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 31 毫秒
1.
A novel dehalogenating/transhalogenating enzyme, halomethane:bisulfide/halide ion methyltransferase, has been isolated from the facultatively methylotrophic bacterium strain CC495, which uses chloromethane (CH(3)Cl) as the sole carbon source. Purification of the enzyme to homogeneity was achieved in high yield by anion-exchange chromatography and gel filtration. The methyltransferase was composed of a 67-kDa protein with a corrinoid-bound cobalt atom. The purified enzyme was inactive but was activated by preincubation with 5 mM dithiothreitol and 0.5 mM CH(3)Cl; then it catalyzed methyl transfer from CH(3)Cl, CH(3)Br, or CH(3)I to the following acceptor ions (in order of decreasing efficacy): I(-), HS(-), Cl(-), Br(-), NO(2)(-), CN(-), and SCN(-). Spectral analysis indicated that cobalt in the native enzyme existed as cob(II)alamin, which upon activation was reduced to the cob(I)alamin state and then was oxidized to methyl cob(III)alamin. During catalysis, the enzyme shuttles between the methyl cob(III)alamin and cob(I)alamin states, being alternately demethylated by the acceptor ion and remethylated by halomethane. Mechanistically the methyltransferase shows features in common with cobalamin-dependent methionine synthase from Escherichia coli. However, the failure of specific inhibitors of methionine synthase such as propyl iodide, N(2)O, and Hg(2+) to affect the methyltransferase suggests significant differences. During CH(3)Cl degradation by strain CC495, the physiological acceptor ion for the enzyme is probably HS(-), a hypothesis supported by the detection in cell extracts of methanethiol oxidase and formaldehyde dehydrogenase activities which provide a metabolic route to formate. 16S rRNA sequence analysis indicated that strain CC495 clusters with Rhizobium spp. in the alpha subdivision of the Proteobacteria and is closely related to strain IMB-1, a recently isolated CH(3)Br-degrading bacterium (T. L. Connell Hancock, A. M. Costello, M. E. Lidstrom, and R. S. Oremland, Appl. Environ. Microbiol. 64:2899-2905, 1998). The presence of this methyltransferase in bacterial populations in soil and sediments, if widespread, has important environmental implications.  相似文献   

2.
The anaerobic veratrol O-demethylase mediates the transfer of the methyl group of the phenyl methyl ether veratrol to tetrahydrofolate. The primary methyl group acceptor is the cobalt of a corrinoid protein, which has to be in the +1 oxidation state to bind the methyl group. Due to the negative redox potential of the cob(II)/cob(I)alamin couple, autoxidation of the cobalt may accidentally occur. In this study, the reduction of the corrinoid to the superreduced [CoI] state was investigated. The ATP-dependent reduction of the corrinoid protein of the veratrol O-demethylase was shown to be dependent on titanium(III) citrate as electron donor and on an activating enzyme. In the presence of ATP, activating enzyme, and Ti(III), the redox potential versus the standard hydrogen electrode (E SHE) of the cob(II)alamin/cob(I)alamin couple in the corrinoid protein was determined to be −290 mV (pH 7.5), whereas E SHE at pH 7.5 was lower than −450 mV in the absence of either activating enzyme or ATP. ADP, AMP, or GTP could not replace ATP in the activation reaction. The ATP analogue adenosine-5′-(β,γ-imido)triphosphate (AMP-PNP, 2–4 mM) completely inhibited the corrinoid reduction in the presence of ATP (2 mM).  相似文献   

3.
We tested a number of inhibitory monovalent anions for their primary site of action on photosystem II(PSII) in chloroplasts. We find that the inhibitory effects of F, HCO2, NO2, NO3, and CH3CO2 are all reversed by addition of a high concentration of HCO3. This class of anions competitively inhibits H14CO3 binding to PSII. All of those anions tested reduced H14CO3 binding more in the light than in the dark. We conclude that the primary inhibitory site of action of a number of monovalent anions is at the HCO3 binding site(s) on the PSII complex. The carbonic anhydrase inhibitor gold cyanide, and also azide, inhibit PSII but at a site other than the HCO3 binding site. We suggest that the unique ability of HCO3 to reverse the effects of inhibitory anions reflects its singular ability to act as a proton donor/acceptor at the anion binding site. A similar role has been proposed for non-substrate-bound HCO3 on carbonic anhydrase by Yeagle et al. (1975 Proc Natl Acad Sci USA 72: 454-458).  相似文献   

