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.
The trinuclear platinum compound [{trans‐PtCl(NH3)2}2(μ‐trans‐Pt(NH3)2{NH2(CH2)6NH2}2)]4+ (BBR3464) belongs to the polynuclear class of platinum‐based anticancer agents. These agents form in DNA long‐range (Pt,Pt) interstrand cross‐links, whose role in the antitumor effects of BBR3464 predominates. Our results show for the first time that the interstrand cross‐links formed by BBR3464 between two guanine bases in opposite strands separated by two base pairs (1,4‐interstrand cross‐links) exist as two distinct conformers, which are not interconvertible, not only if these cross‐links are formed in the 5′‐5′, but also in the less‐usual 3′‐3’ direction. Analysis of the conformers by differential scanning calorimetry, chemical probes of DNA conformation, and minor groove binder Hoechst 33258 demonstrate that each of the four conformers affects DNA in a distinctly different way and adopts a different conformation. The results also support the thesis that the molecule of antitumor BBR3464 when forming DNA interstrand cross‐links may adopt different global structures, including different configurations of the linker chain of BBR3464 in the minor groove of DNA. Our findings suggest that the multiple DNA interstrand cross‐links available to BBR3464 may all contribute substantially to its cytotoxicity.  相似文献   

2.
Metal coordination to N9‐substituted adenines, such as the model nucleobase 9‐methyladenine (9MeA), under neutral or weakly acidic pH conditions in water preferably occurs at N1 and/or N7. This leads, not only to mononuclear linkage isomers with N1 or N7 binding, but also to species that involve both N1 and N7 metal binding in the form of dinuclear or oligomeric species. Application of a trans‐(NH3)2PtII unit and restriction of metal coordination to the N1 and N7 sites and the size of the oligomer to four metal entities generates over 50 possible isomers, which display different feasible connectivities. Slowly interconverting rotamers are not included in this number. Based on 1H NMR spectroscopic analysis, a qualitative assessment of the spectroscopic features of N1,N7‐bridged species was attempted. By studying the solution behavior of selected isolated and structurally characterized compounds, such as trans‐[PtCl(9MeA‐N7)(NH3)2]ClO4 ? 2H2O or trans,trans‐[{PtCl(NH3)2}2(9MeA‐N1,N7)][ClO4]2 ? H2O, and also by application of a 9MeA complex with an (NH3)3PtII entity at N7, [Pt(9MeA‐N7)(NH3)3][NO3]2, which blocks further cross‐link formation at the N7 site, basic NMR spectroscopic signatures of N1,N7‐bridged PtII complexes were identified. Among others, the trinuclear complex trans‐[Pt(NH3)2{μ‐(N1‐9MeA‐N7)Pt(NH3)3}2][ClO4]6 ? 2H2O was crystallized and its rotational isomerism in aqueous solution was studied by NMR spectroscopy and DFT calculations. Interestingly, simultaneous PtII coordination to N1 and N7 acidifies the exocyclic amino group of the two 9MeA ligands sufficiently to permit replacement of one proton each by a bridging heterometal ion, HgII or CuII, under mild conditions in water.  相似文献   

3.
In title an­hydro­us catena‐poly­[[trans‐bis­(ethane‐1,2‐di­amine‐κ2N,N′)copper(II)]‐μ‐di­thionato‐κ2O:O′], [Cu(S2O6)(C2H8N2)2]n or [{H2N(CH2)2NH2}2Cu(O·O2SSO2·O)], successive Cu atoms are bridged by a single doubly charged di­thionate group, forming a one‐dimensional polymer with inversion centres at the metal atoms and the mid‐point of the S—S bond [Cu—O = 2.5744 (15) Å]. In title (hydrated) trans‐di­aqua­bis­(propane‐1,3‐di­amine‐κ2N,N′)copper(II) di­thionate, [Cu(C3H10N2)2(H2O)2](S2O6) or [{H2N(CH2)3NH2}2Cu(OH2)2](S2O6), both ions have imposed 2/m symmetry. The `axial' anion components are displaced by a pair of water ligands [Cu—O = 2.439 (3) Å], the shorter Cu—O distance being compensated by the lengthened Cu—N distance [2.0443 (18), cf. 2.0100 (13) and 2.0122 (16) Å].  相似文献   

