Leptospira spp., the causative agents of leptospirosis, adhere to components of the extracellular matrix, a pivotal role for colonization of host tissues during infection. Previously, we and others have shown that Leptospira immunoglobulin-like proteins (Lig) of Leptospira spp. bind to fibronectin, laminin, collagen, and fibrinogen. In this study, we report that Leptospira can be immobilized by human tropoelastin (HTE) or elastin from different tissues, including lung, skin, and blood vessels, and that Lig proteins can bind to HTE or elastin. Moreover, both elastin and HTE bind to the same LigB immunoglobulin-like domains, including LigBCon4, LigBCen7′–8, LigBCen9, and LigBCen12 as demonstrated by enzyme-linked immunosorbent assay (ELISA) and competition ELISAs. The LigB immunoglobulin-like domain binds to the 17th to 27th exons of HTE (17–27HTE) as determined by ELISA (LigBCon4, KD = 0.50 μm; LigBCen7′–8, KD = 0.82 μm; LigBCen9, KD = 1.54 μm; and LigBCen12, KD = 0.73 μm). The interaction of LigBCon4 and 17–27HTE was further confirmed by steady state fluorescence spectroscopy (KD = 0.49 μm) and ITC (KD = 0.54 μm). Furthermore, the binding was enthalpy-driven and affected by environmental pH, indicating it is a charge-charge interaction. The binding affinity of LigBCon4D341N to 17–27HTE was 4.6-fold less than that of wild type LigBCon4. In summary, we show that Lig proteins of Leptospira spp. interact with elastin and HTE, and we conclude this interaction may contribute to Leptospira adhesion to host tissues during infection.Pathogenic Leptospira spp. are spirochetes that cause leptospirosis, a serious infectious disease of people and animals (1, 2). Weil syndrome, the severe form of leptospiral infection, leads to multiorgan damage, including liver failure (jaundice), renal failure (nephritis), pulmonary hemorrhage, meningitis, abortion, and uveitis (3, 4). Furthermore, this disease is not only prevalent in many developing countries, it is reemerging in the United States (3). Although leptospirosis is a serious worldwide zoonotic disease, the pathogenic mechanisms of Leptospira infection remain enigmatic. Recent breakthroughs in applying genetic tools to Leptospira may facilitate studies on the molecular pathogenesis of leptospirosis (58).The attachment of pathogenic Leptospira spp. to host tissues is critical in the early phase of Leptospira infection. Leptospira spp. adhere to host tissues to overcome mechanical defense systems at tissue surfaces and to initiate colonization of specific tissues, such as the lung, kidney, and liver. Leptospira invade hosts tissues through mucous membranes or injured epidermis, coming in contact with subepithelial tissues. Here, certain bacterial outer surface proteins serve as microbial surface components recognizing adhesive matrix molecules (MSCRAMMs)2 to mediate the binding of bacteria to different extracellular matrices (ECMs) of host cells (9). Several leptospiral MSCRAMMs have been identified (1018), and we speculate that more will be identified in the near future.Lig proteins are distributed on the outer surface of pathogenic Leptospira, and the expression of Lig protein is only found in low passage strains (14, 16, 17), probably induced by environmental cues such as osmotic or temperature changes (19). Lig proteins can bind to fibrinogen and a variety of ECMs, including fibronectin (Fn), laminin, and collagen, thereby mediating adhesion to host cells (2023). Lig proteins also constitute good vaccine candidates (2426).Elastin is a component of ECM critical to tissue elasticity and resilience and is abundant in skin, lung, blood vessels, placenta, uterus, and other tissues (2729). Tropoelastin is the soluble precursor of elastin (28). During the major phase of elastogenesis, multiple tropoelastin molecules associate through coacervation (3032). Because of the abundance of elastin or tropoelastin on the surface of host cells, several bacterial MSCRAMMs use elastin and/or tropoelastin to mediate adhesion during the infection process (3335).Because leptospiral infection is known to cause severe pulmonary hemorrhage (36, 37) and abortion (38), we hypothesize that some leptospiral MSCRAMMs may interact with elastin and/or tropoelastin in these elastin-rich tissues. This is the first report that Lig proteins of Leptospira interact with elastin and tropoelastin, and the interactions are mediated by several specific immunoglobulin-like domains of Lig proteins, including LigBCon4, LigBCen7′–8, LigBCen9, and LigBCen12, which bind to the 17th to 27th exons of human tropoelastin (HTE).  相似文献   

16.