4.
Almost all of the chlorine-containing gas emitted from natural sources is methyl chloride (CH3Cl), which contributes to the destruction of the stratospheric ozone layer. Tropical and subtropical plants emit substantial amounts of CH3Cl. A gene involved in CH3Cl emission from Arabidopsis was previously identified and designated HARMLESS TO OZONE LAYER (hereafter AtHOL1) based on the mutant phenotype. Our previous studies demonstrated that AtHOL1 and its homologs, AtHOL2 and AtHOL3, have S-adenosyl-l-methionine-dependent methyltransferase activities. However, the physiological functions of AtHOLs have yet to be elucidated. In the present study, our comparative kinetic analyses with possible physiological substrates indicated that all of the AtHOLs have low activities toward chloride. AtHOL1 was highly reactive to thiocyanate (NCS), a pseudohalide, synthesizing methylthiocyanate (CH3SCN) with a very high kcat/Km value. We demonstrated in vivo that substantial amounts of NCS were synthesized upon tissue damage in Arabidopsis and that NCS was largely derived from myrosinase-mediated hydrolysis of glucosinolates. Analyses with the T-DNA insertion Arabidopsis mutants (hol1, hol2, and hol3) revealed that only hol1 showed increased sensitivity to NCS in medium and a concomitant lack of CH3SCN synthesis upon tissue damage. Bacterial growth assays indicated that the conversion of NCS into CH3SCN dramatically increased antibacterial activities against Arabidopsis pathogens that normally invade the wound site. Furthermore, hol1 seedlings showed an increased susceptibility toward an Arabidopsis pathogen, Pseudomonas syringae pv. maculicola. Here we propose that AtHOL1 is involved in glucosinolate metabolism and defense against phytopathogens. Moreover, CH3Cl synthesized by AtHOL1 could be considered a byproduct of NCS metabolism.Methyl chloride (CH3Cl) is the most abundant halohydrocarbon emitted into the atmosphere and constitutes about 17% of the chlorine currently in the stratosphere (1). CH3Cl is derived mainly from natural sources and contributes to the destruction of the stratospheric ozone layer. As the total abundance of ozone-depleting gases such as chlorofluorocarbons in the atmosphere has begun to decrease in recent years as a result of The Montreal Protocol on Substances That Deplete the Ozone Layer, the impact of CH3Cl emission from natural sources will become greater on the atmospheric chemistry. CH3Cl emission into the atmosphere has been estimated at 1,700–13,600 Gg/year (1), which underscores the great uncertainty of the estimation. Oceans (2), biomass burning (3), wood-rotting fungi, and coastal salt marshes (4) are the major sources of CH3Cl production. Recently, it was reported that large amounts of CH3Cl are emitted from tropical and subtropical plants, which are hence considered as the major sources of CH3Cl (57). It was estimated that the CH3Cl emission from tropical plants could account for 30–50% of the global CH3Cl emission (8). To accomplish an accurate estimation of CH3Cl production in the atmosphere through “bottom-up” approaches, elucidating the mechanisms and physiological functions of CH3Cl emission from plants will be important.The biological synthesis of methyl halides has been demonstrated mainly by biochemical analyses. The enzymatic activities that transfer a methyl group from S-adenosyl-l-methionine (SAM)2 to halide ions (Cl, Br, I), which synthesize methyl halides, were first discovered in cell-free extracts of Phellinus pomaceus (a white rot fungus), Endocladia muricata (a marine red alga), and Mesembryanthemum crystallinum (ice plant, a halophytic plant) (9). Enzyme purification and cDNA cloning of the methyl chloride transferase (MCT) was first reported with Batis maritima, a halophytic plant that grows abundantly in salt marshes. As high concentrations of ions such as Cl are often detrimental to plants, halophytic plants are considered to possess various salt tolerance mechanisms. MCT was hypothesized to control and regulate the internal concentration of Cl, rich in the habitat in which halophytic plant grows (10, 11).In the meantime, purification of thiol methyltransferase (TMT), which methylates bisulfide (HS) and halide (Cl, Br, I) ions was reported with cabbage, Brassica oleracea (12). The purified and recombinant TMTs were later shown to also methylate the thiocyanate ion (NCS), which is called pseudohalide because of its chemical properties similar to halide ions (13, 14). NCS is a hydrolysis product found in some glucosinolates, which are secondary metabolites found mainly in the order Brassicales including the model plant Arabidopsis thaliana (15). Upon tissue damage such as by insect or herbivore attack, glucosinolates are hydrolyzed by myrosinase (β-thioglucosidase) into biologically active compounds including isothiocyanates. Isothiocyanates derived from indole glucosinolates and 4-hydroxybenzyl glucosinolates are reported to be highly unstable and yield NCS upon reacting with various nucleophiles (1517). Based on the enzymatic activity, the physiological role of TMT was speculated to metabolize glucosinolate breakdown products (14). However, there are no reported studies that examine these MCT and TMT hypotheses through in vivo experiments.An Arabidopsis homolog of MCT was also identified, and its T-DNA insertion Arabidopsis mutants were analyzed (18). Because the gene disruption eliminated almost all of the methyl halide emissions from the mutants, the gene was revealed to be involved in methyl halide synthesis and was designated HOL (HARMLESS TO OZONE LAYER; denoted as AtHOL1 in our studies) based on the mutant phenotype (18). Recently, we identified AtHOL1 homologs AtHOL2 and AtHOL3 in Arabidopsis, and we demonstrated biochemically that the three recombinant AtHOLs have SAM-dependent methyltransferase activities (19). In this study, reverse genetic and biochemical analyses of all AtHOL isoforms revealed that AtHOL1 in vivo is involved in the methylation of NCS produced by glucosinolate hydrolysis. Although there are several studies that have examined the biological activities of glucosinolate hydrolysis products, the mechanisms of NCS synthesis and its methylation to methyl thiocyanate (CH3SCN) have yet to be reported in detail. The biological activity and physiological function of CH3SCN synthesized by AtHOL1 was also examined.  相似文献   

