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 15 毫秒
1.
Paenibacillus A11-2 can efficiently cleave two carbon&bond;sulfur bonds in dibenzothiophene (DBT) and alkyl DBTs, which are refractory by conventional petroleum hydrodesulfurization, to remove sulfur atom at high temperatures. An 8.7-kb DNA fragment containing the genes for the DBT desulfurizing enzymes of A11-2 was cloned in Escherichia coli and characterized. Heterologous expression analysis of the deletion mutants identified three open reading frames that were required for the desulfurization of DBT to 2-hydroxybiphenyl (2-HBP). The three genes were designated tdsA, tdsB, and tdsC (for thermophilic desulfurization). Both the nucleotide sequences and the deduced amino acid sequences show significant homology to dszABC genes of Rhodococcus sp. IGTS8, but there are several local differences between them. Subclone analysis revealed that the product of tdsC oxidizes DBT to DBT-5,5'-dioxide via DBT-5-oxide, the product of tdsA converts DBT-5,5'-dioxide to 2-(2-hydroxyphenyl) benzene sulfinate, and the product of tdsB converts 2-(2-hydroxyphenyl)benzene sulfinate to 2-HBP. Cell-free extracts of a recombinant E. coli harboring all the three desulfurization genes converted DBT to 2-HBP at both 37 and 50 degrees C. In vivo and in vitro exhibition of desulfurization activity of the recombinant genes derived from a Paenibacillus indicates that an E. coli oxidoreductase can be functionally coupled with the monooxygenases of a gram-positive thermophile.  相似文献   

2.
Desulphurising enzymes remove the sulphur moiety from an organosulphur molecule leaving the carbon skeleton intact. Two kinds of desulphurisation reaction are recognised. The dibenzothiophene (DBT)-specific pathway desulphurises DBT to inorganic sulphite and 2- hydroxybiphenyl (HBP), and the benzothiophene (BTH)-specific pathway desulphurises BTH to 2-(2-hydroxyphenyl)ethan 1-al (HPEal) and probably inorganic sulphite. The DBT-desulphurisation pathway was originally identified in Rhodococcus erythropolis strain IGTS8 (ATCC 53968), and the BTH-desulphurisation pathway in Gordonia sp. strain 213E (NCIMB 40816). These organisms do not further metabolise the organic product of desulphurisation.In this article current knowledge of the biochemistry and genetics of the desulphurisation enzymes is reviewed. The need for separate, DBT- and BTH-specific desulphurisation routes is rationalised in terms of the chemical differences between the two compounds. The desulphurisation pathway is compared with other microbial DBT- degrading enzyme systems. Finally some comments are made concerning the application of desulphurisation enzymes for fuel desulphurisation and on the relevance of these enzymes to the ecology of the mycolata (sensu Chun et al, 1996).  相似文献   

3.
The biphenyl-mineralizing bacterium Burkholderia sp. strain LB400 also utilized 3-chloro-, 4-chloro-, 2,3-dichloro- and 2,4′-dichlorobiphenyl for growth. By the attack of the initial enzyme a chlorine was eliminated dioxygenolytically from position 2 of one of the aromatic rings when hydrogens of both were substituted by chlorine. The strain mineralized 3-chloro- and 2,3′-dichlorobiphenyl via the central intermediate 3-chlorobenzoate through its chlorocatechol pathway enzymes, but excreted stoichiometric amounts of 4-chlorobenzoate from 4-chloro- and 2,4-dichlorobiphenyl. These two compounds were mineralized by a co-culture of strain LB400 and a derivative of the (methyl-) benzoate-degrading strain Pseudomonas putida mt-2 (TOL). The complete degradation was achieved upon transfer of a cluster of at least five genes, encoding the regulated chlorocatechol pathway operon, from strain LB400 to strain mt-2. This transfer was demonstrated by the polymerase chain reaction. Received: 15 April 1998 / Received revision: 12 June 1998 / Accepted: 19 June 1998  相似文献   

