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.
Human serum apotransferrin (hTF) binds to Zr(IV) slowly in the presence of nitrilotriacetate (NTA), citrate or ethylenediaminetetraacetate (EDTA) as donor ligands. For Zr(NTA)(2)(2-) as donor, equilibrium was reached in ca. 2 h (pH 7.4, 298 K, 10 mM Hepes, 5 mM bicarbonate) and full loading of the N- and C-lobe sites was achievable to give Zr(2)-hTF. (13)C NMR data suggest that carbonate can bind as a synergistic anion. (1)H and 2D [(1)H,(13)C] (using epsilon-[(13)C]Met-hTF) NMR studies show that there is little lobe-selectively in the order of Zr(IV) uptake. Fe(III) displaced Zr(IV) from the C-lobe of Zr(2)-hTF first, followed by the N-lobe. However, in the presence of a large excess of NTA, Zr(IV) binds to the N-lobe of holo-hTF (Fe(2)-hTF) first followed by the C-lobe. The (1)H and (13)C NMR chemical shift changes for epsilon-[(13)CH(3)] of Met464, which is close to the C-lobe site, are quite distinct from those observed previously for Al(III), Fe(III), Ti(IV), Ga(III) and Bi(III) binding to hTF, suggesting that Zr(IV) binding may not induce lobe closure [as observed previously for Hf(IV)]. This may affect receptor recognition and play a role in the different biological behaviour of Zr(IV) compared to Ti(IV).  相似文献   

2.
Transferrin, the human iron transport protein, binds Ti(IV) even more tightly than it binds Fe(III). However, the fate of titanium bound to transferrin is not well understood. Here we present results which address the fate of titanium once bound to transferrin. We have determined the redox potentials for a series of Ti(IV) complexes and have used these data to develop a linear free energy relationship (LFER) correlating Ti(IV) ? Ti(III) redox processes with Fe(III) ? Fe(II) redox processes. This LFER enables us to compare the redox potentials of Fe(III) complexes and Ti(IV) complexes that mimic the active site of transferrin and allows us to predict the redox potential of titanium-transferrin. Using cyclic voltammetry and discontinuous metalloprotein spectroelectrochemistry (dSEC) in conjunction with the LFER, we report that the redox potential of titanium-transferrin is lower than − 600 mV (lower than that of iron-transferrin) and is predicted to be ca. − 900 mV vs. NHE (normal hydrogen electrode). We conclude that Ti(IV)/Ti(III) reduction in titanium-transferrin is not accessible by biological reducing agents. This observation is discussed in the context of current hypotheses concerning the role of reduction in transferrin mediated iron transport.  相似文献   

3.
Nitrogenase-catalyzed reactions using Ti(III) were examined under a wide variety of conditions to determine the suitability of Ti(III) to serve as a general nitrogenase reductant. Solutions prepared from H2-reduced TiCl3, aluminum-reduced TiCl3, TiCl2, evaporated TiCl3 from an HCl, solution, and TiF3 were evaluated as reductants. Three general types of reactivity were observed. The first showed that, below Ti(III) concentrations of about 0.50 mM, nitrogenase catalysis utilized Ti(III) in a first-order reaction. The second showed that, above 0.50 mM, the rate of nitrogenase catalysis was zero order in Ti(III), indicating the enzyme was saturated with this reductant. Above 2.0-5.0 mM, nitrogenase catalysis was inhibited by Ti(III) depending on the titanium source used for solution preparation. This inhibition was investigated and found to be independent of the buffer type and pH, while high salt and citrate concentrations caused moderate inhibition. [Ti(IV)] above 2.0-3.0 mM and [Ti(III)] above about 5.0 mM were inhibitory. ATP/2e values were 4-5 for [Ti(III)] at or below 1.0-2.0 mM, 2.0 from 5.0 to 7.0 mM Ti(III) where nitrogenase is not inhibited, and 2.0 above 7.0 mM Ti(III) where severe inhibition occurs. For nitrogenase-catalyzed reactions using Ti(III) as reductant, the potential of the solution changes with time as the Ti(III)/Ti(IV) ratio changes. From the change in the rate of product formation (Ti(III) disappearance) with change in solution potential, the rate of nitrogenase catalysis was determined as a function of solution potential. From such experiments, a midpoint turnover potential of -480 mV was determined for nitrogenase catalysis with an associated n = 2 value.  相似文献   