5.
Propionate consumption was studied in syntrophic batch and chemostat cocultures of Syntrophobacter fumaroxidans and Methanospirillum hungatei. The Gibbs free energy available for the H2-consuming methanogens was <−20 kJ mol of CH4−1 and thus allowed the synthesis of 1/3 mol of ATP per reaction. The Gibbs free energy available for the propionate oxidizer, on the other hand, was usually >−10 kJ mol of propionate−1. Nevertheless, the syntrophic coculture grew in the chemostat at steady-state rates of 0.04 to 0.07 day−1 and produced maximum biomass yields of 2.6 g mol of propionate−1 and 7.6 g mol of CH4−1 for S. fumaroxidans and M. hungatei, respectively. The energy efficiency for syntrophic growth of S. fumaroxidans, i.e., the biomass produced per unit of available Gibbs free energy was comparable to a theoretical growth yield of 5 to 12 g mol of ATP−1. However, a lower growth efficiency was observed when sulfate served as an additional electron acceptor, suggesting inefficient energy conservation in the presence of sulfate. The maintenance Gibbs free energy determined from the maintenance coefficient of syntrophically grown S. fumaroxidans was surprisingly low (0.14 kJ h−1 mol of biomass C−1) compared to the theoretical value. On the other hand, the Gibbs free-energy dissipation per mole of biomass C produced was much higher than expected. We conclude that the small Gibbs free energy available in many methanogenic environments is sufficient for syntrophic propionate oxidizers to survive on a Gibbs free energy that is much lower than that theoretically predicted.  相似文献   

6.
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   

7.
Biogenic emissions of halomethanes (CH3CI, CH3Br and CH3I) and methanethiol (CH3SH) are of major significance to atmospheric chemistry, but there is little information on such emissions from higher plants. We present evidence that plants can produce all these gases through an identical methyltransferase reaction. A survey of 118 herbaceous species, based on CH3I production by leaf discs supplied with Kl, detected the presence of in vivo halide methyltransferase activity in 87 species. The activities ranged over nearly 4 orders of magnitude. Plants generally considered salt tolerant had relatively low activities, and salinization of three such species did not increase the activity. The highest activities were found in the family Brassicaceae. Leaf extracts of Brassica oleracea catalysed the S-adenosyl-L-methioninc-dependenl niethylalion of the halides I?, Br? and CI? to the respective halomethanes. In addition, the extract similarly methylated HS? (bisulphide) to CH3SH. These two types of enzyme activity (halide and bisulphide methyltransferase) were also present in all of the 20 species comprising a subsample that represented the range of CH3I emissions observed in the initial survey of in vivo CH3I production ability, and in a marine red alga Endocladia muricata. Moreover, the two activities occurred in approximately the same ratio in all the higher plants tested. These findings highlight the potential of higher plants to contribute to the atmosphericbudget of halomethanes and melhanethiol. The halide and bisulphide methyltransferase activities may also provide a mechanism for the elimination of halide and HS? ions, both of which are known to be phytotoxic.  相似文献   

