首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
The induction of hydroxyl radical (OH) production via quinone redox cycling in white-rot fungi was investigated to improve pollutant degradation. In particular, we examined the influence of 4-methoxybenzaldehyde (anisaldehyde), Mn2+, and oxalate on Pleurotus eryngii OH generation. Our standard quinone redox cycling conditions combined mycelium from laccase-producing cultures with 2,6-dimethoxy-1,4-benzoquinone (DBQ) and Fe3+-EDTA. The main reactions involved in OH production under these conditions have been shown to be (i) DBQ reduction to hydroquinone (DBQH2) by cell-bound dehydrogenase activities; (ii) DBQH2 oxidation to semiquinone (DBQ) by laccase; (iii) DBQ autoxidation, catalyzed by Fe3+-EDTA, producing superoxide (O2) and Fe2+-EDTA; (iv) O2 dismutation, generating H2O2; and (v) the Fenton reaction. Compared to standard quinone redox cycling conditions, OH production was increased 1.2- and 3.0-fold by the presence of anisaldehyde and Mn2+, respectively, and 3.1-fold by substituting Fe3+-EDTA with Fe3+-oxalate. A 6.3-fold increase was obtained by combining Mn2+ and Fe3+-oxalate. These increases were due to enhanced production of H2O2 via anisaldehyde redox cycling and O2 reduction by Mn2+. They were also caused by the acceleration of the DBQ redox cycle as a consequence of DBQH2 oxidation by both Fe3+-oxalate and the Mn3+ generated during O2 reduction. Finally, induction of OH production through quinone redox cycling enabled P. eryngii to oxidize phenol and the dye reactive black 5, obtaining a high correlation between the rates of OH production and pollutant oxidation.The degradation of lignin and pollutants by white-rot fungi is an oxidative and rather nonspecific process based on the production of substrate free radicals (36). These radicals are produced by ligninolytic enzymes, including laccase and three kinds of peroxidases: lignin peroxidase, manganese peroxidase, and versatile peroxidase (VP) (23). The H2O2 required for peroxidase activities is provided by several oxidases, such as glyoxal oxidase and aryl-alcohol oxidase (AAO) (9, 18). This free-radical-based degradative mechanism leads to the production of a broad variety of oxidized compounds. Common lignin depolymerization products are aromatic aldehydes and acids, and quinones (34). In addition to their high extracellular oxidation potential, white-rot fungi show strong ability to reduce these lignin depolymerization products, using different intracellular and membrane-bound systems (4, 25, 39). Since reduced electron acceptors of oxidized compounds are donor substrates for the above-mentioned oxidative enzymes, the simultaneous actions of both systems lead to the establishment of redox cycles (35). Although the function of these redox cycles is not fully understood, they have been hypothesized to be related to further metabolism of lignin depolymerization products that require reduction to be converted in substrates of the ligninolytic enzymes (34). A second function attributed to these redox cycles is the production of reactive oxygen species, i.e., superoxide anion radicals (O2), H2O2, and hydroxyl radicals (OH), where lignin depolymerization products and fungal metabolites act as electron carriers between intracellular reducing equivalents and extracellular oxygen. This function has been studied in Pleurotus eryngii, whose ligninolytic system is composed of laccase (26), VP (24), and AAO (9). Incubation of this fungus with different aromatic aldehydes has been shown to provide extracellular H2O2 on a constant basis, due to the establishment of a redox cycle catalyzed by an intracellular aryl-alcohol dehydrogenase (AAD) and the extracellular AAO (7, 10). The process was termed aromatic aldehyde redox cycling, and 4-methoxybenzaldehyde (anisaldehyde) serves as the main Pleurotus metabolite acting as a cycle electron carrier (13). A second cyclic system, involving a cell-bound quinone reductase activity (QR) and laccase, was found to produce O2 and H2O2 during incubation of P. eryngii with different quinones (11). The process was described as the cell-bound divalent reduction of quinones (Q) by QR, followed by extracellular laccase oxidation of hydroquinones (QH2) into semiquinones (Q), which autoxidized to some extent, producing O2 (Q + O2 ⇆ Q + O2). H2O2 was formed by O2 dismutation (O2 + HO2 + H+ → O2 + H2O2). In an accompanying paper, we describe the extension of this O2 and H2O2 generation mechanism to OH radical production by the addition of Fe3+-EDTA to incubation mixtures of several white-rot fungi with different quinones (6). Among them, those derived from 4-hydroxyphenyl, guaiacyl, and syringyl lignin units were used: 1,4-benzoquinone (BQ), 2-methoxy-1,4-benzoquinone (MBQ), and 2,6-dimethoxy-1,4-benzoquinone (DBQ), respectively. Semiquinone autoxidation under these conditions was catalyzed by Fe3+-EDTA instead of being a direct electron transfer to O2. The intermediate Fe2+-EDTA reduced not only O2, but also H2O2, leading to OH radical production by the Fenton reaction (H2O2 + Fe2+ → OH + OH + Fe3+).Although OH radicals are the strongest oxidants produced by white-rot fungi (2, 14), studies of their involvement in pollutant degradation are quite scarce. In this context, the objectives of this study were to (i) determine possible factors enhancing the production of OH radicals by P. eryngii via quinone redox cycling and (ii) test the validity of this inducible OH production mechanism as a strategy for pollutant degradation. Our selection of possible OH production promoters was guided by two observations (6). First, the redox cycle of benzoquinones working with washed P. eryngii mycelium is rate limited by hydroquinone oxidation, since the amounts of the ligninolytic enzymes that remained bound to the fungus under these conditions were not large. Second, H2O2 is the limiting reagent for OH production by the Fenton reaction.With these considerations in mind, anisaldehyde and Mn2+ were selected to increase H2O2 production. As mentioned above, anisaldehyde induces H2O2 production in P. eryngii via aromatic aldehyde redox cycling (7). Mn2+ has been shown to enhance H2O2 production during the oxidation of QH2 by P. eryngii laccase by reducing the O2 produced in the semiquinone autoxidation reaction (Mn2+ + O2 → Mn3+ + H2O2 + 2 H+) (26). Mn2+ was also selected to increase the hydroquinone oxidation rate, since this reaction has been shown to be propagated by the Mn3+ generated in the latter reaction (QH2 + Mn3+ → Q + Mn2+ + 2 H+). The replacement of Fe3+-EDTA by Fe3+-oxalate was also planned in order to increase the QH2 oxidation rate above that resulting from the action of laccase. Oxalate is a common extracellular metabolite of wood-rotting fungi to which the function of chelating iron and manganese has been attributed (16, 45). The use of Fe3+-oxalate and nonchelated Fe3+, both QH2 oxidants, has been proven to enable quinone redox cycling in fungi that do not produce ligninolytic enzymes, such as the brown-rot fungus Gloeophyllum trabeum (17, 40, 41). Finally, phenol and the azo dye reactive black 5 (RB5) were selected as model pollutants.  相似文献   

2.
Brown rot basidiomycetes initiate wood decay by producing extracellular reactive oxygen species that depolymerize the structural polysaccharides of lignocellulose. Secreted fungal hydroquinones are considered one contributor because they have been shown to reduce Fe3+, thus generating perhydroxyl radicals and Fe2+, which subsequently react further to produce biodegradative hydroxyl radicals. However, many brown rot fungi also secrete high levels of oxalate, which chelates Fe3+ tightly, making it unreactive with hydroquinones. For hydroquinone-driven hydroxyl radical production to contribute in this environment, an alternative mechanism to oxidize hydroquinones is required. We show here that aspen wood undergoing decay by the oxalate producer Postia placenta contained both 2,5-dimethoxyhydroquinone and laccase activity. Mass spectrometric analysis of proteins extracted from the wood identified a putative laccase (Joint Genome Institute P. placenta protein identification number 111314), and heterologous expression of the corresponding gene confirmed this assignment. Ultrafiltration experiments with liquid pressed from the biodegrading wood showed that a high-molecular-weight component was required for it to oxidize 2,5-dimethoxyhydroquinone rapidly and that this component was replaceable by P. placenta laccase. The purified laccase oxidized 2,5-dimethoxyhydroquinone with a second-order rate constant near 104 M−1 s−1, and measurements of the H2O2 produced indicated that approximately one perhydroxyl radical was generated per hydroquinone supplied. Using these values and a previously developed computer model, we estimate that the quantity of reactive oxygen species produced by P. placenta laccase in wood is large enough that it likely contributes to incipient decay.Brown rot basidiomycetes are the principal recyclers of woody biomass in coniferous forest ecosystems and also the chief cause of decay in wooden structures (8, 41). Unlike the closely related white rot fungi, they degrade the cellulose and hemicellulose in wood while mineralizing little of the lignin that shields these structural polysaccharides from enzymatic attack. As a result, extensively brown-rotted wood consists primarily of an oxidized, partially cleaved residue derived from the original lignin (7, 16, 21, 22, 40). This failure to remove lignin efficiently suggests that brown rot systems contain fewer components than white rot systems and, in agreement, the recently published genome sequence of Postia placenta shows that this brown rot fungus lacks the ligninolytic peroxidases generally thought important for white rot (26). Instead, brown rot fungi appear to rely, at least during incipient decay, on small agents that can penetrate the lignin to access the polysaccharides (10, 14).A better understanding of the biodegradative agents produced by P. placenta may provide clues about what constitutes a minimally effective system for the microbial deconstruction of lignocellulose. One potential low-molecular-weight contributor is an extracellular metabolite, 2,5-dimethoxyhydroquinone (2,5-DMHQ), which has been found in cultures of P. placenta and other brown rot fungi and also shown to reduce Fe3+ with concomitant H2O2 production, thus producing hydroxyl radicals (·OH) via the Fenton reaction (Fig. (Fig.1,1, reaction 6) (4, 17, 20, 29, 34, 36). Past work has shown that the chemical changes introduced by brown rot fungi into wood, cellulose, and other polymers are consistent with attack by reactive oxygen species (ROS) such as ·OH (4, 7, 19, 21-23). A second small agent with a proposed role is extracellular oxalic acid, which P. placenta produces in sufficient quantity to acidify colonized wood to pH 2 to 4. Assays in vitro have shown that cellulose is slowly hydrolyzed at these acidities (11).Open in a separate windowFIG. 1.Chemical reactions discussed in the text. For simplicity, the HOO·/O2·− acid/base pair is shown only as HOO·. H2Q, hydroquinone; HQ·, semiquinone; Q, quinone.However, there is an apparent contradiction between these two mechanisms: oxalate is a strong chelator of Fe3+, and the resulting Fe3+ trioxalate complex has too negative a reduction potential to react readily with methoxyhydroquinones such as 2,5-DMHQ (28, 37). In considering this problem, we noted the surprising finding that the P. placenta genome encodes two putative laccases, enzymes that are considered atypical of brown rot fungi (26). Laccases oxidize methoxyhydroquinones to semiquinone radicals, which generally have more negative reduction potentials than their parent hydroquinones (38), and are therefore expected to be better reductants of Fe3+. In addition, methoxysemiquinones reduce O2 to generate perhydroxyl radicals (HOO·) and their conjugate base superoxide (O2·−), which dismutate to produce H2O2. Furthermore, HOO·/O2·− can reduce some Fe3+ chelates to generate additional Fe2+ and can oxidize some Fe2+ chelates to generate additional H2O2 (9, 13, 33, 38). By these routes, a P. placenta laccase could bypass the requirement for the hydroquinone to react directly with Fe3+ and could thus generate a complete Fenton system (Fig. (Fig.1,1, reactions 1 to 7).Here we have expressed one of the P. placenta putative laccase genes heterologously and thus demonstrate that it encodes a typical laccase. In addition, we show that laccase activity and this particular enzyme are present in wood undergoing decay by P. placenta. Furthermore, we report that 2,5-DMHQ is present in the biodegrading wood, that it is a substrate for the P. placenta laccase, and that its oxidation during incipient wood decay requires a macromolecular component that is replaceable by P. placenta laccase. Finally, we show that the oxidation of 2,5-DMHQ by the P. placenta laccase results in significant H2O2 production, and we estimate that the quantity of ROS produced by this route is large enough that it probably contributes to incipient brown rot.  相似文献   