17.
Helicobacter pylori CagA plays a key role in gastric carcinogenesis. Upon delivery into gastric epithelial cells, CagA binds and deregulates SHP-2 phosphatase, a bona fide oncoprotein, thereby causing sustained ERK activation and impaired focal adhesions. CagA also binds and inhibits PAR1b/MARK2, one of the four members of the PAR1 family of kinases, to elicit epithelial polarity defect. In nonpolarized gastric epithelial cells, CagA induces the hummingbird phenotype, an extremely elongated cell shape characterized by a rear retraction defect. This morphological change is dependent on CagA-deregulated SHP-2 and is thus thought to reflect the oncogenic potential of CagA. In this study, we investigated the role of the PAR1 family of kinases in the hummingbird phenotype. We found that CagA binds not only PAR1b but also other PAR1 isoforms, with order of strength as follows: PAR1b > PAR1d ≥ PAR1a > PAR1c. Binding of CagA with PAR1 isoforms inhibits the kinase activity. This abolishes the ability of PAR1 to destabilize microtubules and thereby promotes disassembly of focal adhesions, which contributes to the hummingbird phenotype. Consistently, PAR1 knockdown potentiates induction of the hummingbird phenotype by CagA. The morphogenetic activity of CagA was also found to be augmented through inhibition of non-muscle myosin II. Because myosin II is functionally associated with PAR1, perturbation of PAR1-regulated myosin II by CagA may underlie the defect of rear retraction in the hummingbird phenotype. Our findings reveal that CagA systemically inhibits PAR1 family kinases and indicate that malfunctioning of microtubules and myosin II by CagA-mediated PAR1 inhibition cooperates with deregulated SHP-2 in the morphogenetic activity of CagA.Infection with Helicobacter pylori strains bearing cagA (cytotoxin-associated gene A)-positive strains is the strongest risk factor for the development of gastric carcinoma, the second leading cause of cancer-related death worldwide (13). The cagA gene is located within a 40-kb DNA fragment, termed the cag pathogenicity island, which is specifically present in the genome of cagA-positive H. pylori strains (46). In addition to cagA, there are ∼30 genes in the cag pathogenicity island, many of which encode a bacterial type IV secretion system that delivers the cagA-encoded CagA protein into gastric epithelial cells (710). Upon delivery into gastric epithelial cells, CagA is localized to the plasma membrane, where it undergoes tyrosine phosphorylation at the C-terminal Glu-Pro-Ile-Tyr-Ala motifs by Src family kinases or c-Abl kinase (1114). The C-terminal Glu-Pro-Ile-Tyr-Ala-containing region of CagA is noted for the structural diversity among distinct H. pylori isolates. Oncogenic potential of CagA has recently been confirmed by a study showing that systemic expression of CagA in mice induces gastrointestinal and hematological malignancies (15).When expressed in gastric epithelial cells, CagA induces morphological transformation termed the hummingbird phenotype, which is characterized by the development of one or two long and thin protrusions resembling the beak of the hummingbird. It has been thought that the hummingbird phenotype is related to the oncogenic action of CagA (7, 1619). Pathophysiological relevance for the hummingbird phenotype in gastric carcinogenesis has recently been provided by the observation that infection with H. pylori carrying CagA with greater ability to induce the hummingbird phenotype is more closely associated with gastric carcinoma (2023). Elevated motility of hummingbird cells (cells showing the hummingbird phenotype) may also contribute to invasion and metastasis of gastric carcinoma.In host cells, CagA interacts with the SHP-2 phosphatase, C-terminal Src kinase, and Crk adaptor in a tyrosine phosphorylation-dependent manner (16, 24, 25) and also associates with Grb2 adaptor and c-Met in a phosphorylation-independent manner (26, 27). Among these CagA targets, much attention has been focused on SHP-2 because the phosphatase has been recognized as a bona fide oncoprotein, gain-of-function mutations of which are found in various human malignancies (17, 18, 28). Stable interaction of CagA with SHP-2 requires CagA dimerization, which is mediated by a 16-amino acid CagA-multimerization (CM)2 sequence present in the C-terminal region of CagA (29). Upon complex formation, CagA aberrantly activates SHP-2 and thereby elicits sustained ERK MAP kinase activation that promotes mitogenesis (30). Also, CagA-activated SHP-2 dephosphorylates and inhibits focal adhesion kinase (FAK), causing impaired focal adhesions. It has been shown previously that both aberrant ERK activation and FAK inhibition by CagA-deregulated SHP-2 are involved in induction of the hummingbird phenotype (31).Partitioning-defective 1 (PAR1)/microtubule affinity-regulating kinase (MARK) is an evolutionally conserved serine/threonine kinase originally isolated in C. elegans (3234). Mammalian cells possess four structurally related PAR1 isoforms, PAR1a/MARK3, PAR1b/MARK2, PAR1c/MARK1, and PAR1d/MARK4 (3537). Among these, PAR1a, PAR1b, and PAR1c are expressed in a variety of cells, whereas PAR1d is predominantly expressed in neural cells (35, 37). These PAR1 isoforms phosphorylate microtubule-associated proteins (MAPs) and thereby destabilize microtubules (35, 38), allowing asymmetric distribution of molecules that are involved in the establishment and maintenance of cell polarity.In polarized epithelial cells, CagA disrupts the tight junctions and causes loss of apical-basolateral polarity (39, 40). This CagA activity involves the interaction of CagA with PAR1b/MARK2 (19, 41). CagA directly binds to the kinase domain of PAR1b in a tyrosine phosphorylation-independent manner and inhibits the kinase activity. Notably, CagA binds to PAR1b via the CM sequence (19). Because PAR1b is present as a dimer in cells (42), CagA may passively homodimerize upon complex formation with the PAR1 dimer via the CM sequence, and this PAR1-directed CagA dimer would form a stable complex with SHP-2 through its two SH2 domains.Because of the critical role of CagA in gastric carcinogenesis (7, 1619), it is important to elucidate the molecular basis underlying the morphogenetic activity of CagA. In this study, we investigated the role of PAR1 isoforms in induction of the hummingbird phenotype by CagA, and we obtained evidence that CagA-mediated inhibition of PAR1 kinases contributes to the development of the morphological change by perturbing microtubules and non-muscle myosin II.  相似文献   

18.
19.
Most human genes undergo alternative splicing, but aberrant splice forms are hallmarks of many cancers, usually resulting from mutations initiating abnormal exon skipping, intron retention, or the introduction of a new splice sites. We have identified a family of aberrant splice variants of HAS1 (the hyaluronan synthase 1 gene) in some B lineage cancers, characterized by exon skipping and/or partial intron retention events that occur either together or independently in different variants, apparently due to accumulation of inherited and acquired mutations. Cellular, biochemical, and oncogenic properties of full-length HAS1 (HAS1-FL) and HAS1 splice variants Va, Vb, and Vc (HAS1-Vs) are compared and characterized. When co-expressed, the properties of HAS1-Vs are dominant over those of HAS1-FL. HAS1-FL appears to be diffusely expressed in the cell, but HAS1-Vs are concentrated in the cytoplasm and/or Golgi apparatus. HAS1-Vs synthesize detectable de novo HA intracellularly. Each of the HAS1-Vs is able to relocalize HAS1-FL protein from diffuse cytoskeleton-anchored locations to deeper cytoplasmic spaces. This HAS1-Vs-mediated relocalization occurs through strong molecular interactions, which also serve to protect HAS1-FL from its otherwise high turnover kinetics. In co-transfected cells, HAS1-FL and HAS1-Vs interact with themselves and with each other to form heteromeric multiprotein assemblies. HAS1-Vc was found to be transforming in vitro and tumorigenic in vivo when introduced as a single oncogene to untransformed cells. The altered distribution and half-life of HAS1-FL, coupled with the characteristics of the HAS1-Vs suggest possible mechanisms whereby the aberrant splicing observed in human cancer may contribute to oncogenesis and disease progression.About 70–80% of human genes undergo alternative splicing, contributing to proteomic diversity and regulatory complexities in normal development (1). About 10% of mutations listed so far in the Human Gene Mutation Database (HGMD) of “gene lesions responsible for human inherited disease” were found to be located within splice sites. Furthermore, it is becoming increasingly apparent that aberrant splice variants, generated mostly due to splicing defects, play a key role in cancer. Germ line or acquired genomic changes (mutations) in/around splicing elements (24) promote aberrant splicing and aberrant protein isoforms.Hyaluronan (HA)3 is synthesized by three different plasma membrane-bound hyaluronan synthases (1, 2, and 3). HAS1 undergoes alternative and aberrant intronic splicing in multiple myeloma, producing truncated variants termed Va, Vb, and Vc (5, 6), which predicted for poor survival in a cohort of multiple myeloma patients (5). Our work suggests that this aberrant splicing arises due to inherited predispositions and acquired mutations in the HAS1 gene (7). Cancer-related, defective mRNA splicing caused by polymorphisms and/or mutations in splicing elements often results in inactivation of tumor