4.
Phytosiderophores, mugineic acids, have been demonstrated to be involved in Fe acquisition in gramineous plants. In this study, chromosomal arm locations of genes encoding for biosynthesis of various phytosiderophores were identified in a cultivar of barley (Hordeum vulgare L. cv. Betzes). Using wheat (Triticum aestivum L. cv. Chinese Spring)-barley (cv. Betzes) ditelosomic addition lines for 4HS and 4HL, a gene for hydroxylation of 2′-deoxymugineic acid to mugineic acid was localized to the long arm of barley chromosome 4H. To locate the gene for hydroxylation of mugineic acid to 3-epihydroxymugineic acid, hybrids between the 4H addition line and other wheat-barley addition lines were studied. Only a hybrid between 4H and 7H addition lines produced 3-epihydroxymugineic acid. The gene was further localized to the long arm of chromosome 7H by feeding mugineic acid to ditelosomic addition lines for 7HS and 7HL. A new phytosiderophore was discovered in both 7H and 7HL addition lines, which was identified to be 3-epihydroxy-2′-deoxymugineic acid by detailed nuclear magnetic resonance studies. These results revealed that in barley there are two pathways from 2′-deoxymugineic acid to 3-epihydroxymugineic acid: 2′-deoxymugineic acid → mugineic acid → 3-epihydroxymugineic acid and 2′-deoxymugineic acid → 3-epihydroxy-2′-deoxymugineic acid → 3-epihydroxymugineic acid. Barley genes encoding for the hydroxylations of phytosiderophores are located in different chromosomes and each gene hydroxylates different C-positions: the long arm of chromosome 4H carries the gene for hydroxylating the C-2′ position and the long arm of chromosome 7H carries the gene for hydroxylating the C-3 position of the azetidine ring. Received: 10 August 1998 / Accepted: 30 September 1998  相似文献   

5.
Rhodococcus erythropolis has been widely studied for desulfurization. However, activity levels required for commercial application have not been achieved. A major limitation of the current work in biodesulfurization is inadequate information regarding sulfur metabolism generally, and in particular the metabolism of the sulfur obtained from dibenzothiophene (DBT) metabolism via the 4S pathway. In this work, we have investigated the possible routes taken by the sulfur from DBT to convert into biomass or other metabolites. We propose two alternate hypotheses. In the first, we hypothesize that the cell can convert via sulfite reductase (SR) the sulfite from the metabolism of DBT into sulfide that can be assimilated into biomass. However, in the process, it may convert any excess sulfite into extracellular sulfate via sulfite oxidoreductase (SOR) to avoid the toxic effects of sulfite. In the second, we speculate that the cell cannot assimilate the sulfite directly into biomass via SR. It must first use SOR to produce extracellular sulfate, and then recapture that sulfate into biomass via SR. Thus, either way, we propose that SOR and SR activities, in addition to dsz genes and cofactors, may be critical in increasing desulfurization levels significantly. In particular, we suggest that the simultaneous increase in SOR activity and decrease in SR activity can enable increased desulfurization activity.  相似文献   

6.
Two Rhodococcus erythropolis isolates, named A66 and A69, together with the well-characterized R. erythropolis strain IGTS8 were compared biochemically and genetically. Both isolates, like strain IGTS8, desulfurized DBT to 2-hydroxybiphenyl (2-HBP), following the 4S pathway of desulfurization. Strain IGTS8 showed the highest (81.5%) desulfurization activity in a medium containing DBT at 30 °C. Strain A66 showed approximately the same desulfurization activity either when incubated at 30 °C or at 37 °C, while strain A69 showed an increase of desulfurization efficiency (up to 79%) when incubated at 37 °C. Strains A66 and A69 were also able to grow using various organosulfur or organonitrogen-compounds as the sole sulfur or nitrogen sources. The biological responses of A66, A69 and IGTS8 strains to a series of mutagens and environmental agents were evaluated, trying to mimic actual circumstances involved in exposure/handling of microorganisms during petroleum biorefining. The results showed that strains A69 and IGTS8 were much more resistant to UVC treatment than A66. The three desulfurization genes (dszA, dszB and dszC) present in strains A66 and A69 were partially characterized. They seem to be located on a plasmid, not only in the strain IGTS8, but also in A66 and A69. PCR amplification was observed using specific primers for dsz genes in all the strains tested; however, no amplification product was observed using primers for carbazole (car) or quinoline (qor) metabolisms. All this information contributes to broaden our knowledge concerning both the desulfurization of DBT and the degradation of organonitrogen compounds within the R. erythropolis species.  相似文献   

