首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A series of photosynthetic electron transport (PET) inhibitors from the group of salicylanilide alkylcarbamates was investigated. The compounds were analyzed using RP-HPLC to determine lipophilicity, and their PET inhibition was determined in spinach (Spinacia oleracea L.) chloroplasts. The site of action of the studied compounds is situated at the donor site of photosystem 2 (PS 2). Compounds substituted by chlorine in C′-3 and C′-4 of the aniline ring and the optimal length of the alkyl chain pentyl-heptyl in the carbamate moiety provided the most active PET inhibitors (IC50 inhibition <10 μmol/L). Disubstitution in C′-3,4 by chlorine caused significant PET inhibiting activity decrease. Nevertheless, for all three series of C′-3, C′-4, C′-3,4 compounds, the dependence of PET activity on lipophilicity showed to be quasi-parabolic.  相似文献   

2.
Certain phenolic compounds represent a distinct class of Photosystem (PS) II Q(B) site inhibitors. In this paper, we report a detailed study of the effects of 2,4,6-trinitrophenol (TNP) and other phenolic inhibitors, bromoxynil and dinoseb, on PS II energetics. In intact PS II, phenolic inhibitors bound to only 90-95% of Q(B) sites even at saturating concentrations. The remaining PS II reaction centers (5-10%) showed modified Q(A) to Q(B) electron transfer but were sensitive to urea/triazine inhibitors. The binding of phenolic inhibitors was 30- to 300-fold slower than the urea/triazine class of Q(B) site inhibitors, DCMU and atrazine. In the sensitive centers, the S(2)Q(A)(-) state was 10-fold less stable in the presence of phenolic inhibitors than the urea/triazine herbicides. In addition, the binding affinity of phenolic herbicides was decreased 10-fold in the S(2)Q(A)(-) state than the S(1)Q(A) state. However, removal of the oxygen-evolving complex (OEC) and associated extrinsic polypeptides by hydroxylamine (HA) washing abolished the slow binding kinetics as well as the destabilizing effects on the charge-separated state. The S(2)-multiline electron paramagnetic resonance (EPR) signal and the 'split' EPR signal, originating from the S(2)Y(Z) state showed no significant changes upon binding of phenolic inhibitors at the Q(B) site. We thus propose a working model where Q(A) redox potential is lowered by short-range conformational changes induced by phenolic inhibitor binding at the Q(B) niche. Long-range effects of HA-washing eliminate this interaction, possibly by allowing more flexibility in the Q(B) site.  相似文献   

3.
A series of 3-(1-alkylaminoalkylidene)-6-methyl-2H-pyran-2,4(3H)-diones was newly synthe-sized, and they were assayed as photosynthetic electron transport (PET) inhibitors because of their structural resemblance to cyanoacrylates and 2-alkylaminoalkylidene-1,3-cyclohexanedione derivatives, which are potent PET inhibitors. Some of the compounds synthesized here showed very high PET inhibition.  相似文献   

4.
The Cl- channel blocker NPPB (5-nitro-2-(3-phenylpropylamino) benzoic acid) inhibited photosynthetic oxygen evolution of isolated thylakoid membranes in a pH-dependent manner with a K(i) of about 2 microM at pH 6. Applying different electron acceptors, taking electrons either directly from photosystem II (PS II) or photosystem I (PS I), the site of inhibition was localized within PS II. Measurements of fluorescence induction kinetics and thermoluminescence suggest that the binding of NPPB to the QB binding site of PS II is similar to the herbicide DCMU (3-(3,4-dichlorophenyl)-1,1-dimethylurea). The effects of different arylaminobenzoate derivatives and other Cl- channel inhibitors on photosynthetic electron transport were investigated. The structure--activity relationship of the inhibitory effect on PS II shows interesting parallels to the one observed for the arylaminobenzoate block of mammalian Cl- channels. A molecular modeling approach was used to fit NPPB into the QB binding site and to identify possible molecular interactions between NPPB and the amino acid residues of the binding site in PS II. Taken together, these data give a detailed molecular picture of the mechanism of NPPB binding.  相似文献   