3.
Laccase is a copper-containing phenoloxidase, involved in lignin degradation by white rot fungi. The laccase substrate range can be extended to include nonphenolic lignin subunits in the presence of a noncatalytic cooxidant such as 2,2′-azinobis(3-ethylbenzthiazoline-6-sulfonate) (ABTS), with ABTS being oxidized to the stable cation radical, ABTS·+, which accumulates. In this report, we demonstrate that the ABTS·+ can be efficiently reduced back to ABTS by physiologically occurring organic acids such as oxalate, glyoxylate, and malonate. The reduction of the radical by oxalate results in the formation of H2O2, indicating the formation of O2·− as an intermediate. O2·− itself was shown to act as an ABTS·+ reductant. ABTS·+ reduction and H2O2 formation are strongly stimulated by the presence of Mn2+, with accumulation of Mn3+ being observed. Additionally, 4-methyl-O-isoeugenol, an unsaturated lignin monomer model, is capable of directly reducing ABTS·+. These data suggest several mechanisms for the reduction of ABTS·+ which would permit the effective use of ABTS as a laccase cooxidant at catalytic concentrations.Lignin, the second most abundant renewable organic compound in the biosphere after cellulose, is highly recalcitrant, and therefore its biodegradation is a rate-limiting step in the global carbon cycle (9). White rot fungi have evolved a unique mechanism to accomplish this degradation, which utilizes extracellular enzymes to generate oxidative radical species (16). This degradative system is highly nonspecific, and as a consequence, these fungi can also oxidize a broad spectrum of structurally diverse environmental pollutants (4, 18). Three main groups of enzymes, i.e., lignin peroxidases (LiP), manganese peroxidases (MnP), and laccases, along with their low-molecular-weight cofactors, have been implicated in the lignin degradation process. LiP can oxidize the nonphenolic aromatic moieties that make up approximately 85% of the lignin polymer (21), while MnP uses the Mn2+/Mn3+ couple to oxidize phenolic subunits (19). Laccase, a copper-containing phenoloxidase, catalyzes the four-electron reduction of oxygen to water, and this is accompanied by the oxidation of a phenolic substrate (32).In recent years, however, the laccase substrate range has been extended to include nonphenolic lignin subunits in the presence of readily oxidizable primary substrates. These cooxidants have been denoted mediators because they were previously speculated (but not proven) to act as electron transfer mediators. The most extensively investigated laccase mediator is 2,2′-azinobis(3-ethylbenzthiazoline-6-sulfonate) (ABTS), a synthetic nitrogen-substituted aromatic compound which allows the oxidation of nonphenolic lignin model compounds (6) and the delignification of kraft pulp (8) by laccase. More recent work has also focused on an alternative compound, 1-hydroxybenzotriazole (7, 10). In the presence of these compounds, laccase can also catalyze the oxidation of polycyclic aromatic hydrocarbons (PAH) (12, 23), chemical synthesis (29), and textile dye bleaching (31). ABTS is oxidized by laccase to its corresponding cation radical. In the case of ABTS, the radical (ABTS·+) is highly stable, and it has been suggested that it may act as a diffusible oxidant of the enzyme (7). However, although the redox chemistry of ABTS (22) and its radical has been characterized, the mechanisms by which it interacts with laccase to “mediate” lignin oxidation are still unknown. Potthast et al. (28) have found evidence suggesting that ABTS acts as an activator or cooxidant of the enzyme. The observation that the laccase/ABTS couple can oxidize the nonphenolic veratryl alcohol, while ABTS·+ alone cannot (6), provides a further indication of this activator role for ABTS. If compounds such as ABTS do indeed act as cooxidants of the enzyme, it is necessary that some mechanism(s) exists for the recycling of their cation radicals back to their reduced forms so as to be available for subsequent catalytic cycles.A number of low-molecular-weight compounds have been implicated in the catalysis of MnP during the oxidation of lignin. The most important of these is manganese, which is present in virtually all woody tissues (17). Divalent manganese (Mn2+) is oxidized by the enzyme to the trivalent form (Mn3+), which is capable of oxidizing an extensive range of phenolic compounds (19). To catalyze lignin oxidation, Mn3+ is chelated and stabilized by organic acids, which facilitate its diffusion to act as an oxidant at a distance from the MnP active site (19, 33). A range of these acids are produced by ligninolytic fungi (25, 30, 33), but the most ubiquitous is oxalate, whose production at levels as high as 28 mM by cultures of Pleurotus ostreatus has been observed (1). Oxalate can itself be oxidized by Mn3+, producing the formate anion radical (CO2·−), which can then reduce molecular oxygen to produce superoxide (O2·−) (24), and a role for these radicals as reducing agents in lignin degradation has been suggested (24).In this report, evidence is presented indicating that physiologically occurring organic acids can directly reduce ABTS·+. The rate of reduction is highly stimulated by the presence of manganese, and the results indicate a mechanism involving O2·−.  相似文献   

4.
Microbiologically influenced corrosion of steel in anaerobic environments has been attributed to hydrogenotrophic microorganisms. A sludge sample collected from the bottom plate of a crude-oil storage tank was used to inoculate a medium containing iron (Fe0) granules, which was then incubated anaerobically at 37°C under an N2-CO2 atmosphere to enrich for microorganisms capable of using iron as the sole source of electrons. A methanogen, designated strain KA1, was isolated from the enrichment culture. An analysis of its 16S rRNA gene sequence revealed that strain KA1 is a Methanococcus maripaludis strain. Strain KA1 produced methane and oxidized iron much faster than did the type strain of M. maripaludis, strain JJT, which produced methane at a rate expected from the abiotic H2 production rate from iron. Scanning electron micrographs of iron coupons that had been immersed in either a KA1 culture, a JJT culture, or an aseptic medium showed that only coupons from the KA1 culture had corroded substantially, and these were covered with crystalline deposits that consisted mainly of FeCO3.Iron (Fe0) is an inexpensive metal and is widely used in many industrial processes and industrial/commercial products. When iron contacts an aqueous electrolyte, it readily corrodes. This happens because, as a result of metallurgical and environmental heterogeneities, the electrolytes are not evenly distributed across the surface of the metal and consequently the electric potential is also unevenly distributed. Therefore, electrons flow within the metal from an area of higher electrical potential (the anode) to an area of lower electrical potential (the cathode). At the anode, iron atoms lose electrons and dissolve into ferrous ions (Fe2+), whereas cations or elements dissolved in solution (e.g., H+ under anaerobic conditions or O2 under aerobic conditions) are reduced by electrons at the cathode.The corrosion of structures that contain iron is economically devastating. It has been estimated that in the United States alone, the cost of corrosion is 276 billion dollars annually (17). Iron is corroded not only by physiochemical processes but also by the metabolic activity of microorganisms; this metabolic process is termed microbiologically influenced corrosion (MIC). Some 10% of all corrosion damage may be the result of microbial activity (15), and sulfate-reducing bacteria (SRB) are widely regarded as the causative agents of MIC in anaerobic environments (11, 12, 18, 21). The mechanism by which SRB stimulate iron corrosion may occur via the uptake of electrons at the cathodic surface of iron (cathodic depolarization) in conjunction with sulfate reduction (8e + SO42 + 10H+ → H2S + 4H2O) (27), while at the anionic surface, iron atoms are oxidized to ferrous ions (Fe → Fe2+ + 2e). In fact, certain SRB use not only hydrogen but also iron as a source of electrons for sulfate reduction (1, 9, 22). Because not all SRB grow as fast in the presence of iron as they do in the presence of hydrogen (9), fast-growing SRB on iron may have a specific enzyme(s) that removes electrons from iron.Because some methanogens are viable in a hydrogen atmosphere, as are most SRBs, these methanogens may also cause iron corrosion under anaerobic conditions. Several methanogens have been shown to grow and produce methane in medium containing iron as the sole source of electrons (5). The extent of the corrosion by these methanogens, however, was not substantial (2). Others have reported that methanogens do not increase the rate of iron corrosion in comparison with aseptic solutions (6, 7). Recently Dinh and colleagues (9) isolated a methanogen (strain IM1) that produces methane more rapidly than does Methanococcus maripaludis (DSMZ 2771) when cultured with iron granules. Although the rate of iron oxidation was not measured in their experiments, their results suggests that strain IMI oxidizes iron more rapidly than does strain DSMZ 2771.We report herein that a methanogen that was isolated from the sludge of an oil storage tank can unequivocally oxidize iron.  相似文献   

