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 46 毫秒
1.
Previous studies have shown that deletion of nine residues in the autolysis loop of thrombin produces a mutant with an anticoagulant propensity of potential clinical relevance, but the molecular origin of the effect has remained unresolved. The x-ray crystal structure of this mutant solved in the free form at 1.55 Å resolution reveals an inactive conformation that is practically identical (root mean square deviation of 0.154 Å) to the recently identified E* form. The side chain of Trp215 collapses into the active site by shifting >10 Å from its position in the active E form, and the oxyanion hole is disrupted by a flip of the Glu192–Gly193 peptide bond. This finding confirms the existence of the inactive form E* in essentially the same incarnation as first identified in the structure of the thrombin mutant D102N. In addition, it demonstrates that the anticoagulant profile often caused by a mutation of the thrombin scaffold finds its likely molecular origin in the stabilization of the inactive E* form that is selectively shifted to the active E form upon thrombomodulin and protein C binding.Serine proteases of the trypsin family are responsible for digestion, blood coagulation, fibrinolysis, development, fertilization, apoptosis, and immunity (1). Activation of the protease requires the transition from a zymogen form (2) and formation of an ion pair between the newly formed amino terminus of the catalytic chain and the side chain of the highly conserved residue Asp194 (chymotrypsinogen numbering) next to the catalytic Ser195. This ensures substrate access to the active site and proper formation of the oxyanion hole contributed by the backbone N atoms of Ser195 and Gly193 (3). The zymogen → protease conversion is classically associated with the onset of catalytic activity (3, 4) and provides a useful paradigm for understanding key features of protease function and regulation.Recent kinetic (5) and structural (6, 7) studies of thrombin, the key protease in the blood coagulation cascade (8), have drawn attention to a significant plasticity of the trypsin fold that impacts the function of the enzyme in a decisive manner. The active form of the protease, E, coexists with an inactive form, E*, that is distinct from the zymogen conformation (9). 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 Glu192 and Gly193 that disrupts the oxyanion hole. Importantly, the ion pair between Ile16 and Asp194 remains intact, suggesting that E* is not equivalent to the zymogen form of the protease and that the E*-E equilibrium is established after the conversion from the zymogen form has taken place. Indeed, existing structures of the zymogen forms of trypsin (10), chymotrypsin (11), and chymase (12) feature a broken Ile16–Asp194 ion pair but no collapse of the 215–217 β-strand. Stopped-flow experiments show that the E*-E conversion takes place on a time scale of <10 ms (5), as opposed to the much longer (100–1000 ms) time scale required for the zymogen-protease conversion (13, 14).The E* form is not a peculiarity of thrombin. The collapse of the 215–217 β-strand into the active site is observed in the inactive form of αI-tryptase (15), the high temperature requirement-like protease (16), complement factor D (17), granzyme K (18), hepatocyte growth factor activator (19), prostate kallikrein (20), and prostasin (21). A disrupted oxyanion hole is observed in complement factor B (22) and the arterivirus protease Nsp4 (23). The most likely explanation for the widespread occurrence of inactive conformations of trypsin-like proteases is that the E*-E equilibrium is a basic property of the trypsin fold that fine tunes activity and specificity once the zymogen → protease conversion has taken place (9).The new paradigm established by the E*-E equilibrium has obvious physiological relevance. In the case of complement factors, kallikreins, tryptase, and some coagulation factors must be kept to a minimum until binding of a trigger factor ensues. Stabilization of E* may afford a resting state of the protease waiting for action, as seen for other systems (2428). For example, factor B is mostly inactive until binding of complement factor C3 unleashes catalytic activity at the site where amplification of C3 activation is most needed prior to formation of the membrane attack complex (29). Indeed, the crystal structure of factor B reveals a conformation with the oxyanion hole disrupted by a flip of the 192–193 peptide bond (22), as observed in the E* form of thrombin (6, 7).The allosteric equilibrium as shown in Scheme 1, involves the rates for the E* → E transition, k1, and backward, k1, that define the equilibrium constant r = k1/k1 = [E*]/[E] (5). The value of kcat/Km for an enzyme undergoing the E*-E equilibrium is as shown in Equation 1 (30), where sE is the value of s for the E form, and obviously sE* = 0. Likewise, the binding of an inhibitor to the enzyme undergoing the E*-E equilibrium is shown in Equation 2, where KE is the value of the equilibrium association constant K for the E form, and KE* = 0. As the value of r increases upon stabilization of E*, the values of s and K in Equations 1 and 2 decrease without limits. Hence, stabilization of E* has the potential to completely abrogate substrate hydrolysis (s → 0) or inhibitor binding (K → 0). However, binding of a suitable cofactor could restore activity by triggering the E* → E transition. This suggests a simple explanation for the anticoagulant profile observed in a number of thrombin mutants that have poor activity toward all physiological substrates but retain activity toward the anticoagulant protein C in the presence of the cofactor thrombomodulin (3134). Here we report evidence that stabilization of E* provides a molecular mechanism to turn thrombin into an anticoagulant.  相似文献   

2.
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   

3.
SLC26A7 (human)/Slc26a7 (mouse) is a recently identified chloride-base exchanger and/or chloride transporter that is expressed on the basolateral membrane of acid-secreting cells in the renal outer medullary collecting duct (OMCD) and in gastric parietal cells. Here, we show that mice with genetic deletion of Slc26a7 expression develop distal renal tubular acidosis, as manifested by metabolic acidosis and alkaline urine pH. In the kidney, basolateral Cl/HCO3 exchange activity in acid-secreting intercalated cells in the OMCD was significantly decreased in hypertonic medium (a normal milieu for the medulla) but was reduced only mildly in isotonic medium. Changing from a hypertonic to isotonic medium (relative hypotonicity) decreased the membrane abundance of Slc26a7 in kidney cells in vivo and in vitro. In the stomach, stimulated acid secretion was significantly impaired in isolated gastric mucosa and in the intact organ. We propose that SLC26A7 dysfunction should be investigated as a potential cause of unexplained distal renal tubular acidosis or decreased gastric acid secretion in humans.The collecting duct segment of the distal kidney nephron plays a major role in systemic acid base homeostasis by acid secretion and bicarbonate absorption. The acid secretion occurs via H+-ATPase and H-K-ATPase into the lumen and bicarbonate is absorbed via basolateral Cl/HCO3 exchangers (14). The tubules, which are located within the outer medullary region of the kidney collecting duct (OMCD),2 have the highest rate of acid secretion among the distal tubule segments and are therefore essential to the maintenance of acid base balance (2).The gastric parietal cell is the site of generation of acid and bicarbonate through the action of cytosolic carbonic anhydrase II (5, 6). The intracellular acid is secreted into the lumen via gastric H-K-ATPase, which works in conjunction with a chloride channel and a K+ recycling pathway (710). The intracellular bicarbonate is transported to the blood via basolateral Cl/HCO3 exchangers (1114).SLC26 (human)/Slc26 (mouse) isoforms are members of a conserved family of anion transporters that display tissue-specific patterns of expression in epithelial cells (1524). Several SLC26 members can function as chloride/bicarbonate exchangers. These include SLC26A3 (DRA), SLC26A4 (pendrin), SLC26A6 (PAT1 or CFEX), SLC26A7, and SLC26A9 (2531). SLC26A7 and SLC26A9 can also function as chloride channels (3234).SLC26A7/Slc26a7 is predominantly expressed in the kidney and stomach (28, 29). In the kidney, Slc26a7 co-localizes with AE1, a well-known Cl/HCO3 exchanger, on the basolateral membrane of (acid-secreting) A-intercalated cells in OMCD cells (29, 35, 36) (supplemental Fig. 1). In the stomach, Slc26a7 co-localizes with AE2, a major Cl/HCO3 exchanger, on the basolateral membrane of acid secreting parietal cells (28). To address the physiological function of Slc26a7 in the intact mouse, we have generated Slc26a7 ko mice. We report here that Slc26a7 ko mice exhibit distal renal tubular acidosis and impaired gastric acidification in the absence of morphological abnormalities in kidney or stomach.  相似文献   

4.
