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 312 毫秒
1.
Many wood-rotting fungi, including Phellinus pomaceus, produce chloromethane (CH3Cl). P. pomaceus can be cultured in undisturbed glucose mycological peptone liquid medium to produce high amounts of CH3Cl. The biosynthesis of CH3Cl is catalyzed by a methyl chloride transferase (MCT), which appears to be membrane bound. The enzyme is labile upon removal from its natural location and upon storage at low temperature in its bound state. Various detergents failed to solubilize the enzyme in active form, and hence it was characterized by using a membrane fraction. The enzyme had a sharp pH optimum between 7 and 7.2. Its apparent Km for Cl (ca. 300 mM) was much higher than that for I (250 μM) or Br (11 mM). A comparison of these Km values to the relative in vivo methylation rates for different halides suggests that the real Km for Cl may be much lower, but the calculated value is high because the CH3Cl produced is used immediately in a coupled reaction. Among various methyl donors tested, S-adenosyl-l-methionine (SAM) was the only one that supported significant methylation by MCT. The reaction was inhibited by S-adenosyl-l-homocysteine, an inhibitor of SAM-dependent methylation, suggesting that SAM is the natural methyl donor. These findings advance our comprehension of a poorly understood metabolic sector at the origin of biogenic emissions of halomethanes, which play an important role in atmospheric chemistry.Halogenated organic compounds are ubiquitous in nature (29). They participate in the depletion of stratospheric ozone and have a profound impact on atmospheric chemistry (4, 18, 24). Although the dominant sources of these compounds are biogenic emissions (12, 25, 26, 28), their significance to the emitter organisms is rather poorly understood, with only a few indications of the roles they might play. In fungi, halomethanes serve as methyl group donors for the biosynthesis of esters, anisoles, and veratryl alcohol (9, 11). In algae, halomethanes are by-products of reactions in which scavenging of H2O2 releases HOBr, which is presumed to be a defense molecule against bacteria, fungi, and herbivores (23, 27). A recent report (28) that a marine alga, Endocladia muricata, and a salt-tolerant plant, Mesembryanthemum crystallinum, could methylate Cl ions to chloromethane (CH3Cl) triggered speculation that this may be a mechanism for Cl detoxification and salt tolerance. The S-adenosyl-l-methionine (SAM)-dependent methyl chloride transferase (MCT) that catalyzes this reaction was partially purified from E. muricata (28). The enzyme can also use I and Br as substrates.These results suggest possibilities for engineering a Cl detoxification capability into crop plants, many of which are sensitive to Cl (6, 17). Wood-rotting fungi of the family Hymenochaetaceae are the most efficient producers of CH3Cl (5, 7, 13). Phellinus pomaceus converts Cl to CH3Cl with over 90% efficiency, even at extremely low concentrations of the ion (7). A low MCT activity was detected in cell extracts of this fungus (28).Halomethanes are the primary carriers of halogens between the biosphere and the atmosphere (4, 18) and therefore play pivotal roles in the effect of halogens on atmospheric chemistry and the integrity of the ozone layer (24). Since biogenic sources are major contributors of atmospheric halomethanes (7, 12, 18, 25, 28), attempts to understand atmospheric composition must include an understanding of the metabolic processes underlying the generation of these gases. In addition, engineering a Cl detoxification capability into plants depends on the identification of novel metabolic pathways and an understanding of their regulation. Within this dual context, our objective was to determine the biochemical nature of the CH3Cl-evolving system of P. pomaceus.  相似文献   

2.
All vertebrate cells regulate their cell volume by activating chloride channels of unknown molecular identity, thereby activating regulatory volume decrease. We show that the Ca2+-activated Cl channel TMEM16A together with other TMEM16 proteins are activated by cell swelling through an autocrine mechanism that involves ATP release and binding to purinergic P2Y2 receptors. TMEM16A channels are activated by ATP through an increase in intracellular Ca2+ and a Ca2+-independent mechanism engaging extracellular-regulated protein kinases (ERK1/2). The ability of epithelial cells to activate a Cl conductance upon cell swelling, and to decrease their cell volume (regulatory volume decrease) was dependent on TMEM16 proteins. Activation of ICl,swell was reduced in the colonic epithelium and in salivary acinar cells from mice lacking expression of TMEM16A. Thus TMEM16 proteins appear to be a crucial component of epithelial volume-regulated Cl channels and may also have a function during proliferation and apoptotic cell death.Regulation of cell volume is fundamental to all cells, particularly during cell growth and division. External hypotonicity leads to cell swelling and subsequent activation of volume-regulated chloride and potassium channels, to release intracellular ions and to re-shrink the cells, a process termed regulatory volume decrease (RVD)3 (1). Volume-regulated chloride currents (ICl,swell) have dual functions during cell proliferation as well as apoptotic volume decrease (AVD), preceding apoptotic cell death (2). Although ICl,swell is activated in swollen cells to induce RVD, AVD takes place under normotonic conditions to shrink cells (3, 4). Early work suggested intracellular Ca2+ as an important mediator for activation of ICl,swell and volume-regulated K+ channels (5), whereas subsequent studies only found a permissive role of Ca2+ for activation of ICl,swell (6), reviewed in Ref. 1. In addition, a plethora of factors and signaling pathways have been implicated in activation of ICl,swell, making cell volume regulation an extremely complex process (reviewed in Refs. 1, 3, and 7). These factors include intracellular ATP, the cytoskeleton, phospholipase A2-dependent pathways, and protein kinases such as extracellular-regulated kinase ERK1/2 (reviewed in Refs. 1 and 7). Previous approaches in identifying swelling-activated Cl channels have been unsuccessful or have produced controversial data. Thus none of the previous candidates such as pICln, the multidrug resistance protein, or ClC-3 are generally accepted to operate as volume-regulated Cl channels (reviewed in Refs. 8 and 9). Notably, the cystic fibrosis transmembrane conductance regulator (CFTR) had been shown in earlier studies to influence ICl,swell and volume regulation (1012). The variable properties of ICl,swell suggest that several gene products may affect ICl,swell in different cell types.The TMEM16 transmembrane protein family consists of 10 different proteins with numerous splice variants that contain 8–9 transmembrane domains and have predicted intracellular N- and C-terminal tails (13, 1618). TMEM16A (also called ANO1) is required for normal development of the murine trachea (14) and is associated with different types of tumors, dysplasia, and nonsyndromic hearing impairment (13, 15). TMEM16A has been identified as a subunit of Ca2+-activated Cl channels that are expressed in epithelial and non-epithelial tissues (1618). Interestingly, members of the TMEM16 family have been suggested to play a role in osmotolerance in Saccharomyces cerevisiae (19). Here we show that TMEM16 proteins also contribute to ICl,swell and regulatory volume decrease.  相似文献   

3.
