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.
Rhodococcus globerulus strain P6 contains at least three genes, bphC1, bphC2, and bphC3, coding for 2,3-dihydroxybiphenyl 1,2-dioxygenases; the latter two specify enzymes of the family of one-domain extradiol dioxygenases. In order to assess the importance of these different isoenzymes for the broad catabolic activity of this organism towards the degradation of polychlorinated biphenyls (PCBs), the capacities of recombinant enzymes expressed in Escherichia coli to transform different chlorosubstituted dihydroxybiphenyls formed by the action of R. globerulus P6 biphenyl dioxygenase and biphenyl 2,3-dihydrodiol dehydrogenase were determined. Whereas both BphC2 and BphC3 showed similar activities for 2,3-dihydroxybiphenyl and all monochlorinated 2,3-dihydroxybiphenyls, BphC1 exhibited only weak activity for 2'-chloro-2,3-dihydroxybiphenyl. More highly chlorinated 2'-chlorosubstituted 2,3-dihydroxybiphenyls were also transformed at high rates by BphC2 and BphC3 but not BphC1. In R. globerulus P6, BphC2 was constitutively expressed, BphC1 expression was induced during growth on biphenyl, and BphC3 was not expressed at significant levels under the experimental conditions. Although we cannot rule out the expression of BphC3 under certain environmental conditions, it seems that the contrasting substrate specificities of BphC1 and BphC2 contribute significantly to the versatile PCB-degrading phenotype of R. globerulus P6.  相似文献   

2.
AIMS: Isolation of the genes relative to PCB biodegradation and identification of the bph gene function in Rhodococcus sp. R04. METHODS AND RESULTS: A 8.7-kb fragment carrying the biphenyl catabolic genes bphABCD was isolated from the gene library in Rhodococcus sp. R04. Based on the deduced amino acid sequence homology, seven bph genes, bphA1A2A3A4, bphB, bphC and bphD, were thought to be responsible for the initial four steps of biphenyl degradation. In Escherichia coli, BphA exhibited poor activity for biphenyl transformation, and BphB, BphC and BphD were found to be catalytically active towards 2,3-dihydro-2,3-dihydroxybiphenyl, 2,3-dihydroxybiphenyl and 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoate, respectively (activities of 50, 8.1 and 2.4 micromol l(-1) min(-1) mg(-1)). SDS-PAGE analysis indicated that the sizes of bphA1A2A3A4, bphB, bphC and bphD gene products were 49, 19, 14, 47, 32, 30 and 31 kDa, respectively. After disruption of bph genes, the bphA1 mutants lost the ability to grow on biphenyl, the bphB and bphD mutants were able to transform a little of biphenyl, but hardly grew on biphenyl. CONCLUSION: The cloned bph genes indeed play an important role in the biphenyl catabolism in this strain. SIGNIFICANCE AND IMPACT OF THE STUDY: This bph gene organization in Rhodococcus sp. R04 differs from that of other biphenyl degraders reported previously, indicating it is a novel type of bph gene cluster. Analysis of the phylogenetic tree suggested that BphA1 and BphA2 in Rhodococcus sp. R04 had a different evolutionary relationship with those in the other PCB degraders.  相似文献   

3.
Rhodococcus rhodochrous K37, a Gram-positive bacterium grown under alkaline conditions, was isolated for its ability to metabolize PCBs. Analysis revealed that it has eight genes encoding extradiol dioxygenase, which has 2,3-dihydroxybiphenyl 1,2-dioxygenase activity, and these genes were designated bphC1 to bphC8. According to the classification of extradiol dioxygenases [Eltis, L. D., and Bolin, J. T., J. Bacteriol., 178, 5930-5937 (1996)], BphC3 and BphC6 belong to the type II enzyme group. The other six BphCs were classified as members of the type I extradiol dioxygenase group. BphC4 and BphC8 were classified into a new subfamily of type I, family 3. Two linear plasmids, 200 kb and 270 kb in size, were found in K37, and the bphC6 and bphC8 genes were located in the 200 kb linear plasmid. Northern hybridization analysis revealed that the bphC1, bphC2, and bphC7 genes were induced in the presence of testosterone, the bphC6 gene was induced by fluorene, and the bphC8 gene was induced by biphenyl. All eight BphC products exhibited much higher substrate activity for 2,3-dihydroxybiphenyl than for catechol, 3-methylcatechol, or 4-methylcatechol.  相似文献   