4.
The substitution reactions of the complexes [{trans‐Pt(NH3)2H2O}2(μ‐1,4‐diaminobutane)]4+ ( I ), [{trans‐Pt(NH3)2H2O}2(μ‐1,6‐diaminohexane)]4+ ( II ), and [{trans‐Pt(NH3)2H2O}2(μ‐1,8‐diaminooctane)]4+ ( III ), with nucleophiles L‐cysteine (L‐Cys), glutathione (GSH), guanosine‐5′‐monophosphate (5′‐GMP), L‐histidine (L‐His), and pyridine were studied in 0.1 M NaClO4 aqueous solutions at pH = 2.5. The substitutions were studied under pseudo‐first‐order conditions as a function of concentration and temperature using UV–vis spectrophotometry. At three different temperatures (288, 298, and 308 K) the reactions of the II and III complexes and 5′‐GMP were studied. The order of reactivity of study ligands is L‐Cys > GSH > 5′‐GMP > L‐His > pyridine and the order of reactivity of the complexes is I < II ≈ III . The obtained results indicate that the structure of the alkanediamine linker in the dinuclear Pt(II) complexes controls the substitution process. The negative values reported for entropy of activation confirmed the associative substitution mode. These results are discussed in order to find the connection between structure and reactivity of the dinuclear Pt(II) complexes.  相似文献   

5.
N,N′‐dioxide ligands such as 2, 2′‐bipyridine‐N,N‐dioxide (BPDO‐I) and 4, 4′‐bipyridine‐N,N‐dioxide (BPDO‐II) were used to trap the hydrated dimethyltin cations under controlled hydrolysis. The use of the chelating ligand BPDO‐I leads to the isolation of the discrete monocation [Me2Sn(BPDO‐I)(OH2)(NO3)]+[NO3] ( 2 ), whereas the linear ligand BPDO‐II directs the construction of cationic polymers, [{Me2Sn(OH2)2(μ‐BPDO‐II)}2+{NO3}2 · 2H2O]n ( 3· 2H2O) and [{Me2Sn(μ‐OH)(BPDO‐II)}22+{NO3}2 · H2O]n ( 4· H2O) under different reaction conditions.  相似文献   

6.
Molecular and Crystal Structure of Bis[chloro(μ‐phenylimido)(η5‐pentamethylcyclopentadienyl)tantalum(IV)](Ta–Ta), [{TaCl(μ‐NPh)Cp*}2] Despite the steric hindrance of the central atom in [TaCl2(NPh)Cp*] (Ph = C6H5, Cp* = η5‐C5(CH3)5), caused by the Cp* ligand, the imido‐ligand takes a change in bond structure when this educt is reduced to the binuclear complex [{TaCl(μ‐NPh)Cp*}2] in which tantalum is stabilized in the unusual oxidation state +4.  相似文献   

7.
Reacting stoichiometric amounts of 1‐(diphenylphosphino)ferrocene­carboxylic acid and [Ti(η5‐C5HMe4)22‐Me3SiC[triple‐bond]CSiMe3)] produced the title carboxyl­atotitanocene complex, [{μ‐1κ2O,O′:2(η5)‐C5H4CO2}{2(η5)‐C5H4P(C6H5)2}{1(η5)‐C5H(CH3)4}2FeIITiIII] or [FeTi(C9H13)2(C6H4O2)(C17H14P)]. The angle subtended by the Ti/O/O′ plane, where O and O′ are the donor atoms of the κ2‐carboxy­late group, and the plane of the carboxyl‐substituted ferrocene cyclo­penta­dienyl is 24.93 (6)°.  相似文献   

8.
The reaction of a new heterocyclic bidentate N containing spacer, (ligand) 5,5′‐methylenebis(pyridine) with ruthenium sulphoxide precursors resulted, dinuclear complexes. We herein report three formulations; [{cis,fac‐RuCl2(so)3}2(μ‐mbp)].3so; [{trans,mer‐RuCl2(so)32}2(μ‐mbp)].3so and [{trans‐RuCl4(so)}2(μ‐mbp)]2?[X]2+; where so = dimethyl‐sulfoxide/tetramethylenesulfoxide; mbp = 5,5′‐methylenebis(pyridine) and [X]+ = [(dmso)2H]+, Na+ or [(tmso)H]+. These complexes were characterized on the basis of elemental analyses, molar conductance measurement, magnetic susceptibility, FT‐IR, 1H‐NMR, 13C{1H}‐NMR, electronic spectroscopy and FAB‐Mass spectrometry. Catalytic activity of these complexes has been investigated in hydrolysis of benzonitrile. All the complexes exhibit good antibacterial activity against gram‐negative bacteria Escherichia coli in comparison to Chloramphenicol.  相似文献   