5.
Certain phenolic compounds represent a distinct class of Photosystem (PS) II QB site inhibitors. In this paper, we report a detailed study of the effects of 2,4,6-trinitrophenol (TNP) and other phenolic inhibitors, bromoxynil and dinoseb, on PS II energetics. In intact PS II, phenolic inhibitors bound to only 90-95% of QB sites even at saturating concentrations. The remaining PS II reaction centers (5-10%) showed modified QA to QB electron transfer but were sensitive to urea/triazine inhibitors. The binding of phenolic inhibitors was 30- to 300-fold slower than the urea/triazine class of QB site inhibitors, DCMU and atrazine. In the sensitive centers, the S2QA state was 10-fold less stable in the presence of phenolic inhibitors than the urea/triazine herbicides. In addition, the binding affinity of phenolic herbicides was decreased 10-fold in the S2QA state than the S1QA state. However, removal of the oxygen-evolving complex (OEC) and associated extrinsic polypeptides by hydroxylamine (HA) washing abolished the slow binding kinetics as well as the destabilizing effects on the charge-separated state. The S2-multiline electron paramagnetic resonance (EPR) signal and the ‘split’ EPR signal, originating from the S2YZ state showed no significant changes upon binding of phenolic inhibitors at the QB site. We thus propose a working model where QA redox potential is lowered by short-range conformational changes induced by phenolic inhibitor binding at the QB niche. Long-range effects of HA-washing eliminate this interaction, possibly by allowing more flexibility in the QB site.  相似文献   

6.
The effect of the Photosystem II (PSII) inhibitors dichlorophenyldimethylurea (DCMU) and bromonitrothymol (BNT) on the rate of the high-light induced D1 protein turnover was studied in whole cells of two cyanobacterial strains Synechocystis PCC 6803 and Synechococcus PCC 7942. In Synechocystis the D1 degradation was slowed down to a similar extent in the presence of either inhibitor compared with control cells. This slower degradation corresponded with the retardation of Photosystem II photoinactivation (PSIIPI) measured as a decline of PS II activity in the illuminated cells treated with chloramphenicol (CAP). The ongoing D1 synthesis in the presence of both PS II inhibitors was confirmed by unchanging PS II activity and the steady-state level of D1 during illumination in the absence of CAP. In Synechococcus cells both DCMU and BNT blocked the turnover of the 'low-light' D1 form (D1:1) but did not prevent the exchange of the 'high-light' form D1:2 for the D1:1 form. The similar effect of both herbicides on the D1 exchange was in contrast with their influence on the rate of PSIIPI. While DCMU had a pronounced protective effect, BNT significantly increased the rate of PS II photodamage. The fast BNT-induced decline of PS II activity was also observed in Synechocystis cells treated with azide, an inhibitor of reactive oxygen species scavenging enzymes. Therefore, we assume that the distinct sensitivity of the two cyanobacterial strains to BNT can be caused by different content and/or activity of these enzymes in each strain.  相似文献   

7.
Gamma-secretase is a protease complex of four integral membrane proteins, with presenilin (PS) as the apparent catalytic component, and this enzyme processes the transmembrane domains of a variety of substrates, including the amyloid beta-protein precursor and the Notch receptor. Here we explore the mechanisms of structurally diverse gamma-secretase inhibitors by examining their ability to displace an active site-directed photoprobe from PS heterodimers. Most gamma-secretase inhibitors, including a potent inhibitor of the PS-like signal peptide peptidase, blocked the photoprobe from binding to PS1, indicating that these compounds either bind directly to the active site or alter it through an allosteric interaction. Conversely, some reported inhibitors failed to displace this interaction, demonstrating that these compounds do not interfere with the protease by affecting its active site. Differential effects of the inhibitors with respect to photoprobe displacement and in cell-based and cell-free assays suggest that these compounds are important mechanistic tools for deciphering the workings of this intramembrane-cleaving protease complex and its similarity to other polytopic aspartyl proteases.  相似文献   