SLC26 proteins function as anion exchangers, channels, and sensors. Previous cellular studies have shown that Slc26a3 and Slc26a6 interact with the R-region of the cystic fibrosis transmembrane conductance regulator (CFTR), (R)CFTR, via the Slc26-STAS (sulfate transporter anti-sigma) domain, resulting in mutual transport activation. We recently showed that Slc26a9 has both nCl-HCO3 exchanger and Cl channel function. In this study, we show that the purified STAS domain of Slc26a9 (a9STAS) binds purified (R)CFTR. When Slc26a9 and (R)CFTR fragments are co-expressed in Xenopus oocytes, both Slc26a9-mediated nCl-HCO3 exchange and Cl currents are almost fully inhibited. Deletion of the Slc26a9 STAS domain (a9-ΔSTAS) virtually eliminated the Cl currents with only a modest affect on nCl-HCO3 exchange activity. Co-expression of a9-ΔSTAS and the (R)CFTR fragment did not alter the residual a9-ΔSTAS function. Replacing the Slc26a9 STAS domain with the Slc26a6 STAS domain (a6-a9-a6) does not change Slc26a9 function and is no longer inhibited by (R)CFTR. These data indicate that the Slc26a9-STAS domain, like other Slc26-STAS domains, binds CFTR in the R-region. However, unlike previously reported data, this binding interaction inhibits Slc26a9 ion transport activity. These results imply that Slc26-STAS domains may all interact with (R)CFTR but that the physiological outcome is specific to differing Slc26 proteins, allowing for dynamic and acute fine tuning of ion transport for various epithelia.Slc26 genes and proteins have attracted the attention of physiologists and geneticists. Why? Slc26a1 (Sat-1) was characterized as a Na+-independent SO42− transporter (1). Given the transport characteristics of the founding member of the gene family, Slc26 proteins were assumed to be sulfate transporters. Disease phenotypes, clone characterization, and family additions demonstrate that the Slc26 proteins are anion transporters or channels (24). These proteins have varied tissue expression patterns. At one extreme, Slc26a5 in mammals is found in the hair cells of the inner ear (5), whereas Slc26a2 (DTDST) is virtually ubiquitous in epithelial tissues (2).Several Slc26 proteins are found in the epithelia of the lung, intestine, stomach, pancreas, and kidney, usually in apical membranes. Interestingly these are also tissues and membranes in which the cystic fibrosis transmembrane conductance regulator (CFTR)5 has been found functionally or by immunohistochemistry. Ko and co-workers (68) examined the distribution of Slc26a3 and Slc26a6 in HCO3 secretory epithelia, and asked if an interaction might occur between these Slc26 proteins and CFTR. In particular, these studies indicate that in expression systems, there is a reciprocal-stimulatory interaction of the STAS (sulfate transporter anti-sigma) domains of Slc26a3 and Slc26a6 with the regulatory region (R-region) of CFTR. These investigators hypothesized that this stimulatory interaction could account for the differences in pancreatic insufficiency and sufficiency observed in cystic fibrosis patients. Nevertheless, knock-out Slc26a6 mouse studies reveal more complicated cell and tissue physiology (see “Discussion”).Slc26a9 has been reported to be a Cl-HCO3 exchanger (9, 10) or a large Cl conductance (3, 11, 12). Loriol and co-workers (12) indicated that SLC26A9 has a Cl conductance that may be stimulated by HCO3. Two other groups have indicated that the Cl conductance is not affected by the presence of HCO3 (10, 11). We have recently demonstrated that Slc26a9 functions as both an electrogenic nCl-HCO3 exchanger and a Cl channel (10). Dorwart and colleagues (11) found that WNK kinases inhibited the SLC26A9 Cl conductance but that this effect was independent of kinase activity. One group has a preliminary report indicating that WNK3 decreased Cl uptake, whereas WNK4 increased Cl uptake via Slc26a9 expressed in Xenopus oocytes (13).Slc26a9 and CFTR are also co-expressed in several tissues. Slc26a9 protein has been localized to epithelia of the stomach and lung (9, 10, 14), although mRNA is also detectable in brain, heart, kidney, small intestine, thymus, and ovary (10). The R-region of CFTR was previously shown to increase the activity of Slc26a3 and Slc26a6 by interaction with STAS domains (6, 15, 16). Because Slc26a9 displays several different modes of ion transport, we asked if the R-region of CFTR would also increase the activity of Slc26a9. Our results indicate that the R-region of CFTR does interact with the STAS domain of Slc26a9. However, in the case of Slc26a9 this apparently similar interaction results in inhibition of Slc26a9 ion transport.  相似文献   

4.