9.
The reaction of [{Ir(cod)(μ‐Cl)}2] and K2CO3 or of [{Ir(cod)(μ‐OMe)}2] alone with the non‐natural tetrapyrrole 2,2′‐bidipyrrin (H2BDP) yields, depending on the stoichiometry, the mononuclear complex [Ir(cod)(HBDP)] or the homodinuclear complex [{Ir(cod)}2(BDP)]. Both complexes react readily with carbon monoxide to yield the species [Ir(CO)2(HBDP)] and [{Ir(CO)2}2(BDP)], respectively. The results from NMR spectroscopy and X‐ray diffraction reveal different conformations for the tetrapyrrolic ligand in both complexes. The reaction of [{Ir(coe)2(μ‐Cl)}2] with H2BDP proceeds differently and yields the macrocyclic [4e?,2H+]‐oxidized product [IrCl2(9‐Meic)] (9‐Meic = monoanion of 9‐methyl‐9,10‐isocorrole), which can be addressed as an iridium analog of cobalamin.  相似文献   

10.
Synthesis and Molecular Structure of [{Cp′(μ‐η1 : η5‐C5H3Me)Mo(μ‐AlRH)}2] (Cp′ = C5H4Me, R = iBu, Et) [Cp′2MoH2] reacts with HAlR2 to give [{Cp′(μ‐η1 : η5‐C5H3Me)Mo(μ‐AlRH)}2] (Cp′ = C5H4Me, R = iBu ( 1 ), Et ( 2 )). Crystal structure determinations were carried out on [Cp′2MoH2] and 1 . 1 exhibits a direct Mo–Al bond (2.636(2) Å).  相似文献   

11.
On the Reactivity of Titanocene Complexes [Ti(Cp′)22‐Me3SiC≡CSiMe3)] (Cp′ = Cp, Cp*) towards Benzenedicarboxylic Acids Titanocene complexes [Ti(Cp′)2(BTMSA)] ( 1a , Cp′ = Cp = η5‐C5H5; 1b , Cp′ = Cp* = η5‐C5Me5; BTMSA = Me3SiC≡CSiMe3) were found to react with iodine and methyl iodide yielding [Ti(Cp′)2(μ‐I)2] ( 2a / b ; a refers to Cp′ = Cp and b to Cp′ = Cp*), [Ti(Cp′)2I2] ( 3a / b ) and [Ti(Cp′)2(Me)I] ( 4a / b ), respectively. In contrast to 2a , complex 2b proved to be highly moisture sensitive yielding with cleavage of HCp* [{Ti(Cp*)I}2(μ‐O)] ( 7 ). The corresponding reactions of 1a / b with p‐cresol and thiophenol resulted in the formation of [Ti(Cp′)2{O(p‐Tol)}2] ( 5a / b ) and [Ti(Cp′)2(SPh)2] ( 6a / b ), respectively. Reactions of 1a and 1b with 1,n‐benzenedicarboxylic acids (n = 2–4) resulted in the formation of dinuclear titanium(III) complexes of the type [{Ti(Cp′)2}2{μ‐1,n‐(O2C)2C6H4}] (n = 2, 8a / b ; n = 3, 9a / b ; n = 4, 10a / b ). All complexes were fully characterized analytically and spectroscopically. Furthermore, complexes 7 , 8b , 9a ·THF, 10a / b were also be characterized by single‐crystal X‐ray diffraction analyses.  相似文献   

12.
Synthesis and Structural Studies of Aluminum Dialkylamines and Dialkylamides: N‐Chirality of (CH3)3AlNHRR′ and cis‐trans ‐Isomerism at X2AlNRR′ (X = CH3, Cl, H) Aluminum dialkylamines and dialkylamides were prepared from Al(CH3)3 and NH(CH3)R′ (R′: –C2H5, –tC4H9) and characterized by elemental analyses, 1H‐, 13C‐, and 27Al‐NMR spectroscopy. The crystal structures of [(CH3)2AlN(CH3)(–tC4H9)]2 ( IV ), [Cl2AlN(CH3)(C2H5)]2 ( V ), and [H2AlN(CH3)(C2H5)] ( VI‐trans and VI‐cis ) are discussed.  相似文献   