4.
The 2,3-dihydroxybiphenyl 1,2-dioxygenase from Sphingomonas xenophaga strain BN6 (BphC1) oxidizes 3-chlorocatechol by a rather unique distal ring cleavage mechanism. In an effort to improve the efficiency of this reaction, bphC1 was randomly mutated by error-prone PCR. Mutants which showed increased activities for 3-chlorocatechol were obtained, and the mutant forms of the enzyme were shown to contain two or three amino acid substitutions. Variant enzymes containing single substitutions were constructed, and the amino acid substitutions responsible for altered enzyme properties were identified. One variant enzyme, which contained an exchanged amino acid in the C-terminal part, revealed a higher level of stability during conversion of 3-chlorocatechol than the wild-type enzyme. Two other variant enzymes contained amino acid substitutions in a region of the enzyme that is considered to be involved in substrate binding. These two variant enzymes exhibited a significantly altered substrate specificity and an about fivefold-higher reaction rate for 3-chlorocatechol conversion than the wild-type enzyme. Furthermore, these variant enzymes showed the novel capability to oxidize 3-methylcatechol and 2,3-dihydroxybiphenyl by a distal cleavage mechanism.  相似文献   

5.
S Kosono  M Maeda  F Fuji  H Arai    T Kudo 《Applied microbiology》1997,63(8):3282-3285
Rhodococcus erythropolis TA421, a polychlorinated biphenyl and biphenyl degrader isolated from a termite ecosystem, has seven bphC genes expressing 2,3-dihydroxybiphenyl dioxygenase activity. R. erythropolis TA421 harbored a large and probably linear plasmid on which three (bphC2, bphC3, and bphC4) of the seven bphC genes were located. A non-biphenyl-degrading mutant, designated strain TA422, was obtained spontaneously from R. erythropolis TA421. TA422 lacked the plasmid, suggesting that the three bphC genes were involved in the degradation of biphenyl. Southern blot analyses showed that R. erythropolis TA421 and Rhodococcus globerulus P6 have a similar set of bphC genes and that the genes for biphenyl catabolism are located on plasmids of different sizes. These results indicated that the genes encoding the biphenyl catabolic pathway in Rhodococcus strains are borne on plasmids.  相似文献   

6.
The 2,3-dihydroxybiphenyl dioxygenase from Sphingomonas sp. strain BN6 (BphC1-BN6) differs from most other extradiol dioxygenases by its ability to oxidize 3-chlorocatechol to 3-chloro-2-hydroxymuconic semialdehyde by a distal cleavage mechanism. The turnover of different substrates and the effects of various inhibitors on BphC1-BN6 were compared with those of another 2,3-dihydroxybiphenyl dioxygenase from the same strain (BphC2-BN6) as well as with those of the archetypical catechol 2,3-dioxygenase (C23O-mt2) encoded by the TOL plasmid. Cell extracts containing C23O-mt2 or BphC2-BN6 converted the relevant substrates with an almost constant rate for at least 10 min, whereas BphC1-BN6 was inactivated significantly within the first minutes during the turnover of all substrates tested. Furthermore, BphC1-BN6 was much more sensitive than the other two enzymes to inactivation by the Fe(II) ion-chelating compound o-phenanthroline. The reason for inactivation of BphC1-BN6 appeared to be the loss of the weakly bound ferrous ion, which is the cofactor in the catalytic center. A mutant enzyme of BphC1-BN6 constructed by site-directed mutagenesis showed a higher stability to inactivation by o-phenanthroline and an increased catalytic efficiency for the conversion of 2,3-dihydroxybiphenyl and 3-methylcatechol but was still inactivated during substrate oxidation.  相似文献   