The present study tests the hypothesis that the structure of extracellular domain Loop 2 can markedly affect ethanol sensitivity in glycine receptors (GlyRs) and γ-aminobutyric acid type A receptors (GABAARs). To test this, we mutated Loop 2 in the α1 subunit of GlyRs and in the γ subunit of α1β2γ2GABAARs and measured the sensitivity of wild type and mutant receptors expressed in Xenopus oocytes to agonist, ethanol, and other agents using two-electrode voltage clamp. Replacing Loop 2 of α1GlyR subunits with Loop 2 from the δGABAAR (δL2), but not the γGABAAR subunit, reduced ethanol threshold and increased the degree of ethanol potentiation without altering general receptor function. Similarly, replacing Loop 2 of the γ subunit of GABAARs with δL2 shifted the ethanol threshold from 50 mm in WT to 1 mm in the GABAA γ-δL2 mutant. These findings indicate that the structure of Loop 2 can profoundly affect ethanol sensitivity in GlyRs and GABAARs. The δL2 mutations did not affect GlyR or GABAAR sensitivity, respectively, to Zn2+ or diazepam, which suggests that these δL2-induced changes in ethanol sensitivity do not extend to all allosteric modulators and may be specific for ethanol or ethanol-like agents. To explore molecular mechanisms underlying these results, we threaded the WT and δL2 GlyR sequences onto the x-ray structure of the bacterial Gloeobacter violaceus pentameric ligand-gated ion channel homologue (GLIC). In addition to being the first GlyR model threaded on GLIC, the juxtaposition of the two structures led to a possible mechanistic explanation for the effects of ethanol on GlyR-based on changes in Loop 2 structure.Alcohol abuse and dependence are significant problems in our society, with ∼14 million people in the United States being affected (1, 2). Alcohol causes over 100,000 deaths in the United States, and alcohol-related issues are estimated to cost nearly 200 billion dollars annually (2). To address this, considerable attention has focused on the development of medications to prevent and treat alcohol-related problems (35). The development of such medications would be aided by a clear understanding of the molecular structures on which ethanol acts and how these structures influence receptor sensitivity to ethanol.Ligand-gated ion channels (LGICs)2 have received substantial attention as putative sites of ethanol action that cause its behavioral effects (612). Research in this area has focused on investigating the effects of ethanol on two large superfamilies of LGICs: 1) the Cys-loop superfamily of LGICs (13, 14), whose members include nicotinic acetylcholine, 5-hydroxytryptamine3, γ-aminobutyric acid type A (GABAA), γ-aminobutyric acid type C, and glycine receptors (GlyRs) (10, 11, 1520) and 2) the glutamate superfamily, including N-methyl d-aspartate, α-amino-3-hydroxyisoxazolepropionic acid, and kainate receptors (21, 22). Recent studies have also begun investigating ethanol action in the ATP-gated P2X superfamily of LGICs (2325).A series of studies that employed chimeric and mutagenic strategies combined with sulfhydryl-specific labeling identified key regions within Cys-loop receptors that appear to be initial targets for ethanol action that also can determine the sensitivity of the receptors to ethanol (712, 18, 19, 2630). This work provides several lines of evidence that position 267 and possibly other sites in the transmembrane (TM) domain of GlyRs and homologous sites in GABAARs are targets for ethanol action and that mutations at these sites can influence ethanol sensitivity (8, 9, 26, 31).Growing evidence from GlyRs indicates that ethanol also acts on the extracellular domain. The initial findings came from studies demonstrating that α1GlyRs are more sensitive to ethanol than are α2GlyRs despite the high (∼78%) sequence homology between α1GlyRs and α2GlyRs (32). Further work found that an alanine to serine exchange at position 52 (A52S) in Loop 2 can eliminate the difference in ethanol sensitivity between α1GlyRs and α2GlyRs (18, 20, 33). These studies also demonstrated that mutations at position 52 in α1GlyRS and the homologous position 59 in α2GlyRs controlled the sensitivity of these receptors to a novel mechanistic ethanol antagonist (20). Collectively, these studies suggest that there are multiple sites of ethanol action in α1GlyRs, with one site located in the TM domain (e.g. position 267) and another in the extracellular domain (e.g. position 52).Subsequent studies revealed that the polarity of the residue at position 52 plays a key role in determining the sensitivity of GlyRs to ethanol (20). The findings with polarity in the extracellular domain contrast with the findings at position 267 in the TM domain, where molecular volume, but not polarity, significantly affected ethanol sensitivity (9). Taken together, these findings indicate that the physical-chemical parameters of residues at positions in the extracellular and TM domains that modulate ethanol effects and/or initiate ethanol action in GlyRs are not uniform. Thus, knowledge regarding the physical-chemical properties that control agonist and ethanol sensitivity is key for understanding the relationship between the structure and the actions of ethanol in LGICs (19, 31, 3440).GlyRs and GABAARs, which differ significantly in their sensitivities to ethanol, offer a potential method for identifying the structures that control ethanol sensitivity. For example, α1GlyRs do not reliably respond to ethanol concentrations less than 10 mm (32, 33, 41). Similarly, γ subunit-containing GABAARs (e.g. α1β2γ2), the most predominantly expressed GABAARs in the central nervous system, are insensitive to ethanol concentrations less than 50 mm (42, 43). In contrast, δ subunit-containing GABAARs (e.g. α4β3δ) have been shown to be sensitive to ethanol concentrations as low as 1–3 mm (4451). Sequence alignment of α1GlyR, γGABAAR, and δGABAAR revealed differences between the Loop 2 regions of these receptor subunits. Since prior studies found that mutations of Loop 2 residues can affect ethanol sensitivity (19, 20, 39), the non-conserved residues in Loop 2 of GlyR and GABAAR subunits could provide the physical-chemical and structural bases underlying the differences in ethanol sensitivity between these receptors.The present study tested the hypothesis that the structure of Loop 2 can markedly affect the ethanol sensitivity of GlyRs and GABAARs. To accomplish this, we performed multiple mutations that replaced the Loop 2 region of the α1 subunit in α1GlyRs and the Loop 2 region of the γ subunit of α1β2γ2 GABAARs with corresponding non-conserved residues from the δ subunit of GABAAR and tested the sensitivity of these receptors to ethanol. As predicted, replacing Loop 2 of WT α1GlyRs with the homologous residues from the δGABAAR subunit (δL2), but not the γGABAAR subunit (γL2), markedly increased the sensitivity of the receptor to ethanol. Similarly, replacing the non-conserved residues of the γ subunit of α1β2γ2 GABAARs with δL2 also markedly increased ethanol sensitivity of GABAARs. These findings support the hypothesis and suggest that Loop 2 may play a role in controlling ethanol sensitivity across the Cys-loop superfamily of receptors. The findings also provide the basis for suggesting structure-function relationships in a new molecular model of the GlyR based on the bacterial Gloeobacter violaceus pentameric LGIC homologue (GLIC).  相似文献   

5.