13.
The ready availability of rare parent amido d8 complexes of the type [{M(μ‐NH2)(cod)}2] (M=Rh ( 1 ), Ir ( 2 ); cod=1,5‐cyclooctadiene) through the direct use of gaseous ammonia has allowed the study of their reactivity. Both complexes 1 and 2 exchanged the di‐olefines by carbon monoxide to give the dinuclear tetracarbonyl derivatives [{M(μ‐NH2)(CO)2}2] (M=Rh or Ir). The diiridium(I) complex 2 reacted with chloroalkanes such as CH2Cl2 or CHCl3, giving the diiridium(II) products [(Cl)(cod)Ir(μ‐NH2)2Ir(cod)(R)] (R=CH2Cl or CHCl2) as a result of a two‐center oxidative addition and concomitant metal–metal bond formation. However, reaction with ClCH2CH2Cl afforded the symmetrical adduct [{Ir(μ‐NH2)(Cl)(cod)}2] upon release of ethylene. We found that the rhodium complex 1 exchanged the di‐olefines stepwise upon addition of selected phosphanes (PPh3, PMePh2, PMe2Ph) without splitting of the amido bridges, allowing the detection of mixed COD/phosphane dinuclear complexes [(cod)Rh(μ‐NH2)2Rh(PR3)2], and finally the isolation of the respective tetraphosphanes [{Rh(μ‐NH2)(PR3)2}2]. On the other hand, the iridium complex 2 reacted with PMe2Ph by splitting the amido bridges and leading to the very rare terminal amido complex [Ir(cod)(NH2)(PMePh2)2]. This compound was found to be very reactive towards traces of water, giving the more stable terminal hydroxo complex [Ir(cod)(OH)(PMePh2)2]. The heterocyclic carbene IPr (IPr=1,3‐bis(2,6‐diisopropylphenyl)imidazol‐2‐ylidene) also split the amido bridges in complexes 1 and 2 , allowing in the case of iridium to characterize in situ the terminal amido complex [Ir(cod)(IPr)(NH2)]. However, when rhodium was involved, the known hydroxo complex [Rh(cod)(IPr)(OH)] was isolated as final product. On the other hand, we tested complexes 1 and 2 as catalysts in the transfer hydrogenation of acetophenone with iPrOH without the use of any base or in the presence of Cs2CO3, finding that the iridium complex 2 is more active than the rhodium analogue 1 .  相似文献   

14.
Reaction of [U(TrenTIPS)(PH2)] ( 1 , TrenTIPS=N(CH2CH2NSiPri3)3) with C6H5CH2K and [U(TrenTIPS)(THF)][BPh4] ( 2 ) afforded a rare diuranium parent phosphinidiide complex [{U(TrenTIPS)}2(μ‐PH)] ( 3 ). Treatment of 3 with C6H5CH2K and two equivalents of benzo‐15‐crown‐5 ether (B15C5) gave the diuranium μ‐phosphido complex [{U(TrenTIPS)}2(μ‐P)][K(B15C5)2] ( 4 ). Alternatively, reaction of [U(TrenTIPS)(PH)][Na(12C4)2] ( 5 , 12C4=12‐crown‐4 ether) with [U{N(CH2CH2NSiMe2But)2CH2CH2NSi(Me)(CH2)(But)}] ( 6 ) produced the diuranium μ‐phosphido complex [{U(TrenTIPS)}(μ‐P){U(TrenDMBS)}][Na(12C4)2] [ 7 , TrenDMBS=N(CH2CH2NSiMe2But)3]. Compounds 4 and 7 are unprecedented examples of uranium phosphido complexes outside of matrix isolation studies, and they rapidly decompose in solution underscoring the paucity of uranium phosphido complexes. Interestingly, 4 and 7 feature symmetric and asymmetric UPU cores, respectively, reflecting their differing steric profiles.  相似文献   