5.
It was recently found that the cytoplasmic butyryl-coenzyme A (butyryl-CoA) dehydrogenase-EtfAB complex from Clostridium kluyveri couples the exergonic reduction of crotonyl-CoA to butyryl-CoA with NADH and the endergonic reduction of ferredoxin with NADH via flavin-based electron bifurcation. We report here on a second cytoplasmic enzyme complex in C. kluyveri capable of energetic coupling via this novel mechanism. It was found that the purified iron-sulfur flavoprotein complex NfnAB couples the exergonic reduction of NADP+ with reduced ferredoxin (Fdred) and the endergonic reduction of NADP+ with NADH in a reversible reaction: Fdred2− + NADH + 2 NADP+ + H+ = Fdox + NAD+ + 2 NADPH. The role of this energy-converting enzyme complex in the ethanol-acetate fermentation of C. kluyveri is discussed.Clostridium kluyveri is unique in fermenting ethanol and acetate to butyrate, caproate, and H2 (reaction 1) and in deriving a large (30%) portion of its cell carbon from CO2. Both the energy metabolism and the pathways of biosynthesis have therefore been the subject of many investigations (for relevant literature, see references 12 and 27). (1)During growth of C. kluyveri on ethanol and acetate, approximately five ethanol and four acetate molecules are converted to three butyrate molecules and one caproate molecule (reaction 1a), and one ethanol molecule is oxidized to one acetate, one H+, and two H2 (reaction 1b) molecules (23, 31). How exergonic reaction 1a is coupled with endergonic reaction 1b and with ATP synthesis from ADP and Pi (ΔGo′ = +32 kJ/mol) has remained unclear for many years. (1a) (1b)We recently showed (12) that, in Clostridium kluyveri, the exergonic reduction of crotonyl-coenzyme A (crotonyl-CoA) (Eo′ = −10 mV) with NADH (Eo′ = −320 mV) involved in reaction 1a is coupled with the endergonic reduction of ferredoxin (Fdox) (Eo′ = −420 mV) with NADH (Eo′ = −320 mV) involved in reaction 1b via the recently proposed mechanism of flavin-based electron bifurcation (7). The coupling reaction is catalyzed by the cytoplasmic butyryl-CoA dehydrogenase-EtfAB complex (reaction 2) (12): (2)The reduced ferredoxin (Fdred2−) is assumed to be used for rereduction of NAD+ via a membrane-associated, proton-translocating ferredoxin:NAD oxidoreductase (RnfABCDEG) (reaction 3), and the proton motive force thus generated is assumed to drive the phosphorylation of ADP via a membrane-associated F1F0 ATP synthetase (reaction 4): (3) (4)The novel coupling mechanism represented by reactions 2 and 3 allowed for the first time the possibility of formulating a metabolic scheme for the ethanol-acetate fermentation that could account for the observed fermentation products and growth yields and thus for the observed ATP gains (27). One issue, however, remained open, namely, why the formation of butyrate from ethanol and acetate in the fermentation involves both an NADP+- and an NAD+-specific β-hydroxybutyryl-CoA dehydrogenase (16), considering that, in the oxidative part of the fermentation (ethanol oxidation to acetyl-CoA), only NADH is generated (8, 9, 13).The presence of a reduced ferredoxin:NADP+ oxidoreductase was proposed based on results of enzymatic studies performed 40 years ago. Cell extracts of Clostridium kluyveri were found to catalyze the formation of H2 from NADPH in a ferredoxin- and NAD+-dependent reaction (34). The results were interpreted to indicate that C. kluyveri contains a ferredoxin-dependent hydrogenase and an NADPH:ferredoxin oxidoreductase with transhydrogenase activity. H2 formation from NADPH was strictly dependent on the presence of NAD+ and was inhibited by NADH, inhibition being competitive with the presence of NAD+, indicating that ferredoxin reduction with NADPH is under the allosteric control of the NAD+/NADH couple. The cell extracts also catalyzed the NADH-dependent reduction of NADP+ with reduced ferredoxin (21, 34). Purification of the enzyme catalyzing these reactions was not achieved, and no function in the energy metabolism of C. kluyveri was assigned.In this communication, we report on the properties of the recombinant enzyme that catalyzes the NAD+-dependent reduction of ferredoxin with NADPH and the NADH-dependent reduction of NADP+ with reduced ferredoxin and show that the cytoplasmic heterodimeric enzyme couples the exergonic reduction of NADP+ with reduced ferredoxin with the endergonic reduction of NADP+ with NADH in a fully reversible reaction. The transhydrogenation reaction is endergonic, because in vivo the NADH/NAD+ ratio is generally near 0.3 and the NADPH/NADP+ ratio is generally above 1 (2, 30). (5)NADP+ reduction is most probably the physiological function of the enzyme, which is why we chose the abbreviation NfnAB (for NADH-dependent reduced ferredoxin:NADP+ oxidoreductase).  相似文献   

6.
The biochemical and molecular mechanisms used by alkaliphilic bacteria to acquire iron are unknown. We demonstrate that alkaliphilic (pH > 9) Bacillus species are sensitive to artificial iron (Fe3+) chelators and produce iron-chelating molecules. These alkaliphilic siderophores contain catechol and hydroxamate moieties, and their synthesis is stimulated by manganese(II) salts and suppressed by FeCl3 addition. Purification and mass spectrometric characterization of the siderophore produced by Caldalkalibacillus thermarum failed to identify any matches to previously observed fragmentation spectra of known siderophores, suggesting a novel structure.Iron is an abundant element in nature; however, in most aqueous aerobic environments iron forms insoluble ferric hydroxide, Fe(OH)3. This poses a major problem for most aerobic bacteria, as ferric hydroxide has a solubility constant of 10−39 M, therefore limiting the concentration of ferric ions to 10−18 M at pH 7.0. For example, bacteria living in seawater (approximate pH 8.0) require iron, yet dissolved iron is only present at 0.02 to 2.0 nM (5). Despite this apparent lack of bioavailability, iron has been repeatedly demonstrated to be an essential element for aerobic bacterial growth (1).With the lack of readily accessible iron at physiological pH, most bacteria have evolved systems to deal with the incumbent problem of iron acquisition. Under iron-rich conditions, Fe2+ uptake receptors, such as FeoAB, are synthesized in bacteria, which passively import iron in the immediate vicinity of the cell (1, 23). No equivalent system has been identified for Fe3+ transport. To acquire Fe3+ under aqueous aerobic conditions, bacteria commonly have import systems involving the synthesis, secretion, and regathering of a group of secondary metabolites known as siderophores (1, 11). Siderophores are low-molecular-weight chemical moieties that chelate Fe3+ and typically have complex formation (Kf) constants in the range of 1023 to 1052 (11). Siderophores, like other chelators, are known to increase the solubility of iron by hindering the formation of Fe-oxyhydroxides at high pH, at which the Fe-oxyhydroxides are the dominating inorganic species (27). Siderophores are also known to facilitate the dissolution of Fe from minerals (3). Siderophore-iron complexes can either be transported through cellular membranes using dedicated transport systems or if the Fe(III) central atom is reduced, making the iron bioavailable for cellular processes (10, 14). Three major groups of siderophores have been described in bacteria: hydroxamates, catecholates, and carboxylates. Hydroxamates and catechols are commonly produced by aerobic bacteria living at neutral to alkaline pH, whereas carboxylates are significantly more common in bacteria living in mildly acidic pH (11-13). In the genus Bacillus, Bacillus megaterium and Bacillus subtilis are producers of schizokinen and bacillibactin, respectively (6, 20). Bacillus anthracis produces both a catechol and a hydroxamate siderophore (7, 34), and B. licheniformis strain VK21 is the only known example of a thermoresistant catecholate-producing Gram-positive bacterium (32).Although there is extensive literature on iron capture mechanisms in bacteria that thrive at neutral pH, there is little information at a biochemical or molecular level on how aerobic bacteria growing at extreme alkaline pHs (i.e., pH 9 to 11) acquire iron. At alkaline pH, the solubility constant for iron decreases far below the requirement for living cells, and the concentration of bioavailable iron is estimated to be approximately 10−23 M at pH 10 (11). Taking this extreme lack of iron into account, the sequestering mechanisms of alkaliphilic bacteria must be powerful, yet there has been little analysis of the types of iron-chelating molecules these bacteria produce.  相似文献   