4.
Abstract Reactions between the anticancer drug titanocene dichloride (Cp2TiCl2) and various nucleotides and their constituents in aqueous solution or N,N-dimethylformamide (DMF) have been investigated by 1H and 31P NMR spectroscopy and in the solid state by IR spectroscopy. In aqueous solution over the pH* (pH meter reading in D2O) range 2.3-6.5, CMP forms one new species with Ti(IV) bound only to the phosphate group. In acidic media at pH*<4.6, three species containing titanocene bound to the phosphate group of dGMP, AMP, dTMP and UMP are formed rapidly. The bases also appear to influence titanocene binding. Only one of these Ti(IV)-bound species can be detected in the pH* range of 4.6-6.5 in each case. The order of reactivity towards Cp2TiCl2(aq) at pH* ca. 3 is GMP>TMP approximately AMP > CMP. At pH* > 7.0, hydrolysis of Cp2TiCl2 predominated and little reaction with the nucleotides was observed. Binding of deoxyribose 5'-phosphate and 4-nitrophenyl phosphate to Cp2TiCl2(aq) via their phosphate groups was detected by 31P NMR spectroscopy, but no reaction between Cp2TiCl2(aq) and deoxyguanosine, 9-ethylguanine or deoxy-D-ribose was observed in aqueous solution. The nucleoside phosphodiesters 3',5'-cyclic GMP and 2',3'-cyclic CMP did not react with Cp2TiCl2(aq) in aqueous solution; however, in the less polar solvent DMF, 3',5'-cyclic GMP coordination to [Cp2Ti]2+ via its phosphodiester group was readily observed. Binding of titanocene to the phosphodiester group of the dinucleotide GpC was also observed in DMF by 31P NMR. The nucleoside triphosphates ATP and GTP reacted more extensively with Cp2TiCl2(aq) than their monophosphates; complexes with bound phosphate groups were formed in acidic media and to a lesser extent at neutral pH. Cleavage of phosphate bonds in ATP (and GTP) by Cp2TiCl2(aq) to form inorganic phosphate, AMP (or GMP) and ADP (or GDP) was observed in aqueous solutions. In addition, titanocene binding to ATP was not inhibited by Mg(II), but the ternary complex titanocene-ATP-Mg appeared to form. These reactions contrast markedly with those of the drug cisplatin, which binds predominantly to the base nitrogen atoms of nucleotides and only weakly to the phosphate groups. The high affinity of Ti(IV) for phosphate groups may be important for its biological activity.  相似文献   

5.
The ferric form of the N-lobe of human serum transferrin (Fe(III)-hTF/2N) has been expressed at high levels in Pichia pastoris. The Fe(III)-hTF/2N was crystallized in the space group P41212, and X-ray crystallography was used to solve the structure of the recombinant protein at 2.5 A resolution. This represents only the second P. pastoris-derived protein structure determined to date, and allows the comparison of the structures of recombinant Fe(III)-hTF/2N expressed in P. pastoris and mammalian cells with serum-derived transferrin. The polypeptide folding pattern is essentially identical in all of the three proteins. Mass spectroscopic analyses of P. pastoris- hTF/2N and proteolytically derived fragments revealed glycosylation of Ser-32 with a single hexose. This represents the first localization of an O-linked glycan in a P. pastoris-derived protein. Because of its distance from the iron-binding site, glycosylation of Ser-32 should not affect the iron-binding properties of hTF/2N expressed in P. pastoris, making this an excellent expression system for the production of hTF/2N.  相似文献   

6.
《FEBS letters》1999,442(2-3):157-161
The reaction of human serum apotransferrin with titanium(IV) citrate under physiological conditions results in the formation of a specific bis-titanium(IV) transferrin adduct (Ti2Tf hereafter) with two titanium(IV) ions loaded at the iron binding sites. The same specific Ti2Tf complex is formed by reacting apotransferrin with titanium(III) chloride and exposing the sample to air. The derivative thus obtained was characterized by spectroscopic techniques, including absorption, UV difference, circular dichroism and 13C NMR spectroscopies, and shown to be stable within the pH range 5.5–9.0. Surprisingly, the reaction of apoTf with titanium(IV) nitrilotriacetate (NTA) does not lead to formation of appreciable amounts of Ti2Tf, even after long incubation times, although some weak interactions of Ti(IV)NTA with apoTf are spectroscopically detected. Implications of the present results for a role of transferrin in the uptake, transport and delivery of soluble titanium(IV) compounds under physiological conditions are discussed.  相似文献   