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   

5.
6.
7.
Fibroblasts degrade type I collagen, the major extracellular protein found in mammals, during events ranging from bulk tissue resorption to invasion through the three-dimensional extracellular matrix. Current evidence suggests that type I collagenolysis is mediated by secreted as well as membrane-anchored members of the matrix metalloproteinase (MMP) gene family. However, the roles played by these multiple and possibly redundant, degradative systems during fibroblast-mediated matrix remodeling is undefined. Herein, we use fibroblasts isolated from Mmp13−/−, Mmp8−/−, Mmp2−/−, Mmp9−/−, Mmp14−/− and Mmp16−/− mice to define the functional roles for secreted and membrane-anchored collagenases during collagen-resorptive versus collagen-invasive events. In the presence of a functional plasminogen activator-plasminogen axis, secreted collagenases arm cells with a redundant collagenolytic potential that allows fibroblasts harboring single deficiencies for either MMP-13, MMP-8, MMP-2, or MMP-9 to continue to degrade collagen comparably to wild-type fibroblasts. Likewise, Mmp14−/− or Mmp16−/− fibroblasts retain near-normal collagenolytic activity in the presence of plasminogen via the mobilization of secreted collagenases, but only Mmp14 (MT1-MMP) plays a required role in the collagenolytic processes that support fibroblast invasive activity. Furthermore, by artificially tethering a secreted collagenase to the surface of Mmp14−/− fibroblasts, we demonstrate that localized pericellular collagenolytic activity differentiates the collagen-invasive phenotype from bulk collagen degradation. Hence, whereas secreted collagenases arm fibroblasts with potent matrix-resorptive activity, only MT1-MMP confers the focal collagenolytic activity necessary for supporting the tissue-invasive phenotype.In the postnatal state, fibroblasts are normally embedded in a self-generated three-dimensional connective tissue matrix composed largely of type I collagen, the major extracellular protein found in mammals (13). Type I collagen not only acts as a structural scaffolding for the associated mesenchymal cell populations but also regulates gene expression and cell function through its interactions with collagen binding integrins and discoidin receptors (2, 4). Consistent with the central role that type I collagen plays in defining the structure and function of the extracellular matrix, the triple-helical molecule is resistant to almost all forms of proteolytic attack and can display a decades-long half-life in vivo (46). Nonetheless, fibroblasts actively remodel type I collagen during wound healing, inflammation, or neoplastic states (2, 713).To date type I collagenolytic activity is largely confined to a small subset of fewer than 10 proteases belonging to either the cysteine proteinase or matrix metalloproteinase (MMP)2 gene families (4, 1418). As all collagenases are synthesized as inactive zymogens, complex proteolytic cascades involving serine, cysteine, metallo, and aspartyl proteinases have also been linked to collagen turnover by virtue of their ability to mediate the processing of the pro-collagenases to their active forms (13, 15, 19). After activation, each collagenase can then cleave native collagen within its triple-helical domain, thus precipitating the unwinding or “melting” of the resulting collagen fragments at physiologic temperatures (4, 15). In turn, the denatured products (termed gelatin) are susceptible to further proteolysis by a broader class of “gelatinases” (4, 15). Collagen fragments are then either internalized after binding to specific receptors on the cell surface or degraded to smaller peptides with potent biological activity (2024).Previous studies by our group as well as others have identified MMPs as the primary effectors of fibroblast-mediated collagenolysis (20, 25, 26). Interestingly, adult mouse fibroblasts express at least six MMPs that can potentially degrade type I collagen, raising the possibility of multiple compensatory networks that are designed to preserve collagenolytic activity (25). Four of these collagenases belong to the family of secreted MMPs, i.e. MMP-13, MMP-8, MMP-2, and MMP-9, whereas the other two enzymes are members of the membrane-type MMP subgroup, i.e. MMP-14 (MT1-MMP) and MMP-16 (MT3-MMP) (13, 2729). From a functional perspective, the specific roles that can be assigned to secreted versus membrane-anchored collagenases remain undefined. As such, fibroblasts were isolated from either wild-type mice or mice harboring loss-of-function deletions in each of the major secreted and membrane-anchored collagenolytic genes, and the ability of the cells to degrade type I collagen was assessed. Herein, we demonstrate that fibroblasts mobilize either secreted or membrane-anchored MMPs to effectively degrade type I collagen in qualitatively and quantitatively distinct fashions. However, under conditions where fibroblasts use either secreted and membrane-anchored MMPs to exert quantitatively equivalent collagenolytic activity, only MT1-MMP plays a required role in supporting a collagen-invasive phenotype. These data establish a new paradigm wherein secreted collagenases are functionally limited to bulk collagenolytic processes, whereas MT1-MMP uniquely arms the fibroblast with a focalized degradative activity that mediates subjacent collagenolysis as well as invasion.  相似文献   

8.