7.
8.
Values of Δ34S (, where δ34SHS and indicate the differences in the isotopic compositions of the HS and SO42− in the eluent, respectively) for many modern marine sediments are in the range of −55 to −75‰, much greater than the −2 to −46‰ ɛ34S (kinetic isotope enrichment) values commonly observed for microbial sulfate reduction in laboratory batch culture and chemostat experiments. It has been proposed that at extremely low sulfate reduction rates under hypersulfidic conditions with a nonlimited supply of sulfate, isotopic enrichment in laboratory culture experiments should increase to the levels recorded in nature. We examined the effect of extremely low sulfate reduction rates and electron donor limitation on S isotope fractionation by culturing a thermophilic, sulfate-reducing bacterium, Desulfotomaculum putei, in a biomass-recycling culture vessel, or “retentostat.” The cell-specific rate of sulfate reduction and the specific growth rate decreased progressively from the exponential phase to the maintenance phase, yielding average maintenance coefficients of 10−16 to 10−18 mol of SO4 cell−1 h−1 toward the end of the experiments. Overall S mass and isotopic balance were conserved during the experiment. The differences in the δ34S values of the sulfate and sulfide eluting from the retentostat were significantly larger, attaining a maximum Δ34S of −20.9‰, than the −9.7‰ observed during the batch culture experiment, but differences did not attain the values observed in marine sediments.Dissimilatory SO42− reduction is a geologically ancient, anaerobic, energy-yielding metabolic process during which SO42−-reducing bacteria (SRB) reduce SO42− to H2S while oxidizing organic molecules or H2. SO42− reduction is a dominant pathway for organic degradation in marine sediments (23) and in terrestrial subsurface settings where sulfur-bearing minerals dominate over Fe3+-bearing minerals. For example, at depths greater than 1.5 km below land surface in the fractured sedimentary and igneous rocks of the Witwatersrand Basin of South Africa, SO42− reduction is the dominant electron-accepting process (3, 26, 46, 48, 61).The enrichment of 32S in biogenic sulfides, with respect to the parent SO42−, imparted by SRB, is traceable through the geologic record (10, 54). The magnitude of the Δ34S (= δ34Spyrite − δ34Sbarite/gypsum, where δ34Spyrite and δ34Sbarite/gypsum are the isotopic compositions of pyrite and barite or gypsum) increases from −10‰ in the 3.47-billion-year-old North Pole deposits to −30‰ in late-Archaean deposits (55), to −75‰ in Neoproterozoic to modern sulfide-bearing marine sediments (13).The kinetic isotopic enrichment, ɛ34S, deduced from trends in the δ34S values of SO42− and HS in batch culture microbial SO42− reduction experiments using the Rayleigh relationship, ranges from −2‰ to −46‰ (6, 7, 11, 17, 22, 27, 28, 30, 31, 38, 39). The variation in ɛ34S values has been attributed to the SO42− concentration, the type of electron donor and its concentration, the SO42− reduction rate per cell (csSRR) (22), temperature, and species-specific isotope enrichment effects. In these laboratory experiments, doubling times are on the order of hours and csSRRs range from to 0.1 to 18 fmol cell−1 h−1 (7, 12, 17, 22, 30, 32, 39, 40).Experiments performed during the 1960s found that the magnitude of ɛ34S was inversely proportional to the csSRR for organic electron donors (16, 31, 38, 39) when SO42− was not limiting. More-recent batch culture experiments on 3 psychrophilic (optimum growth temperature, <20°C) and mesophilic (optimum growth temperature, between 20°C and 45°C) SRB strains (7) and on 32 psychrophilic to thermophilic SRB strains (22), however, have failed to reproduce such a relationship. In 2001, Canfield (11) reported an inverse correlation between ɛ34S and reduction rate using a flowthrough sediment column and demonstrated that ɛ34S values of approximately −35 to −40‰ were produced when organic substrates added by way of amendment were limited with respect to SO42−. Because it was not possible to readily evaluate changes in biomass in the sediment column with changes in temperature or substrate provision rate, it was inferred that changes in ɛ34S were related to changes in the csSRRs. More recently, Canfield et al. (12) observed a 6‰ variation in ɛ34S values related to the temperature of the batch culture experiments relative to the optimum growth temperature. The few early experiments that were performed using H2 as the electron donor yielded ɛ34S values ranging from −3 to −19‰ (22, 38, 39), which appear to correlate with the csSRR (39). Hoek et al. (32) also found that the ɛ34S values for the thermophilic SO42− reducer Thermodesulfatator indicus increased from between −1.5‰ and −10‰ in batch cultures with high H2 concentrations to between −24‰ and −37‰ in batch cultures grown under H2 limitation with respect to SO42−. Detmers et al. (22) found that the average ɛ34S of SRB that oxidize their organic carbon electron donor completely to CO2 averaged −25‰, versus −9.5‰ for SRB that release acetate during their oxidation of their organic carbon electron donor. Detmers et al. (22) speculated that the greater free energy yield per mole of SO42− from incomplete carbon oxidation relative to that for complete carbon oxidation promotes complete SO42− reduction and hinders isotopic enrichment due to isotopic exchange of the intracellular sulfur species pools.None of these experiments, however, have yielded ɛ34S factors capable of producing the Δ34S values of −55 to −75‰ observed in the geological record from ∼1.0 billion years ago to today. Various schemes have been hypothesized, and observations that involve either the disproportionation of S2O32− (36), the disproportionation of S0 produced by oxidation of either H2S or S2O32− (15), or the disproportionation of SO32− (29) have been made. Attribution of the increasing Δ34S values recorded for Achaean to Neoproterozoic sediments to the increasing role of H2S oxidative pathways makes sense in the context of increasing O2 concentrations in the atmosphere (14) but is not consistent with the lack of significant fractionation observed during oxidative reactions (29). To explain the Δ34S values of −55 to −77‰ reported to occur in interstitial pore waters from 100- to 300-m-deep, hypersulfidic ocean sediments (51, 64, 67), where the presence of a S-oxidative cycle is unlikely, an alternative, elaborate model of the SO42−-reducing pathway has been proposed by Brunner and Bernasconi (9). This model attributes the large Δ34S values to a multistep, reversible reduction of SO32− to HS involving S3O62− and S2O32− (20, 25, 41, 42, 52, 66). The conditions under which the maximum ɛ34S values might be expressed are a combination of elevated HS concentrations, electron donor limitations, nonlimiting SO42− concentrations, and a very low csSRR. The csSRR for subsurface environments has been estimated from biogeochemical-flux modeling to be 10−6 to 10−7 fmol cell−1 h−1 (23), with a corresponding cell turnover rate greater than 1,000 years (37).Batch and chemostat culture systems, despite low growth rates, cannot completely attain a state of zero growth with constant substrate provision and therefore do not accurately reflect the in situ nutritional states of microbes in many natural settings. Retentostats, or recycling fermentor vessels, recycle 100% of biomass to the culturing vessel, allowing experimenters to culture microbial cells to a large biomass with a constant nutrient supply rate until the substrate supply rate itself becomes the growth-limiting factor and cells enter a resting state in which their specific growth rate approaches zero and they carry on maintenance metabolism (1, 47, 53, 58, 59, 62, 63). Utilizing this approach, Colwell et al. (18) were able to obtain a cell-specific respiration rate of 7 × 10−4 fmol of CH4 cell−1 h−1 for a mesophilic marine methanogen, a rate that is comparable to that estimated for methanogenic communities in deep marine sediments off the coast of Peru (49).In this study, the conditions that Brunner and Bernasconi (9) hypothesized would lead to the large Δ34S values seen in nature were recreated in the laboratory by limiting the electron donor supply rate with respect to the SO42− supply rate in a retentostat vessel. The S isotopic enrichment by a resting culture of Desulfotomaculum putei at an extremely low csSRR was compared to that of a batch culture experiment to determine whether the ɛ34S values produced under the former conditions approach the Δ34S seen in nature.  相似文献   

9.
16S rRNA gene libraries from the lithoautotrophic Fe(II)-oxidizing, nitrate-reducing enrichment culture described by Straub et al. (K. L. Straub, M. Benz, B. Schink, and F. Widdel, Appl. Environ. Microbiol. 62:1458-1460, 1996) were dominated by a phylotype related (95% 16S rRNA gene homology) to the autotrophic Fe(II) oxidizer Sideroxydans lithotrophicus. The libraries also contained phylotypes related to known heterotrophic nitrate reducers Comamonas badia, Parvibaculum lavamentivorans, and Rhodanobacter thiooxidans. The three heterotrophs were isolated and found to be capable of only partial (12 to 24%) Fe(II) oxidation, suggesting that the Sideroxydans species has primary responsibility for Fe(II) oxidation in the enrichment culture.A variety of microorganisms oxidize Fe(II) with nitrate under anaerobic, circumneutral pH conditions (29) and may contribute to an active microbially driven anoxic Fe redox cycle (1, 27-29, 31, 32). Straub et al. (28) obtained the first Fe(II)-oxidizing, nitrate-reducing (enrichment) culture capable of fully autotrophic growth by a reaction such as 5Fe2+ + NO3 + 12H2O → 5Fe(OH)3 + 0.5N2 + 9H+. This process has since been demonstrated in detail with the hyperthermophilic archaeon Ferroglobus placidus (9) and with the mesophilic Proteobacteria Chromobacterium violacens strain 2002 (34) and Paracoccus ferrooxidans strain BDN-1 (16). Nitrate-dependent Fe(II) oxidation in the presence of fixed carbon has been documented for Dechlorosoma suillum strain PS (4), Geobacter metallireducens (7), Desulfitobacterium frappieri (23), and Acidovorax strain BoFeN1 (15). In addition to oxidizing insoluble Fe(II)-bearing minerals (33), the enrichment culture described by Straub et al. (28) is the only autotrophic Fe(II)-oxidizing, nitrate-reducing culture capable of near-complete oxidation of uncomplexed Fe(II) with reduction of nitrate to N2. During Fe(II) oxidation, F. placidus reduces nitrate to nitrite, which may play a significant role in overall Fe(II) oxidation. Although both C. violacens and Paracoccus ferrooxidans reduce nitrate to N2, C. violacens oxidizes only 20 to 30% of the initial Fe(II), and P. ferrooxidans uses FeEDTA2− but not free (uncomplexed) Fe(II) in medium analogous to that used for cultivation of the enrichment culture described by Straub et al. (28). The enrichment culture described by Straub et al. (28) is thus the most robust culture capable of autotrophic growth coupled to nitrate-dependent Fe(II) oxidation available at present. The composition and activity of this culture was investigated with molecular and cultivation techniques. The culture examined is one provided by K. L. Straub to E. E. Roden in 1998 for use in studies of nitrate-dependent oxidation of solid-phase Fe(II) compounds (33) and has been maintained in our laboratory since that time.  相似文献   