6.
The flesh-eating bacterium group A Streptococcus (GAS) binds and activates human plasminogen, promoting invasive disease. Streptococcal surface enolase (SEN), a glycolytic pathway enzyme, is an identified plasminogen receptor of GAS. Here we used mass spectrometry (MS) to confirm that GAS SEN is octameric, thereby validating in silico modeling based on the crystal structure of Streptococcus pneumoniae α-enolase. Site-directed mutagenesis of surface-located lysine residues (SENK252 + 255A, SENK304A, SENK334A, SENK344E, SENK435L, and SENΔ434–435) was used to examine their roles in maintaining structural integrity, enzymatic function, and plasminogen binding. Structural integrity of the GAS SEN octamer was retained for all mutants except SENK344E, as determined by circular dichroism spectroscopy and MS. However, ion mobility MS revealed distinct differences in the stability of several mutant octamers in comparison with wild type. Enzymatic analysis indicated that SENK344E had lost α-enolase activity, which was also reduced in SENK334A and SENΔ434–435. Surface plasmon resonance demonstrated that the capacity to bind human plasminogen was abolished in SENK252 + 255A, SENK435L, and SENΔ434–435. The lysine residues at positions 252, 255, 434, and 435 therefore play a concerted role in plasminogen acquisition. This study demonstrates the ability of combining in silico structural modeling with ion mobility-MS validation for undertaking functional studies on complex protein structures.Streptococcus pyogenes (group A Streptococcus, GAS)8 is a common bacterial pathogen, causing over 700 million human disease episodes each year (1). These range from serious life-threatening invasive diseases including necrotizing fasciitis and streptococcal toxic shock-like syndrome to non-invasive infections like pharyngitis and pyoderma. Invasive disease, in combination with postinfection immune sequelae including rheumatic heart disease and acute poststreptococcal glomerulonephritis, account for over half a million deaths each year (1). Although a resurgence of GAS invasive infections has occurred in western countries since the mid-1980s, disease burden is much greater in developing countries and indigenous populations of developed nations, where GAS infections are endemic (24).GAS is able to bind human plasminogen and activate the captured zymogen to the serine protease plasmin (517). The capacity of GAS to do this plays a critical role in virulence and invasive disease initiation (3, 1719). The plasminogen activation system in humans is an important and highly regulated process that is responsible for breakdown of extracellular matrix components, dissolution of blood clots, and cell migration (20, 21). Plasminogen is a 92-kDa zymogen that circulates in human plasma at a concentration of 2 μm (22). It consists of a binding region of five homologous triple loop kringle domains and an N-terminal serine protease domain that flank the Arg561–Val562 site (23), where it is cleaved by tissue plasminogen activator and urokinase plasminogen activator to yield the active protease plasmin (20, 23). GAS also has the ability to activate human plasminogen by secreting the virulence determinant streptokinase. Streptokinase forms stable complexes with plasminogen or plasmin, both of which exhibit plasmin activity (20, 24). Activation of plasminogen by the plasmin(ogen)-streptokinase complex circumvents regulation by the host plasminogen activation inhibitors, α2-antiplasmin and α2-macroglobulin (11, 20). GAS can bind the plasmin(ogen)-streptokinase complex and/or plasmin(ogen) directly via plasmin(ogen) receptors at the bacterial cell surface (6). These receptors include the plasminogen-binding group A streptococcal M-like protein (PAM) (25), the PAM-related protein (19), glyceraldehyde-3-phosphate dehydrogenase (GAPDH; also known as streptococcal plasmin receptor, Plr, or streptococcal surface dehydrogenase) (9, 26), and streptococcal surface enolase (SEN or α-enolase) (27). Interactions with these GAS receptors occurs via lysine-binding sites within the kringle domains of plasminogen (6).In addition to its ability to bind human plasminogen, SEN is primarily the glycolytic enzyme that converts 2-phosphoglycerate to phosphoenolpyruvate (2729). SEN is abundantly expressed in the cytosol of most bacterial species but has also been identified as a surface-located protein in GAS and other bacteria including pneumococci, despite lacking classical cell surface protein motifs such as a signal sequence, membrane-spanning domain, or cell-wall anchor motif (27, 28, 30, 31). The interaction between SEN and plasminogen is reported to be facilitated by the two C-terminal lysine residues at positions 434 and 435 (27, 32). In contrast, an internal binding motif containing lysines at positions 252 and 255 in the closely related α-enolase of Streptococcus pneumoniae has been shown to play a pivotal role in the acquisition of plasminogen in this bacterial species (33). The octameric pneumococcal α-enolase structure consists of a tetramer of dimers. Hence, potential binding sites could be buried in the interface between subunits. In fact, the crystal structure of S. pneumoniae α-enolase revealed that the two C-terminal lysine residues are significantly less exposed than the internal plasminogen-binding motif (34).In this study, we constructed an in silico model of GAS SEN, based on the pneumococcal octameric α-enolase crystal structure, and validated this model using ion mobility (IM) mass spectrometry (MS). Site-directed mutagenesis followed by structural and functional analyses revealed that Lys344 plays a crucial role in structural integrity and enzymatic function. Furthermore, we demonstrate that the plasminogen-binding motif residues Lys252 and Lys255 and the C-terminal Lys434 and Lys435 residues are located adjacently in the GAS SEN structure and play a concerted role in the binding of human plasminogen.  相似文献   

7.
8.
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   

9.
10.
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   

11.
