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.
Pyrimidine (pym) ligands with their two endocyclic N‐donor atoms provide 120° angles for molecular constructs, which, with the 90° angle metal fragments cis‐a2MII (M=Pt, Pd; a=NH3 or a2=diamine), form cyclic complexes known as metallacalix[n]arenes (with n=3, 4, 6, 8, …?). The number of possible isomers of these species depends on the symmetry of the pym ligand. Although highly symmetrical (C2v) pym ligands form a single linkage isomer for any n and can adopt different conformations (e.g., cone, partial cone, 1,3‐alternate, and 1,2‐alternate in the case of n=4), low‐symmetry pym ligands (Cs) can produce a higher number of linkage isomers (e.g., four in the case of n=4) and a large number of different conformers. In the absence of any self‐sorting bias, the number of possible species derived from a self‐assembly process between cis‐a2MII and a Cs‐symmetrical pym ligand can thus be very high. By using the Cs‐symmetric pym nucleobase cytosine, we have demonstrated that the number of feasible isomers for n=4 can be reduced to one by applying preformed building blocks such as cis‐[a2M(cytosine‐N3)2]n+ or cis‐[a2M(cytosinate‐N1)2] (for the latter, see the accompanying paper: A. Khutia, P. J. Sanz Miguel, B. Lippert, Chem. Eur. J. 2011 , 17, DOI: 10.1002/chem.2010002723) and treating them with additional cis‐a2MII. Moreover, intramolecular hydrogen‐bonding interactions between the O2 and N4H2 sites of the cytosine ligands reduce the number of possible rotamers to one. This approach of the “directed” assembly of a defined metallacalix[4]arene is demonstrated.  相似文献   

2.
The preparation and X‐ray crystal structure analysis of {trans‐[Pt(MeNH2)2(9‐MeG‐N1)2]} ? {3 K2[Pt(CN)4]} ? 6 H2O ( 3 a ) (with 9‐MeG being the anion of 9‐methylguanine, 9‐MeGH) are reported. The title compound was obtained by treating [Pt(dien)(9‐MeGH‐N7)]2+ ( 1 ; dien=diethylenetriamine) with trans‐[Pt(MeNH2)2(H2O)2]2+ at pH 9.6, 60 °C, and subsequent removal of the [(dien)PtII] entities by treatment with an excess amount of KCN, which converts the latter to [Pt(CN)4]2?. Cocrystallization of K2[Pt(CN)4] with trans‐[Pt(MeNH2)2(9‐MeG‐N1)2] is a consequence of the increase in basicity of the guanine ligand following its deprotonation and Pt coordination at N1. This increase in basicity is reflected in the pKa values of trans‐[Pt(MeNH2)2(9‐MeGH‐N1)2]2+ (4.4±0.1 and 3.3±0.4). The crystal structure of 3 a reveals rare (N7,O6 chelate) and unconventional (N2,C2,N3) binding patterns of K+ to the guaninato ligands. DFT calculations confirm that K+ binding to the sugar edge of guanine for a N1‐platinated guanine anion is a realistic option, thus ruling against a simple packing effect in the solid‐state structure of 3 a . The linkage isomer of 3 a , trans‐[Pt(MeNH2)2(9‐MeG‐N7)2] ( 6 a ) has likewise been isolated, and its acid–base properties determined. Compound 6 a is more basic than 3 a by more than 4 log units. Binding of metal entities to the N7 positions of 9‐MeG in 3 a has been studied in detail for [(NH3)3PtII], trans‐[(NH3)2PtII], and [(en)PdII] (en=ethylenediamine) by using 1H NMR spectroscopy. Without exception, binding of the second metal takes place at N7, but formation of a molecular guanine square with trans‐[(Me2NH2)PtII] cross‐linking N1 positions and trans‐[(NH3)2PtII] cross‐linking N7 positions could not be confirmed unambiguously, despite the fact that calculations are fully consistent with its existence.  相似文献   

3.
Acid‐base and ligating properties of three bis(substituted)pyrazine (pz) and pyrimidine (pym) ligands (pyrazine‐2, 5‐dicarboxylic acid, 2, 5‐pzdcH2, 2, 3‐bis(pyridine‐2‐yl)pyrazine, 2, 3‐bppz, pyrimidine‐4, 6‐dicarboxylic acid, 4, 6‐pmdcH2) toward cis‐PtIIa2 (a = NH3, a2 = en, a2 = 2, 2′‐bpy) have been studied. Combinations of pz‐N/pym‐N with donor atoms of the substituents lead to 5‐membered platinum chelates, but exclusive N, N‐coordination through the pyridyl substituents of 2, 3‐bppz can lead to a 7‐membered platinum chelate with a characteristic L‐shape of the resulting cation. It is observed for PtII(2, 2′‐bpy), yet not for PtII(en), and is a consequence of differences in sterical interactions between the 2, 3‐bppz ligand and the coligands of PtII.  相似文献   