8.
Kinetics of Denitrifying Growth by Fast-Growing Cowpea Rhizobia   总被引:3,自引:2,他引:1       下载免费PDF全文
Two fast-growing strains of cowpea rhizobia (A26 and A28) were found to grow anaerobically at the expense of NO3, NO2, and N2O as terminal electron acceptors. The two major differences between aerobic and denitrifying growth were lower yield coefficients (Y) and higher saturation constants (Ks) with nitrogenous oxides as electron acceptors. When grown aerobically, A26 and A28 adhered to Monod kinetics, respectively, as follows: Ks, 3.4 and 3.8 μM; Y, 16.0 and 14.0 g · cells eq−1; μmax, 0.41 and 0.33 h−1. Yield coefficients for denitrifying growth ranged from 40 to 70% of those for aerobic growth. Only A26 adhered to Monod kinetics with respect to growth on all three nitrogenous oxides. The apparent Ks values were 41, 270, and 460 μM for nitrous oxide, nitrate, and nitrite, respectively; the Ks for A28 grown on nitrate was 250 μM. The results are kinetically and thermodynamically consistent in explaining why O2 is the preferred electron acceptor. Although no definitive conclusions could be drawn regarding preferential utilization of nitrogenous oxides, nitrite was inhibitory to both strains and effected slower growth. However, growth rates were identical (μmax, 0.41 h−1) when A26 was grown with either O2 or NO3 as an electron acceptor and were only slightly reduced when A28 was grown with NO3 (0.25 h−1) as opposed to O2 (0.33 h−1).  相似文献   

9.
A series of [ω-(adenosin-5′-O-yl)alkyl]cobalamins were examined for their inhibitory properties of ribonucleoside triphosphate reductase (RTPR) from Lactobacillus leichmannii in the presence of 5′-deoxyadenosylcobalamin (AdoCbl, Coenzyme B12). These AdoCbl analogs, in which oligomethylene chains (C3-C7) were inserted between the corrin Co-atom and a 5′-O-atom of the adenosine moiety, were designed to probe the Co-C bond posthomolysis state in AdoCbl-dependent enzymes, a state in which the Co and 5′-C distance is believed to be significantly increased. Experimentally, all five analogs were competitive inhibitors, with Ki in the range of 8–56 μM. The [ω-(adenosin-5′-O-yl)alkyl]cobalamin analog with C5 methylene carbons was the strongest inhibitor. This same pattern of inhibition, in which the C5-analog is the strongest inhibitor, was previously observed in the AdoCbl-dependent eliminase enzyme systems, diol dehydratase and glycerol dehydratase. However, in methylmalonyl CoA mutase, the strongest inhibition is by the C6-analog. This supports the hypothesis that the cobalamin posthomolysis intermediate in the eliminase enzymes differs from that in the mutase enzymes. These findings led, in turn, to an examination of the visible spectra of enzyme-bound cob(II)alamin in these two subclasses of AdoCbl-dependent enzymes. The results reveal an additional insight into the difference between the two classes: in the eliminases, the γ-band of bound cob(II)alamin is shifted from the 473 nm for free cob(II)alamin to longer wavelengths, 475–480 nm. However, in mutases, the γ-band of bound cob(II)alamin is shifted to shorter wavelengths, 465–470 nm. Overall, the results (a) provide strong evidence that two subclasses of AdoCbl-dependent enzymes exist, (b) give insights into the probable posthomolysis state in RTPR and other eliminases, and (c) identifies the C5-analog as the tightest-binding analog for crystallization and other biophysical studies.  相似文献   

