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.
The mechanism of the denitrification and nitrosation reactions catalyzed by the heme cd-containing nitrite reductase from Pseudomonas stutzeri JM 300 has been studied with whole cell suspensions using H2(18)O, 15NO, and 15NO-2. The extent of H2(18)O exchange with the enzyme-bound nitrosyl intermediate, as determined by the 18O content of product N2O, decreased with increasing nitrite concentration, which is consistent with production of N2O by sequential reaction of two nitrite ions with the enzyme. Reaction of NO with whole cells in H2(18)O gave amounts of 18O in the N2O product consistent with equilibration of nitric oxide with a small pool of free nitrite. Using 15NO and NH2OH, competition between denitrification and nitrosation reactions was demonstrated, as is required if the enzyme-nitrosyl complex is an intermediate in both nitrosation and denitrification reactions. The first evidence for exchange of 18O between H2(18)O and a nitrosation intermediate occurring after the enzyme-nitrosyl complex, presumably an enzyme-bound nitrosamine, has been obtained. The collective results are most consistent with denitrification N2O originating via attack of NO-2 on a coordinated nitrosyl, as proposed earlier (Averill, B. A., and Tiedje, J. M. (1982) FEBS Lett. 138, 8-11).  相似文献   

2.
Two denitrifying bacteria, Pseudomonas stutzeri and Achromobacter cycloclastes, were incubated with Na15NO2 and NaN3 under conditions that allowed catalysis of nitrosyl transfer from nitrite to azide. This transfer, which is presumed to be mediated by the heme- and copper-containing nitrite reductase of P. stutzeri and A. cycloclastes, respectively, leads to formation of isotopically mixed 14,15N2O, whereas denitrification leads to 15N2O. The conditions that emphasized nitrosyl transfer also partially inhibited the nitric oxide reductase system and led to accumulation of 15NO. Absorption of NO from the gas phase by acidic CrSO4 in a sidewell largely abolished nitrosyl transfer to azide. With these two organisms, which are thought to be representative of denitrifiers generally, catalysis of nitrosyl transfer seemed to depend on NO.  相似文献   

3.
Bacterial denitrification reverses nitrogen fixation in the global N-cycle by transforming nitrate or nitrite to dinitrogen. Both nitrite and nitric oxide (NO) are considered as the chemical species within the denitrification pathway, that precede nitrous oxide (N2O), the first recognized intermediate with N,N-bonds antecedent to N2. Molecular cloning of the structural genes for NO reductase from Pseudomonas stutzeri has allowed us to generate the first mutants defective in NO utilization (Nor- phenotype) by marker exchange of the norCB genes with a gene cassette for gentamicin resistance. Nitric oxide reductase was found to be an indispensable component for denitrification; its loss constituted a conditionally lethal mutation. NO as the sole product accumulated from nitrite by mutant cells induced for nitrite respiration (denitrification). The Nor- mutant lost the capability to reduce NO and did not grow anymore anaerobically on nitrate. A Nir-Nor- double mutation, that inactivated also the respiratory nitrite reductase cytochrome cd1 rendered the bacterium again viable under anaerobiosis. Our observations provide evidence for a denitrification pathway in vivo of NO2(-)----NO----N2O, and N,N-bond formation catalyzed by NO reductase and not by cytochrome cd1.  相似文献   

4.
The dissimilatory nitrite reductase (cytochrome c,d1) from Pseudomonas aeruginosa was observed at pH 7.5 to catalyze nitrosyl transfer (nitrosation) between [15N]nitrite and several N-nucleophiles or H2 18O, with rate enhancement of the order of 10(8) relative to analogous chemical reactions. The reducing system (ascorbate, N,N,N',N'-tetramethylphenylenediamine) could reduce nitrite (but not NO) enzymatically and had essentially no direct chemical reactivity toward nitrite or NO. The N-nitrosations showed saturation kinetics with respect to the nucleophile and, while exhibiting Vmax values which varied by about 40-fold, nevertheless showed little or no dependence of Vmax on nucleophile pKa. The N-nitrosations and NO-2/H2O-18O exchange required the reducing system, whereas NO/H2O-18O exchange was inhibited by the reducing system. NO was not detected to serve as a nitrosyl donor to N-nucleophiles. These and other kinetic observations suggest that the enzymatic nitrosyl donor is an enzyme-bound species derived from reduced enzyme and one molecule of nitrite, possibly a heme-nitrosyl compound (E-FeII X NO+) for which there is precedence. Nitrosyl transfer to N-nucleophiles may occur within a ternary complex of enzyme, nitrite, and nucleophile. Catalysis of nitrosyl transfer by nitrite reductase represents a new class of enzymatic reactions and may present another example of electrophilic catalysis by a metal center. The nitrosyl donor trapped by these reactions is believed to represent an intermediate in the reduction of nitrite by cytochrome c,d1.  相似文献   