7.
Gram-positive Rhodococcus sp. strain RHA1 possesses strong polychlorinated biphenyl-degrading capabilities. An RHA1 bphC gene mutant, strain RDC1, had been previously constructed (E. Masai, A. Yamada, J. M. Healy, T. Hatta, K. Kimbara, M. Fukuda, and K. Yano, Appl. Environ. Microbiol. 61:2079-2085, 1995). An alternative 2,3-dihydroxybiphenyl 1,2-dioxygenase (2,3-DHBD), designated EtbC, was identified in RDC1 cells grown on ethylbenzene. EtbC contained the broadest substrate specificity of any meta cleavage dioxygenase identified in a Rhodococcus strain to date, including RHA1 BphC. EtbC was purified to near homogeneity from RDC1 cells grown on ethylbenzene, and a 58-amino-acid NH2-terminal sequence was determined. The NH2-terminal amino acid sequence was used for the identification of the etbC gene from an RDC1 chromosomal DNA 2,3-DHBD expression library. The etbC gene was successfully cloned, and we report here the determination of its nucleotide sequence. The substrate specificity patterns of cell extract and native nondenaturing polyacrylamide gel electrophoresis analysis identified the coexpression of two 2,3-DHBDs (BphC and EtbC) in RHA1 cells grown on either biphenyl or ethylbenzene. The possible implication of coexpressed BphC extradiol dioxygenases in the strong polychlorinated-biphenyl degradation activity of RHA1 was suggested.  相似文献   

8.
AIMS: Biphenyl-degrading bacteria are able to metabolize dibenzofuran via lateral dioxygenation and meta-cleavage of the dihydroxylated dibenzofuran produced. This degradation was considered to be incomplete because accumulation of a yellow-orange ring-cleavage product was observed. In this study, we want to characterize the 1,2-dihydroxydibenzofuran cleaving enzyme which is involved in dibenzofuran degradation in the bacterium Ralstonia sp. SBUG 290. METHODS AND RESULTS: In this strain, complete degradation of dibenzofuran was observed after cultivation on biphenyl. The enzyme shows a wide substrate utilization spectrum, including 1,2-dihydroxydibenzofuran, 2,3-dihydroxybiphenyl, 1,2-dihydroxynaphthalene, 3- and 4-methylcatechol and catechol. MALDI-TOF analysis of the protein revealed a strong homology to the bphC gene products. We therefore cloned a 3.2 kb DNA fragment containing the bphC gene of Ralstonia sp. SBUG 290. The deduced amino acid sequence of bphC is identical to that of the corresponding gene in Pseudomonas sp. KKS102. The bphC gene was expressed in Escherichia coli and the meta-fission activity was detected using either 2,3-dihydroxybiphenyl or 1,2-dihydroxydibenzofuran as substrate. CONCLUSIONS: These results demonstrate that complete degradation of dibenzofuran by biphenyl degraders can occur after initial oxidation steps catalysed by gene products encoded by the bph-operon. The ring fission of 1,2-dihydroxydibenzofuran is catalysed by BphC. Differences found in the metabolism of the ring fission product of dibenzofuran among biphenyl degrading bacteria are assumed to be caused by different substrate specificities of BphD. SIGNIFICANCE AND IMPACT OF THE STUDY: This study shows for the first time that the gene products of the bph-operon are involved in the mineralization of dibenzofuran in biphenyl degrading bacteria.  相似文献   

9.
A novel thermostable Mn(II)-dependent 2,3-dihydroxybiphenyl-1,2-dioxygenase (BphC_JF8) catalyzing the meta-cleavage of the hydroxylated biphenyl ring was purified from the thermophilic biphenyl and naphthalene degrader, Bacillus sp. JF8, and the gene was cloned. The native and recombinant BphC enzyme was purified to homogeneity. The enzyme has a molecular mass of 125 +/- 10 kDa and was composed of four identical subunits (35 kDa). BphC_JF8 has a temperature optimum of 85 degrees C and a pH optimum of 7.5. It exhibited a half-life of 30 min at 80 degrees C and 81 min at 75 degrees C, making it the most thermostable extradiol dioxygenase studied. Inductively coupled plasma mass spectrometry analysis confirmed the presence of 4.0-4.8 manganese atoms per enzyme molecule. The EPR spectrum of BphC_JF8 exhibited g = 2.02 and g = 4.06 signals having the 6-fold hyperfine splitting characteristic of Mn(II). The enzyme can oxidize a wide range of substrates, and the substrate preference was in the order 2,3-dihydroxybiphenyl > 3-methylcatechol > catechol > 4-methylcatechol > 4-chlorocatechol. The enzyme is resistant to denaturation by various chelators and inhibitors (EDTA, 1,10-phenanthroline, H2O2, 3-chlorocatechol) and did not exhibit substrate inhibition even at 3 mm 2,3-dihydroxybiphenyl. A decrease in Km accompanied an increase in temperature, and the Km value of 0.095 microm for 2,3-dihydroxybiphenyl (at 60 degrees C) is among the lowest reported. The kinetic properties and thermal stability of the native and recombinant enzyme were identical. The primary structure of BphC_JF8 exhibits less than 25% sequence identity to other 2,3-dihydroxybiphenyl 1,2-dioxygenases. The metal ligands and active site residues of extradiol dioxygenases are conserved, although several amino acid residues found exclusively in enzymes that preferentially cleave bicyclic substrates are missing in BphC_JF8. A three-dimensional homology model of BphC_JF8 provided a basis for understanding the substrate specificity, quaternary structure, and stability of the enzyme.  相似文献   