7.
Functionalization of cyclopentadienyl (Cp) ligands and incorporation of these into a Ti(IV) center require careful design and selection of the appropriate synthetic routes to obtain the desired product in reasonably good yields. As part of our research efforts in the area of titanocene antitumor agents, we have revisited the synthesis of Cp rings with electron-withdrawing groups and their corresponding titanocene dichlorides, (Cp-R)2TiCl2 and (Cp-R)CpTiCl2, where R is CO2CH3 and CO2CH2CH3. These complexes were characterized by elemental analysis and 1H and 13C NMR and IR spectroscopies. This report presents the first detailed synthetic route for (Cp-CO2CH2CH3)CpTiCl2 and provides an alternate route for synthesis of (Cp-R)2TiCl2 complexes. The ability of these complexes to deliver Ti(IV) to apotransferrin was investigated to elucidate how the functionalized Cp ligands affect the titanium intake by apotransferrin. The subject complexes transfer Ti(IV) to human apotransferrin, loading both N- and C-lobes. The antitumor activity of these complexes against HT-29 cancer colon cells was determined using a 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide assay. Carboethoxy Cp functionalization results in complexes with a toxicity comparable to that of titanocene dichloride. The carbomethoxy-functionalized complexes proved to be nonactive at the time intervals studied here, regardless of their ability to donate the titanium atom to human apotransferrin.  相似文献   

8.
Titanium(II) solutions, prepared by dissolving titanium wire in triflic acid + HF, contain equimolar quantities of Ti(IV). Treatment of such solutions with excess Fe(III) or Ru(III) complexes yield Ti(IV), but reactions with Ti(II) in excess give Ti(III). Oxidations by (NH3)5Ru(III) complexes, but not by Fe(III) species, are catalyzed by titanium(IV) and by fluoride. Stoichiometry is unchanged. The observed rate law for the Ru(III)-Ti(II)-Ti(IV) reactions in fluoride media points to competing reaction paths differing by a single F, with both routes involving a Ti(II)-Ti(IV) complex which is activated by deprotonation. It is suggested that coordination of Ti(IV) to TiII(aq) minimizes the mismatch of Jahn-Teller distortions which would be expected to lower the Ti(II,III) self-exchange rate.  相似文献   

9.
The ability of antitumor active metallocenes Cp2MCl2, (M=Ti, V, Mo, Nb) and the biologically inactive derivative (MeCp)2TiCl2, to inhibit the relaxation of supercoiled plasmid DNA pBR322 by human topoisomerase II has been studied by gel electrophoresis. All metallocenes inhibit the enzyme with maximum inhibition observed at 2.0 mM (Cp2TiCl2), 3.0 mM (Cp2MoCl2), 0.2 mM (Cp2NbCl2), 0.25 mM (Cp2VCl2) and 2.0 mM (MeCpTiCl2). The implications for the mechanism of antitumor activity of the metallocene dihalides are discussed.  相似文献   