7.
Desulfurization of a model fuel system consisting of hexadecane and dibenzothiophene (DBT) by Rhodococcus rhodochrous IGTS8 was demonstrated in a 2-L continuous stirred tank reactor (CSTR). The reactor was operated in a semicontinuous and continuous mode with and without recycling of the model fuel. A constant volumetric desulfurization activity A(t), (in mg HBP L(-1) h(-1)) was maintained in the reactor with a feeding strategy of fresh cell suspension based on a first-order decay of the biocatalyst. Maximum desulfurization rates, as measured by specific desulfurization activity, of 1.9 mg HBP/g DCW h were attained. Rates of biocatalyst decay were on the order of 0.072 h(-1). Theoretical predictions of a respiratory quotient (RQ) associated with this biotransformation reaction agree well with experimental data from off-gas analysis. In addition, the ratio of the specific desulfurization activity a(t), (in mg HBP/g DCW h) of recycled and fresh biocatalyst was determined and evaluated.  相似文献   

8.
Dibenzothiophene (DBT) monooxygenase (DszC) catalysis, the first and also the key step in the microbial DBT desulfurization, is the conversion of DBT to DBT sulfone (DBTO2). In this study, dszC of a DBT-desulfurizing bacterium Rhodococcus sp. DS-3 was cloned by PCR. The sequence cloned was 99% homologous to Rhodococcus erythropolis IGTS8 that was reported in the Genebank. The gene dszC could be overexpressed effectively after being inserted into plasmid pET28a and transformed into E. coli BL21 strain. The expression amount of DszC was about 20% of total supernatant at low temperature. The soluble DszC in the supernatant was purified by Ni2+ chelating His-Tag resin column and sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) to electronics purity. Only one band was detected by Western-blotting, which is for the antibody released in mouse against purified DszC in the expression product of BL21 (DE3, paC5) and Rhodococcus sp. DS-3. The activity of purified DszC was 0.36 U. DszC can utilize the organic compound such as DBT and methyl-DBT, but not DBT derivates such as DBF, which has no sulfur or inorganic sulfur. __________ Translated from Acta Scientiarum Naturalium Universitatis Nankaiensis, 2005, 38(6): 1–6 [译自: 南开大学学报 (自然科学版), 2005, 38(6): 1–6]  相似文献   

9.
Microbial desulfurization has been extensively studied as a promising alternative to the widely applied chemical desulfurization process. Sulfur removal from petroleum and its products becomes essential, as the environmental regulations become increasingly stringent. Rhodococcus qingshengii IGTS8 has gained ground as a naturally occurring model biocatalyst, due to its superior specific activity for desulfurization of dibenzothiophene (DBT). Recalcitrant organic sulfur compounds—DBT included—are preferentially removed by selective carbon-sulfur bond cleavage to avoid a reduction in the calorific value of the fuel. The process, however, still has not reached economically sustainable levels, as certain limitations have been identified. One of those bottlenecks is the repression of catalytic activity caused by ubiquitous sulfur sources such as inorganic sulfate, methionine, or cysteine. Herein, we report an optimized culture medium for wild-type stain IGTS8 that completely alleviates the sulfate-mediated repression of biodesulfurization activity without modification of the natural biocatalyst. Medium C not only promotes growth in the presence of several sulfur sources, including DBT, but also enhances biodesulfurization of resting cells grown in the presence of up to 5 mM sulfate. Based on the above, the present work can be considered as a step towards the development of a more viable commercial biodesulfurization process.  相似文献   