10.
A lithotrophic freshwater Beggiatoa strain was enriched in O2-H2S gradient tubes to investigate its ability to oxidize sulfide with NO3 as an alternative electron acceptor. The gradient tubes contained different NO3 concentrations, and the chemotactic response of the Beggiatoa mats was observed. The effects of the Beggiatoa sp. on vertical gradients of O2, H2S, pH, and NO3 were determined with microsensors. The more NO3 that was added to the agar, the deeper the Beggiatoa filaments glided into anoxic agar layers, suggesting that the Beggiatoa sp. used NO3 to oxidize sulfide at depths below the depth that O2 penetrated. In the presence of NO3 Beggiatoa formed thick mats (>8 mm), compared to the thin mats (ca. 0.4 mm) that were formed when no NO3 was added. These thick mats spatially separated O2 and sulfide but not NO3 and sulfide, and therefore NO3 must have served as the electron acceptor for sulfide oxidation. This interpretation is consistent with a fourfold-lower O2 flux and a twofold-higher sulfide flux into the NO3-exposed mats compared to the fluxes for controls without NO3. Additionally, a pronounced pH maximum was observed within the Beggiatoa mat; such a pH maximum is known to occur when sulfide is oxidized to S0 with NO3 as the electron acceptor.  相似文献   

11.
1. The influence of Cl, Br, NO3 and F ions on the visible-absorption spectrum of deionized aspartate aminotransferase was investigated. 2. Except for F, these anions caused an increase of the extinction at 430mμ with a concomitant decrease of that at 362mμ. 3. The affinity constants for Cl and NO3 ions were calculated by a procedure based on the assumption that the anion stabilizes the protonated form of the enzyme chromophore (λmax. 430mμ). 4. The true pK of the chromophore of the enzyme was found to be 5·25.  相似文献   

12.
The H+-ATPase of tonoplast vesicles isolated from red beet (Beta vulgaris L.) storage tissue was studied with respect to the kinetic effects of Cl and NO3. N-Ethylmaleimide (NEM) was employed as a probe to investigate substrate binding and gross conformational changes of the enzyme. Chloride decreased the Km of the enzyme for ATP but caused relatively little alteration of the Vmax. Nitrate increased Km only. Michaelis-Menten kinetics applied throughout with respect to ATP concentration. Nitrate yielded similar kinetics of inhibition in both the presence and absence of Cl. Other monovalent anions that specifically increased the Km of the ATPase for ATP were, in order of increasing Ki, SCN, ClO4, and ClO3. Sulfate, although inhibitory, manifested noncompetitive kinetics with respect to ATP concentration. ADP, like NO3, was a competitive inhibitor of the ATPase but ADP and NO3 did not interact cooperatively nor did either interfere with the inhibitory action of the other. It is concluded that NO3 does not show competitive kinetics because of its stereochemical similarity to the terminal phosphoryl group of ATP. NEM was an irreversible inhibitor of the tonoplast ATPase. Both Mg·ADP and Mg·ATP protected the enzyme from inactivation by NEM but Mg·ADP was the more potent of the two. Chloride and NO3 exerted little or no effect on the protective actions of Mg·ADP and Mg·ATP suggesting that neither Cl nor NO3 are involved in substrate binding.  相似文献   

13.
A modified 3-hydroxypropionate cycle has been proposed as the autotrophic CO2 fixation pathway for the thermoacidophilic crenarchaeon Metallosphaera sedula. The cycle requires the reductive conversion of 3-hydroxypropionate to propionyl-coenzyme A (propionyl-CoA). The specific activity of the 3-hydroxypropionate-, CoA-, and MgATP-dependent oxidation of NADPH in autotrophically grown cells was 0.023 μmol min−1mg protein−1. The reaction sequence is catalyzed by at least two enzymes. The first enzyme, 3-hydroxypropionyl-CoA synthetase, catalyzes the following reaction: 3-hydroxypropionate + ATP + CoA → 3-hydroxypropionyl-CoA + AMP + PPi. The enzyme was purified 95-fold to a specific activity of 18 μmol min−1 mg protein−1 from autotrophically grown M. sedula cells. An internal peptide sequence was determined and a gene encoding a homologous protein identified in the genome of Sulfolobus tokodaii; similar genes were found in S. solfataricus and S. acidocaldarius. The gene was heterologously expressed in Escherichia coli, and the His-tagged protein was purified. Both the native enzyme from M. sedula and the recombinant enzyme from S. tokodaii not only activated 3-hydroxypropionate to its CoA ester but also activated propionate, acrylate, acetate, and butyrate; however, with the exception of propionate, the affinities for these substrates were reduced. 3-Hydroxypropionyl-CoA synthetase is up-regulated eightfold in autotrophically versus heterotrophically grown M. sedula, supporting its proposed role during CO2 fixation in this archaeon and possibly other members of the Sulfolobaceae family.  相似文献   