10.
11.
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   

12.
A database search of the Paramecium genome reveals 34 genes related to Ca2+-release channels of the inositol-1,4,5-trisphosphate (IP3) or ryanodine receptor type (IP3R, RyR). Phylogenetic analyses show that these Ca2+ release channels (CRCs) can be subdivided into six groups (Paramecium tetraurelia CRC-I to CRC-VI), each one with features in part reminiscent of IP3Rs and RyRs. We characterize here the P. tetraurelia CRC-IV-1 gene family, whose relationship to IP3Rs and RyRs is restricted to their C-terminal channel domain. CRC-IV-1 channels localize to cortical Ca2+ stores (alveolar sacs) and also to the endoplasmic reticulum. This is in contrast to a recently described true IP3 channel, a group II member (P. tetraurelia IP3RN-1), found associated with the contractile vacuole system. Silencing of either one of these CRCs results in reduced exocytosis of dense core vesicles (trichocysts), although for different reasons. Knockdown of P. tetraurelia IP3RN affects trichocyst biogenesis, while CRC-IV-1 channels are involved in signal transduction since silenced cells show an impaired release of Ca2+ from cortical stores in response to exocytotic stimuli. Our discovery of a range of CRCs in Paramecium indicates that protozoans already have evolved multiple ways for the use of Ca2+ as signaling molecule.Ca2+ is an important component of cell activity in all organisms, from protozoa to mammals. Thereby Ca2+ may originate from the outside medium and/or from internal stores (7, 18). Ca2+ release from internal stores is mediated by various Ca2+ release channels (CRCs), of which the inositol-1,4,5-trisphosphate receptor (IP3R) and ryanodine receptor (RyR) families have been studied most extensively (8, 9, 29, 63). IP3Rs and RyRs have been identified in various metazoan organisms (reviewed in references 9, 28, and 104). According to these reviews, there exist three genetically distinct isoforms of each receptor type in mammals and orthologues have been identified in various nonmammalian vertebrates, e.g., frogs, chickens, and fish. RyRs and IP3Rs were also cloned and sequenced in the invertebrates Drosophila melanogaster and Caenorhabditis elegans, which possess one copy of each receptor type.Functional evidence for Ca2+ release in response to ryanodine or IP3 receptor agonists has been described in several unicellular systems. Treatment of permeabilized Plasmodium chabaudi parasites with IP3 results in Ca2+ release, which is inhibited by the IP3 receptor antagonist heparin (69). Another apicomplexan parasite, Toxoplasma gondii, responds to agonists and antagonists of both, ryanodine and IP3 receptors, by mediating increases in intracellular Ca2+ concentration ([Ca2+]i) (56). Stimulation of Trypanosoma cruzi with carbachol results in increased [Ca2+]i and IP3 (59). IP3 and cyclic ADP-ribose induces Ca2+ release in Euglena gracilis microsome fractions in a dose-dependent manner (61). In the giant algae Chara corallina and Nitrella translucens, IP3 produces action potentials involving increased [Ca2+]i (93). Treatment of vacuolar membrane vesicles from Candida albicans with IP3 results in Ca2+ release, blocked by heparin and ruthenium red (14). IP3 generates and maintains a Ca2+ gradient in the hyphal tip of Neurospora crassa and the IP3-sensitive channels have been reconstituted and characterized with the planar bilayer method (87). In summary, these publications suggest that IP3-dependent signaling pathways are conserved among unicellular organisms, including protozoa.Despite these data, the molecular characterization of IP3 or ryanodine receptors in low eukaryotes is currently a challenge since the identification of orthologues has not been possible thus far, probably because of evolutionary sequence divergence (66). Traynor et al. (96) identified an IP3 receptor-like protein, IplA, in Dictyostelium discoideum, which possesses regions related to IP3R sequences, but thus far no evidence for IP3 interaction exists. We have recently described an IP3R in the ciliated protozoa Paramecium tetraurelia (referred to here as P. tetraurelia IP3RN) (53), with features characteristic of mammalian IP3Rs in terms of topology and ability for IP3 binding. The expression level of P. tetraurelia IP3RN is modulated by extracellular Ca2+ concentrations ([Ca2+]o) and immunofluorescence studies reveal an unexpected localization to the contractile vacuole complex (CVC), the major organelle involved in osmoregulation (2). The ionic composition of the contractile vacuole fluid by ion-selective microelectrodes (91) suggests that the organelle plays a major role in expelling an excess of cytosolic Ca2+. Therefore, these IP3Rs may here mediate a latent, graded reflux of Ca2+ for fine-tuning of [Ca2+]i and thus serve [Ca2+] homeostasis (53).Besides [Ca2+] homeostasis, the Paramecium cell has to regulate a variety of well-characterized processes (75). This includes exocytosis of dense-core secretory vesicles (trichocysts) (71, 74, 99). Each cell possesses up to 1,000 trichocysts attached to the cell membrane. Their contents can be extruded synchronously in response to natural stimuli, i.e., predators (34, as confirmed by Knoll et al. [49]), to artificial polyamine secretagogues such as aminoethyldextran (AED) (78), to caffeine (48) or to the ryanodine substitute, 4-chloro-meta-cresol (4-CmC) (46). Their expulsion strictly depends on Ca2+ (10) and is accompanied by an increase of intracellular [Ca2+]i (24, 47). This Ca2+ signal originates from rapid mobilization of cortical stores, the alveolar sacs (33, 64, 74), superimposed by Ca2+ influx (46, 72). It thus represents a SOC-type mechanism (SOC, store-operated Ca2+ entry) known from mammalian systems (81).Upon exocytosis stimulation ∼60% of their total Ca2+ is released from alveolar sacs (33). These are Ca2+ stores (90) represented by flat membrane compartments tightly attached at the cell membrane surrounding each trichocyst docking site. They possess a SERCA-type pump located at the membrane facing the cell center (36, 37) and a luminal high-capacity/low-affinity CaBP of the calsequestrin type (73). Thus far, Ca2+ release channels of these stores were identified only indirectly as cells respond by exocytosis to the RyR activators caffeine (54, 48) and 4-CmC (46). However, an involvement of conserved RyRs has remained questionable as ryanodine is not able to activate Ca2+ release from alveolar sacs, as is the case with IP3 (54). Therefore, one of the most intriguing questions is the elucidation of the molecular nature of the channels mediating Ca2+ release from alveolar sacs upon stimulated exocytosis.In the present work we describe a novel family of CRCs (P. tetraurelia CRC-IV-1), whose members display several properties of the channels postulated above. In detail, the identified CRC-IV-1 channels localize to the alveolar sacs. Functional and fluorochrome analyses after gene silencing reveal that they are essential for mediating Ca2+ release and exocytosis in response to AED, caffeine, or 4-CmC. Their classification as “novel” CRC type is based on a restricted relationship to the C-terminal channel domains of IP3Rs and RyRs. The overall size and the number of putative transmembrane domains resemble IP3Rs, but N-terminal parts of CRC-IV-1 channels do not show any conservation, such as an IP3-binding domain. Therefore, CRC-IV-1 channels represent distant relatives of IP3Rs and RyRs and may belong to an ancestral Ca2+ signaling pathway.  相似文献   