The extracellular domain of the epithelial sodium channel ENaC is exposed to a wide range of Cl concentrations in the kidney and in other epithelia. We tested whether Cl alters ENaC activity. In Xenopus oocytes expressing human ENaC, replacement of Cl with SO42−, H2PO4, or SCN produced a large increase in ENaC current, indicating that extracellular Cl inhibits ENaC. Extracellular Cl also inhibited ENaC in Na+-transporting epithelia. The anion selectivity sequence was SCN < SO42− < H2PO4 < F < I < Cl < Br. Crystallization of ASIC1a revealed a Cl binding site in the extracellular domain. We found that mutation of corresponding residues in ENaC (αH418A and βR388A) disrupted the response to Cl, suggesting that Cl might regulate ENaC through an analogous binding site. Maneuvers that lock ENaC in an open state (a DEG mutation and trypsin) abolished ENaC regulation by Cl. The response to Cl was also modulated by changes in extracellular pH; acidic pH increased and alkaline pH reduced ENaC inhibition by Cl. Cl regulated ENaC activity in part through enhanced Na+ self-inhibition, a process by which extracellular Na+ inhibits ENaC. Together, the data indicate that extracellular Cl regulates ENaC activity, providing a potential mechanism by which changes in extracellular Cl might modulate epithelial Na+ absorption.The epithelial Na+ channel ENaC2 is a heterotrimer of homologous α, β, and γ subunits (1, 2). ENaC functions as a pathway for Na+ absorption across epithelial cells in the kidney collecting duct, lung, distal colon, and sweat duct (reviewed in Refs. 3 and 4). Na+ transport is critical for the maintenance of Na+ homeostasis and for the control of the composition and quantity of the fluid on the apical membrane of these epithelia. ENaC mutations and defects in its regulation cause inherited forms of hypertension and hypotension (5) and may contribute to the pathogenesis of lung disease in cystic fibrosis (6).ENaC is a member of the DEG/ENaC family of ion channels. A common structural feature of these channels is a large extracellular domain that plays a critical role in channel gating. For example, in ASICs, the extracellular domain functions as a receptor for protons, which transiently activate the channel by titrating residues that form an acidic pocket (7). FaNaCh is a ligand-gated family member in Helix aspersa, activated by the peptide FMRFamide (8). In Caenorhabditis elegans MEC family members, the extracellular domain is thought to respond to mechanical signals (9).ENaC differs from other family members because it is constitutively active in the absence of a ligand/stimulus. However, a convergence of data indicate that ENaC gating is modulated by a variety of molecules that bind to or modify its extracellular domains, including proteases (1012), Na+ (1315), protons (16), and the divalent cations Zn2+ and Ni2+ (17, 18). These findings suggest that the ENaC extracellular domain might regulate epithelial Na+ transport by sensing and integrating diverse signals in the extracellular environment.In the current study, we tested the hypothesis that ENaC activity is regulated by changes in the extracellular Cl concentration. Several observations suggested that Cl might be a strong candidate to regulate the channel. First, transport of Na+ and Cl are often coupled to maintain electroneutrality. Second, ENaC is exposed to large changes in extracellular Cl concentration. For example, in the kidney collecting duct, the urine Cl concentration varies widely (19). As the predominant anion, its concentration parallels that of Na+ in most clinical states. However, under conditions of metabolic alkalosis and metabolic acidosis, the Na+ and Cl concentrations can become dissociated as a result of increased urinary bicarbonate (alkalosis) or ammonium (acidosis) (19). Thus, ENaC is well positioned to respond to changes in Cl concentration. Third, crystallization of ASIC1a revealed a binding site for a Cl ion at the base of the thumb domain (7). The Cl is coordinated by Arg-310 and Glu-314 from one subunit and Lys-212 from an adjacent subunit. Although the functional role of Cl binding to ASIC1a is unknown, it supports the hypothesis that extracellular Cl might regulate the activity of DEG/ENaC ion channels.  相似文献   

9.