15.
Investigation of the Hydrolytic Build‐up of Iron(III)‐Oxo‐Aggregates The synthesis and structures of five new iron/hpdta complexes [{FeIII4(μ‐O)(μ‐OH)(hpdta)2(H2O)4}2FeII(H2O)4]·21H2O ( 2 ), (pipH2)2[Fe2(hpdta)2]·8H2O ( 4 ), (NH4)4[Fe6(μ‐O)(μ‐OH)5(hpdta)3]·20.5H2O ( 5 ), (pipH2)1.5[Fe4(μ‐O)(μ‐OH)3(hpdta)2]·6H2O ( 7 ), [{Fe6(μ3‐O)2(μ‐OH)2(hpdta)2(H4hpdta)2}2]·py·50H2O ( 9 ) are described and the formation of these is discussed in the context of other previously published hpdta‐complexes (H5hpdta = 2‐Hydroxypropane‐1, 3‐diamine‐N, N, N′, N′‐tetraacetic acid). Terminal water ligands are important for the successive build‐up of higher nuclearity oxy/hydroxy bridged aggregates as well as for the activation of substrates such as DMA and CO2. The formation of the compounds under hydrolytic conditions formally results from condensation reactions. The magnetic behaviour can be quantified analogously up to the hexanuclear aggregate 5 . The iron(III) atoms in 1 ‐ 7 are antiferromagnetically coupled giving rise to S = 0 spin ground states. In the dodecanuclear iron(III) aggregate 9 we observe the encapsulation of inorganic ionic fragments by dimeric{M2hpdta}‐units as we recently reported for AlIII/hpdta‐system.  相似文献   

16.
Self‐assembly of Cd(phen)2+ and Cu(phen)2+ (phen = 1,10‐phenanthroline) building blocks with the bent ligand 4,4′‐dithiodipyridine (dtdp) has been investigated. Both building blocks serve as corner units with constrained cis‐geometry. The arched chain coordination polymer [{Cd(phen)(μ‐dtdp)(dtdp)(H2O)}(ClO4)2·2CH3OH·1.5H2O]n ( 1 ) crystallised from a mixture of Cd(ClO4)2·H2O, phen and dtdp in methanol. The reaction of [Cu(phen)(H2O)2](CF3SO3)2 ( 2 ) with dtdp in an ethanol/water mixture yielded a chair‐like metallamacrocycle, [{Cu(phen)(CF3SO3)2}2(μ‐dtpd)2] ( 3 ). The crystal structure of the precursor complex 2 is also reported.  相似文献   

17.
Metallacyclic complex [(Me2N)3Ta(η2‐CH2SiMe2NSiMe3)] ( 3 ) undergoes C?H activation in its reaction with H3SiPh to afford a Ta/μ‐alkylidene/hydride complex, [(Me2N)2{(Me3Si)2N}Ta(μ‐H)2(μ‐C‐η2‐CHSiMe2NSiMe3)Ta(NMe2)2] ( 4 ). Deuterium‐labeling studies with [D3]SiPh show H–D exchange between the Ta?D ?Ta unit and all methyl groups in [(Me2N)2{(Me3Si)2N}Ta(μ‐D)2(μ‐C‐η2‐CHSiMe2NSiMe3)Ta(NMe2)2] ([D2]‐ 4 ) to give the partially deuterated complex [Dn]‐ 4 . In addition, 4 undergoes β‐H abstraction between a hydride and an NMe2 ligand and forms a new complex [(Me2N){(Me3Si)2N}Ta(μ‐H)(μ‐N‐η2‐C,N‐CH2NMe)(μ‐C‐η2‐C,N‐CHSiMe2NSiMe3)Ta(NMe2)2] ( 5 ) with a cyclometalated, η2‐imine ligand. These results indicate that there are two simultaneous processes in [Dn]‐ 4 : 1) H–D exchange through σ‐bond metathesis, and 2) H?D elimination through β‐H abstraction (to give [Dn]‐ 5 ). Both 4 and 5 have been characterized by single‐crystal X‐ray diffraction studies.  相似文献   