14.
Rapid Methane Oxidation in a Landfill Cover Soil   总被引:28,自引:5,他引:28       下载免费PDF全文
Methane oxidation rates observed in a topsoil covering a retired landfill are the highest reported (45 g m−2 day−1) for any environment. This microbial community had the capacity to rapidly oxidize CH4 at concentrations ranging from <1 ppm (microliters per liter) (first-order rate constant [k] = −0.54 h−1) to >104 ppm (k = −2.37 h−1). The physiological characteristics of a methanotroph isolated from the soil (characteristics determined in aqueous medium) and the natural population, however, were similar to those of other natural populations and cultures: the Q10 and optimum temperature were 1.9 and 31°C, respectively, the apparent half-saturation constant was 2.5 to 9.3 μM, and 19 to 69% of oxidized CH4 was assimilated into biomass. The CH4 oxidation rate of this soil under waterlogged (41% [wt/vol] H2O) conditions, 6.1 mg liter−1 day−1, was near rates reported for lake sediment and much lower than the rate of 116 mg liter−1 day−1 in the same soil under moist (11% H2O) conditions. Since there are no large physiological differences between this microbial community and other CH4 oxidizers, we attribute the high CH4 oxidation rate in moist soil to enhanced CH4 transport to the microorganisms; gas-phase molecular diffusion is 104-fold faster than aqueous diffusion. These high CH4 oxidation rates in moist soil have implications that are important in global climate change. Soil CH4 oxidation could become a negative feedback to atmospheric CH4 increases (and warming) in areas that are presently waterlogged but are projected to undergo a reduction in summer soil moisture.  相似文献   

15.
Shewanella oneidensis MR-1, a facultatively anaerobic gammaproteobacterium, respires a variety of anaerobic terminal electron acceptors, including the inorganic sulfur compounds sulfite (SO32−), thiosulfate (S2O32−), tetrathionate (S4O62−), and elemental sulfur (S0). The molecular mechanism of anaerobic respiration of inorganic sulfur compounds by S. oneidensis, however, is poorly understood. In the present study, we identified a three-gene cluster in the S. oneidensis genome whose translated products displayed 59 to 73% amino acid similarity to the products of phsABC, a gene cluster required for S0 and S2O32− respiration by Salmonella enterica serovar Typhimurium LT2. Homologs of phsA (annotated as psrA) were identified in the genomes of Shewanella strains that reduce S0 and S2O32− yet were missing from the genomes of Shewanella strains unable to reduce these electron acceptors. A new suicide vector was constructed and used to generate a markerless, in-frame deletion of psrA, the gene encoding the putative thiosulfate reductase. The psrA deletion mutant (PSRA1) retained expression of downstream genes psrB and psrC but was unable to respire S0 or S2O32− as the terminal electron acceptor. Based on these results, we postulate that PsrA functions as the main subunit of the S. oneidensis S2O32− terminal reductase whose end products (sulfide [HS] or SO32−) participate in an intraspecies sulfur cycle that drives S0 respiration.Microbial reduction of inorganic sulfur compounds is central to the biogeochemical cycling of sulfur and other elements such as carbon and metals (29). The ability to reduce elemental sulfur (S0) is found in members of both prokaryotic domains (20), including mesophilic deltaproteobacteria (Desulfovibrio vulgaris, Pelobacter carbinolicus, Geobacter sulfurreducens) (6, 9, 36, 51), thermophilic deltaproteobacteria (Desulfurella acetivorans) (39), gammaproteobacteria (Shewanella putrefaciens) (41), epsilonproteobacteria (Wolinella succinogenes) (49), cyanobacteria (“Oscillatoria limnetica”) (45), and hyperthermophilic archaea (1, 53). Partially reduced inorganic sulfur compounds such as tetrathionate (S4O62−), thiosulfate (S2O32−), and sulfite (SO32−) are also important electron acceptors in the biogeochemical cycling of sulfur (29, 51). S4O62−-reducing bacteria, for example, may produce S2O32− as a metabolic end product of S4O62− reduction, while S2O32− disproportionation is a key reaction catalyzed by sulfate-reducing bacteria, resulting in the formation of sulfate (SO42−) and sulfide (S2−) (26).Shewanella oneidensis MR-1, a facultatively anaerobic gammaproteobacterium, respires a variety of compounds as an anaerobic electron acceptor, including the inorganic sulfur compounds S0, SO32−, S2O32−, and S4O62−; transition metals [e.g., Fe(III) and Mn(IV)]; and radionuclides [e.g., U(VI) and Tc(VII)] (8, 21, 41, 44, 50, 55, 56). The majority of studies of anaerobic respiration by S. oneidensis have focused on the mechanism of electron transport to transition metals and radionuclides (11, 14, 34, 46, 58, 59), while the mechanism of electron transport to inorganic sulfur compounds has not been thoroughly examined.Microbial S0 respiration is postulated to occur via two pathways, both of which are based on an intraspecies sulfur cycle. In the first pathway (catalyzed by members of the genus Salmonella [20]), S2O32− is reduced, yielding HS and SO32− (24). SO32− diffuses from the cell and reacts chemically with extracellular S0 to form S2O32−, which reenters the periplasm and is rereduced, thereby sustaining an intraspecies sulfur cycle. In the second pathway (catalyzed by W. succinogenes [24]), water-soluble polysulfides (Sn2; n > 2), formed by chemical interactions of S0 at pHs >7 (52), are reduced stepwise in the periplasm to Sn 12− and HS. Similarly to what occurs with the first pathway, microbially produced HS diffuses from the cell and reacts chemically with S0 to produce additional Sn2, which reenters the periplasm and is rereduced to sustain an analogous intraspecies sulfur cycle (24).Genetic analyses of S2O32− reduction-deficient mutants of Salmonella enterica serovar Typhimurium have demonstrated that phsA (denoting production of hydrogen sulfide) is required for HS production during S2O32− respiration (10, 17, 22). In addition, phsA-deficient mutants are unable to reduce S0 as an electron acceptor (24). The phsA homolog of W. succinogenes (annotated as psrA, for polysulfide reduction) is required for S0 respiration (32, 37). W. succinogenes psrA is the first gene of a three-gene cluster (including psrA, psrB, and psrC) whose products encode a polysulfide reductase, a quinol oxidase, and a membrane anchor, respectively (15). In addition, the structure of the polysulfide reductase complex (PsrABC) from Thermus thermophilus has recently been solved, and results indicate that PsrC acts as a quinol oxidase that transfers electrons stepwise via PsrB and PsrA to Sn2 during anaerobic S0 respiration (27). The main objectives of the present study were to (i) identify the S. Typhimurium phsA homolog in the S. oneidensis genome, (ii) employ a newly constructed suicide cloning vector for in-frame gene deletion mutagenesis in S. oneidensis to delete the S. Typhimurium phsA homolog of S. oneidensis, and (iii) test the S. oneidensis psrA deletion mutant for respiratory activity on a combination of two electron donors and 11 electron acceptors, including the inorganic sulfur compounds S4O62−, S2O32−, and S0.  相似文献   