10.
The dibenzofuran-degrading bacterial strain DPO360 represents a new species of the genus Terrabacter together with the previously described dibenzofuran-mineralizing bacterial strain DPO1361 (K.-H. Engesser, V. Strubel, K. Christoglou, P. Fischer, and H. G. Rast, FEMS Microbiol. Lett. 65:205-210, 1989; V. Strubel, Ph.D. thesis, University of Stuttgart, Stuttgart, Germany, 1991; V. Strubel, H. G. Rast, W. Fietz, H.-J. Knackmuss, and K.-H. Engesser, FEMS Microbiol. Lett. 58:233-238, 1989). Two 2,3-dihydroxybiphenyl-1,2-dioxygenases (BphC1 and BphC2) and one catechol-2,3-dioxygenase (C23O) were shown to be expressed in Terrabacter sp. strain DPO360 growing with dibenzofuran as a sole source of carbon and energy. These enzymes exhibited strong sensitivity to oxygen. They were purified to apparent homogeneity as homodimers (BphC and BphC2) and as a homotetrameric catechol-2,3-dioxygenase (C23O). According to their specificity constants kcat/Km, both BphC1 and BphC2 were shown to be responsible for the cleavage of 2,2',3-trihydroxybiphenyl, the first metabolite in dibenzofuran mineralization along the angular dioxygenation pathway. With this substrate, BphC2 exhibited a considerably higher kcat/Km, value (183 microM/min) than BphC1 (29 microM/min). Catechol-2,3-dioxygenase was recognized to be not involved in the ring cleavage of 2,2',3-trihydroxybiphenyl (kcat/Km, 1 microM/min). Analysis of deduced amino acid sequence data of bphC1 revealed 36% sequence identity to nahC from Pseudomonas putida PpG7 (S. Harayama and M. Rekik, J. Biol. Chem. 264:15328-15333, 1989) and about 40% sequence identity to various bphC genes from different Pseudomonas and Rhodococcus strains. In addition, another 2,3-dihydroxybiphenyl-1,2-dioxygenase gene (bphC3) was cloned from the genome of Terrabacter sp. strain DPO360. Expression of this gene, however, could not be detected in Terrabacter sp. strain DPO360 after growth with dibenzofuran.  相似文献   

11.
M Maeda  S Y Chung  E Song    T Kudo 《Applied microbiology》1995,61(2):549-555
Rhodococcus erythropolis TA421 was isolated from a termite ecosystem and is able to degrade a wide range of polychlorinated biphenyl (PCB) congeners. Genetic and biochemical analyses of the PCB catabolic pathway of this organism revealed that there are four different bphC genes (bphC1, bphC2, bphC3, and bphC4) which encode 2,3-dihydroxybiphenyl dioxygenases. As determined by Southern hybridization, none of the bphC genes exhibits homology to any other bphC gene. bphC1, bphC2, and bphC4 encode enzymes that have narrow substrate specificities and cleave the first aromatic ring in the meta position. In contrast, bphC3 encodes a meta cleavage dioxygenase with broad substrate specificity. Asturias et al. have shown that the closely related organism Rhodococcus globerulus P6 contains three different bphC genes (bphC1, bphC2, and bpHC3) which encode meta cleavage dioxygenases. The data suggest that there is a diverse family of bphC genes which encode PCB meta cleavage dioxygenases in members of the genus Rhodococcus.  相似文献   