13.
In aerobic, circumneutral environments, the essential element Fe occurs primarily in scarcely soluble mineral forms. We examined the independent and combined effects of a siderophore, a reductant (ascorbate), and a low-molecular-weight carboxylic acid (oxalate) on acquisition of Fe from the mineral hematite (α-Fe2O3) by the obligate aerobe Pseudomonas mendocina ymp. A site-directed ΔpmhA mutant that was not capable of producing functional siderophores (i.e., siderophore phenotype) did not grow on hematite as the only Fe source. The concentration of an added exogenous siderophore (1 μM desferrioxamine B [DFO-B]) needed to restore wild-type (WT)-like growth kinetics to the siderophore strain was ∼50-fold less than the concentration of the siderophore secreted by the WT organism grown under the same conditions. The roles of a reductant (ascorbate) and a simple carboxylic acid (oxalate) in the Fe acquisition process were examined in the presence and absence of the siderophore. Addition of ascorbate (50 μM) alone restored the growth of the siderophore culture to the WT levels. A higher concentration of oxalate (100 μM) had little effect on the growth of a siderophore culture; however, addition of 0.1 μM DFO-B and 100 μM oxalate restored the growth of the mutant to WT levels when the oxalate was prereacted with the hematite, demonstrating that a metabolizing culture benefits from a synergistic effect of DFO-B and oxalate.Iron (Fe) is essential for almost all life. However, in aerobic, circumneutral environments, Fe is bound primarily in scarcely soluble minerals and amorphous solids [e.g., the solubility product (KSP) for amorphous Fe(OH)3 is 10−38] (53) and is therefore poorly bioavailable. Aerobic microorganisms directly transform mineral-bound Fe(III) into soluble, highly bioavailable forms (1), overcoming significant kinetic and thermodynamic barriers to mineral dissolution and serving as primary transporters of Fe from the geosphere into global biogeochemical cycles.A primary means by which aerobic microorganisms enhance Fe mobility and bioavailability is by secreting siderophores, which are structurally diverse, low-molecular-weight chelating agents with extremely high affinities for Fe(III) (12, 27, 37, 40). Fe(III)-siderophore stability constants can be as high as 1052 (1, 40), which is many orders of magnitude higher than the stability constants for low-molecular-weight organic acids, such as oxalic acid [for Fe(III) + 3 oxalate ⇆ Fe(oxalate)3, K = 1018.6] (45). While their high affinity for Fe(III) is clearly important for helping siderophores mobilize Fe from Fe(III) (hydr)oxides in the aqueous phase, the mechanisms of Fe mobilization appear to be complex and are the subject of much recent study (14, 17, 18, 26, 28, 49). In particular, the role of siderophores in ligand-promoted dissolution mechanisms has undergone careful evaluation in vitro. The model is described simply here as follows for amorphous Fe(OH)3 and has been described in detail by Kraemer (26): Fe(OH)3 + 3H+ ⇆ Fe(III) + 3H2O (KSP) (equation 1); Fe(III) + H3L ⇆ FeL + 3H+ (KFeL) (equation 2); and Fe(OH)3 + H3L ⇆ FeL + 3H2O (Keq = KSPKFeL = [FeL]/[H3L]) (equation 3). The concentration of the solubilized FeL complex, according to equation 3, is determined as follows: [FeL] = [H3L]KSP × KFeL. The estimated concentration of siderophores in carbonic soil (∼10−8 to 10−7 M), combined with their strong affinities for Fe(III) (39), suggests that [FeL] could in principle easily be micromolar or higher and could support vigorous bacterial growth. However, the trishydroxamate siderophores that have been studied most to date adsorb only weakly to Fe(III) (hydr)oxide minerals, likely due to steric constraints, although charge repulsion may also play a role for positively charged siderophores, such as desferrioxamine B (DFO-B) (6, 26, 41, 42). Therefore, it has been proposed that siderophores act primarily in conjunction with other molecules, such as simple plant-derived carboxylic acids or reductants, which interact more strongly with mineral surfaces and release Fe directly through ligand-promoted and/or reductive mechanisms (52). This proposed “synergistic effect,” in which the combined effect of various elements is greater than the sum of the individual effects, suggests that an interaction of biogenic molecules may overcome kinetic and thermodynamic barriers to the release of Fe from minerals in the presence of siderophores. The role of the siderophore in such a synergistic system is not a direct role in surface processes; rather, the siderophore maintains a low concentration of aqueous Fe in equilibrium with the mineral (an Fe sink), thus driving the reaction toward more dissolution (26, 41). Only a low concentration of a siderophore relative to the concentrations of surface-reacting organic species is required to promote efficient dissolution (26).The synergistic effect has been observed directly in in vitro, abiotic experiments using combinations of microbe-derived siderophores and simple organic acids. A combination of environmentally relevant concentrations of oxalate (1 to 80 μM) and DFO-B (40 μM), for example, doubled the rate of Fe(III) hydroxide mineral (goethite) dissolution compared with the rate when only oxalate or DFO-B was present in a recent in vitro study (6). Actively metabolizing aerobic bacteria, which can move Fe from solution into cells and recycle or release new siderophores back into the medium, might be expected to promote the synergistic siderophore-carboxylic acid interaction even further in a batch system. Likewise, it has been suggested that organic reductants may work synergistically with siderophores. In particular, a recent study showed that exogenously added reductants significantly enhance the bioavailability of Fe to an aerobic siderophore-producing bacterium, Pseudomonas mendocina ymp (15), isolated from the Nevada Test Site and used in the work described here.As an obligate aerobe, P. mendocina ymp does not have dissimilatory reduction pathways, so that its use of iron (hydr)oxide minerals is only for acquisition of nutritional Fe and not for cellular respiration. In contrast to dissimilatory Fe-reducing bacteria, which require millimolar concentrations of Fe (2, 29-31, 36, 43, 50), P. mendocina ymp requires micromolar concentrations (19, 20, 24, 32-34). Previously, this strain''s ability to dissolve and use various mineral forms of Fe was quantified in a series of microbial growth studies (23, 24, 32-34). P. mendocina ymp is known to produce hydroxamate-containing siderophores that increase the rate of dissolution of the Fe oxide mineral hematite. A recent study demonstrated that reductants significantly enhanced the bioavailability of Fe-(hydr)oxide minerals to P. mendocina (15). The ymp strain was also shown to have endogenous Fe(III)-reducing activity, which Hersman et al. suggested could be involved in solubilizing ferric minerals (24). Likewise, closely related strains of Pseudomonas stutzeri have been shown to produce pyridine-2,6-bis(thiocarboxylic acid) (PDTC), which they can use in the reduction, transport, and detoxification of metals and metalloids (11, 16, 51). However, control experiments showed that P. mendocina did not secrete molecules that exhibited a significant amount of reducing activity under the conditions used in this study (see the supplemental material). Notably, this species does not appear to contain a set of PDTC biosynthesis genes.In this work we used the wild-type (WT) strain P. mendocina ymp along with a mutant with a site-directed markerless mutation that was not capable of producing siderophores (ΔpmhA mutant with the siderophore phenotype) (3) in a series of experiments examining siderophore use and potential synergistic effects with either a simple carboxylic acid (oxalate) or an exogenous reductant (ascorbate). Both ascorbate and oxalate are plant products that are frequently found in the shallow subsurface; their effects on in vitro Fe (hydr)oxide dissolution have been well described (6).  相似文献   

14.
Microorganisms play a significant role in the speciation and mobility of arsenic in the environment. In this study, the oxidation of arsenite [As(III)] to arsenate [As(V)] linked to chlorate (ClO3) reduction was shown to be catalyzed by sludge samples, enrichment cultures (ECs), and pure cultures incubated under anaerobic conditions. No activity was observed in treatments lacking inoculum or with heat-killed sludge, or in controls lacking ClO3. The As(III) oxidation was linked to the complete reduction of ClO3 to Cl, and the molar ratio of As(V) formed to ClO3 consumed approached the theoretical value of 3:1 assuming the e equivalents from As(III) were used to completely reduce ClO3. In keeping with O2 as a putative intermediate of ClO3 reduction, the ECs could also oxidize As(III) to As(V) with O2 at low concentrations. Low levels of organic carbon were essential in heterotrophic ECs but not in autotrophic ECs. 16S rRNA gene clone libraries indicated that the ECs were dominated by clones of Rhodocyclaceae (including Dechloromonas, Azospira, and Azonexus phylotypes) and Stenotrophomonas under autotrophic conditions. Additional phylotypes (Alicycliphilus, Agrobacterium, and Pseudoxanthomonas) were identified in heterotrophic ECs. Two isolated autotrophic pure cultures, Dechloromonas sp. strain ECC1-pb1 and Azospira sp. strain ECC1-pb2, were able to grow by linking the oxidation of As(III) to As(V) with the reduction of ClO3. The presence of the arsenite oxidase subunit A (aroA) gene was demonstrated with PCR in the ECs and pure cultures. This study demonstrates that ClO3 is an alternative electron acceptor to support the microbial oxidation of As(III).The contamination of drinking water with arsenic (As) is a global public health issue. Arsenic is a human carcinogenic compound (2), which poses a risk to millions of people around the world (31). The most common oxidation states of As in aqueous environments are arsenite [As(III), H3AsO3] or arsenate [As(V), H2AsO4, and HAsO42−]. Microbial processes play critical roles in controlling the fate and transformation of As in subsurface systems (22). As(V) binds to aluminum oxides more extensively than As(III) under circumneutral pH conditions (12, 16). Both As(III) and As(V) are strongly adsorbed on iron oxides (9). However, As(III) is more rapidly desorbed compared to As(V) (35).Aerobic bacteria can oxidize As(III) forming As(V) (14, 28), which potentially is less mobile in the subsurface environment. Also, in environments with dissolved ferrous iron [Fe(II)] the oxidation of Fe(II) (both abiotic and biotic) would result in formation of Fe(III) (hydr)oxides such as ferrihydrite which adsorb As. Oxidation processes, therefore, can decrease the mobilization of As in groundwater. However, oxygen (O2) is poorly soluble in groundwater and may become consumed by microbial activity, creating anaerobic zones. Alternative oxidants aside from O2 also have the potential to support the microbial oxidation of As(III). Recently, several studies have demonstrated that nitrate-dependent As(III) oxidation is carried out by anaerobic microorganisms to gain energy from As(III) oxidation. As(III)-oxidizing denitrifying bacteria have been isolated from various environments including As-contaminated lakes and soil (21, 25), as well as enrichment cultures (ECs), and isolates from pristine sediments and sludge samples (33, 34). 16S rRNA gene clone library characterization of the ECs indicates that the predominant phylotypes were from the genus Azoarcus and the family Comamonadaceae (34).Beside nitrate, chlorate (ClO3) can also be considered as a possible alternative oxidant for microorganisms to promote the bioremediation of contaminated plumes (6, 17). (Per)chlorate is commonly used as a terminal electron acceptor by anaerobic bacteria; as a result, it is completely degraded to the benign end product, chloride (Cl). Microbial reduction of perchlorate proceeds via a three-step process of ClO4 → ClO3→ ClO2 → O2 + Cl (6). Reduction of perchlorate to chlorate, and chlorate to chlorite is catalyzed by respiratory (per)chlorate reductases (3). Subsequent disproportionation of chlorite into Cl and O2 is catalyzed by chlorite dismutase, which is the fastest step, and the O2 produced is immediately consumed for energy of cell synthesis (6). Although organic compounds are the most well studied electron donors for (per)chlorate reduction, Fe(II) oxidation has also been shown to be linked to microbial ClO3 reduction (36).The main objective of the present study is to explore the potential use of ClO3 as an electron acceptor for the microbial oxidation of As(III) by anaerobic bacteria. The theoretical stoichiometry of the reaction is presented below: (1) Based on bioenergetic considerations, the reaction is feasible as indicated by the highly exergonic standard change in Gibbs free energy [ΔG0′ = −92.4 kJ mol−1 As(III)] calculated from E0′ values of 0.618 and 0.139 V for ClO3/Cl (6) and As(V)/As(III) (18), respectively.  相似文献   