suppressor activity (e.g. HRPT2 (8, 9), PTEN (10), MLHI (1114), and ATR (15)) or generation of dominant negative inhibitors (e.g. CHEK2 (16) and VWOX (17)). In breast cancer, aberrantly spliced forms of progesterone and estrogen receptors are found (reviewed in Ref. 3). Intronic mutations inactivate p53 through aberrant splicing and intron retention (18). Somatic mutations with the potential to alter splicing are frequent in some cancers (1925). Single nucleotide polymorphisms in the cyclin D1 proto-oncogene predispose to aberrant splicing and the cyclin D1b intronic splice variant (2629). Cyclin D1b confers anchorage independence, is tumorogenic in vivo, and is detectable in human tumors (30), but as yet no clinical studies have confirmed an impact on outcome. On the other hand, aberrant splicing of HAS1 shows an association between aberrant splice variants and malignancy, suggesting that such variants may be potential therapeutic targets and diagnostic indicators (19, 3133). Increased HA expression has been associated with malignant progression of multiple tumor types, including breast, prostate, colon, glioma, mesothelioma, and multiple myeloma (34). The three mammalian HA synthase (HAS) isoenzymes synthesize HA and are integral transmembrane proteins with a probable porelike structural assembly (3539). Although in humans, the three HAS genes are located on different chromosomes (hCh19, hCh8, and hCh16, respectively) (40), they share a high degree of sequence homology (41, 42). HAS isoenzymes synthesize a different size range of HA molecules, which exhibit different functions (43, 44). HASs contribute to a variety of cancers (4555). Overexpression of HASs promotes growth and/or metastatic development in fibrosarcoma, prostate, and mammary carcinoma, and the removal of the HA matrix from a migratory cell membrane inhibits cell movement (45, 53). HAS2 confers anchorage independence (56). Our work has shown aberrant HAS1 splicing in multiple myeloma (5) and Waldenstrom''s macroglobulinemia (6). HAS1 is overexpressed in colon (57), ovarian (58), endometrial (59), mesothelioma (60), and bladder cancers (61). A HAS1 splice variant is detected in bladder cancer (61).Here, we characterize molecular and biochemical characteristics of HAS1 variants (HAS1-Vs) (5), generated by aberrant splicing. Using transient transfectants and tagged HAS1 family constructs, we show that HAS1-Vs differ in cellular localization, de novo HA localization, and turnover kinetics, as compared with HAS1-FL, and dominantly influence HAS1-FL when co-expressed. HAS1-Vs proteins form intra- and intermolecular associations among themselves and with HAS1-FL, including covalent interactions and multimer formation. HAS1-Vc supports vigorous cellular transformation of NIH3T3 cells in vitro, and HAS1-Vc-transformed NIH3T3 cells are tumorogenic in vivo.  相似文献   

20.
The lysine acetylation of proteins is a reversible post-translational modification that plays a critical regulatory role in both eukaryotes and prokaryotes. Mycobacterium tuberculosis is a facultative intracellular pathogen and the causative agent of tuberculosis. Increasing evidence shows that lysine acetylation may play an important role in the pathogenesis of M. tuberculosis. However, only a few acetylated proteins of M. tuberculosis are known, presenting a major obstacle to understanding the functional roles of reversible lysine acetylation in this pathogen. We performed a global acetylome analysis of M. tuberculosis H37Ra by combining protein/peptide prefractionation, antibody enrichment, and LC-MS/MS. In total, we identified 226 acetylation sites in 137 proteins of M. tuberculosis H37Ra. The identified acetylated proteins were functionally categorized into an interaction map and shown to be involved in various biological processes. Consistent with previous reports, a large proportion of the acetylation sites were present on proteins involved in glycolysis/gluconeogenesis, the citrate cycle, and fatty acid metabolism. A NAD+-dependent deacetylase (MRA_1161) deletion mutant of M. tuberculosis H37Ra was constructed and its characterization showed a different colony morphology, reduced biofilm formation, and increased tolerance of heat stress. Interestingly, lysine acetylation was found, for the first time, to block the immunogenicity of a peptide derived from a known immunogen, HspX, suggesting that lysine acetylation plays a regulatory role in immunogenicity. Our data provide the first global survey of lysine acetylation in M. tuberculosis. The dataset should be an important resource for the functional analysis of lysine acetylation in M. tuberculosis and facilitate the clarification of the entire metabolic networks of this life-threatening pathogen.Mycobacterium tuberculosis was responsible