8.
In our search for new natural photosynthetic inhibitors that could lead to the development of “green herbicides” less toxic to environment, the diterpene labdane-8α,15-diol (1) and its acetyl derivative (2) were isolated for the first time from Croton ciliatoglanduliferus Ort. They inhibited photophosphorylation, electron transport (basal, phosphorylating and uncoupled) and the partial reactions of both photosystems in spinach thylakoids. Compound 1 inhibits the photosystem II (PS II) partial reaction from water to Na+ Silicomolibdate (SiMo) and has no effect on partial reaction from diphenylcarbazide (DPC) to 2,6-dichlorophenol indophenol (DCPIP), therefore 1 inhibits at the water splitting enzyme and also inhibits PS I partial reaction from reduced phenylmetasulfate (PMS) to methylviologen (MV). Thus, it also inhibits in the span of P700 to Iron sulfur center X (FX). Compound 2 inhibits both, the PS II partial reactions from water to SiMo and from DPC to DCPIP; besides this, it inhibits the photosystem I (PS I) partial reaction from reduced PMS to MV. With these results, we concluded that the targets of the natural product 2 are located at the water splitting enzyme, and at P680 in PS II and at the span of P700 to FX in PS I. The results of compounds 1 and 2 on PS II were corroborated by chlorophyll a fluorescence.  相似文献   

9.
The cytosolic human carbonic anhydrase (hCA, EC 4.2.1.1) isozyme III (hCA III) has been cloned and purified by the GST-fusion protein method. Recombinant pure hCA III had the following kinetic parameters for the CO(2) hydration reaction at 20 degrees C and pH 7.5: k(cat) of 1.3 x 10(4) s(-1) and k(cat)/K(M) of 2.5 x 10(5) M(-1) s(-1), being a slower catalyst for the physiological reaction as compared to the genetically related cytosolic isoforms hCA I and II. An inhibition study with a library of sulfonamides and one sulfamate, some which are clinically used compounds, is reported. hCA III is less prone to be inhibited by these compounds as compared to hCA I and II for which many low nanomolar inhibitors were detected earlier. The best hCA III inhibitors were prontosil, sulpiride, indisulam, benzolamide, aminobenzolamide, and 4-amino-6-chloro-benzene-1,3-disulfonamide which showed K(I)s in the range of 2.3-18.1 microM. Clinically used compounds such as acetazolamide, methazolamide, ethoxzolamide, dorzolamide, brinzolamide, topiramate, zonisamide, celecoxib, and valdecoxib were less effective hCA III inhibitors, with affinities in the range of 154-2200 microM. This is the first study in which low micromolar hCA III inhibitors are reported.  相似文献   

10.
The interaction of five human carbonic anhydrase (hCA, EC 4.2.1.1) isozymes, that is, hCA I, II, IV, V, and IX with a small library of phosphonic acids/organic phosphates, including methylphosphonic acid, MPA; phenylphosphonic acid, PPA; N-(phosphonoacetyl)-L-aspartic acid, PALA, methylene diphosphonic acid MDPA, the O-phosphates of serine (Ser-OP) and threonine (Thr-OP) as well as the antiviral phosphonate foscarnet has been studied. hCA I was activated by all these compounds, with the best activators being MPA and PPA (K(A)s of 0.10-1.20 microM). MPA and PPA were on the other hand nanomolar inhibitors of hCA II (K(I)s of 98-99 nM). PALA showed an affinity of 7.8 microM, whereas the other compounds were weak, millimolar inhibitors of this isozyme. The best hCA IV inhibitors were PALA (79 nM) and PPA (5.4 microM), whereas the other compounds showed K(I)s in the range of 0.31-5.34 mM. The mitochondrial isozyme was weakly inhibited by all these compounds (K(I)s in the range of 0.09-41.7 mM), similarly to the transmembrane, tumor-associated isozyme (K(I)s in the range of 0.86-2.25 mM). Thus, phosphonates may lead to CA inhibitors with selectivity against two physiologically relevant isozymes, the cytosolic hCA II or the membrane-bound hCA IV.  相似文献   