5.
6.
The mechanism of anaerobic reduction of NO2? to N2O in a photodenitrifier, Rhodopseudomonas sphaeroides forma sp. denitrificans, was investigated. With ascorbate-reduced phenazine methosulfate (PMS) as the electron donor, the nitrite reductase of this photodenitrifier reduced NO2? to NO and a trace amount of N2O. With dithionite-reduced benzyl viologen as the electron donor, the major product of NO2? reduction was NH2OH, and a trace amount of N2O was also produced. The nitrate reductase itself had no NO reductase activity with ascorbate-reduced PMS. It was concluded that the essential product of NO2? reduction by the purified nitrite reductase is NO. Chromatophore membranes stoichiometrically produced N2O from NO2? with any electron donor, such as dithionite-redduced benzyl viologen, ascorbate-reduced PMS or NADH/FMN. The membranes also contrained activity of NO reduction of N2O with either ascorbate-reduced PMS or duroquinol. The NO reductase activity with duroquinol was inhibited by antimycin A. Stoichiometric production of N2O from N2? also was observed in the reconstituted NO2? reduction system which contained the cytochrome bc1 complex, cytochrome c2, the nitrite reductase and duroquinol as the electron donor. The preparation of the cytochrome bc1 complex itself contianed NO reductase activity. From these results the mechanism of NO2? reduction to N2O in this photodenitrifier was determined as the nitrite reductase reducing NO2? to NO with electrons from the cytochrome bc1 complex, and NO subsequently being reduced, without release, to N2O with electrons from the cytochrome bc1 complex by the NO reductase, which is closely associated with the complex.  相似文献   

7.
The ambidentate ligand nitrite (NO2) binds to transition metal centers through the N (nitro) or O (nitrito) atom. In metal porphyrin complexes, the energy difference between the two linkage isomers is small and hence slight differences in reaction conditions and/or ligand design give rise to formation of the isomers in different ratios. In the present work, similar behavior has been observed in case of the {Fe-NO}6 nitrosyl [(Me2bpb)Fe(NO)(NO2)] (2), derived from a non-heme planar dicarboxamide ligand N,N′-bispyridinecarboxamido-4,5-dimethylbenzenediamine (H2Me2bpb). Under anaerobic conditions, reaction of the Fe(III) complex [(Me2bpb)Fe(py)2]ClO4 (1) with NO(g) in MeCN affords 2, a product that contains both the N- and O-bound isomer in different ratios depending on the reaction conditions. In protic solvents, the same reaction affords the {Fe-NO}7 nitrosyl [(Me2bpb)Fe(NO)] (3). Both nitrosyls have been characterized by infrared spectroscopy and X-ray diffraction studies.  相似文献   

8.
Nitrate and nitrite concentrations in the water and nitrous oxide and nitrite fluxes across the sediment-water interface were measured monthly in the River Colne estuary, England, from December 1996 to March 1998. Water column concentrations of N2O in the Colne were supersaturated with respect to air, indicating that the estuary was a source of N2O for the atmosphere. At the freshwater end of the estuary, nitrous oxide effluxes from the sediment were closely correlated with the nitrite concentrations in the overlying water and with the nitrite influx into the sediment. Increases in N2O production from sediments were about 10 times greater with the addition of nitrite than with the addition of nitrate. Rates of denitrification were stimulated to a larger extent by enhanced nitrite than by nitrate concentrations. At 550 μM nitrite or nitrate (the highest concentration used), the rates of denitrification were 600 μmol N · m−2 · h−1 with nitrite but only 180 μmol N · m−2 · h−1 with nitrate. The ratios of rates of nitrous oxide production and denitrification (N2O/N2 × 100) were significantly higher with the addition of nitrite (7 to 13% of denitrification) than with nitrate (2 to 4% of denitrification). The results suggested that in addition to anaerobic bacteria, which possess the complete denitrification pathway for N2 formation in the estuarine sediments, there may be two other groups of bacteria: nitrite denitrifiers, which reduce nitrite to N2 via N2O, and obligate nitrite-denitrifying bacteria, which reduce nitrite to N2O as the end product. Consideration of free-energy changes during N2O formation led to the conclusion that N2O formation using nitrite as the electron acceptor is favored in the Colne estuary and may be a critical factor regulating the formation of N2O in high-nutrient-load estuaries.  相似文献   