for 1.3 million deaths and 8.6 million new cases of tuberculosis (TB)1 worldwide in 2012 (1). This global public health crisis remains a serious problem, with the emergence of drug-resistant M. tuberculosis, especially multidrug-resistant and extensively drug-resistant M. tuberculosis, and also the emergence of coinfections of TB and human immunodeficiency virus (2, 3). To counter the increasing threat of TB, it is critical to understand fundamental aspects of TB-related biology. Such studies will not only provide new drug targets for the design of novel therapeutic agents, but also facilitate the development of novel diagnostic tools and new vaccines.Acetylation is one of the important protein modifications and occurs both co- and post-translationally on the α-amino group at the N terminus of the protein, so-called “N-terminal acetylation,” or on the ε-amino group on the side chain of lysine (4). Lysine acetylation is one of the most common post-translational modifications to proteins in both eukaryotes and prokaryotes. As a dynamic and reversible process, protein acetylation plays important roles in many cellular physiological processes, including cell-cycle regulation and apoptosis, cell morphology (5), metabolic pathways (68), protein interactions (9), and enzymatic activity (8, 10). In recent years, great advances have been made in proteomic studies, and a large number of lysine-acetylated proteins have been identified in many eukaryotes, including human (5, 11, 12), rat (13), mouse (11), Drosophila (14), Arabidopsis (15, 16), Saccharomyces cerevisiae (17), and protozoans (18, 19). The global analysis of lysine acetylation has also been reported in bacteria, including Escherichia coli (2022), Erwinia amylovora (23), Bacillus subtilis (24), and Salmonella enterica (6). These acetylome studies have generated large datasets of bacterial proteins acetylated on lysine residues and have demonstrated the diverse cellular functions of lysine acetylation in bacteria.Increasing evidence shows that protein acetylation occurs and plays an important regulatory role in mycobacteria (8, 2531). For example, Lange et al. reported the N-terminal acetylation of early secreted antigenic target 6 (ESAT-6) protein (31). Rv1151c is reported to be an NAD+-dependent protein deacetylase in M. tuberculosis that deacetylates and thus regulates the activity of acetyl-CoA synthase (25, 32). Two cyclic adenosine monophosphate (cAMP)-binding proteins in M. smegmatis and M. tuberculosis (MSMEG_5458 and Rv0998, respectively) show similarity to the GNAT family of acetyltransferases and could acetylate a universal stress protein (USP, MSMEG_4207) (30). Subsequent structural studies revealed the fine mechanisms of how cAMP regulates the protein lysine acetyltransferase in mycobacteria (27, 28). Very recently, reversible lysine acetylation was shown to regulate the activity of several fatty acyl-CoA synthetases in M. tuberculosis (8, 26), and also to regulate acetate and propionate metabolism in M. smegmatis (8, 26). However, to the best of our knowledge, only a few acetylated proteins in M. tuberculosis have been identified, presenting a major obstacle to further understanding the regulatory roles of reversible lysine acetylation in this life-threatening pathogen.To fill this gap in our knowledge, we undertook a systematic study of the functional roles of lysine acetylation in M. tuberculosis. We performed an acetylomic analysis of M. tuberculosis H37Ra using high-accuracy MS combined with the identification of 226 unique lysine acetylation sites on 137 proteins. This set of M. tuberculosis proteins acetylated on lysine residues supports the emerging view that lysine acetylation is a general and fundamental regulatory process, and is not restricted to eukaryotes. It also opens the way for its detailed functional and evolutionary analysis of lysine acetylation in M. tuberculosis. The identified acetylated proteins that are involved in several important biological processes were functionally categorized into an interaction map. This is the first time that an interaction network of acetylated proteins in M. tuberculosis has been constructed, and should allow us to better understand the significance of acetylation in key cellular mechanisms in M. tuberculosis. To further explore the effects of lysine acetylation on the physiology of M. tuberculosis H37Ra, MRA_1161, the gene encoding the only known protein deacetylase in this bacterium, was deleted. The roles of MRA_1161 in the colony morphology, carbon source utilization, heat stress tolerance, and biofilm formation of M. tuberculosis were analyzed. The effect of lysine acetylation on the immunogenicity of a known immunogen, HspX, was also tested.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号