15.
Industrial production of lactic acid with the current pyruvate decarboxylase-negative Saccharomyces cerevisiae strains requires aeration to allow for respiratory generation of ATP to facilitate growth and, even under nongrowing conditions, cellular maintenance. In the current study, we observed an inhibition of aerobic growth in the presence of lactic acid. Unexpectedly, the cyb2Δ reference strain, used to avoid aerobic consumption of lactic acid, had a specific growth rate of 0.25 h−1 in anaerobic batch cultures containing lactic acid but only 0.16 h−1 in identical aerobic cultures. Measurements of aerobic cultures of S. cerevisiae showed that the addition of lactic acid to the growth medium resulted in elevated levels of reactive oxygen species (ROS). To reduce the accumulation of lactic acid-induced ROS, cytosolic catalase (CTT1) was overexpressed by replacing the native promoter with the strong constitutive TPI1 promoter. Increased activity of catalase was confirmed and later correlated with decreased levels of ROS and increased specific growth rates in the presence of high lactic acid concentrations. The increased fitness of this genetically modified strain demonstrates the successful attenuation of additional stress that is derived from aerobic metabolism and may provide the basis for enhanced (micro)aerobic production of organic acids in S. cerevisiae.Lactic acid is an organic acid with a wide range of applications. In the food industry, lactic acid has traditionally been used as an antimicrobial as well as a flavor enhancer. Besides having applications in textile, cosmetic, and pharmaceutical industries (5), lactic acid has been applied for the manufacture of lactic acid polymers (11, 40). These polymers have properties that are similar to those of petroleum-derived plastics. Skyrocketing oil prices caused by dwindling fossil fuel reserves coupled with pressures to tackle environmental issues are creating increased demand for bioderived, and often biodegradable, polymers, such as poly-lactic acid.Current industrial lactic acid fermentations are based on different species of lactic acid bacteria. These bacteria have complex nutrient requirements due to their limited ability to synthesize B vitamins and amino acids (8) and are intolerant to acidic conditions with a pH between 5.5 and 6.5 required for growth (40). Acidification of the growth medium during lactic acid fermentation is typically counteracted by the addition of neutralizing agents (e.g., CaCO3), resulting in the formation of large quantities of insoluble salts, such as gypsum, during downstream processing.Saccharomyces cerevisiae has received attention as a possible alternative biocatalyst. This organism is relatively tolerant to low pH and has simple nutrient requirements. The production of lactic acid with metabolically engineered S. cerevisiae was achieved by introducing a NAD+-dependent lactate dehydrogenase, leading to the simultaneous formation of both ethanol and lactate (1a, 12, 31, 32, 36). Further improvements were made by constructing a pyruvate decarboxylase-negative (Pdc) S. cerevisiae strain (1a, 31, 44) that converted glucose to lactic acid as the sole fermentation product.Although the redox balance and ATP generation in lactic acid fermentation are analogous to those in alcoholic fermentation, engineered homolactic S. cerevisiae strains could not sustain anaerobic growth (44). In addition, the lactate formation rate under anaerobic conditions in the presence of excess glucose was significantly lower than the specific ethanol production rate of the wild-type strain. Moreover, exposure of the anaerobic cell suspension to oxygen immediately led to a 2.5-fold increase in the lactate formation rate. The stimulatory effect of oxygen on lactic acid fermentation may reflect an energetic constraint in lactate fermentation, probably as a consequence of energy-dependent product export (42, 44). In agreement with this hypothesis, intracellular ATP concentrations and the related energy charge decrease rapidly during anaerobic homolactic fermentation by S. cerevisiae (1). Consequently, industrial production of lactic acid with S. cerevisiae may require (micro)aerobic conditions to allow for the generation of sufficient ATP to enable cell growth and, even under nongrowing conditions, maintenance.The formation of reactive oxygen species (ROS) during cellular respiration is an unavoidable side effect of aerobic life relying on oxygen as the final electron acceptor. At least 2% of oxygen consumed in mitochondrial respiration undergoes only one electron reduction, mainly by the semiquinone form of coenzyme Q, generating superoxide radicals (O2) (26). In addition, the prooxidant effects of organic acids have been demonstrated using sod mutants (30). An in vitro study by Ali et al. (3) also linked ROS formation to weak organic acids and showed enhanced hydroxy radical (OH) generation in the presence of lactic acid.Among different ROS, the hydroxy radical that originates from H2O2 in the metal-mediated Fenton/Haber-Weiss reactions is especially reactive. It indiscriminately oxidizes intracellular proteins, nucleic acids, and lipids in the cell membranes (4, 38). Lactate interacts with the ferric ion (Fe3+) to form a stable complex of Fe3+-lactate at a molar ratio of 1:2. This complex then reacts with H2O2 to enhance the OH generation via the Fenton reaction (2, 3). Although a similar in vivo mechanism has not yet been proven, previous research indicates that lactic acid and other weak organic acids enhance oxidative stress of aerobic yeast cultures.Like other eukaryotic organisms, S. cerevisiae possesses enzymatic defense mechanisms, including several crucial antioxidant enzymes, such as catalase and superoxide dismutase (SOD). SOD removes O2 by converting it to H2O2, which, in turn, can be disproportionated to water by catalase or glutathione peroxidase. Cytosolic catalase, Ctt1p, is thought to play a general role, as CTT1 expression is regulated by various stresses, including oxidative stress, osmotic stress, and starvation (15, 23, 33). More recently, catalase has also been implicated in response to acetic acid tolerance and acetic acid-induced programmed cell death (17, 47).The goals of the present study were to assess the in vivo relevance of lactate-mediated oxidative stress in S. cerevisiae and to investigate whether its effects could be ameliorated by enhanced expression of catalase.  相似文献   

16.
Despite many efforts to develop AIDS vaccines eliciting virus-specific T-cell responses, whether induction of these memory T cells by vaccination before human immunodeficiency virus (HIV) exposure can actually contribute to effective T-cell responses postinfection remains unclear. In particular, induction of HIV-specific memory CD4+ T cells may increase the target cell pool for HIV infection because the virus preferentially infects HIV-specific CD4+ T cells. However, virus-specific CD4+ helper T-cell responses are thought to be important for functional CD8+ cytotoxic-T-lymphocyte (CTL) induction in HIV infection, and it has remained unknown whether HIV-specific memory CD8+ T cells induced by vaccination without HIV-specific CD4+ T-cell help can exert effective responses after virus exposure. Here we show the impact of CD8+ T-cell memory induction without virus-specific CD4+ T-cell help on the control of a simian immunodeficiency virus (SIV) challenge in rhesus macaques. We developed a prophylactic vaccine by using a Sendai virus (SeV) vector expressing a single SIV Gag241-249 CTL epitope fused with enhanced green fluorescent protein (EGFP). Vaccination resulted in induction of SeV-EGFP-specific CD4+ T-cell and Gag241-249-specific CD8+ T-cell responses. After a SIV challenge, the vaccinees showed dominant Gag241-249-specific CD8+ T-cell responses with higher effector memory frequencies in the acute phase and exhibited significantly reduced viral loads. These results demonstrate that virus-specific memory CD8+ T cells induced by vaccination without virus-specific CD4+ T-cell help could indeed facilitate SIV control after virus exposure, indicating the benefit of prophylactic vaccination eliciting virus-specific CTL memory with non-virus-specific CD4+ T-cell responses for HIV control.Virus-specific T-cell responses are crucial for controlling human immunodeficiency virus (HIV) and simian immunodeficiency virus (SIV) replication (3, 4, 12, 20, 28, 36, 37). Therefore, a great deal of effort has been exerted to develop AIDS vaccines eliciting virus-specific T-cell responses (23, 27, 30, 47), but whether this approach actually results in HIV control remains unclear (1, 6). It is important to determine which T-cell responses need to be induced by prophylactic vaccination for HIV control after virus exposure.Because HIV preferentially infects HIV-specific CD4+ T cells (5), induction of HIV-specific memory CD4+ T cells by vaccination may increase the target cell pool for HIV infection and could enhance viral replication (42). However, CD4+ helper T-cell responses are important for functional CD8+ cytotoxic-T-lymphocyte (CTL) induction (11, 40, 43, 46), and it has remained unknown whether HIV-specific memory CD8+ T cells induced by vaccination with non-virus-specific CD4+ T-cell help (but without HIV-specific CD4+ T-cell help) can exert effective responses after virus exposure. Indeed, the real impact of prophylactic induction of CTL memory itself on HIV replication has not been well documented thus far.We previously developed a prophylactic AIDS vaccine consisting of DNA priming followed by boosting with a recombinant Sendai virus (SeV) vector expressing SIVmac239 Gag (26). Evaluation of this vaccine''s efficacy against a SIVmac239 challenge in Burmese rhesus macaques showed that some vaccinees contained SIV replication whereas unvaccinated animals developed AIDS (15, 27). In particular, vaccination consistently resulted in control of SIV replication in those animals possessing the major histocompatibility complex class I (MHC-I) haplotype 90-120-Ia. Gag206-216 (IINEEAADWDL) and Gag241-249 (SSVDEQIQW) epitope-specific CD8+ T-cell responses were shown to be involved in SIV control in these vaccinated macaques (14, 16).In the present study, focusing on CD8+ T-cell responses directed against one of these epitopes, we have evaluated the efficacy of a vaccine expressing the Gag241-249 epitope fused with enhanced green fluorescent protein (EGFP) against a SIVmac239 challenge in 90-120-Ia-positive rhesus macaques. The animals exhibited this single-epitope-specific CD8+ T-cell response and SeV-EGFP-specific CD4+ T-cell responses after vaccination and showed rapid, dominant induction of potent secondary Gag241-249-specific CD8+ T-cell responses after a SIV challenge. Plasma viral loads in these vaccinees were significantly reduced compared to those of naive controls. These results indicate that induction of CD8+ T-cell memory without virus-specific CD4+ T-cell help by prophylactic vaccination can result in effective CD8+ T-cell responses after virus exposure.  相似文献   