10.
Microbial desulfurization, or biodesulfurization (BDS), of fuels is a promising technology because it can desulfurize compounds that are recalcitrant to the current standard technology in the oil industry. One of the obstacles to the commercialization of BDS is the reduction in biocatalyst activity concomitant with the accumulation of the end product, 2-hydroxybiphenyl (HBP), during the process. BDS experiments were performed by incubating Rhodococcus erythropolis IGTS8 resting-cell suspensions with hexadecane at 0.50 (vol/vol) containing 10 mM dibenzothiophene. The resin Dowex Optipore SD-2 was added to the BDS experiments at resin concentrations of 0, 10, or 50 g resin/liter total volume. The HBP concentration within the cytoplasm was estimated to decrease from 1,100 to 260 μM with increasing resin concentration. Despite this finding, productivity did not increase with the resin concentration. This led us to focus on the susceptibility of the desulfurization enzymes toward HBP. Dose-response experiments were performed to identify major inhibitory interactions in the most common BDS pathway, the 4S pathway. HBP was responsible for three of the four major inhibitory interactions identified. The concentrations of HBP that led to a 50% reduction in the enzymes'' activities (IC50s) for DszA, DszB, and DszC were measured to be 60 ± 5 μM, 110 ± 10 μM, and 50 ± 5 μM, respectively. The fact that the IC50s for HBP are all significantly lower than the cytoplasmic HBP concentration suggests that the inhibition of the desulfurization enzymes by HBP is responsible for the observed reduction in biocatalyst activity concomitant with HBP generation.  相似文献   

11.
The gram-positive bacterium Rhodococcus sp. strain IGTS8 is able to remove sulfur from certain aromatic compounds without breaking carbon-carbon bonds. In particular, sulfur is removed from dibenzothiophene (DBT) to give the final product, 2-hydroxybiphenyl. A genomic library of IGTS8 was constructed in the cosmid vector pLAFR5, but no desulfurization phenotype was imparted to Escherichia coli. Therefore, IGTS8 was mutagenized, and a new strain (UV1) was selected that had lost the ability to desulfurize DBT. The genomic library was transferred into UV1, and several colonies that had regained the desulfurization phenotype were isolated, though free plasmid could not be isolated. Instead, vector DNA had integrated into either the chromosome or a large resident plasmid. DNA on either side of the inserted vector sequences was cloned and used to probe the original genomic library in E. coli. This procedure identified individual cosmid clones that, when electroporated into strain UV1, restored desulfurization. When the origin of replication from a Rhodococcus plasmid was inserted, the efficiency with which these clones transformed UV1 increased 20- to 50-fold and they could be retrieved as free plasmids. Restriction mapping and subcloning indicated that the desulfurization genes reside on a 4.0-kb DNA fragment. Finally, the phenotype was transferred to Rhodococcus fascians D188-5, a species normally incapable of desulfurizing DBT. The mutant strain, UV1, and R. fascians produced 2-hydroxybiphenyl from DBT when they contained appropriate clones, indicating that the genes for the entire pathway have been isolated.  相似文献   

12.
Strain SY1, identified as a Corynebacterium sp., was isolated on the basis of the ability to utilize dibenzothiophene (DBT) as a sole source of sulfur. Strain SY1 could utilize a wide range of organic and inorganic sulfur compounds, such as DBT sulfone, dimethyl sulfide, dimethyl sulfoxide, dimethyl sulfone, CS2, FeS2, and even elemental sulfur. Strain SY1 metabolized DBT to dibenzothiophene-5-oxide, DBT sulfone, and 2-hydroxybiphenyl, which was subsequently nitrated to produce at least two different hydroxynitrobiphenyls during cultivation. These metabolites were separated by silica gel column chromatography and identified by nuclear magnetic resonance, UV, and mass spectral techniques. Resting cells of SY1 desulfurized toluenesulfonic acid and released sulfite anion. On the basis of these results, a new DBT degradation pathway is proposed.  相似文献   