The structure of the membrane integral rotor ring of the proton translocating F1F0 ATP synthase from spinach chloroplasts was determined to 3.8 Å resolution by x-ray crystallography. The rotor ring consists of 14 identical protomers that are symmetrically arranged around a central pore. Comparisons with the c11 rotor ring of the sodium translocating ATPase from Ilyobacter tartaricus show that the conserved carboxylates involved in proton or sodium transport, respectively, are 10.6–10.8 Å apart in both c ring rotors. This finding suggests that both ATPases have the same gear distance despite their different stoichiometries. The putative proton-binding site at the conserved carboxylate Glu61 in the chloroplast ATP synthase differs from the sodium-binding site in Ilyobacter. Residues adjacent to the conserved carboxylate show increased hydrophobicity and reduced hydrogen bonding. The crystal structure reflects the protonated form of the chloroplast c ring rotor. We propose that upon deprotonation, the conformation of Glu61 is changed to another rotamer and becomes fully exposed to the periphery of the ring. Reprotonation of Glu61 by a conserved arginine in the adjacent a subunit returns the carboxylate to its initial conformation.ATP synthases found in the energy-transducing membranes of bacteria, mitochondria, and chloroplasts catalyze ATP synthesis and ATP hydrolysis coupled with transmembrane proton or sodium ion transport. The enzymes are multi-subunit complexes composed of an extra-membranous catalytic F1 domain and an interconnected integral membrane F0 domain. The hydrophilic F1 domain consists of five different polypeptides with a stoichiometry of α3β3γδϵ. Detailed structural information obtained with the mitochondrial enzyme (13) in combination with biochemical (4), biophysical (5), and single molecule studies (69) revealed that synthesis or hydrolysis of ATP in the F1 domain is accomplished via a rotary catalytic mechanism. In addition to information on the catalytic mechanism, structure analysis and single molecule studies of the mitochondrial or the chloroplast F1 complex have also unraveled the molecular mechanism of several F1-specific inhibitors (1014). Less detailed information is available on the integral membrane F0 domain, which consists of three different polypeptides (a, b, and c) and mediates the transfer of protons or sodium ions across the membrane. Subunits a and b were shown to reside at the periphery of a cylindrical complex formed by multiple copies of the c subunit (1518). The number of c subunits in the cylindrical subcomplex shows substantial variation in different organisms. Ten protomers are found in ATP synthases from yeast, Escherichia coli and Bacillus PS3 (1921), 11 in Ilyobacter tartaricus, Propionigenium modestum, and Clostridium paradoxum (2224), 13 in the thermoalkalophilic Bacillus TA2.TA1 (25), 14 in spinach chloroplasts (26), and 15 in the cyanobacterium Spirulina platensis (27). The structure of isolated subunits a, b, and c from E. coli has been studied by mutagenesis analysis and by NMR spectroscopy in a mixed solvent that was suggested to mimic the membrane environment (2832). These studies showed that subunit a folds with five membrane-spanning helices. The fourth of these helices directly interacts with subunit c and contains a conserved arginine (Arg210), which is thought to be involved in proton transfer (33). Subunit b, which is present in two copies in the intact F0, contains a single transmembrane helix. Cross-linking data support a direct interaction of the two copies of the b subunit (29). Subunit c was studied at two different pH values to obtain the protonated and deprotonated form of a conserved carboxylate (Asp61 in E. coli) that was shown to be essential for proton transport (34). NMR spectroscopy revealed that the isolated c subunit consists of two long hydrophobic membrane spanning segments connected by a short hydrophilic loop (30, 35). This loop is located close to the γ and ϵ subunit on the F1 side of the membrane (36, 37). Low resolution x-ray crystallography, cryo-electron microscopy, and atomic force microscopy showed that the membrane-spanning helices of the multiple copies of subunit c in the intact F0 complex are tightly packed in two concentric rings (19, 22, 26). Atomic resolution of the c ring was recently provided for the Na+-translocating F-type ATPase from I. tartaricus (38) and the related Na+-translocating V-type ATPase from Enterococcus hirae (39). Rotation of the c ring was demonstrated by cross-linking (18), fluorescence studies (40), and single molecule visualization (41, 42). Based on the structural and biochemical information on F1 and F0, different mechanical models have been proposed describing how the rotation of the c ring is coupled to the rotation of the F1 rotor subunits. This rotation in turn drives sequential conformational shifts at the three catalytic β subunits that result in ATP synthesis (4345). Vice versa hydrolysis of ATP in the F1 domain is thought to drive rotation of the γϵc10–15 subcomplex and transports protons or sodium ions across the membrane.Here we describe the crystal structure of the chloroplast c14 rotor, which is the first structure of an isolated c ring rotor from a proton driven ATPase. The structure was solved by molecular replacement using a tetradecameric search model that was generated from a monomer taken from the I. tartaricus c11 structure. The imposition of noncrystallographic symmetry restraints during refinement substantially improved electron density and structure determination.  相似文献   

12.
Codon optimization was used to synthesize the blh gene from the uncultured marine bacterium 66A03 for expression in Escherichia coli. The expressed enzyme cleaved β-carotene at its central double bond (15,15′) to yield two molecules of all-trans-retinal. The molecular mass of the native purified enzyme was ∼64 kDa as a dimer of 32-kDa subunits. The Km, kcat, and kcat/Km values for β-carotene as substrate were 37 μm, 3.6 min−1, and 97 mm−1 min−1, respectively. The enzyme exhibited the highest activity for β-carotene, followed by β-cryptoxanthin, β-apo-4′-carotenal, α-carotene, and γ-carotene in decreasing order, but not for β-apo-8′-carotenal, β-apo-12′-carotenal, lutein, zeaxanthin, or lycopene, suggesting that the presence of one unsubstituted β-ionone ring in a substrate with a molecular weight greater than C35 seems to be essential for enzyme activity. The oxygen atom of retinal originated not from water but from molecular oxygen, suggesting that the enzyme was a β-carotene 15,15′-dioxygenase. Although the Blh protein and β-carotene 15,15′-monooxygenases catalyzed the same biochemical reaction, the Blh protein was unrelated to the mammalian β-carotene 15,15′-monooxygenases as assessed by their different properties, including DNA and amino acid sequences, molecular weight, form of association, reaction mechanism, kinetic properties, and substrate specificity. This is the first report of in vitro characterization of a bacterial β-carotene-cleaving enzyme.Vitamin A (retinol) is a fat-soluble vitamin and important for human health. In vivo, the cleavage of β-carotene to retinal is an important step of vitamin A synthesis. The cleavage can proceed via two different biochemical pathways (1, 2). The major pathway is a central cleavage catalyzed by mammalian β-carotene 15,15′-monooxygenases (EC 1.14.99.36). β-Carotene is cleaved by the enzyme symmetrically into two molecules of all-trans-retinal, and retinal is then converted to vitamin A in vivo (35). The second pathway is an eccentric cleavage that occurs at double bonds other than the central 15,15′-double bond of β-carotene to produce β-apo-carotenals with different chain lengths, which are catalyzed by carotenoid oxygenases from mammals, plants, and cyanobacteria (6). These β-apo-carotenals are degraded to one molecule of retinal, which is subsequently converted to vitamin A in vivo (2).