Molecular identification of the Ca2+-dependent chloride channel TMEM16A (ANO1) provided a fundamental step in understanding Ca2+-dependent Cl secretion in epithelia. TMEM16A is an intrinsic constituent of Ca2+-dependent Cl channels in cultured epithelia and may control salivary output, but its physiological role in native epithelial tissues remains largely obscure. Here, we demonstrate that Cl secretion in native epithelia activated by Ca2+-dependent agonists is missing in mice lacking expression of TMEM16A. Ca2+-dependent Cl transport was missing or largely reduced in isolated tracheal and colonic epithelia, as well as hepatocytes and acinar cells from pancreatic and submandibular glands of TMEM16A−/− animals. Measurement of particle transport on the surface of tracheas ex vivo indicated largely reduced mucociliary clearance in TMEM16A−/− mice. These results clearly demonstrate the broad physiological role of TMEM16A−/− for Ca2+-dependent Cl secretion and provide the basis for novel treatments in cystic fibrosis, infectious diarrhea, and Sjöegren syndrome.Electrolyte secretion in epithelial tissues is based on the major second messenger pathways cAMP and Ca2+, which activate the cystic fibrosis transmembrane conductance regulator (CFTR)2 Cl channels and Ca2+-dependent Cl channels, respectively (13). CFTR conducts Cl in epithelial cells of airways, intestine, and the ducts of pancreas and sweat gland, while Ca2+-dependent Cl channels secrete Cl in pancreatic acini and salivary and sweat glands (46). Controversy exists as to the contribution of these channels to Cl secretion in submucosal glands of airways and the relevance for cystic fibrosis (79). While cAMP-dependent Cl secretion by CFTR is well examined, detailed analysis of epithelial Ca2+-dependent Cl secretion is hampered by the lack of a molecular counterpart. Although bestrophins may form Ca2+-dependent Cl channels and facilitate Ca2+-dependent Cl secretion in epithelial tissues (10, 11), they are unlikely to form secretory Cl channels in the apical cell membrane, because Ca2+-dependent Cl secretion is still present in epithelia of mice lacking expression of bestrophin (12). Bestrophins may rather have an intracellular function by facilitating receptor mediated Ca2+ signaling and activation of membrane localized channels (13). With the discovery that TMEM16A produces Ca2+-activated Cl currents with biophysical and pharmacological properties close to those in native epithelial tissues, these proteins are now very likely candidates for endogenous Ca2+-dependent Cl channels (1417). In cultured airway epithelial cells, small interfering RNA knockdown of endogenous TMEM16A largely reduced calcium-dependent chloride secretion (16). However, apart from preliminary studies of airways and salivary glands, the physiological significance of TMEM16A in native epithelia, particularly in glands, is unclear (14, 17).  相似文献   

10.
11.
12.
Reduced lipolysis in hormone-sensitive lipase-deficient mice is associated with impaired glucose-stimulated insulin secretion (GSIS), suggesting that endogenous β-cell lipid stores provide signaling molecules for insulin release. Measurements of lipolysis and triglyceride (TG) lipase activity in islets from HSL−/− mice indicated the presence of other TG lipase(s) in the β-cell. Using real time-quantitative PCR, adipose triglyceride lipase (ATGL) was found to be the most abundant TG lipase in rat islets and INS832/13 cells. To assess its role in insulin secretion, ATGL expression was decreased in INS832/13 cells (ATGL-knockdown (KD)) by small hairpin RNA. ATGL-KD increased the esterification of free fatty acid (FFA) into TG. ATGL-KD cells showed decreased glucose- or Gln + Leu-induced insulin release, as well as reduced response to KCl or palmitate at high, but not low, glucose. The KATP-independent/amplification pathway of GSIS was considerably reduced in ATGL-KD cells. ATGL−/− mice were hypoinsulinemic and hypoglycemic and showed decreased plasma TG and FFAs. A hyperglycemic clamp revealed increased insulin sensitivity and decreased GSIS and arginine-induced insulin secretion in ATGL−/− mice. Accordingly, isolated islets from ATGL−/− mice showed reduced insulin secretion in response to glucose, glucose + palmitate, and KCl. Islet TG content and FFA esterification into TG were increased by 2-fold in ATGL−/− islets, but glucose usage and oxidation were unaltered. The results demonstrate the importance of ATGL and intracellular lipid signaling for fuel- and non-fuel-induced insulin secretion.Free fatty acids (FFA)5 and other lipid molecules are important for proper glucose-stimulated insulin secretion (GSIS) by β-cells. Thus, deprivation of fatty acids (FA) in vivo (1) diminishes GSIS, whereas a short term exposure to FFA enhances it (13). In contrast, a sustained provision of FA, particularly in the presence of high glucose in vitro, is detrimental to β-cells in that it reduces insulin gene expression (4) and secretion (5) and induces β-cell apoptosis (6). The FA supply to the β-cells can be from exogenous sources, such as plasma FFAs and lipoproteins, or endogenous sources, such as intracellular triglyceride (TG) stores. Studies from our laboratory (710) and others (11, 12) support the concept that the hydrolysis of endogenous TG plays an important role in fuel-induced insulin secretion because TG depletion with leptin (13) or inhibition of TG lipolysis by lipase inhibitors such as 3,5-dimethylpyrazole (7) or orlistat (11, 12) markedly curtail GSIS in rat islets. Furthermore, mice with β-cell-specific knock-out of hormone-sensitive lipase (HSL), which hydrolyzes both TG and diacylglycerol (DAG), show defective first phase GSIS in vivo and in vitro (14).Lipolysis is an integral part of an essential metabolic pathway, the TG/FFA cycle, in which FFA esterification onto a glycerol backbone leading to the synthesis of TG is followed by its hydrolysis with the release of the FFA that can then be re-esterified. Intracellular TG/FFA cycling is known to occur in adipose tissue of rats and humans (15, 16) and also in liver and skeletal muscle (17). It is generally described as a “futile cycle” as it leads to the net hydrolysis of ATP with the generation of heat (18). However, several studies have shown that this cycle has important functions in the cell. For instance, in brown adipose tissue, it contributes to overall thermogenesis (17, 19). In islets from the normoglycemic, hyperinsulinemic, obese Zucker fatty rat, increased GSIS is associated with increased glucose-stimulated lipolysis and FA esterification, indicating enhanced TG/FFA cycling (10). Stimulation of lipolysis by glucose has also been observed in isolated islets from normal rats (12) and HSL−/− mice (8) indicating the