10.
As part of our research efforts in the area of titanium-based antitumor agents, we have investigated the cytotoxic activity of [Ti(4)(maltolato)(8)(mu-O)(4)], (Cp-R)(2)TiCl(2) and (Cp-R)CpTiCl(2) (R = CO(2)CH(3) and CO(2)CH(2)CH(3)), and three water-soluble titanocene-amino acid complexes-[Cp(2)Ti(aa)(2)]Cl(2) (aa = L: -cysteine, L: -methionine, and D: -penicillamine)-on the human colon adenocarcinoma cell line, HT-29. The capacity of [Ti(4)(maltolato)(8)(mu-O)(4)] to donate Ti(IV) to human apo-transferrin and its hydrolytic stability have been investigated and compared to the previously reported data on modified titanocenes with either hydrophilic ancillary ligands or the functionalized cyclopentadienyl ligands. Notably, the titanium-maltolato complex does not transfer Ti(VI) to human apo-transferrin at any time within the first seven days of its interaction, demonstrating the inert character of this species. Stability studies on these complexes have shown that titanocene complexes decompose at physiological pH while the [Ti(4)(maltolato)(8)(mu-O)(4)] complex is stable at this pH without any notable decomposition for a period of ten days. The antitumor activity of these complexes against colon cancer HT-29 cells was determined using an MTT cell viability assay at 72 and 96 h. The titanocene-amino acid and the (Cp-R)(2)TiCl(2)/(Cp-R)CpTiCl(2) (R = CO(2)CH(3)) complexes were not biologically active when human transferrin was absent; they also were inactive when human transferrin was present at dose-equivalent concentrations. (Cp-R)(2)TiCl(2) and (Cp-R)CpTiCl(2) (R = CO(2)CH(2)CH(3)) showed cytotoxic activity in HT-29 cells comparable to that which is displayed by titanocene dichloride. The titanium-maltolato complex had higher levels of cytotoxic activity than any other titanocene complex investigated. Transferrin may be important in protecting the titanium center from hydrolysis, but this may be achieved by selecting ligands that could result in hydrolytically stable, yet active, complexes.  相似文献   

11.
L Zhang  K Y Szeto  W B Wong  T T Loh  P J Sadler  H Sun 《Biochemistry》2001,40(44):13281-13287
Several bismuth compounds are currently used as antiulcer drugs, but the mechanism of action still remains unclear. The antimicrobial activity of Bi(III) complexes toward Gram-negative bacteria is reported to be dependent on the iron uptake system [Domenico, P., et al. (1996) J. Antimicrob. Chemother. 38, 1031-1040]. Electronic absorption and 13C NMR spectroscopic data show that Bi(III) binds to human lactoferrin at the specific Fe(III) sites along with either carbonate or oxalate as the synergistic anion. The uptake of Bi(III) by apo-hLF was rapid [minutes in 10 mM Hepes buffer and 5 mM bicarbonate (pH 7.4)], and almost equal in both lobes. The presence of ATP facilitates the release of Bi(III) from the Bi2-hLF complex when the pH is lowered. The Bi2-hLF complex blocked the uptake of the radiolabeled 59Fe-hLF complex into rat IEC-6 cells. Surprisingly, apo-hLF (but not apotransferrin) was almost as effective in blocking 59Fe uptake as bismuth-loaded lactoferrin. These results suggest that Bi(III)-loaded hLF might be recognized by the lactoferrin receptor and be taken up into cells.  相似文献   

12.
F J Carver  E Frieden 《Biochemistry》1978,17(1):167-172
The release of iron from transferrin was investigated by incubating the diferric protein in the presence of potential iron-releasing agents. The effective chemical group appears to be pyrophosphate, which is present in blood cells as nucleoside di- and triphosphates, notably adenosine triphosphate (ATP). An alternative structure with comparable activity is represented by 2,3-diphosphoglycerate. Neither 1 mM adenosine monophosphate (AMP) nor 1 mM orthophosphate released iron from transferrin. The ATP-induced iron-releasing activity was dependent on weak acidic conditions and was sensitive to temperature and sodium chloride concentration. The rate of iron release rapidly increased as transferrin was titrated with HCl from pH 6.8 to 6.1 in the presence of 1 mM ATP and 160 mM NaCl at 20 degrees C. Iron release from transferrin without ATP was observed below pH 5.5. Ascorbate (10(-4) M) reduced Fe(III), but only after iron release from transferrin by a physiological concentration of ATP. A proposal for the mechanism of iron release from transferrin by ATP and the utilization of reduced iron by erythroid cells is described.  相似文献   