17.
18.
19.
Dissimilatory NO3 reduction in sediments is often measured in bulk incubations that destroy in situ gradients of controlling factors such as sulfide and oxygen. Additionally, the use of unnaturally high NO3 concentrations yields potential rather than actual activities of dissimilatory NO3 reduction. We developed a technique to determine the vertical distribution of the net rates of dissimilatory nitrate reduction to ammonium (DNRA) with minimal physical disturbance in intact sediment cores at millimeter-level resolution. This allows DNRA activity to be directly linked to the microenvironmental conditions in the layer of NO3 consumption. The water column of the sediment core is amended with 15NO3 at the in situ 14NO3 concentration. A gel probe is deployed in the sediment and is retrieved after complete diffusive equilibration between the gel and the sediment pore water. The gel is then sliced and the NH4+ dissolved in the gel slices is chemically converted by hypobromite to N2 in reaction vials. The isotopic composition of N2 is determined by mass spectrometry. We used the combined gel probe and isotopic labeling technique with freshwater and marine sediment cores and with sterile quartz sand with artificial gradients of 15NH4+. The results were compared to the NH4+ microsensor profiles measured in freshwater sediment and quartz sand and to the N2O microsensor profiles measured in acetylene-amended sediments to trace denitrification.Nitrate accounts for the eutrophication of many human-affected aquatic ecosystems (19, 21). Sediment bacteria may mitigate NO3 pollution by denitrification and anaerobic ammonium oxidation (anammox), which produce N2 (13, 18). However, inorganic nitrogen is retained in aquatic ecosystems when sediment bacteria reduce NO3 to NH4+ by dissimilatory nitrate reduction to ammonium (DNRA) (5, 12, 16, 39). Hence, DNRA contributes to rather than counteracts eutrophication (23). DNRA may be the dominant pathway of dissimilatory NO3 reduction in sediments that are rich in electron donors, such as labile organic carbon and sulfide (4, 8, 17, 38, 55). High rates of DNRA are thus found in sediments affected by coastal aquaculture (8, 36) and settling algal blooms (16).DNRA, denitrification, and the chemical factors that control the partitioning between them (e.g., sulfide) should ideally be investigated in undisturbed sediments. The redox stratification of sediments involves vertical concentration gradients of pore water solutes. These gradients are often very steep, and their measurement requires high-resolution techniques, such as microsensors (26, 42) and gel probes (9, 54). If, for instance, the influence of sulfide on DNRA and denitrification is to be investigated, one wants to know exactly the sulfide concentration in the layers of DNRA and denitrification activity, as well as the flux of sulfide into these layers. This information can easily be obtained using H2S and pH microsensors (22, 43). It is less trivial to determine the vertical distribution of DNRA and denitrification activity in undisturbed sediments. Denitrification activity can be traced using a combination of the acetylene inhibition technique (51) and N2O microsensors (1). Acetylene inhibits the last step of denitrification, and therefore, N2O accumulates in the layer of denitrification activity (44). This method underestimates the denitrification activity in sediments with high rates of coupled nitrification-denitrification because acetylene also inhibits nitrification (50).The vertical distribution of DNRA activity in undisturbed sediment has, to the best of our knowledge, never been determined; thus, the microenvironmental conditions in the layer of DNRA activity remain unknown. Until now, the influence of chemical factors on DNRA and denitrification in sediments has been assessed by slurry incubations (4, 12, 30), by flux measurements with sealed sediment cores (7, 47) or flowthrough sediment cores (16, 27, 37), and in one case, in reconstituted sediment cores sliced at centimeter-level resolution (39). Here, we present a new method, the combined gel probe and isotope labeling technique, to determine the vertical distribution of the net rates of DNRA in sediments. The sediments remain largely undisturbed and the NO3 amendments are within the range of in situ concentrations. The DNRA measurements can be related to the microprofiles of potential influencing factors measured in close vicinity of the gel probe. This allows DNRA activity to be directly linked with the microenvironmental conditions in the sediment.  相似文献   

20.
Iron oxidation at neutral pH by the phototrophic anaerobic iron-oxidizing bacterium Rhodobacter sp. strain SW2 leads to the formation of iron-rich minerals. These minerals consist mainly of nano-goethite (α-FeOOH), which precipitates exclusively outside cells, mostly on polymer fibers emerging from the cells. Scanning transmission X-ray microscopy analyses performed at the C K-edge suggest that these fibers are composed of a mixture of lipids and polysaccharides or of lipopolysaccharides. The iron and the organic carbon contents of these fibers are linearly correlated at the 25-nm scale, which in addition to their texture suggests that these fibers act as a template for mineral precipitation, followed by limited crystal growth. Moreover, we evidence a gradient of the iron oxidation state along the mineralized fibers at the submicrometer scale. Fe minerals on these fibers contain a higher proportion of Fe(III) at cell contact, and the proportion of Fe(II) increases at a distance from the cells. All together, these results demonstrate the primordial role of organic polymers in iron biomineralization and provide first evidence for the existence of a redox gradient around these nonencrusting, Fe-oxidizing bacteria.Fe(II) can serve as a source of electrons for phylogenetically diverse microorganisms that precipitate iron minerals as products of their metabolism (see, e.g., references 3, 5, 25, and 30). For example, mixotrophic or autotrophic bacteria can couple the oxidation of Fe(II) to the reduction of nitrate in anoxic and neutral-pH environments. With Fe(III) being highly insoluble at neutral pH, this metabolism leads to the formation of poorly to well-crystallized iron minerals (3, 18, 26, 27) that precipitate partly within the cell periplasm for some strains (22). Similar Fe minerals are also synthesized by autotrophic bacteria that perform anoxygenic photosynthesis, using Fe(II) as an electron donor and light as a source of energy for CO2 fixation (8, 12, 30), according to the equation HCO3 + 4 Fe2+ + 10 H2O ⇆ <CH2O> + 4 Fe(OH)3 + 7 H+.However, the biological mechanisms of iron oxidation in these bacteria and in particular the way they cope with the formation of minerals within their ultrastructures are still not fully understood. Indeed, iron minerals are potentially lethal since their precipitation may alter cellular ultrastructures but also catalyze the production of free radicals (2). Recent genetic studies of the phototrophic, iron-oxidizing bacteria Rhodobacter sp. strain SW2 (6) and Rhodopseudomonas palustris strain TIE-1 (16) have identified genes (fox and pio operons, respectively) encoding proteins specific for iron oxidation. Interestingly, Jiao and Newman (16) suggested that one of these proteins could have a periplasmic localization. However, in contrast to what has been observed in some other phototrophic iron oxidizers (25) and in some nitrate-reducing, iron-oxidizing bacteria (22), no iron-mineral precipitation occurs within the periplasm of the purple nonsulfur iron-oxidizing bacterium Rhodobacter sp. strain SW2 (3). Similarly to some other anaerobic neutrophilic (22, 25) and microaerobic iron-oxidizing bacteria (5, 10), this strain seems indeed to have the ability to localize iron biomineralization at a distance from the cells, leaving large areas of the cells free of precipitates (17, 25). While it has been shown that the Gallionella and Leptothrix genera, for example, produce extracellular polymers that facilitate the nucleation of iron minerals outside cells (see, e.g., references 5 and 9), only a little is known about the existence and function of such polymers in anaerobic, neutrophilic iron-oxidizing bacteria and particularly in the phototrophic strain SW2. In the present study, we investigate iron biomineralization by the photoautotrophic iron-oxidizing bacterium Rhodobacter sp. strain SW2. We use scanning transmission X-ray microscopy (STXM) to map and identify organic polymers produced by the cells as well as the redox state of iron at the 25-nanometer scale regularly during a 2 week-period. These results demonstrate the primordial role of organic polymers in iron biomineralization and provide the first evidence for the existence of a redox gradient around SW2 cells.  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