β-Carotene 15,15′-monooxygenase was first isolated as a cytosolic enzyme by identifying the product of β-carotene cleavage as retinal (7). The characterization of the enzyme and the reaction pathway from β-carotene to retinal were also investigated (4, 8). The enzyme activity has been found in mammalian intestinal mucosa, jejunum enterocytes, liver, lung, kidney, and brain (5, 9, 10). Molecular cloning, expression, and characterization of β-carotene 15,15′-monooxygenase have been reported from various species, including chickens (11), fruit flies (12), humans (13), mice (14), and zebra fishes (15).Other proteins thought to convert β-carotene to retinal include bacterioopsin-related protein (Brp) and bacteriorhodopsin-related protein-like homolog protein (Blh) (16). Brp protein is expressed from the bop gene cluster, which encodes the structural protein bacterioopsin, consisting of at least three genes as follows: bop (bacterioopsin), brp (bacteriorhodopsin-related protein), and bat (bacterioopsin activator) (17). brp genes were reported in Haloarcula marismortui (18), Halobacterium sp. NRC-1 (19), Halobacterium halobium (17), Haloquadratum walsbyi, and Salinibacter ruber (20). Blh protein is expressed from the proteorhodopsin gene cluster, which contains proteorhodopsin, crtE (geranylgeranyl-diphosphate synthase), crtI (phytoene dehydrogenase), crtB (phytoene synthase), crtY (lycopene cyclase), idi (isopentenyl diphosphate isomerase), and blh gene (21). Sources of blh genes were previously reported in Halobacterium sp. NRC-1 (19), Haloarcula marismortui (18), Halobacterium salinarum (22), uncultured marine bacterium 66A03 (16), and uncultured marine bacterium HF10 49E08 (21). β-Carotene biosynthetic genes crtE, crtB, crtI, crtY, ispA, and idi encode the enzymes necessary for the synthesis of β-carotene from isopentenyl diphosphate, and the Idi, IspA, CrtE, CrtB, CrtI, and CrtY proteins have been characterized in vitro (2328). Blh protein has been proposed to catalyze or regulate the conversion of β-carotene to retinal (29, 30), but there is no direct proof of the enzymatic activity.In this study, we used codon optimization to synthesize the blh gene from the uncultured marine bacterium 66A03 for expression in Escherichia coli, and we performed a detailed biochemical and enzymological characterization of the expressed Blh protein. In addition, the properties of the enzyme were compared with those of mammalian β-carotene 15,15′-monooxygenases.  相似文献   

13.
The reduction of nitrite (NO2) into nitric oxide (NO), catalyzed by nitrite reductase, is an important reaction in the denitrification pathway. In this study, the catalytic mechanism of the copper-containing nitrite reductase from Alcaligenes xylosoxidans (AxNiR) has been studied using single and multiple turnover experiments at pH 7.0 and is shown to involve two protons. A novel steady-state assay was developed, in which deoxyhemoglobin was employed as an NO scavenger. A moderate solvent kinetic isotope effect (SKIE) of 1.3 ± 0.1 indicated the involvement of one protonation to the rate-limiting catalytic step. Laser photoexcitation experiments have been used to obtain single turnover data in H2O and D2O, which report on steps kinetically linked to inter-copper electron transfer (ET). In the absence of nitrite, a normal SKIE of ∼1.33 ± 0.05 was obtained, suggesting a protonation event that is kinetically linked to ET in substrate-free AxNiR. A nitrite titration gave a normal hyperbolic behavior for the deuterated sample. However, in H2O an unusual decrease in rate was observed at low nitrite concentrations followed by a subsequent acceleration in rate at nitrite concentrations of >10 mm. As a consequence, the observed ET process was faster in D2O than in H2O above 0.1 mm nitrite, resulting in an inverted SKIE, which featured a significant dependence on the substrate concentration with a minimum value of ∼0.61 ± 0.02 between 3 and 10 mm. Our work provides the first experimental demonstration of proton-coupled electron transfer in both the resting and substrate-bound AxNiR, and two protons were found to be involved in turnover.Denitrification is an anaerobic respiration pathway found in bacteria, archaea, and fungi, in which ATP synthesis is coupled to the sequential reduction of nitrate (NO3) and nitrite (NO2) (NO3 → NO2 → NO → N2O → N2) (13).3 The first committed step in this reaction cascade is the formation of gaseous NO by nitrite reductase (NiR), the key enzyme of this pathway. Two distinct classes of periplasmic NiR are found in denitrifying bacteria, one containing cd1 hemes as prosthetic groups (46) and the other utilizing two copper centers to catalyze the one-electron reduction of nitrite (7). Copper-containing NiRs are divided into two main groups according to the color of their oxidized type 1 copper center (T1Cu), with shades ranging from blue to green (3, 7). NiR from Alcaligenes xylosoxidans subsp. xylosoxidans (NCIMB 11015, AxNiR), which is analyzed in this study, is a member of the blue CuNiR group. The blue and green subclasses show a high degree of sequence similarity (70%) (8) and have similar trimeric structures with each monomer (∼36.5 kDa in AxNiR) consisting of two greek key β-barrel cupredoxin-like motifs as well as one long and two short α-helical regions (7, 9).Each NiR monomer contains two copper-binding sites per catalytic unit. One is a T1Cu center, which receives electrons from a physiological redox partner protein and is buried 7 Å beneath the protein surface (10), and the other copper is a type 2 center (T2Cu), constituting the catalytically active substrate-binding site (11). The physiological electron donor for the blue NiRs are the small copper protein azurin (14 kDa) (7) and cytochrome c551 (7, 12, 13). The T1Cu, which is responsible for the color of NiR, serves as the electron delivery center and is coordinated by two histidine residues as well as one cysteine and one methionine residue. The catalytic T2Cu, which like all T2Cu centers has very weak optical bands, is ligated to three His residues and an H2O/OH ligand in the resting state. This H2O/OH ligand is held in place by hydrogen bonds to the active site residues, Asp-92 (AxNiR numbering) and His-249, and gets displaced by the substrate during catalytic turnover (14). The T2Cu is located at the base of a 13–14-Å substrate access channel at the interface of two monomers with one of the three His residues being part of the adjacent subunit (15, 16). The two copper centers are connected by a 12.6-Å covalent bridge provided by the T1Cu-coordinating Cys and by one of the T2Cu His ligands (17, 18). This linkage has been suggested to constitute the electron transfer (ET) pathway from the T1Cu center to the catalytically active T2Cu center via 11 covalent bonds (19).Intramolecular ET from T1- to T2Cu has been extensively examined using pulse radiolysis studies (7, 1924). In a variety of NiR species, ET could be measured, both in the presence and absence of substrate, with observed ET rate constants (kET(obs)) ranging from ∼150 to ∼2000 s−1. According to the Marcus semi-classical ET theory (25), the redox potentials (E0, redox midpoint potential at pH 7.0) of the copper centers affect both the thermodynamic equilibrium and the ET kinetics. In the absence of substrate, the difference in the redox potentials has been found to be insignificant at pH 7 (E0 (T1Cu) ∼240 mV and E0 (T2Cu) ∼230 mV (20)), implying a thermodynamically equal electron distribution between the two metal centers. From an enzymatic point of view, however, approaching this equilibrium position on such a fast time scale (≥150 s−1) is unfavorable in the absence of substrate, as NiR has been shown to form an inactive species with a reduced T2Cu that is devoid of the H2O/OH ligand and unable to bind nitrite (26, 27). Substrate binding has been proposed to induce a favorable shift in the T2Cu redox potential, which would be expected to result in an accelerated ET compared with the substrate-free reaction (7, 16, 25, 2730). However, kET(obs) values in AxgNiR (GIFU1051) have been demonstrated to be lower in the nitrite-bound than in the substrate-free enzyme between pH 7.7 and 5.5 (21). Below pH 5.5, the ET rate constants were observed to be similar in the nitrite-free and -bound enzyme (21).In addition to changes in the redox potentials and thus in the driving force of the ET reaction, several structural changes in the redox centers have been reported as a result of substrate binding, which may also influence the inter-copper ET rate by changing the reorganization energy (16, 25, 30, 31). These rearrangements include subtle changes in the Cys-His bridge linking T1- and T2Cu (32) and conformational transitions of the catalytically relevant active site residue Asp-92 (see below and Ref. 29). Moreover, the presence of nitrite has been postulated to be relayed