16.
Growth chamber studies with soybeans (Glycine max [L.] Merr.) were designed to determine the relative limitations of NO3, NADH, and nitrate reductase (NR) per se on nitrate metabolism as affected by light and temperature. Three NR enzyme assays (+NO3in vivo, −NO3in vivo, and in vitro) were compared. NR activity decreased with all assays when plants were exposed to dark. Addition of NO3 to the in vivo NR assay medium increased activity (over that of the −NO3in vivo assay) at all sampling periods of a normal day-night sequence (14 hr-30 C day; 10 hr-20 C night), indicating that NO3 was rate-limiting. The stimulation of in vivo NR activity by NO3 was not seen in plants exposed to extended dark periods at elevated temperatures (16 hr-30 C), indicating that under those conditions, NO3 was not the limiting factor. Under the latter condition, in vitro NR activity was appreciable (19 μmol NO2 [g fresh weight, hr]−1) suggesting that enzyme level per se was not the limiting factor and that reductant energy might be limiting.  相似文献   

17.
18.
A complex system of muddy fluid-discharging and methane (CH4)-releasing seeps was discovered in a valley of the river Mukhrinskaya, one of the small rivers of the Irtysh Basin, West Siberia. CH4 flux from most (90%) of these gas ebullition sites did not exceed 1.45 g CH4 h−1, while some seeps emitted up to 5.54 g CH4 h−1. The δ13C value of methane released from these seeps varied between −71.1 and −71.3‰, suggesting its biogenic origin. Although the seeps were characterized by low in situ temperatures (3.5 to 5°C), relatively high rates of methane oxidation (15.5 to 15.9 nmol CH4 ml−1 day−1) were measured in mud samples. Fluorescence in situ hybridization detected 107 methanotrophic bacteria (MB) per g of mud (dry weight), which accounted for up to 20.5% of total bacterial cell counts. Most (95.8 to 99.3%) methanotroph cells were type I (gammaproteobacterial) MB. The diversity of methanotrophs in this habitat was further assessed by pyrosequencing of pmoA genes, encoding particulate methane monooxygenase. A total of 53,828 pmoA gene sequences of seep-inhabiting methanotrophs were retrieved and analyzed. Nearly all of these sequences affiliated with type I MB, including the Methylobacter-Methylovulum-Methylosoma group, lake cluster 2, and several as-yet-uncharacterized methanotroph clades. Apparently, microbial communities attenuating methane fluxes from these local but strong CH4 sources in floodplains of high-latitude rivers have a large proportion of potentially novel, psychrotolerant methanotrophs, thereby providing a challenge for future isolation studies.  相似文献   

19.
The kinetic parameters Km, Vmax, Tt (turnover time), and v (natural velocity) were determined for H2 and acetate conversion to methane by Wintergreen Lake sediment, using short-term (a few hours) methods and incubation temperatures of 10 to 14°C. Estimates of the Michaelis-Menten constant, Km, for both the consumption of hydrogen and the conversion of hydrogen to methane by sediment microflora averaged about 0.024 μmol g−1 of dry sediment. The maximal velocity, Vmax, averaged 4.8 μmol of H2 g−1 h−1 for hydrogen consumption and 0.64 μmol of CH4 g−1 h−1 for the conversion of hydrogen to methane during the winter. Estimated natural rates of hydrogen consumption and hydrogen conversion to methane could be calculated from the Michaelis-Menten equation and estimates of Km, Vmax, and the in situ dissolved-hydrogen concentration. These results indicate that methane may not be the only fate of hydrogen in the sediment. Among several potential hydrogen donors tested, only formate stimulated the rate of sediment methanogenesis. Formate conversion to methane was so rapid that an accurate estimate of kinetic parameters was not possible. Kinetic experiments using [2-14C]acetate and sediments collected in the summer indicated that acetate was being converted to methane at or near the maximal rate. A minimum natural rate of acetate conversion to methane was estimated to be about 110 nmol of CH4 g−1 h−1, which was 66% of the Vmax (163 nmol of CH4 g−1 h−1). A 15-min preincubation of sediment with 5.0 × 10−3 atm of hydrogen had a pronounced effect on the kinetic parameters for the conversion of acetate to methane. The acetate pool size, expressed as the term Km + Sn (Sn is in situ substrate concentration), decreased by 37% and Tt decreased by 43%. The Vmax remained relatively constant. A preincubation with hydrogen also caused a 37% decrease in the amount of labeled carbon dioxide produced from the metabolism of [U-14C]valine by sediment heterotrophs.  相似文献   

20.
Human ATP:cob(I)alamin adenosyltransferase (ATR) is a mitochondrial enzyme that catalyzes an adenosyl transfer to cob(I)alamin, synthesizing 5′-deoxyadenosylcobalamin (AdoCbl) or coenzyme B12. ATR is also a chaperone that escorts AdoCbl, transferring it to methylmalonyl-CoA mutase, which is important in propionate metabolism. Mutations in ATR lead to methylmalonic aciduria type B, an inborn error of B12 metabolism. Our previous studies have furnished insights into how ATR protein dynamics influence redox-linked cobalt coordination chemistry, controlling its catalytic versus chaperone functions. In this study, we have characterized three patient mutations at two conserved active site residues in human ATR, R190C/H, and E193K and obtained crystal structures of R190C and E193K variants, which display only subtle structural changes. All three mutations were found to weaken affinities for the cob(II)alamin substrate and the AdoCbl product and increase KM(ATP). 31P NMR studies show that binding of the triphosphate product, formed during the adenosylation reaction, is also weakened. However, although the kcat of this reaction is significantly diminished for the R190C/H mutants, it is comparable with the WT enzyme for the E193K variant, revealing the catalytic importance of Arg-190. Furthermore, although the E193K mutation selectively impairs the chaperone function by promoting product release into solution, its catalytic function might be unaffected at physiological ATP concentrations. In contrast, the R190C/H mutations affect both the catalytic and chaperoning activities of ATR. Because the E193K mutation spares the catalytic activity of ATR, our data suggest that the patients carrying this mutation are more likely to be responsive to cobalamin therapy.  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