11.
Boichenko VA  Hou JM  Mauzerall D 《Biochemistry》2001,40(24):7126-7132
The volume and enthalpy changes for charge transfer in the 0.1-10 micros time window in photosynthetic reaction centers of the intact cells of Synechocystis PCC 6803 were determined using pulsed, time-resolved photoacoustics. This required invention of a method to correct for the cell artifact at the temperature of maximum density of water caused by the heterogeneous system. Cells grown under either white or red light had different PS I/PS II molar ratios, approximately 3 and approximately 1.7, respectively, but invariable action spectra and effective antenna sizes of the photosystems. In both cultures, the photoacoustic measurements revealed that their thermodynamic parameters differed strongly in the spectral regions of predominant excitation of PS I (680 nm) and PS II (625 nm). On correcting for contribution of the two photosystems at these wavelengths, the volume change was determined to be -27 +/- 3 and -2 +/- 3 A3 for PS I and PS II, respectively. The energy storage on the approximately 1 micros time scale was estimated to be 80 +/- 15% and 45 +/- 10% per trap in PS I and PS II, respectively. These correspond to enthalpies of -0.33 +/- 0.2 and -1 +/- 0.2 eV for the assumed formation of ion radical pairs P700+F(AB-) and Y(Z*)P680Q(A-), respectively. Taking the free energy of the above reactions as the differences of their redox potentials in situ, apparent entropy changes were estimated to be +0.4 +/- 0.2 and -0.2 +/- 0.2 eV for PS I and PS II, respectively. These values are similar to that obtained in vitro for the purified reaction center complexes on the microsecond time scale [Hou et al. (2001) Biochemistry 40, 7109-7116, 7117-7125]. The constancy of these thermodynamic values over a 2-fold change of the ratio of PS I/PS II is support for this method of in vivo analysis. Our pulsed PA method can correct the "cell" or heterogeneous artifact and thus opens a new route for studying the thermodynamics of electron transfer in vivo.  相似文献   

12.
《Mutation Research Letters》1995,346(3):135-144
Carcinogenic polycyclic aromatic hydrocarbons (PAHs) and nitropolycyclic aromatic hydrocarbons (nitro-PAHs) have been identified in airborne particulate organic matter extracts. The pollutant sources were generally contributed by motor vehicles and industrial activity. Massive quantities of urban solid wastes, containing plastic materials such as PVC, PET, PS, and PE, burnt in the open air in local garbage dumps are frequently found in developing countries. In this study, the smog particulates from the combustion of these synthetic polymers were produced in a laboratory combustion chamber. The mutagenicity of acetone extracts from the smog particulates was evaluated with Salmonella typhimurium TA98 and TA100 in the presence and absence of S9 mix. Four samples in TA98 exhibited higher mutagenicity than those in TA100. The greatest mutagenicity was observed from the extracts of particulates from combustion of PVC followed by that of PS, PET, and PE. To determine the major mutagenic compounds in these samples, mutagens were partially purified through TLC and their mutagenicity was monitored with TA98. 1-NP and DNPs in the above samples were also determined by HPLC. The amounts of 1-NP and DNPs generally corresponded with their mutagenicity. Higher levels of 1-NP and DNPS from the combustion of PVC, PET, and PS. the combustion of synthetic polymer wastes might be responsible for the presence of high levels of 1-NP and DNPs in Taiwan urban air.  相似文献   

13.
Silicomolybdate (SiMo) and its effects on thylakoids have been characterized to evaluate its use as a probe for Photosystem II (PS II). It can accept electrons at two places in the electron transport chain: one at PS II and the other at PS I. In the presence of 1 M 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone (DBMIB) only the site at PS II is available. It is suggested that SiMo must disp;ace bicarbonate from its binding site to be able to function as an electron acceptor. This displacement is non-competitive. The binding of SiMo is inhibited differentially by PS II inhibitors: dinoseb>ioxynil> diuron. This difference is determined by the different positions of the inhibitors within the QB binding niche and their interaction with bicarbonate. The experimental results show that the SiMo-binding niche is located between the parallel helices of the D1 and D2 proteins of PS II, close to the non-heme iron. We conclude that SiMo is an electron acceptor with unique characteristics useful as a probe of the acceptor side of PS II.  相似文献   