presence of glucose-responsive TG/FFA cycling in pancreatic β-cells.The identity of the key lipases involved in the TG/FFA cycle in pancreatic islets is uncertain. HSL is expressed in islets (20), is up-regulated by long term treatment with elevated glucose (21), and is associated with insulin secretory granules (22). In addition, our earlier results suggested that elevated HSL expression correlates with augmented TG/FFA cycling in islets of Zucker fatty rats (10). However, it appears that other lipases may contribute to lipolysis and the regulation of GSIS in islet tissue. Thus, results from studies using HSL−/− mice showed unaltered GSIS (8, 23), except in fasted male mice (8, 9) in which lipolysis was decreased but not abolished. Furthermore, HSL−/− mice show residual TG lipase activity (8) indicating the presence of other TG lipases.Recently, adipocyte triglyceride lipase (ATGL; also known as Desnutrin, TTS-2, iPLA2-ζ, and PNPLA2) (2426) was found to account for most if not all of the residual lipolysis in HSL−/− mice (26, 27). Two homologues of ATGL, Adiponutrin and GS2, have been described in adipocytes (24). All three enzymes contain a patatin-like domain with broad lipid acyl-hydrolase activity. However, it is not known if adiponutrin and GS2 are actually TG hydrolases. An additional lipase, TG hydrolase or carboxylesterase-3, has been identified in rat adipose tissue (28, 29). Although the hydrolysis of TG is catalyzed by all these lipases, HSL can hydrolyze both TG and DAG, the latter being a better substrate (30).In this study, we observed that besides HSL, ATGL (31), adiponutrin, and GS2 are expressed in rat islets and INS832/13 cells, with ATGL being the most abundant. We then focused on the role of ATGL in fuel-stimulated insulin secretion in two models, INS832/13 β-cells in which ATGL expression was reduced by RNA interference-knockdown (ATGL-KD) and ATGL−/− mice.  相似文献   

13.
To understand the mechanism for ion transport through the sodium/bicarbonate transporter SLC4A4 (NBCe1), we examined amino acid residues, within transmembrane domains, that are conserved among electrogenic Na/HCO3 transporters but are substituted with residues at the corresponding site of all electroneutral Na/HCO3 transporters. Point mutants were constructed and expressed in Xenopus oocytes to assess function using two-electrode voltage clamp. Among the mutants, D555E (charge-conserved substitution of the aspartate at position 555 with a glutamate) produced decreasing HCO3 currents at more positive membrane voltages. Immunohistochemistry showed D555E protein expression in oocyte membranes. D555E induced Na/HCO3-dependent pH recovery from a CO2-induced acidification. Current-voltage relationships revealed that D555E produced an outwardly rectifying current in the nominally CO2/HCO3-free solution that was abolished by Cl removal from the bath. In the presence of CO2/HCO3, however, the outward current produced by D555E decreased only slightly after Cl removal. Starting from a Cl-free condition, D555E produced dose-dependent outward currents in response to a series of chloride additions. The D555E-mediated chloride current decreased by 70% in the presence of CO2/HCO3. The substitution of Asp555 with an asparagine also produced a Cl current. Anion selectivity experiments revealed that D555E was broadly permissive to other anions including NO3. Fluorescence measurements of chloride transport were done with human embryonic kidney HEK 293 cells expressing NBCe1 and D555E. A marked increase in chloride transport was detected in cells expressing D555E. We conclude that Asp555 plays a role in HCO3 selectivity.The electrogenic Na/HCO3 cotransporter NBCe1 (SLC4A4) is one of the SLC4A gene family members transporting HCO3 across the plasma membrane (13). NBCe1 plays a role in transepithelial HCO3 movement and pHi regulation in many tissues (46). NBCe1 is responsible for HCO3 reabsorption in the proximal tubules of the kidney (7). The proximal tubule cells reclaim HCO3 from the lumen through a series of reactions involving titration of HCO3 by H+ secretion via the apical Na/H exchanger, production of CO2, and regeneration of HCO3 and H+ in the tubule cells. HCO3 then moves to the interstitium via the basolateral NBCe1. The essential feature driving this basolateral Na+/HCO3 exit is the stoichiometry of 1:3 Na+:HCO3, which makes the equilibrium potential for NBCe1 more positive than the resting membrane potential of the proximal tubule cells (8). The stoichiometry of 1Na+:1HCO3 or 1Na+:2HCO3 causes both ions to move into cells in other tissues such as pancreas, brain, and cardiovascular tissues (9, 10).Despite the importance of NBCe1 for basolateral HCO3 reabsorption in the proximal tubules, the mechanism of electrogenic Na/HCO3 transport via the transporter is not well understood. Ion movement depends on loading ions at their translocation or binding sites that likely reside within the membrane field at some distance from the bath solution (11). This implies that the transmembrane domains (TMs)2 of NBCe1 and amino acid residues within TMs play critical roles in ion transport.Sequence analysis of different SLC4A proteins shows similar hydropathy plots, predicting that these proteins share structural elements of transport function (12). Such similarities have facilitated structure/function studies to define molecular domains or motifs responsible for conferring Na/HCO3 transport of NBCe1. Abuladze et al. (13) performed a large scale mutagenesis on acidic and basic amino acids in non-TMs and found many residues affecting Na+-dependent base flux. McAlear et al. (14) identified amino acids in TM8 involving ion translocation. By a systematic approach of chimeric transporters between NBCe1 and the electroneutral Na/HCO3 cotransporter NBCn1 (SLC4A7) (15), we and our colleagues (16) demonstrated that electrogenic Na/HCO3 transport of NBCe1 requires interactions between the regions TM1–5 and TM6–13 of the protein. Zhu et al. (17) recently proposed TM1 as a domain lining the ion translocation pathway. On the other hand, Chang et al. (18) reported that the cytoplasmic N-terminal domain might contribute to HCO3 permeation.In the present study, we searched amino acid residues that are highly conserved among electrogenic Na/HCO3 transporters but not among electroneutral Na/HCO3 transporters and examined their role in electrogenic Na/HCO3 transport. Nine candidate residues in human renal NBCe1-A (5, 19) were selected and mutated by replacement with the amino acids at the corresponding sites of NBCn1. Mutant transporters were expressed in Xenopus oocytes and assessed via two-electrode voltage clamp. Our data show that Asp555 of NBCe1 plays an important role in HCO3 selectivity.  相似文献   

14.