9.
China is the world's largest producer and consumer of fertilizer N, and decades of overuse has caused nitrate leaching and possibly soil acidification. We hypothesized that this would enhance the soils' propensity to emit N2O from denitrification by reducing the expression of the enzyme N2O reductase. We investigated this by standardized oxic/anoxic incubations of soils from five long‐term fertilization experiments in different regions of China. After adjusting the nitrate concentration to 2 mM, we measured oxic respiration (R), potential denitrification (D), substrate‐induced denitrification, and the denitrification product stoichiometry (NO, N2O, N2). Soils with a history of high fertilizer N levels had high N2O/(N2O+N2) ratios, but only in those field experiments where soil pH had been lowered by N fertilization. By comparing all soils, we found a strong negative correlation between pH and the N2O/(N2O+N2) product ratio (r2 = 0.759, P < 0.001). In contrast, the potential denitrification (D) was found to be a linear function of oxic respiration (R), and the ratio D/R was largely unaffected by soil pH. The immediate effect of liming acidified soils was lowered N2O/(N2O+N2) ratios. The results provide evidence that soil pH has a marginal direct effect on potential denitrification, but that it is the master variable controlling the percentage of denitrified N emitted as N2O. It has been known for long that low pH may result in high N2O/(N2O+N2) product ratios of denitrification, but our documentation of a pervasive pH‐control of this ratio across soil types and management practices is new. The results are in good agreement with new understanding of how pH may interfere with the expression of N2O reductase. We argue that the management of soil pH should be high on the agenda for mitigating N2O emissions in the future, particularly for countries where ongoing intensification of plant production is likely to acidify the soils.  相似文献   

10.
Nitrite reductase (NirK) and nitric oxide reductase (NorB) have long been thought to play an essential role in nitrous oxide (N2O) production by ammonia-oxidizing bacteria. However, essential gaps remain in our understanding of how and when NirK and NorB are active and functional, putting into question their precise roles in N2O production by ammonia oxidizers. The growth phenotypes of the Nitrosomonas europaea ATCC 19718 wild-type and mutant strains deficient in expression of NirK, NorB, and both gene products were compared under atmospheric and reduced O2 tensions. Anoxic resting-cell assays and instantaneous nitrite (NO2) reduction experiments were done to assess the ability of the wild-type and mutant N. europaea strains to produce N2O through the nitrifier denitrification pathway. Results confirmed the role of NirK for efficient substrate oxidation of N. europaea and showed that NorB is involved in N2O production during growth at both atmospheric and reduced O2 tensions. Anoxic resting-cell assays and measurements of instantaneous NO2 reduction using hydrazine as an electron donor revealed that an alternate nitrite reductase to NirK is present and active. These experiments also clearly demonstrated that NorB was the sole nitric oxide reductase for nitrifier denitrification. The results of this study expand the enzymology for nitrogen metabolism and N2O production by N. europaea and will be useful to interpret pathways in other ammonia oxidizers that lack NirK and/or NorB genes.  相似文献   