13.
T Omori  L Monna  Y Saiki    T Kodama 《Applied microbiology》1992,58(3):911-915
Strain SY1, identified as a Corynebacterium sp., was isolated on the basis of the ability to utilize dibenzothiophene (DBT) as a sole source of sulfur. Strain SY1 could utilize a wide range of organic and inorganic sulfur compounds, such as DBT sulfone, dimethyl sulfide, dimethyl sulfoxide, dimethyl sulfone, CS2, FeS2, and even elemental sulfur. Strain SY1 metabolized DBT to dibenzothiophene-5-oxide, DBT sulfone, and 2-hydroxybiphenyl, which was subsequently nitrated to produce at least two different hydroxynitrobiphenyls during cultivation. These metabolites were separated by silica gel column chromatography and identified by nuclear magnetic resonance, UV, and mass spectral techniques. Resting cells of SY1 desulfurized toluenesulfonic acid and released sulfite anion. On the basis of these results, a new DBT degradation pathway is proposed.  相似文献   

14.
Sinorhizobium sp. KT55 was the first Gram-negative isolate to be capable of utilizing benzothiophene as the sole source of sulfur. By GC-MS analysis of metabolites of benzothiophene by this strain, benzothiophene sulfone, benzo[e][1,2]oxathiin S-oxide and o-hydroxystyrene were detected, suggesting that the benzothiophene desulfurization pathway of this strain is benzothiophene → benzothiophene sulfoxide → benzothiophene sulfone → benzo[e][1,2]oxathiin S-oxide →o-hydroxystyrene. Desulfurization activity of this strain was significantly repressed by methionine, cysteine, sulfate, dimethyl sulfoxide, and Casamino acids. Received: 5 January 2001/Accepted: 6 February 2001  相似文献   

15.
Accumulation of elemental sulfur during pyrite oxidation lowers the efficiency of coal desulfurization and bioleaching. In the case of pyrite bioleaching by Leptospirillum ferrooxidans, an iron(II)-ion-oxidizing organism without sulfur-oxidizing capacity, from the pyritic sulfur moiety about 10% elemental sulfur, 2% pentathionate, and 1% tetrathionate accumulated by a recently described cyclic pyrite oxidation mechanism. In the case of pure cultures of Thiobacillus ferrooxidans and mixed cultures of L. ferrooxidans and T. thiooxidans, pyrite was nearly completely oxidized to sulfate because of the capacity of these cultures to oxidize both iron(II) ions and sulfur compounds. Pyrite oxidation in acidic solutions, mediated chemically by iron(III) ion, resulted in an accumulation of similar amounts of sulfur compounds as obtained with L. ferrooxidans. Changes of pH to values below 2 or in the iron ion concentration are not decisive for diverting the flux of sulfur compounds. The literature on pyrite bioleaching is in agreement with the findings indicating that the chemistry of direct and indirect pyrite leaching is identical. Received: 20 April 1998 / Received revision: 27 August 1998 / Accepted: 3 September 1998  相似文献   

16.
Dibenzothiophene (DBT), a model compound for sulfur-containing organic molecules found in fossil fuels, can be desulfurized to 2-hydroxybiphenyl (2-HBP) by Rhodococcus sp. strain IGTS8. Complementation of a desulfurization (dsz) mutant provided the genes from Rhodococcus sp. strain IGTS8 responsible for desulfurization. A 6.7-kb TaqI fragment cloned in Escherichia coli-Rhodococcus shuttle vector pRR-6 was found to both complement this mutation and confer desulfurization to Rhodococcus fascians, which normally is not able to desulfurize DBT. Expression of this fragment in E. coli also conferred the ability to desulfurize DBT. A molecular analysis of the cloned fragment revealed a single operon containing three open reading frames involved in the conversion of DBT to 2-HBP. The three genes were designated dszA, dszB, and dszC. Neither the nucleotide sequences nor the deduced amino acid sequences of the enzymes exhibited significant similarity to sequences obtained from the GenBank, EMBL, and Swiss-Prot databases, indicating that these enzymes are novel enzymes. Subclone analyses revealed that the gene product of dszC converts DBT directly to DBT-sulfone and that the gene products of dszA and dszB act in concert to convert DBT-sulfone to 2-HBP.  相似文献   