to the T1Cu site via the so-called substrate sensor loop (via His-94, Asp-92, and His-89 in AxNiR), thereby triggering ET to the T2Cu (19, 27, 29, 32). The tight coupling of ET to the presence of substrate has been argued to prevent the formation of a deactivated enzyme species with a prematurely reduced T2Cu (14, 16, 19, 26, 27, 33). In accordance with such a feedback mechanism, in a combined crystallographic and single-crystal spectroscopic study, inter-copper ET could only be detected in crystals where nitrite was bound to the T2Cu site, whereas in the absence of substrate no such ET was observed (34). This finding, however, contradicts the pulse radiolysis results at room temperature (see above), and the apparent discrepancy between solution studies and x-ray crystallographic data collected at cryogenic temperature remains to be resolved.The one-electron reduction of nitrite to NO involves two protons according to the chemical net equation NO2 + 2H+ + e → NO + H2O, if the T2Cu is ligated by an H2O molecule in the resting state rather than an OH ion. Although the exact enzymatic mechanism is still somewhat controversial (35, 36), one suggested reaction sequence is given in Scheme 1. The potential participation of active site residues in catalyzing the proton transfer (PT) steps has been investigated by studying the pH dependence of NiR under steady-state conditions as well as by pulse radiolysis. The trends obtained for kcat and kET(obs), are similar with pH optima between 5.2 and 6, indicating the involvement of two amino acid residues (21, 22, 37). Asp-92 and His-249 have been proposed as acid-base catalysts (18, 21, 22, 28, 38), and the abrupt drop in rates at increasing pH may indicate that OH can act as a competitive inhibitor for nitrite (39). The relevance of these active site residues, however, as well as the timing of the two protonation steps is still a matter of debate (35, 40, 41).4Open in a separate windowSCHEME 1.A potential reaction mechanism proposed for CuNiRs. Adapted from Ref. 36. Nitrite is shown to bind to the oxidized T2Cu as nitrous acid, thus involving the first protonation step. It coordinates to the oxidized T2Cu center in a bidentate fashion. Following inter-copper ET yielding a reduced T2Cu center, the initially deprotonated Asp-92 accepts a proton, which is subsequently transferred to the substrate. His-249 may be a potential source of this second proton. PT and ET reactions may be reversible and they may be concerted rather than sequential as suggested by the arrows. See text for further information.There are no experimental studies that have been aimed at directly examining the kinetic coupling of PT and ET steps in AxNiR. In this study of the blue AxNiR, our aims were to gain further insight into the mechanism of nitrite reduction by combining multiple turnover experiments with laser photoexcitation studies to measure the (single turnover) inter-copper ET. An extensive analysis of the solvent kinetic isotope effect (SKIE) has been employed as a means of determining whether solvent-exchangeable protons and/or water molecules play a rate-limiting role in the catalytic turnover and/or in inter-copper ET.  相似文献   

14.
15.
16.
Leptospira interrogans is a pathogenic spirochete that causes disease in both humans and animals. LigB (Leptospiral immunoglobulin-like protein B) contributes to the binding of Leptospira to extracellular matrix proteins such as fibronectin (Fn), fibrinogen, laminin, and collagen. A high affinity Fn-binding region of LigB has been recently localized to LigBCen2, which contains the partial eleventh and full twelfth immunoglobulin-like repeats (LigBCen2R) and 47 amino acids of the non-repeat region (LigBCen2NR) of LigB. In this study, LigBCen2NR was shown to bind to the N-terminal domain (NTD) of Fn (KD = 379 nm) by an enzyme-linked immunosorbent assay and isothermal titration calorimetry. Interestingly, this sequence was not observed to adopt secondary structure by far UV circular dichroism or by differential scanning calorimetry, in agreement with computer-based secondary structure predictions. A low partition coefficient (Kav) measured with gel permeation chromatography, a high hydrodynamic radius (Rh) measured with dynamic light scattering, and the insensitivity of the intrinsic viscosity to guanidine hydrochloride treatment all suggest that LigBCen2NR possesses an extended and disordered structure. Two-dimensional 15N-1H HSQC NMR spectra of intact LigBCen2 in the absence and presence of NTD are consistent with these observations, suggesting the presence of both a β-rich region and an unstructured region in LigBCen2 and that the latter of these selectively interacts with NTD. Upon binding to NTD, LigBCen2NR was observed by CD to adopt a β-strand-rich structure, suggestive of the known β-zipper mode of NTD binding.Leptospira interrogans is a pathogenic spirochete that causes leptospirosis throughout the world, especially in developing countries but also in regions of the United States where it has reemerged (1). Weil''s syndrome, a severe form of this disease, is an acute febrile illness associated with multiorgan damage, including liver failure (jaundice), renal failure (nephritis), pulmonary hemorrhage, and meningitis (1), and has a 15% mortality rate if not treated (2). The molecular pathogenesis of leptospirosis is poorly understood, and the bacterial virulence factors involved are largely unknown. Recently, several potential Leptospira virulence factors have been described, including sphingomyelinases, serine proteases, zinc-dependent proteases, and collagenase (3); LipL32 (4); lipopolysaccharide (5); a novel factor H, laminin, and Fn-binding protein (Lsa24 or Len) (68); Loa 22 (9); and Lig (Leptospiral immunoglobulin-like) proteins (1012).Lig proteins, including LigA, LigB, and LigC, contain multiple immunoglobulin-like repeat domains (13 in LigA, 12 in LigB and LigC) (1012). Interestingly, the first 630 residues, from the N terminus to the first half of the seventh immunoglobulin-like domain, are conserved between LigA and LigB, but the rest of the immunoglobulin-like domains are variable (1012) between the two proteins. Also, a non-immunoglobulin-like repeat region found on the C-terminal tail of LigB is not found in LigA (1012). Lig proteins are categorized as microbial surface components recognizing adhesive matrix molecules (MSCRAMMs)2 due to their ability to bind to eukaryotic cells (13) through their interactions with extracellular matrix components, including fibronectin (Fn), laminin, collagens, elastin, and tropoelastin (13, 14, 45). Previously, a high affinity Fn-binding region was localized to LigBCen2, which includes the partial eleventh and complete twelfth immunoglobulin-like repeat region and the first 47 amino acids of the non-repeat regions of LigB (15). LigBCen2 was shown to bind to both the N-terminal domain (NTD) and the gelatin binding domain (GBD) of Fn. The addition of calcium induces a conformational change in LigBCen2 and enhances binding between LigBCen2 and the NTD of Fn (15).The first step in the process of bacterial infection is cellular adhesion, mediated by bacterial adhesins interacting with various components of the extracellular matrix (16). Known interaction modes between Fn and bacterial Fn-binding proteins include the β-zipper (17, 18) and the cationic cradle (19). It was recently discovered that the Fn-binding domains in certain Fn-binding proteins are disordered and extended but gain structure upon binding to the NTD of Fn (2022).We have performed a fine-mapping study of the NTD-binding site on LigBCen2 and identified this site as LigBCen2NR, a portion of the non-repeat region (amino acids 1119–1165). The addition of NTD promotes the folding of LigBCen2NR from a disordered and extended structure to a folded structure. This finding is notable, since LigBCen2NR is located in the non-immunoglobulin-like region of LigB, as compared with other Fn-binding proteins, such as Staphylococcus aureus FnbpA and FnbpB (23), Streptococcus dysgalactiae FnBB (17), and Streptococcus pyogenes SfbI and SfbII (24). Thus, the binding mode appears to be similar to the known β-zipper mechanism but unique in sequence-specific interactions. This finding provides the fundamental groundwork for the development of a therapeutic agent to target this interaction in order to prevent or treat Leptospira infection.  相似文献   

17.