11.
Nitrous oxide (N2O) was previously deemed as a potent greenhouse gas but is actually an untapped energy source, which can accumulate during the microbial denitrification of nitric oxide (NO). Compared with the organic electron donor required in heterotrophic denitrification, elemental sulfur (S0) is a promising electron donor alternative due to its cheap cost and low biomass yield in sulfur-driven autotrophic denitrification. However, no effort has been made to test N2O recovery from sulfur-driven denitrification of NO so far. Therefore, in this study, batch and continuous experiments were carried out to investigate the NO removal performance and N2O recovery potential via sulfur-driven NO-based denitrification under various Fe(II)EDTA-NO concentrations. Efficient energy recovery was achieved, as up to 35.5%–40.9% of NO was converted to N2O under various NO concentrations. N2O recovery from Fe(II)EDTA-NO could be enhanced by the low bioavailability of sulfur and the acid environment caused by sulfur oxidation. The NO reductase (NOR) and N2O reductase (N2OR) were inhibited distinctively at relatively low NO levels, leading to efficient N2O accumulation, but were suppressed irreversibly at NO level beyond 15 mM in continuous experiments. Such results indicated that the regulation of NO at a relatively low level would benefit the system stability and NO removal capacity during long-term system operation. The continuous operation of the sulfur-driven Fe(II)EDTA-NO-based denitrification reduced the overall microbial diversity but enriched several key microbial community. Thauera, Thermomonas, and Arenimonas that are able to carry out sulfur-driven autotrophic denitrification became the dominant organisms with their relative abundance increased from 25.8% to 68.3%, collectively.  相似文献   

12.
Nitric oxide (NO) and NO-derived reactive nitrogen species (RNS) are present in the food vacuole (FV) of Plasmodium falciparum trophozoites. The product of PFL1555w, a putative cytochrome b5, localizes in the FV membrane, similar to what was previously observed for the product of PF13_0353, a putative cytochrome b5 reductase. These two gene products may contribute to NO generation by denitrification chemistry from nitrate and/or nitrite present in the erythrocyte cytosol. The possible coordination of NO to heme species present in the food vacuole was probed by resonance Raman spectroscopy. The spectroscopic data revealed that in situ generated NO interacts with heme inside the intact FVs to form ferrous heme nitrosyl complexes that influence intra-vacuolar heme solubility. The formation of heme nitrosyl complexes within the FV is a previously unrecognized factor that could affect the equilibrium between soluble and crystallized heme within the FV in vivo.  相似文献   

13.
A mixed beech and spruce forest soil was incubated under potential denitrification assay (PDA) condition with 10% acetylene (C2H2) in the headspace of soil slurry bottles. Nitrous oxide (N2O) concentration in the headspace, as well as nitrate, nitrite and ammonium concentrations in the soil slurries were monitored during the incubation. Results show that nitrate disappearance rate was higher than N2O production rate with C2H2 blockage during the incubation. Sum of nitrate, nitrite, and N2O with C2H2 blockage could not recover the original soil nitrate content, showing an N imbalance in such a closed incubation system. Changes in nitrite and ammonium concentration during the incubation could not account for the observed faster nitrate disappearance rate and the N imbalance. Non-determined nitric oxide (NO) and N2 production could be the major cause, and the associated mechanisms could vary for different treatments. Commonly applied PDA measurement likely underestimates the nitrate removal capacity of a system. Incubation time and organic matter/nitrate ratio are the most critical factors to consider using C2H2 inhibition technique to quantify denitrification. By comparing the treatments with and without an antibiotic, the results suggest that microbial N uptake probably played a minor role in N balance, and other denitrifying enzymes but nitrate reductase could be substantially synthesized during the incubation.  相似文献   

14.
The emissions of nitrous oxide (N2O) and nitric oxide (NO) from biological nitrogen removal (BNR) operations via nitrification and denitrification is gaining increased prominence. While many factors relevant to the operation of denitrifying reactors can influence N2O and NO emissions from them, the role of different organic carbon sources on these emissions has not been systematically addressed or interpreted. The overall goal of this study was to evaluate the impact of three factors, organic carbon limitation, nitrite concentrations, and dissolved oxygen concentrations on gaseous N2O and NO emissions from two sequencing batch reactors (SBRs), operated, respectively, with methanol and ethanol as electron donors. During undisturbed ultimate‐state operation, emissions of both N2O and NO from either reactor were minimal and in the range of <0.2% of influent nitrate‐N load. Subsequently, the two reactors were challenged with transient organic carbon limitation and nitrite pulses, both of which had little impact on N2O or NO emissions for either electron donor. In contrast, transient exposure to oxygen led to increased production of N2O (up to 7.1% of influent nitrate‐N load) from ethanol grown cultures, owing to their higher kinetics and potentially lower susceptibility to oxygen inhibition. A similar increase in N2O production was not observed from methanol grown cultures. These results suggest that for dissolved oxygen, but not for carbon limitation or nitrite exposure, N2O emission from heterotrophic denitrification reactors can vary as a function of the electron donor used. Biotechnol. Bioeng. 2010; 106: 390–398. © 2010 Wiley Periodicals, Inc.  相似文献   