14.
T Noguchi  Y Inoue  X S Tang 《Biochemistry》1999,38(31):10187-10195
Fourier transform infrared (FTIR) signals of a histidine side chain were identified in flash-induced S(2)/S(1) difference spectra of the oxygen-evolving complex (OEC) of photosystem II (PS II) using PS II membranes from globally (15)N-labeled spinach and PS II core complexes from Synechocystis cells in which both the imidazole nitrogens of histidine were selectively labeled with (15)N. A negative band at 1113-1114 cm(-1) was downshifted by 7 cm(-1) upon both global (15)N-labeling and selective [(15)N]His labeling, and assigned to the C-N stretching mode of the imidazole ring. This band was unaffected by H-D exchange in the PS II preparations. In addition, several peaks observed at 2500-2850 cm(-1) all downshifted upon global and selective (15)N-labeling. These were ascribed to Fermi resonance peaks on a hydrogen-bonding N-H stretching band of the histidine side chain. FTIR measurements of model compounds of the histidine side chain showed that the C-N stretching band around 1100 cm(-)(1) can be a useful IR marker of the protonation form of the imidazole ring. The band appeared with frequencies in the following order: Npi-protonated (>1100 cm(-1)) > imidazolate > imidazolium > Ntau-protonated (<1095 cm(-1)). The frequency shift upon N-deuteration was occurred in the following order: imidazolium (15-20 cm(-1)) > Ntau-protonated (5-10 cm(-1)) > Npi-protonated approximately imidazolate ( approximately 0 cm(-1)). On the basis of these findings together with the Fermi resonance peaks at >2500 cm(-1) as a marker of N-H hydrogen-bonding, we concluded that the histidine residue in the S(2)/S(1) spectrum is protonated at the Npi site and that this Npi-H is hydrogen bonded. This histidine side chain probably ligated the redox-active Mn ion at the Ntau site, and thus, oxidation of the Mn cluster upon S(2) formation perturbed the histidine vibrations, causing this histidine to appear in the S(2)/S(1) difference spectrum.  相似文献   

15.
Chemical modification of bovine prothrombin fragment 1 according to the procedure of D. J. Welsch and G. L. Nelsestuen (1988) [Biochemistry 27, 4946-4952 and ealier papers] provided a series of fragment 1 derivatives in which various nitrogen-containing side chains were N-acetylated and/or N-2,4,6-trinitrophenylated. In addition the des-[Ala-1,Asn-2]- and des-[Ala-1,Asn-2,Lys-3]-fragment 1 derivatives were prepared by limited enzymatic hydrolysis of fragment 1 using cathepsin C and plasmin, respectively. Quantitative studies on the Ca(II) binding of these proteins have been accomplished using 45Ca(II) equilibrium dialysis. Binding of these fragment 1 derivatives to phosphatidylserine/phosphatidylcholine (PS/PC) vesicles (25:75) in the presence of Ca(II) ions has been studied using the light-scattering technique. Acylation of the 5 lysine residues of fragment 1 by the action of acetic anhydride (500-fold molar excess) in the presence of 75 mM Ca(II), pH 8.0, results in loss of positive cooperativity in Ca(II) binding (Scatchard plot) and an increase in the number of Ca(II) ions bound. The Ca(II)-dependent PS/PC binding of the acylated protein is reduced. Removal of 2 and 3 residues from the amino terminus likewise leads to loss of positive cooperativity in Ca(II) binding and reduced binding affinity to PS/PC vesicles. The important role of the amino-terminal 1-10 sequence is discussed. We conclude that positive cooperativity in Ca(II) binding is not a prerequisite for the Ca(II)-dependent binding of bovine prothrombin fragment 1 to PS/PC vesicles.  相似文献   

16.
A series of 2-(hydrazinocarbonyl)-3-aryl-1H-indole-5-sulfonamides possessing various 2-, 3- or 4- substituted phenyl groups with methyl-, halogeno- and methoxy-functionalities, or a perfluorophenyl moiety, has been derivatized by reaction with 2,4,6-trimethylpyrylium perchlorate. The new sulfonamides were evaluated as inhibitors of four mammalian carbonic anhydrase (CA, EC 4.2.1.1) isoforms, that is, CA I, II (cytosolic), CA IX and XII (transmembrane, tumor-associated forms). Excellent inhibitory activity was observed against hCA IX with most of these sulfonamides, and against hCA XII with some of the new compounds. These compounds were generally less effective inhibitors of hCA II. Being membrane impermeant, these positively-charged sulfonamides are interesting candidates for targeting the tumor-associated CA IX and XII, as possible diagnostic tools or therapeutic agents.  相似文献   