13.
This study was undertaken to investigate the conformational states of the two metal sites in the human serum transferrin molecule. The 9.2 GHz electron paramagnetic resonance spectra of frozen solutions of divanadyl(IV) transferrin consist of a superposition of two sets of resonances, A and B, due to the magnetically nonequivalent binding environments of the VO2+ ion. Examination of the intensities of the A and B resonances as a function of pH from 6.0 to 10.7 reveals that they arise from two conformational states of the metal sites in which the geometrical arrangement and/or identity of one or more ligands in the first coordination sphere are different. From pH 7.5 to 9.0, the metal sites exist in A and B conformations but above pH 9.0 the A conformation. This transformation is coupled to the ionization of an apparently noncoordinating protein functional group with a pK - 10.0 +/- 0.1. Below pH 7.0, binding in the B conformation is rapidly lost, driven in part by the protonation of a functional group, possibly the anion, with a pK - 6.6 +/- 0.1. In 90% D2O, this pK is elevated to 7.8 +/- 0.1. At pH 6.0 in H2O, essentially one VO2+ ion remains bound to the protein with the metal site in the A conformation. Experiments with mixed VO2+ -Fe3+ transferrin complexes indicate that the same may be true of Fe3+. At pH 10.7, a new set of VO2+ resonances, labeled C, are observed; they possibly arise from a third conformation of the metal site. One bicarbonate or corbonate is required per VO2+ ion bound to the protein. 2.7 H+ are released per VO2+ bound in either the A or B conformations. The above results are discussed in terms of the "equivalence" and "nonequivalence" of the metal sites.  相似文献   

14.
The action mechanism of vanadocene dichloride, Cp2VCl2 (Cp=eta5-C5H5), has been investigated by interaction with human serum transferrin for its promising antitumor activities. Our results have shown that Cp2VCl2 binds to transferrin and form a new complex, and the calculated apparent association constant is 1.37 x 10(5)M(-1) from the fluorescence quenching. Simultaneously, the variation of the secondary structure of transferrin occurs, most probably due to the coordination of the amino residues of protein with VIV. It was evidenced that Cp is released free in solution after VIV binding to transferrin by 1H NMR measurements. These results have shown that Cp2VCl2 forms a complex with transferrin, which may provide a possible pathway in the transport and targeted delivery of the antitumor agent.  相似文献   

15.
The binding of iron (Fe) to human serum transferrin (Tf) was analyzed with an HPLC system equipped with an anion exchange column and directly connected with a high-resolution inductively coupled plasma mass spectrometer for metal detection. The (56)Fe level in the eluate was monitored at resolution m/Deltam=3000. Two monoferric Tfs were assigned based on the results of urea-PAGE and desferrioxamine experiments. When Fe was added as Fe-citrate stepwise to an apo-Tf solution in the presence of bicarbonate, the N-lobe site was the preferential Fe-binding site, while the C-lobe site was preferred in the absence of bicarbonate. In both cases, the Fe-peak areas of the preferential site and Fe(2)-Tf increased up to an Fe/Tf molar ratio of 1, and then the peak area of the monoferric Tf decreased while the peak area of Fe(2)-Tf increased. When the Fe/Tf molar ratio was below 1, the amount of Fe bound to the lobe with a weaker affinity was higher in Fe(2)-Tf than in the monoferric Tf in each case. Namely, Fe(2)-Tf was the preferential binding state of Fe to human serum Tf. The preference is reasonable for transferring Fe ions effectively to Tf-receptors.  相似文献   

16.
J C Cannon  N D Chasteen 《Biochemistry》1975,14(21):4573-4577
Vanadyl ion, VO(IV), has been used as an electron paramagnetic resonance (EPR) spin label to study the metal-binding properties of human serum transferrin in the presence of bicarbonate. Iron-saturated transferrin does not bind the vanadyl ion. Room temperature titrations of apotransferrin with VO(IV) as monitored by EPR indicate the extent of binding to be pH dependent, with a full 0.2 VO(IV) ions per transferrin molecule bound at pH 7.5 and 9, but only about 1.2 VO(IV) ions bound at pH 6. The EPR spectra of frozen solutions with or without 0.1 M NaCUO4 at 77 K show that there are two spectroscopically nonequivalent binding sites (A and B) with a slight difference in binding constants. One site (A site) exhibits essentially constant binding capacity in the pH range 6-9, but the other (B site) becomes less avialable as the pH is reduced below 7. Results with mixed Fe(III)-VO(IV) transferrin complexes suggest that iron shows a slight tendency to bind at the B site over the A site pH 7.5 and 9.0. Only the B site in both vanadyl and iron transferrins is perturbed by the presence of perchlorate.  相似文献   