15.
16.
17.
NH4+ inhibition kinetics for CH4 oxidation were examined at near-atmospheric CH4 concentrations in three upland forest soils. Whether NH4+-independent salt effects could be neutralized by adding nonammoniacal salts to control samples in lieu of deionized water was also investigated. Because the levels of exchangeable endogenous NH4+ were very low in the three soils, desorption of endogenous NH4+ was not a significant factor in this study. The Km(app) values for water-treated controls were 9.8, 22, and 57 nM for temperate pine, temperate hardwood, and birch taiga soils, respectively. At CH4 concentrations of ≤15 μl liter−1, oxidation followed first-order kinetics in the fine-textured taiga soil, whereas the coarse-textured temperate soils exhibited Michaelis-Menten kinetics. Compared to water controls, the Km(app) values in the temperate soils increased in the presence of NH4+ salts, whereas the Vmax(app) values decreased substantially, indicating that there was a mixture of competitive and noncompetitive inhibition mechanisms for whole NH4+ salts. Compared to the corresponding K+ salt controls, the Km(app) values for NH4+ salts increased substantially, whereas the Vmax(app) values remained virtually unchanged, indicating that NH4+ acted by competitive inhibition. Nonammoniacal salts caused inhibition to increase with increasing CH4 concentrations in all three soils. In the birch taiga soil, this trend occurred with both NH4+ and K+ salts, and the slope of the increase was not affected by the addition of NH4+. Hence, the increase in inhibition resulted from an NH4+-independent mechanism. These results show that NH4+ inhibition of atmospheric CH4 oxidation resulted from enzymatic substrate competition and that additional inhibition that was not competitive resulted from a general salt effect that was independent of NH4+.Atmospheric CH4 contributes substantially to the greenhouse effect, and the concentration of atmospheric CH4 has increased dramatically in the past century because of human activity associated with agriculture, land use changes, and industry (34, 35). Bacterial oxidation of atmospheric CH4 in well-drained soils is an important regulator of atmospheric CH4 concentration, yet the organisms responsible remain unidentified and the physiology of the process is poorly understood (9, 35, 36). Although soil CH4 consumption is inhibited by a wide variety of anthropogenic disturbances, such as agriculture, N deposition, and forestry (12, 17, 22, 23, 32, 43, 44), predictable inhibition patterns have failed to emerge, which has made it difficult to predict the effects of disturbance on soil CH4 flux in various ecosystems. The most commonly reported disturbance effect is that of NH4+ fertilizers, which can suppress soil CH4 consumption by up to 70% (1, 8, 10, 17, 22, 32, 33, 37, 38, 43). In the field, inhibition may occur immediately following fertilization, may be delayed for months to years, or may never occur despite years of chronic fertilization (9, 17). This variety of responses may stem at least in part from the distribution of physiologically diverse methane oxidizer populations across sites (17, 18, 20).Of the various NH4+ inhibition patterns, immediate inhibition is the best documented. As in field studies, however, physiological laboratory studies have produced variable results, suggesting that there may be multiple inhibition mechanisms (15, 17, 2628, 36, 39). Physicochemical similarities between CH4 and NH3 may permit these two compounds to compete for enzyme active sites so that fortuitous NH3 oxidation competitively inhibits CH4 oxidation (38). Although this mechanism has been demonstrated to occur in pure cultures of methanotrophic bacteria (6) and in a CH4-producing agricultural soil (15), it has not been demonstrated to occur in well-drained, nonagricultural mineral soils, which comprise the dominant terrestrial sink for atmospheric CH4 (14, 38, 45), nor has it been demonstrated to occur at near-atmospheric CH4 concentrations. In many cases, the kinetics of immediate NH4+ inhibition in soil cannot be reconciled easily with substrate competition (15, 16, 2628, 39). An alternative mechanism has been proposed, whereby the toxicity of NO2 or NH2OH produced by fortuitous NH4+ oxidation suppresses methanotrophic activity (26, 27, 39). Hence, multiple inhibition mechanisms may be involved, and these mechanisms may vary with the physiology of different CH4 oxidizer populations (17).Two physiologically distinct communities of CH4 oxidizers apparently exist in soil. One group, generally associated with atmospheric CH4 consumption, exhibits an extremely high affinity for CH4. Representatives of this group have yet to be cultivated or otherwise identified (9). The second group exhibits a much lower affinity for CH4 and is generally associated with common methanotrophs, such as those that have been studied in pure culture for many years (2, 9). In upland mineral soils, only high-affinity activity is usually detectable without artificial enrichment with high CH4 concentrations in