Rotary catalysis in F1F0 ATP synthase is powered by proton translocation through the membrane-embedded F0 sector. Proton binding and release occur in the middle of the membrane at Asp-61 on transmembrane helix (TMH) 2 of subunit c. Previously the reactivity of Cys substituted into TMH2 revealed extensive aqueous access at the cytoplasmic side as probed with Ag+ and other thiolate-directed reagents. The analysis of aqueous accessibility of membrane-embedded regions in subunit c was extended here to TMH1 and the periplasmic side of TMH2. The Ag+ sensitivity of Cys substitutions was more limited on the periplasmic versus cytoplasmic side of TMH2. In TMH1, Ag+ sensitivity was restricted to a pocket of four residues lying directly behind Asp-61. Aqueous accessibility was also probed using Cd2+, a membrane-impermeant soft metal ion with properties similar to Ag+. Cd2+ inhibition was restricted to the I28C substitution in TMH1 and residues surrounding Asp-61 in TMH2. The overall pattern of inhibition, by all of the reagents tested, indicates highest accessibility on the cytoplasmic side of TMH2 and in a pocket of residues around Asp-61, including proximal residues in TMH1. Additionally subunit a was shown to mediate access to this region by the membrane-impermeant probe 2-(trimethylammonium)ethyl methanethiosulfonate. Based upon these results and other information, a pocket of aqueous accessible residues, bordered by the peripheral surface of TMH4 of subunit a, is proposed to extend from the cytoplasmic side of cTMH2 to Asp-61 in the center of the membrane.F1F0 ATP synthase utilizes the energy stored in an H+ or Na+ electrochemical gradient to synthesize ATP in bacteria, mitochondria, and chloroplasts (14). The ATP synthase complex is composed of two sectors, i.e. a water-soluble F1 sector that is bound to a membrane-embedded F0 sector. In bacteria, F1 is composed of five subunits in an α3β3γδϵ ratio and contains three catalytic sites for ATP synthesis and/or hydrolysis centered at the α-β subunit interfaces. F0 is composed of three subunits in an a1b2c10–15 ratio and functions as the ion-conducting pathway (59). Ion translocation through F0 drives rotation of a cylindrical ring of c-subunits that is coupled to rotation of the γ subunit within the (αβ)3 hexamer of F1 to force conformational changes in the three active sites and in turn drive synthesis of ATP by the binding change mechanism (14, 1013).Subunit c of F0 folds in the membrane as a hairpin of two extended α-helices. In Escherichia coli, 10 copies of subunit c pack together to form a decameric ring with TMH12 on the inside and TMH2 on the periphery (6, 14). An atomic resolution structure of the Na+-translocating c11-ring from Ilyobacter tartaricus was recently published by Meier et al. (8). In the c11 structure, the Na+ binding site is formed by two interacting c subunits. The essential Na+-binding Glu residue, which corresponds to Asp-61 in E. coli, is located in TMH2 at the middle of the lipid bilayer. Subunit a consists of five transmembrane helices, four of which likely interact as a four-helix bundle (1518). Subunit a lies on the periphery of the c-ring with TMHs 4 and 5 from subunit a and TMH2 from subunit c forming the a-c interface (1821). During ion translocation through F0, the essential Arg-210 on TMH4 of subunit a is postulated to facilitate the protonation/deprotonation cycle at Asp-61 of subunit c and cause the rotation of the c-ring past the stationary subunit a (3, 4, 19).Chemical modification of cysteine-substituted transmembrane proteins has been widely used as a means of probing the aqueous accessible regions (2224). The reactivity of a substituted cysteine to thiolate-directed probes provides an indication of aqueous accessibility because the reactive thiolate species is preferentially formed in an aqueous environment. The aqueous accessibility of the five TMHs in subunit a of E. coli F0 has been probed using Ag+ and NEM (19, 2527). The results suggest the presence of an aqueous accessible channel in subunit a in the center of TMHs 2–5 extending from the periplasm to the center of the membrane. Protons entering through this periplasmic access channel are postulated to bind to the essential Asp-61 residues of the c-ring and exit to the cytoplasm by a still uncertain pathway at the peripheral face of aTMH4 with protonation/deprotonation of Asp-61 driving c-ring rotation.During H+-driven ATP synthesis, two models for the pathway by which H+ or Na+ exit to the cytoplasm have been proposed. The first model proposes that the ions bound at Asp-61 exit to the cytoplasm via a half-channel composed at least partially by residues in TMH4 of subunit a (2527). Chemical modification studies of Cys-substituted subunit a of E. coli revealed an aqueous accessible surface of TMH4 that includes the essential Arg-210 residue, which extended from the center of the membrane to the cytoplasm, suggesting that the ion exit channel may lie at the a-c interface (19, 25). Alternatively studies of the c-ring from the I. tartaricus enzyme indicate that Na+ can access Glu-65 in the absence of other F0 subunits, suggesting an intrinsic channel in subunit c (28, 29). However, no such channel was apparent in the crystal structure of the c11-ring (8). In a previous study (30), we probed the thiolate reactivity of Cys substitutions in the cytoplasmic half of TMH2 in subunit c. These experiments revealed extensive reactivity to sulfhydryl-directed reagents on the peripheral face of cTMH2, supporting the presence of the cytoplasmic exit channel at the a-c interface. In this study, we extended the survey of aqueous accessibility in transmembrane regions by probing thiolate reactivity of Cys substitutions in TMH1 and in the periplasmic half of TMH2. The reactivity of Cys substituted into these regions proved to be more limited. Only a small region of TMH1, lying directly behind Asp-61, was reactive with Ag+. In addition to Ag+, we used Cd2+ as a complementary, membrane-impermeant probe for aqueous accessibility. The survey of Cd2+ sensitivity confirmed that aqueous accessibility from the cytoplasm is much greater for residues packing at the periphery of the c-ring. The experiments reported here distinguish the aqueous accessible and inaccessible regions of the c-ring and strengthen evidence that the cytoplasmic H+ exit channel is situated at the a-c interface.  相似文献   

18.