18.
Both of the title compounds, catena‐poly­[[[tetra­aqua­magnesium(I)]‐μ‐4,4′‐bi­pyridine‐κ2N:N′] diiodide bis(4,4′‐bi­pyridine) solvate], {[Mg(C10H8N2)(H2O)4]I2·2C10H8N2}n, (I), and catena‐poly­[[[μ‐4,4′‐bi­pyridine‐bis­[di­iodo­bis­(propan‐1‐ol)­strontium(I)]]‐di‐μ‐4,4′‐bi­pyridine‐κ4N:N′] bis(4,4′‐bi­pyri­dine) solvate], {[Sr2I4(C10H8N2)3(C3H8O)4]·2C10H8N2}n, (II), are one‐dimensional polymers which are single‐ and double‐stranded, respectively, the metal atoms being linked by the 4,4′‐bi­pyridine moieties. The Mg complex, (I), is [cis‐{(H2O)4Mg(N‐4,4′‐bi­pyridine‐N′)(2/2)}](∞|∞)I2·4,4′‐bi­pyridine and Mg has a six‐coordinate quasi‐octahedral coordination environment. The Sr complex, (II), is isomorphous with its previously defined Ba counterpart [Kepert, Waters & White (1996). Aust. J. Chem. 49 , 117–135], being [(propan‐1‐ol)2I2Sr(N‐4,4′‐bi­pyridine‐N′)(3/2)](∞|∞)·4,4′‐bi­pyridine, with the I atoms trans‐axial in a seven‐coordinate pentagonal–bipyramidal Sr environment.  相似文献   

19.
In the crystal structure of the l ‐His–cIMP complex, i.e.l ‐histidinium inosine 3′:5′‐cyclic phosphate [systematic name: 5‐(2‐amino‐2‐carboxyethyl)‐1H‐imidazol‐3‐ium 7‐hydroxy‐2‐oxo‐6‐(6‐oxo‐6,9‐dihydro‐1H‐purin‐9‐yl)‐4a,6,7,7a‐tetrahydro‐4H‐1,3,5,2λ5‐furo[3,2‐d][1,3,2λ5]dioxaphosphinin‐2‐olate], C6H10N3O2+·C10H10N4O7P, the Hoogsteen edge of the hypoxanthine (Hyp) base of cIMP and the Hyp face are engaged in specific amino acid–nucleotide (His...cIMP) recognition, i.e. by abutting edge‐to‐edge and by π–π stacking, respectively. The Watson–Crick edge of Hyp and the cIMP phosphate group play a role in nonspecific His...cIMP contacts. The interactions between the cIMP anions (anti/C3′–endo/transgauche/chair conformers) are realized mainly between riboses and phosphate groups. The results for this l ‐His–cIMP complex, compared with those for the previously reported solvated l ‐His–IMP crystal structure, indicate a different nature of amino acid–nucleotide recognition and interactions upon the 3′:5′‐cyclization of the nucleotide phosphate group.  相似文献   

20.
The structure of the title compound, [NiCu(CN)4(C10H8N2)(H2O)2]n or [{Cu(H2O)2}(μ‐C10H8N2)(μ‐CN)2{Ni(CN)2}]n, was shown to be a metal–organic cyanide‐bridged framework, composed essentially of –Cu–4,4′‐bpy–Cu–4,4′‐bpy–Cu– chains (4,4′‐bpy is 4,4′‐bipyridine) linked by [Ni(CN)4]2− anions. Both metal atoms sit on special positions; the CuII atom occupies an inversion center, while the NiII atom of the cyanometallate sits on a twofold axis. The 4,4′‐bpy ligand is also situated about a center of symmetry, located at the center of the bridging C—C bond. The scientific impact of this structure lies in the unique manner in which the framework is built up. The arrangement of the –Cu–4,4′‐bpy–Cu–4,4′‐bpy–Cu– chains, which are mutually perpendicular and non‐intersecting, creates large channels running parallel to the c axis. Within these channels, the [Ni(CN)4]2− anions coordinate to successive CuII atoms, forming zigzag –Cu—N[triple‐bond]C—Ni—C[triple‐bond]N—Cu– chains. In this manner, a three‐dimensional framework structure is constructed. To the authors' knowledge, this arrangement has not been observed in any of the many copper(II)–4,4′‐bipyridine framework complexes synthesized to date. The coordination environment of the CuII atom is completed by two water molecules. The framework is further strengthened by O—H...N hydrogen bonds involving the water molecules and the symmetry‐equivalent nonbridging cyanide N atoms.  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