17.
Rhodococcus erythropolis IGTS8 that possesses dibenzothiophene sulfone monooxygenase mutated at residue 345 (Q345A), can degrade octyl sulfide on which the wild strain cannot grow. Residue 345 and the neighbouring residues were changed by site-directed mutagenesis. Only DszA changed at residue 345 gave an altered C-S bond cleavage pattern of 3-methyl DBT sulfone. This residue is therefore involved in C-S bond cleavage specifically for alkylated DBT sulfone.  相似文献   

18.
An organism, identified as Mycobacterium phlei GTIS10, was isolated based on its ability to use dibenzothiophene (DBT) as a sole source of sulfur for growth at 30-52 degrees C. Similar to other biodesulfurization-competent organisms, M. phlei GTIS10 converts DBT to 2-hydroxybiphenyl (2-HBP), as detected by HPLC. The specific desulfurization activity of the 50 degrees C M. phlei GTIS10 culture was determined to be 1.1+/-0.07 micromol 2-HBP min(-1) (g dry cell)(-1). M. phlei GTIS10 can also utilize benzothiophene and thiophene as sulfur sources for growth. The dszABC operon of M. phlei GTIS10 was cloned and sequenced and was found to be identical to that of Rhodococcus erythropolis IGTS8. The presence of the R. erythropolis IGTS8 120-kb plasmid pSOX, which encodes the dszABC operon, has been demonstrated in M. phlei GTIS10. Even though identical dsz genes are contained in both cultures, the temperature at which resting cells of R. erythropolisIGTS8 reach the highest rate of DBT metabolism is near 30 degrees C whereas the temperature that shows the highest activity in resting cell cultures of M. phlei GTIS10 is near 50 degrees C, and activity is detectable at temperatures as high as 57 degrees C. In M. phlei GTIS10, the rate-limiting step in vivo appears to be the conversion of DBT to dibenzothiophene sulfone catalyzed by the product of the dszC gene, DBT monooxygenase. The thermostability of individual desulfurization enzymes was determined and 2-hydroxybiphenyl-2-sulfinate sulfinolyase, encoded by dszB, was found to be the most thermolabile. These results demonstrate that the thermostability of individual enzymes determined in vitro is not necessarily a good predictor of the functional temperature range of enzymes in vivo.  相似文献   

19.
The erythromycin producer, Saccharopolyspora erythraea ER720, was genetically engineered to produce 6,12-dideoxyerythromycin A, a novel erythromycin derivative, as the major macrolide in the fermentation broth. Inspection of the biosynthetic pathway for erythromycin would suggest that production of this compound could be achieved simply through the disruption of two genes, that encoding the erythromycin C-6 hydroxylase (eryF ) and that encoding the erythromycin C-12 hydroxylase (eryK ). The double mutant, however, was found to produce a mixture of 6,12-dideoxyerythromycin A and the precursor, 6-deoxyerythromycin D. Complete conversion to the desired product (to the limit of detection by TLC) was achieved by inserting an additional copy of the eryG gene, encoding the erythromycin 3′′-O-methyltransferase and driven by the ermE* promoter, into the S. erythraea chromosome. Received: 6 October 1997 / Received revision: 27 January 1998 / Accepted: 24 February 1998  相似文献   

20.
 Metal-mediated hydrolysis of phosphate esters is a common catalytic pathway in nucleic acid biochemistry. Two distinct models are principally invoked in mechanistic discussions of these reactions for magnesium-dependent nuclease activation: namely, the one-versus two-metal-ion pathways. The 3′-5′ exonuclease domain of the Klenow fragment of Escherichia coli DNA polymerase I is a paradigm for the two-metal-ion mechanism; however, this reaction model is principally based on structural and kinetics experiments employing high concentrations of transition metal analogues and high concentrations of background ammonium sulfate during doping experiments. This prompted us to re-evaluate the metal cofactor stoichiometry of the 3′-5′ exonuclease mechanism for the Klenow fragment by solution kinetics and isothermal titration calorimetry using the natural Mg2+ cofactor and salt conditions. Both solution calorimetric and kinetics experiments strongly indicate binding of only one metal ion to the exonuclease active site. Comparative studies with Mn2+ also indicate a requirement for one metal ion to effect 3′-5′ exonuclease activity. Received, accepted: 16 March 1998  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