19.
Lysophosphatidic acid (LPA), a bioactive phospholipid, induces a wide range of cellular effects, including gene expression, cytoskeletal rearrangement, and cell survival. We have previously shown that LPA stimulates secretion of pro- and anti-inflammatory cytokines in bronchial epithelial cells. This study provides evidence that LPA enhances pulmonary epithelial barrier integrity through protein kinase C (PKC) δ- and ζ-mediated E-cadherin accumulation at cell-cell junctions. Treatment of human bronchial epithelial cells (HBEpCs) with LPA increased transepithelial electrical resistance (TER) by ∼2.0-fold and enhanced accumulation of E-cadherin to the cell-cell junctions through Gαi-coupled LPA receptors. Knockdown of E-cadherin with E-cadherin small interfering RNA or pretreatment with EGTA (0.1 mm) prior to LPA (1 μm) treatment attenuated LPA-induced increases in TER in HBEpCs. Furthermore, LPA induced tyrosine phosphorylation of focal adhesion kinase (FAK) and overexpression of the FAK inhibitor, and FAK-related non-kinase-attenuated LPA induced increases in TER and E-cadherin accumulation at cell-cell junctions. Overexpression of dominant negative protein kinase δ and ζ attenuated LPA-induced phosphorylation of FAK, accumulation of E-cadherin at cell-cell junctions, and an increase in TER. Additionally, lipopolysaccharide decreased TER and induced E-cadherin relocalization from cell-cell junctions to cytoplasm in a dose-dependent fashion, which was restored by LPA post-treatment in HBEpCs. Intratracheal post-treatment with LPA (5 μm) reduced LPS-induced neutrophil influx, protein leak, and E-cadherin shedding in bronchoalveolar lavage fluids in a murine model of acute lung injury. These data suggest a protective role of LPA in airway inflammation and remodeling.The airway epithelium is the site of first contact for inhaled environmental stimuli, functions as a physical barrier to environmental insult, and is an essential part of innate immunity. Epithelial barrier disruption is caused by inhaled allergens, dust, and irritants, resulting in inflammation, bronchoconstriction, and edema as seen in asthma and other respiratory diseases (14). Furthermore, increased epithelial permeability also results in para-cellular leakage of large proteins, such as albumin, immunoglobulin G, and polymeric immunoglobulin A, into the airway lumen (5, 6). The epithelial cell-cell junctional complex is composed of tight junctions, adherens junctions, and desmosomes. These adherens junctions play a pivotal role in regulating the activity of the entire junctional complex because the formation of adherens junctions subsequently leads to the formation of other cell-cell junctions (79). The major adhesion molecules in the adherens junctions are the cadherins. E-cadherin is a member of the cadherin family that mediates calcium-dependent cell-cell adhesion. The N-terminal ectodomain of E-cadherin contains homophilic interaction specificity, and the cytoplasmic domain binds to catenins, which interact with actin (1013). Plasma membrane localization of E-cadherin is critical for the maintenance of epithelial cell-cell junctions and airway epithelium integrity (7, 10, 14). A decrease of adhesive properties of E-cadherin is related to the loss of differentiation and the subsequent acquisition of a higher motility and invasiveness of epithelial cells (10, 14, 15). Dislocation or shedding of E-cadherin in the airway epithelium induces epithelial shedding and increases airway permeability in lung airway diseases (10, 14, 16). In an ovalbumin-challenged guinea pig model of asthma, it has been demonstrated that E-cadherin is dislocated from the lateral margins of epithelial cells (10). Histamine increases airway para-cellular permeability and results in an increased susceptibility of airway epithelial cells to adenovirus infection by interrupting E-cadherin adhesion (14). Serine phosphorylation of E-cadherin by casein kinase II, GSK-3β, and PKD1/PKC2 μ enhanced E-cadherin-mediated cell-cell adhesion in NIH3T3 fibroblasts and LNCaP prostate cancer cells (11, 17). However, the regulation and mechanism by which E-cadherin is localized within the pulmonary epithelium is not fully known, particularly during airway remodeling.LPA, a naturally occurring bioactive lipid, is present in body fluids, such as plasma, saliva, follicular fluid, malignant effusions, and bronchoalveolar lavage (BAL) fluids (1820). Six distinct high affinity cell-surface LPA receptors, LPA-R1–6, have been cloned and described in mammals (2126). Extracellular activities of LPA include cell proliferation, motility, and cell survival (2730). LPA exhibits a wide range of effects on differing cell types, including pulmonary epithelial, smooth muscle, fibroblasts, and T cells (3135). LPA augments migration and cytokine synthesis in lymphocytes and induces chemotaxis of Jurkat T cells through Matrigel membranes (34). LPA induces airway smooth muscle cell contractility, proliferation, and airway repair and remodeling (35, 36). LPA also potently stimulates IL-8 (31, 3739), IL-13 receptor α2 (IL-13Rα2) (40), and COX-2 gene expression and prostaglandin E2 release (41) in HBEpCs. Prostaglandin E2 and IL-13Rα2 have anti-inflammatory properties in pulmonary inflammation (42, 43). These results suggest that LPA may play a protective role in lung disease by stimulating an innate immune response while simultaneously attenuating the adaptive immune response. Furthermore, intravenous injection with LPA attenuated bacterial endotoxin-induced plasma tumor necrosis factor-α production and myeloperoxidase activity in the lungs of mice (44), suggesting an anti-inflammatory role for LPA in a murine model of sepsis.We have reported that LPA induces E-cadherin/c-Met accumulation in cell-cell contacts and increases TER in HBEpCs (45). Here, for the first time, we report that LPA-induced increases in TER are dependent on PKCδ, PKCζ, and FAK-mediated E-cadherin accumulation at cell-cell junctions. Furthermore, we demonstrate that post-treatment of LPA rescues LPS-induced airway epithelial disruption in vitro and reduces E-cadherin shedding in a murine model of ALI. This study identifies the molecular mechanisms linking the LPA and LPA receptors to maintaining normal pulmonary epithelium barrier function, which is critical in developing novel therapies directed at ameliorating pulmonary diseases.  相似文献   

20.
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