the laboratory. However, the only prior study in which kinetic constants for NH4+ inhibition of soil CH4 oxidation were reported was conducted in a periodically moist, organic matter-rich agricultural soil with demonstrable methanogenesis (15, 16). Such a soil potentially harbors a rich community of CH4 oxidizers representing a continuum from low-affinity organisms to high-affinity organisms. Although this important investigation demonstrated that NH4+ inhibits CH4 oxidation via enzymatic substrate competition in an agricultural humisol, it is unclear to what extent its results apply to well-drained mineral soils lacking endogenous CH4 sources. Physiological studies of soil CH4 oxidation typically derive kinetic constants from oxidation rates at CH4 concentrations ranging from atmospheric levels (∼1.7 μl liter−1) to ≫Km for high-affinity CH4 oxidizers. Even in soil in which only high-affinity organisms are active, the CH4-oxidizing enzyme(s) could respond differently to NH4+ at high CH4 concentrations than at near-atmospheric concentrations (15, 39). Thus, to study NH4+ inhibition of high-affinity CH4 oxidizers per se, it would be preferable to examine inhibition kinetics at near-atmospheric CH4 concentrations in a soil with no apparent endogenous CH4 source.A common shortcoming of NH4+ inhibition studies, regardless of the organisms involved, has been a lack of attention to nonammoniacal salt effects despite numerous reports of substantial inhibition by such salts (1, 10, 15, 17, 24). King and Schnell (28) examined the effects of several Cl and SO42− salts and concluded that nonammoniacal salts indirectly inhibit CH4 oxidation by desorbing endogenous NH4+ from cation exchange sites in the soil, which then directly inhibit CH4 oxidation. Many N-limited soils, however, have extremely low concentrations of exchangeable NH4+, yet are substantially inhibited by nonammoniacal salts (17), suggesting that these salts have NH4+-independent effects on atmospheric CH4 oxidizers. Additional mechanisms may alter inhibition kinetics, thus hindering the diagnosis of NH4+-specific inhibition.With the limitations described above in mind, we used a simple steady-state kinetics approach to assess the mechanism of NH4+ inhibition of CH4 oxidation at near-atmospheric concentrations (1.8 to 15 μl liter−1) in three well-drained, N-limited forest soils that lack known endogenous CH4 sources. In addition, we examined the effects of nonammoniacal salts in parallel samples to judge the utility of these salts as experimental controls for neutralizing NH4+-independent salt effects.  相似文献   

18.
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   

19.
20.
Of the eight known nickel enzymes, all but glyoxylase I catalyze the use and/or production of gases central to the global carbon, nitrogen, and oxygen cycles. Nickel appears to have been selected for its plasticity in coordination and redox chemistry and is able to cycle through three redox states (1+, 2+, 3+) and to catalyze reactions spanning ∼1.5 V. This minireview focuses on the catalytic mechanisms of nickel enzymes, with an emphasis on the role(s) of the metal center. The metal centers vary from mononuclear to complex metal clusters and catalyze simple hydrolytic to multistep redox reactions.Seven of the eight known nickel enzymes (1). CODH2 interconverts CO and CO2; ACS utilizes CO; the nickel ARD produces CO; hydrogenase generates/utilizes hydrogen gas; MCR generates methane; urease produces ammonia; and SOD generates O2.

TABLE 1

Nickel-containing enzymes
EnzymeReactionRef.
Glx I (EC 4.4.1.5)Methylglyoxal → lactate + H2O (Reaction 1)7
ARD (EC 1.13.11.54)1,2-Dihydroxy-3-oxo-5-(methylthio)pent-1-ene + O2 → HCOOH + methylthiopropionate + CO (Reaction 2)9, 10
Ni-SOD (EC 1.15.1.1)2H+ + 2O2−̇ → H2O2 + O2 (Reaction 3)16, 19
Urease (EC 3.5.1.5)H2N-CO-NH2 + 2H2O → 2NH3 + H2CO3 (Reaction 4)22, 23
Hydrogenase (EC 1.12.X.X)2H+ + 2e ⇌ H2E0′ = −414 mV) (Reaction 5)25, 26
MCR (EC 2.8.4.1)CH3-CoM + CoBSH → CH4 + CoM-SS-CoB (Reaction 6)42, 43
CODH (EC 1.2.99.2)2e + 2H+ + CO2 ⇌ CO + H2O (E0′ = −558 mV) (Reaction 7)30, 31
ACS (EC 2.3.1.169)CH3-CFeSP + CoASH + CO → CH3-CO-SCoA + CFeSP (Reaction 8)30, 31
Open in a separate windowThe nickel sites in enzymes exhibit extreme plasticity in nickel coordination and redox chemistry. The metal center in SOD must be able to redox processes with potentials that span from +890 to −160 mV (2), whereas in MCR and CODH, it must be able to reach potentials as low as −600 mV (3); thus, nickel centers in proteins perform redox chemistry over a potential range of ∼1.5 V!Because natural environments contain only trace amounts of soluble Ni2+, attaining sufficiently high intracellular nickel concentrations to meet the demand of the nickel enzymes requires a high affinity nickel uptake system(s) (4), molecular and metallochaperones (5), and sensors and regulators of the levels of enzymes involved in nickel homeostasis (6). However, space limitations prevent coverage of these pre-catalytic events.  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