12.
Rhodococcus globerulus P6 (previously designated Acinetobacter sp. strain P6, Arthrobacter sp. strain M5, and Corynebacterium sp. strain MB1) is able to degrade a wide range of polychlorinated biphenyl (PCB) congeners. The genetic and biochemical analyses of the PCB catabolic pathway reported here have revealed the existence of a PCB gene cluster--bphBC1D--and two further bphC genes--bphC2 and bphC3--that encode three narrow-substrate-specificity enzymes (2,3-dihydroxybiphenyl dioxygenases) that meta cleave the first aromatic ring. None of the bphC genes show by hybridization homology to each other or to bphC genes in other bacteria, and the three bphC gene products have different kinetic parameters and sensitivities to inactivation by 3-chlorocatechol. This suggests that there exists a wide diversity in PCB meta cleavage enzymes.  相似文献   

13.
Mycobacterium bovis BCG and Mycobacterium tuberculosis possess a single arylamine N-acetyltransferase whose gene is predicted to occur within a six-gene operon. Deletion of the nat gene caused an extended lag phase in M. bovis BCG and a cell morphology associated with an altered pattern of cell wall mycolates. Analysis of cDNA from M. bovis BCG shows that during in vitro growth all the genes in the putative nat operon are expressed and the open reading frames are contiguous, supporting the existence of an operon. Two genes in the operon, Mb3599c and Mb3600c, are predicted to encode homologues of enzymes annotated as a 2,3-dihydroxybiphenyl 1,2-dioxygenase (bphC5) and a 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoate hydrolase (bphD2), respectively, in Rhodococcus RHA1. As predicted, M. bovis BCG cell lysates metabolized the BphC substrate 2,3-dihydroxybiphenyl (2,3-DHB) to 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA), a BphD substrate, which was subsequently hydrolysed. Immunoprecipitation of the BphD homologue from these lysates led to an accumulation of HOPDA. M. bovis BCG growth on both solid and liquid media was inhibited with either 2,3-DHB or an inhibitor of BphC, 3-chlorocatechol (3-CC). In addition, incubation with 2,3-DHB affects the lipid composition of the cell wall resulting in a diminished level of mycolates and an altered cell morphology similar to the Deltanat strain. We propose the enzymes encoded by the putative operon have a similar endogenous role to that of the NAT enzyme and are part of a pathway important for cell wall synthesis.  相似文献   

14.
A bphC gene (915 bp) encoding 2,3-dihydroxybiphenyl 1,2-dioxygenase (BphC) was amplified by PCR from Dyella ginsengisoli LA-4, which was heterologously expressed in Escherichia coli . The purified His-Tag BphC was able to catalyze the meta -cleavage reaction of the dihydroxylated aromatic rings. According to the specificity constant ( K cat/ K m) of BphC_LA-4, the specificity of BphC_LA-4 was determined in the following order: 2,3-dihydroxybiphenyl>3-methylcatechol>catechol>4-chlorocatechol>4-methylcatechol. The experimental data were consistent with the prediction of enzyme–substrate complexes. The highest specific activity of BphC_LA-4 was 118.3 U mg−1 for 2,3-dihydroxybiphenyl.  相似文献   

15.
Summary 3,4-dihydroxybiphenyl is not a substrate for the 2,3-dihydroxybiphenyl 1,2-dioxygenase (BphC) from biphenyldegradingPseudomonas sp. strain CB406. It acts as both a reversible inhibitor and a potent inactivator of the enzyme. The inactivation process requires the presence of O2 and can be reversed by the removal of the 3,4-dihydroxybiphenyl followed by incubation of the enzyme in the presence of dithioerythritol and Fe2+ under anaerobic conditions. Two other extradiol dioxygenases behave similarly, the catechol 2,3-dioxygenase (BphE) from strain CB406 and the BphC fromPseudomonas sp. strain LB400. The BphC fromP. testosteroni B-356 also did not cleave 3,4-dihydroxybiphenyl but it was not inactivated.Abbreviations C23o Catechol 2,3-dioxygenase - 34DHBP 3,4-dihydroxybiphenyl  相似文献   