15.
16.
Peatlands cover more than 30% of the Finnish land area and impact N2O fluxes. Denitrifiers release N2O as an intermediate or end product. In situ N2O emissions of a near pH neutral pristine fen soil in Finnish Lapland were marginal during gas chamber measurements. However, nitrate and ammonium fertilization significantly stimulated in situ N2O emissions. Stimulation with nitrate was stronger than with ammonium. N2O was produced and subsequently consumed in gas chambers. In unsupplemented anoxic microcosms, fen soil produced N2O only when acetylene was added to block nitrous oxide reductase, suggesting complete denitrification. Nitrate and nitrite stimulated denitrification in fen soil, and maximal reaction velocities (vmax) of nitrate or nitrite dependent denitrification where 18 and 52 nmol N2O h-1 gDW -1, respectively. N2O was below 30% of total produced N gases in fen soil when concentrations of nitrate and nitrite were <500 μM. vmax for N2O consumption was up to 36 nmol N2O h-1 gDW -1. Denitrifier diversity was assessed by analyses of narG, nirK/nirS, and nosZ (encoding nitrate-, nitrite-, and nitrous oxide reductases, respectively) by barcoded amplicon pyrosequencing. Analyses of ~14,000 quality filtered sequences indicated up to 25 species-level operational taxonomic units (OTUs), and up to 359 OTUs at 97% sequence similarity, suggesting diverse denitrifiers. Phylogenetic analyses revealed clusters distantly related to publicly available sequences, suggesting hitherto unknown denitrifiers. Representatives of species-level OTUs were affiliated with sequences of unknown soil bacteria and Actinobacterial, Alpha-, Beta-, Gamma-, and Delta-Proteobacterial sequences. Comparison of the 4 gene markers at 97% similarity indicated a higher diversity of narG than for the other gene markers based on Shannon indices and observed number of OTUs. The collective data indicate (i) a high denitrification and N2O consumption potential, and (ii) a highly diverse, nitrate limited denitrifier community associated with potential N2O fluxes in a pH-neutral fen soil.  相似文献   

17.
The aqueous decomposition of the iridium coordinated nitrosothiols (RSNOs) trans-K[IrCl4(CH3CN)NOSPh] (1), and K2[IrCl5(NOECyS)] (2, ECyS = cysteine ethyl ester), was studied by MS analysis of the gaseous products, ESI-MS, NMR, and UV-Vis spectroscopy. Bent NO (NO, nitroxyl anion), sulfenic acids and nitrite were observed as coordinated products in solution, while nitrous oxide (N2O) and nitrogen were detected in the gas phase. The formation of coordinated NO and N2O, a nitroxyl dimerization product, allows us to propose the formation of free nitroxyl (HNO) as an intermediate. Complex 1 decomposes 300 times slower than free PhSNO does. In both cases (1 and 2) kinetic results show a first order decomposition behavior and a very negative ΔS, which strongly indicates an associative rate-determining step. A proposed decomposition mechanism, supported by the experimental data and DFT calculations, involves, as the first step, nucleophilic attack of H2O on to the sulfur atom of the coordinated RSNO, producing an NO complex and free sulfenic acid, followed by two competing reactions: a ligand exchange reaction of this NO with the sulfenic acid or, to a minor extent, coordination of N2O to produce an NO/N2O complex which finally renders free N2 and coordinated NO2. Some of the produced NO is likely to be released from the metal center producing nitroxyl by protonation and finally N2O by dimerization and loss of H2O. In conclusion, the decomposition of these coordinated RSNOs occurs through a different mechanism than for the decomposition of free RSNOs. It involves the formation of sulfenic acids and coordinated NO, which is released from the complexes and protonated at the reaction pH producing nitroxyl (HNO), and ultimately N2O.  相似文献   