17.
The complexes of the type Cp2M(3-TC)Cl, Cp2M(3-TC)2, Cp2M(3-TA)Cl, Cp2M(3-TA)2, Cp2M(2-TB)Cl, Cp2M(2-TB)2 [where Cp = cyclopentadienyl, M = Zr or Ti] were synthesized by the reactions of dichlorobis(cyclopentadienyl)zirconium(IV) and dichlorobis(cyclopentadienyl)titanium(IV) with 3-thiophenecarboxylic acid (3-TCH), 3-thiopheneacetic acid (3-TAH) and 2-thiophenebutyric acid (2-TBH) respectively in different stoichiometric ratios. The new complexes were characterized by their elemental analysis, 1H NMR, IR, and electronic spectral data.  相似文献   

18.
Vanadium (V) is an essential metal for mammals and has different valence states. In blood, V is bound to serum transferrin (Tf), a glycoprotein which has two metal-binding sites, and carbonate is generally required for the binding. In this study, the binding patterns of V(III), V(IV), and V(V) to human serum Tf (hTf) were analyzed using an HPLC system equipped with an anion-exchange column and directly connected to a high-resolution inductively coupled plasma-mass spectrometer for metal detection (51V). In affinity to hTf, the three ions were ranked V(III)>V(IV)>V(V) in the presence of bicarbonate and V(III) reverse congruent V(IV)>V(V) in the absence. Intermediates in the "open forms" binding to the respective sites were detected at the initial stage. V(IV) and V(V) were bound to the N-lobe site in the "closed form" and "open form," respectively. In the absence of bicarbonate, V ions with respective valence states were bound to hTf in the "open form." In terms of binding to hTf, tri-valent V was most favorable in the presence of bicarbonate.  相似文献   

19.
Serum transferrin (Tf) is an iron binding glycoprotein that plays a central role in the metabolism of this essential metal but it also binds other metal ions. Four main transferrin forms containing different iron binding states can be distinguished in human serum samples: monoferric (C-site or N-site), holotransferrin (with two Fe atoms) and apotransferrin (with no metal). Recently, it has been reported that Tf binds also Ti even more tightly than does Fe, in artificially Ti(iv) spiked solutions. However, very limited work has been done on the Ti binding to Tf at physiological concentrations in patients carrying intramedullary Ti nails. Here we report the chemical association of Ti to Tf "in vivo" under different chromatographic conditions by elemental mass spectrometry using double focusing inductively coupled plasma (DF-ICP-MS) as detector. For the separation of the Ti/Fe-Tf forms different gradient conditions have been explored. The observed results reveal that human serum Ti (from patients carrying intramedullary Ti nails) is uniquely associated to the N-lobe of Tf. The investigation of the influence of sialic acid in the carbohydrate chain of human serum Tf, studied by incubating the protein with neuraminidase (sialidase) to obtain the monosialilated species, revealed that the binding affinity of Ti was similar for monosialo-Tf and for native-Tf and occurs in the N-lobe. These results suggest that the species Fe(C)Ti(N)-TF might provide a route for Ti entry into cells via the transferrin receptors after the release of the metal from its implants.  相似文献   

20.
The differential properties of anion-mediated Fe(3+) release between the N- and C-lobes of transferrins have been a focus in transferrin biochemistry. The structural and kinetic characteristics for isolated lobe have, however, been documented with the N-lobe only. Here we demonstrate for the first time the quantitative Fe(3+) release kinetics and the anion-binding structure for the isolated C-lobe of ovotransferrin. In the presence of pyrophosphate, sulfate, and nitrilotriacetate anions, the C-lobe released Fe(3+) with a decelerated rate in a single exponential progress curve, and the observed first order rate constants displayed a hyperbolic profile as a function of the anion concentration. The profile was consistent with a newly derived single-pathway Fe(3+) release model in which the holo form is converted depending on the anion concentration into a "mixed ligand" intermediate that releases Fe(3+). The apo C-lobe was crystallized in ammonium sulfate solution, and the structure determined at 2.3 A resolution demonstrated the existence of a single bound SO(4)(2-) in the interdomain cleft, which interacts directly with Thr(461)-OG1, Tyr(431)-OH, and His(592)-NE2 and indirectly with Tyr(524)-OH. The latter three groups are Fe(3+)-coordinating ligands, strongly suggesting the facilitated Fe(3+) release upon the anion occupation at this site. The SO(4)(2-) binding structure supported the single-pathway kinetic model.  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