16.
17.
2,3-Dihydroxybiphenyl dioxygenase, involved in biphenyl and polychlorinated biphenyl degradation, was purified from cell extracts of polychlorinated biphenyl-degrading Pseudomonas pseudoalcaligenes KF707 and Pseudomonas aeruginosa PAO1161 carrying the cloned bphC gene (encoding 2,3-dihydroxybiphenyl dioxygenase). The purified enzyme contained ferrous iron as a prosthetic group. The specific activities decreased with the loss of ferrous iron from the enzyme, and the activity was restored by incubation with ferrous iron in the presence of cysteine. Addition of ferric iron caused the complete inactivation of the enzyme. The molecular weight was estimated to be 250,000. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis showed a single band with a molecular weight of 31,000, indicating that the enzyme consists of eight identical subunits. The enzyme was specific only for 2,3-dihydroxybiphenyl with a Km value of 87 microM. No significant activity was observed for 3,4-dihydroxybiphenyl, catechol, or 3-methyl- and 4-methylcatechol. The molecular weight, subunit structure, ferrous iron requirement, and NH2-terminal sequence (starting with serine up to 12 residues) were the same between the two enzymes obtained from KF707 and PAO1161 (bphC).  相似文献   

18.
嗜吡啶红球菌R04的联苯降解途径的研究   总被引:3,自引:0,他引:3  
通过GC-MS测定出嗜吡啶红球菌R04菌降解联苯的中间代谢物2,3-二氢二羟基联苯、2,3-二羟基联苯和苯甲酸,并测定了该菌的2,3-二羟基联苯双加氧酶、2-羟基-6-酮基-6-苯基-2,3-己二烯酸(HOPDA)水解酶和苯甲酸双加氧酶活性。最终确定了R04菌降解联苯的途径为2,3-二羟基联苯双加氧酶途径。  相似文献   

19.
The green fluorescent protein gene (gfp) was inserted into the chromosome of Pseudomonas spp. Cam-1 and Sag-50G, two psychrotolerant polychlorinated biphenyl (PCB)-degrading bacteria. The gfp-transformed microorganisms, designated Cam-1-gfp1, Cam-1-gfp2, Sag-50G-gfp1, and Sag-50G-gfp2, exhibited green fluorescence under an epifluorescent microscope. The gfp was inserted into the chromosome of each psychrotolerant strain and was stable with no apparent adverse affects on the metabolism and growth of each organism. Activity of gfp-transformed microorganisms against biphenyl and 2,3-dichlorobiphenyl was determined by assaying for BphC activity and by resting cell assays. The patterns of BphC activity at two different growth temperatures in batch cultures were similar for each of the gfp-transformed microorganisms. Resting cell assays of both the parent strains (Cam-1, Sag-50G) and the gfp-transformed strains (Cam-1-gfp1, Cam-1-gfp2, Sag-50G-gfp1, Sag-50G-gfp2), grown on glycerol or glucose, exhibited BphC activity to a lesser extent and at a slower rate than those observed for biphenyl grown cells. In addition, all gfp-transformed microorganisms degraded 2,3-dichlorobiphenyl (2,3-DCB) in broth to the same extent as the parent strains. When Cam-1-gfp1 and Sag-50G-gfp1 were used as a bioremediation amendment in soil microcosms spiked with 2,3-DCB, both strains survived in high numbers (5.6 to 7.9 log cfu g?1 and 5.6 to 8.0 log cfu g?1) when inoculated into nonsterilized soil over 16 weeks at 22 degrees C and 18 weeks at 4 degrees C, respectively. Biodegradation of 2,3-DCB was enhanced with the microbial amendment; however, the addition of sunflower oil did not help the PCB degrading bacteria and may have enhanced the growth of the indigenous population, thereby decreasing the amended PCB-degrading population.  相似文献   

20.
Crystals have been obtained for a 2,3-dihydroxybiphenyl dioxygenase (conventionally called BphC) from a polychlorinated biphenyl (PCB)-degrader, Pseudomonas sp. strain KKS1O2. The crystals were grown using both ammonium sulfate and MPD as the precipitating agents. The crystals belonged to a tetragonal space group (I422) and diffracted to 2.5 Å. © 1995 Wiley-Liss, Inc.  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