18.
Wastewater treatment plants are known to be important point sources for nitrous oxide (N2O) in the anthropogenic N cycle. Biofilm based treatment systems have gained increasing popularity in the treatment of wastewater, but the mechanisms and controls of N2O formation are not fully understood. Here, we review functional groups of microorganism involved in nitrogen (N) transformations during wastewater treatment, with emphasis on potential mechanism of N2O production in biofilms. Biofilms used in wastewater treatment typically harbour aerobic and anaerobic zones, mediating close interactions between different groups of N transforming organisms. Current models of mass transfer and biomass interactions in biofilms are discussed to illustrate the complex regulation of N2O production. Ammonia oxidizing bacteria (AOB) are the prime source for N2O in aerobic zones, while heterotrophic denitrifiers dominate N2O production in anoxic zones. Nitrosative stress ensuing from accumulation of NO2 ? during partial nitrification or denitrification seems to be one of the most critical factors for enhanced N2O formation. In AOB, N2O production is coupled to nitrifier denitrification triggered by nitrosative stress, low O2 tension or low pH. Chemical N2O production from AOB intermediates (NH2OH, HNO, NO) released during high NH3 turnover seems to be limited to surface-near AOB clusters, since diffusive mass transport resistance for O2 slows down NH3 oxidation rates in deeper biofilm layers. The proportion of N2O among gaseous intermediates (NO, N2O, N2) in heterotrophic denitrification increases when NO or nitrous acid (HNO2) accumulates because of increasing NO2 ?, or when transient oxygen intrusion impairs complete denitrification. Limited electron donor availability due to mass transport limitation of organic substrates into anoxic biofilm zones is another important factor supporting high N2O/N2 ratios in heterotrophic denitrifiers. Biofilms accommodating Anammox bacteria release less N2O, because Anammox bacteria have no known N2O producing metabolism and reduce NO2 ? to N2, thereby lowering nitrosative stress to AOB and heterotrophs.  相似文献   

19.
The ability to use δ18O values of nitrous oxide (N2O) to apportion environmental emissions is currently hindered by a poor understanding of the controls on δ18O–N2O from nitrification (hydroxylamine oxidation to N2O and nitrite reduction to N2O). In this study fertilized agricultural soils and unfertilized temperate forest soils were aerobically incubated with different 18O/16O waters, and conceptual and mathematical models were developed to systematically explain the δ18O–N2O formed by nitrification. Modeling exercises used a set of defined input parameters to emulate the measured soil δ18O–N2O data (Monte Carlo approach). The Monte Carlo simulations implied that abiotic oxygen (O) exchange between nitrite (NO2?) and H2O is important in all soils, but that biological, enzyme‐controlled O‐exchange does not occur during the reduction of NO2? to N2O (nitrifier‐denitrification). Similarly, the results of the model simulations indicated that N2O consumption is not characteristic of aerobic N2O formation. The results of this study and a synthesis of the published literature data indicate that δ18O–N2O formed in aerobic environments is constrained between +13‰ and +35‰ relative to Vienna Standard Mean Ocean Water (VSMOW). N2O formed via hydroxylamine oxidation and nitrifier‐denitrification cannot be separated using δ18O unless 18O tracers are employed. The natural range of nitrifier δ18O–N2O is discussed and explained in terms of our conceptual model, and the major and minor controls that define aerobically produced δ18O–N2O are identified. Despite the highly complex nature of δ18O–N2O produced by nitrification this δ18O range is narrow. As a result, in many situations δ18O values may be used in conjunction with δ15N–N2O data to apportion nitrifier‐ and denitrifier‐derived N2O. However, when biological O‐exchange during denitrification is high and N2O consumption is low, there may be too much overlap in δ18O values to distinguish N2O formed by these pathways.  相似文献   

20.
Many actinomycete strains are able to convert nitrate or nitrite to nitrous oxide (N2O). As a representative of actinomycete denitrification systems, the system of Streptomyces thioluteus was investigated in detail. S. thioluteus attained distinct cell growth upon anaerobic incubation with nitrate or nitrite with concomitant and stoichiometric conversion of nitrate or nitrite to N2O, suggesting that the denitrification acts as anaerobic respiration. Furthermore, a copper-containing, dissimilatory nitrite reductase (CuNir) and its physiological electron donor, azurin, were isolated. This is the first report to show that denitrification generally occurs among actinomycetes.  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