17.
Bacterial neuraminidase has been highlighted as a key enzyme for pathogenic infection and sepsis. Six pterocarpans displaying significant levels of neuraminidase inhibitory activity were isolated from the root bark of Lespedeza bicolor. The isolated compounds were identified as three new pterocarpans (1-3) together with known compounds erythrabyssin II (4), lespebuergine G4 (5), and 1-methoxyerythrabyssin II (6). The new compounds were characterized as bicolosin A (1), bicolosin B (2), and bicolosin C (3). All compounds inhibited bacterial neuraminidase in a dose-dependent manner with significant inhibition (IC(50)=0.09-3.25 μM). All neuraminidase inhibitors screened were found to exhibit noncompetitive kinetics. The three most potent neuraminidase inhibitors (1, 3 and 6) feature a methoxy substitution on C-1.  相似文献   

18.
Abstract

Reaction of three aromatic sulfonamides possessing a primary amino group, i.e., sulfanilamide, homosulfanilamide and p-aminoethyl-benzenesulfonamide with heterocyclic and aromatic aldehydes afforded a series of Schiff bases. Metal complexes of some of these Schiff bases with divalent transition ions such as Zn(II), Cu(II), Co(II) and Ni(II) have also been obtained. The new compounds were assayed as inhibitors of three isozymes of carbonic anhydrase (CA). Several of the new compounds showed a modest selectivity for the membrane-bound (bovine) isozyme CA IV (bCA IV) as compared to the cytosolic human isozymes hCA I and II, in contrast to classical inhibitors which generally possess a 17-33 times lower affinity for bCA IV. This greater selectivity toward bCA IV is due mainly to a slightly decreased potency against hCA II relative to classical inhibitors. However, metal complexes of these Schiff bases possessed an increased affinity for hCA II, being less inhibitory against bCA IV. The first type of compounds reported here (i.e., the Schiff bases of aromatic sulfonamides with heterocyclic aldehydes) might thus lead to the development of low molecular weight isozyme specific CA IV inhibitors. The difference in affinity for the three isozymes of the inhibitors reported by us here is tentatively explained on the basis of recent X-ray crystallographic studies of these isozymes and their adducts with substratesiinhibitors  相似文献   

19.
Novel dioxoacridine sulfonamide compounds were synthesized from reaction of cyclic 1,3-diketones, sulfanilamide (4-amino benzene sulfonamide) and aromatic aldehydes. The structures of these compounds were confirmed by using spectral analysis (IR, H-NMR, (13)C-NMR, and mass). Human carbonic anhydrase isoenzymes (hCA I and hCA II) were purified from erythrocyte cells by affinity chromatography. The inhibitory effects of sulfanilamide, acetazolamide (AAZ), and newly synthesized sulfonamides on hydratase and esterase activities of these isoenzymes have been studied in vitro. The IC(50) values of compounds for esterase activity are 0.71-0.11 μM for hCA I and 0.45-0.12 μM for hCA II, respectively. The K(i) values of these inhibitors were determined as 0,38-0,008 μM for hCA I and 0,19-0,001 μM for hCA II, respectively.  相似文献   

20.
以“津春4号”黄瓜为试材,通过测定黄瓜叶片叶绿素荧光快速诱导动力学曲线和对820 nm光的吸收曲线,结合叶绿素荧光淬灭分析,研究低温光胁迫(4℃,200 μmol·m-2·s-1)6 h后,黄瓜叶片在常温(25℃)不同光强(0、15、200μmol·m-2·s-1)下PS Ⅰ和PS Ⅱ活性的恢复,以及恢复过程中PS Ⅰ与PS Ⅱ的相互作用.结果表明:低温光胁迫6h后,PS Ⅰ和PS Ⅱ发生不同程度的光抑制.在常温恢复阶段,PS Ⅱ活性快速恢复且对光强不敏感;PS Ⅰ活性在弱光下(15 μmol·m-2·s-1)快速恢复,在较强光(200 μmol·m-2·s-1)下恢复较慢.在低温光抑制恢复过程中,常温下PS Ⅱ活性恢复较快可能导致PS Ⅱ向PS Ⅰ的线性电子传递过快,进而抑制PS Ⅰ的活性恢复.因此,在进行黄瓜抗冷性育种时,不应该仅追求较高的PS Ⅱ抗性和较快的PS Ⅱ恢复速度,还应该注意两个光系统活性的协调.在生产中,应当在低温逆境发生及其之后较长一段时间内采取措施降低叶表面光照强度,以利于对植株光合机构的保护和光合活性的恢复.  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