4.
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   

5.
Complexation of 1,4‐phenylenebis(methylene) diisonicotinate, L1 , with cis‐protected PdII components, [Pd( L′ )(NO3)2], in an equimolar ratio yielded binuclear complexes, 1 a – d of [Pd2( L′ )2( L1 )2](NO3)4 formulation where L′ stands for ethylenediamine (en), tetramethylethylenediamine (tmeda), 2,2′‐bipyridine (bpy), and phenanthroline (phen). The combination of 4,4′‐bipyridine, L2 , with the cis‐protected PdII units is known to yield molecular squares, 2 a – d . However, 2 b – d coexist with the corresponding molecular triangles, 3 b – d . Combination of an equivalent each of the ligands L1 and L2 with two equivalents of cis‐protected PdII components in DMSO resulted in the D ‐shaped heteroligated complexes [Pd2( L′ )2( L1 )( L2 )](NO3)4, 4 a – d . Two units of the D ‐shaped complexes interlock, in a concentration dependent fashion, to form the corresponding [2]catenanes [Pd2( L′ )2( L1 )( L2 )]2(NO3)8, 5 a – d under aqueous conditions. Crystal structures of the macrocycle [Pd2(tmeda)2( L1 )( L2 )](PF6)4, 4 b′′ , and the catenane [Pd2(bpy)2( L1 )( L2 )]2(NO3)8, 5 c , provide unequivocal support for the proposed molecular architectures.  相似文献   

6.
The syntheses and crystal structures of the title Pt2II and Pt2III dimers doubly bridged with N,N‐dimethyl­guanidinate ligands, namely bis­(μ‐N,N‐dimethyl­guanidinato)bis­[(2,2′‐bipyridine)platinum(II)](Pt—Pt) bis­(hexa­fluoro­phosphate) acetonitrile disolvate, [Pt2II(C3H8N3)2(C10H8N2)2](PF6)2·2CH3CN, (I), and guanidinium bis­(μ‐N,N‐dimethyl­guanidinato)bis­[(2,2′‐bipyridine)sulfatoplatinum(III)](Pt—Pt) bis­(hexa­fluoro­phosphate) nitrate hexa­hydrate, (C3H10N3)[PtIII2(C3H8N3)2(SO4)2(C10H8N2)2]NO3·6H2O, (II), are reported. The oxidation of the Pt2II dimer into the Pt2III dimer results in a marked shortening of the Pt—Pt distance from 2.8512 (6) to 2.5656 (4) Å. The change is mainly compensated for by the change in the dihedral angle between the two Pt coordination planes upon oxidation, from 21.9 (2) to 16.9 (3)°. We attribute the relatively strong one‐dimensional stack of dimers achieved in the Pt2II compound in part to the strong PtII⋯C(bpy) associations (bpy is 2,2′‐bipyridine) in the crystal structure [Pt⋯C = 3.416 (10) and 3.361 (12) Å].  相似文献   

7.
The Schiff base ligand in the title complex, [Pt(C9H8BrN2S2)2], is deprotonated from its tautomeric thiol form and coordinated to PtIIvia the mercapto S and β–N atoms. The configuration about PtII is a perfect square‐planar, with two equivalent Pt—N [2.023 (3) Å] and Pt—S [2.293 (1) Å] bonds. The phenyl ring is twisted against the coordination moiety Pt1/N1/N1′/S2′/S2 by 31.8 (2)°, due to the steric hindrance induced by ortho‐substituted bulky Br atom.  相似文献   

8.
The title compound, catena‐poly[[[bis(ethylenediamine‐κ2N,N′)platinum(II)]‐ μ‐chlorido‐[bis(ethylenediamine)platinum(IV)]‐μ‐chlorido] tetrakis{4‐[(4‐hydroxyphenyl)diazenyl]benzenesulfonate} dihydrate], {[PtIIPtIVCl2(C2H8N2)4](HOC6H4N=NC6H4SO3)4·2H2O}n, has a linear chain structure composed of square‐planar [Pt(en)2]2+ (en is ethylenediamine) and elongated octahedral trans‐[PtCl2(en)2]2+ cations stacked alternately, bridged by Cl atoms, along the b axis. The Pt atoms are located on an inversion centre, while the Cl atoms are disordered over two sites and form a zigzag ...Cl—PtIV—Cl...PtII... chain, with a PtIV—Cl bond length of 2.3140 (14) Å, an interatomic PtII...Cl distance of 3.5969 (15) Å and a PtIV—Cl...PtII angle of 170.66 (6)°. The structural parameter indicating the mixed‐valence state of the Pt atom, expressed by δ = (PtIV—Cl)/(PtII...Cl), is 0.643.  相似文献   

9.
Two PtIV and two PtII complexes containing a 2,2′‐bipyridine ligand were treated with a short DNA oligonucleotide under light irradiation at 37 °C or in the dark at 37 and 50 °C. Photolysis and thermolysis of the PtIV complexes led to spontaneous reduction of the PtIV to the corresponding PtII complexes and to binding of PtII 2,2′‐bipyridine complexes to N7 of guanine. When the reduction product was [Pt(bpy)Cl2], formation of bis‐oligonucleotide adducts was observed, whereas [Pt(bpy)(MeNH2)Cl]+ gave monoadducts, with chloride ligands substituted in both cases. Neither in the dark nor under light irradiation was the reductive elimination process of these PtIV complexes accompanied by oxidative DNA damage. This work raises the question of the stability of photoactivatable PtIV complexes toward moderate heating conditions.  相似文献   

10.
A macrocyclic tetranuclear platinum(II) complex [Pt(en)(4,4′‐bpy)]4(NO3)8 ( 1 ?(NO3)8; en=ethylenediamine, 4,4′‐bpy=4,4′‐bipyridine) and a mononuclear platinum(IV) complex [Pt(en)2Br2]Br2 ( 2 ?Br2) formed two kinds of PtII/PtIV mixed valence assemblies when reacted: a discrete host–guest complex 1 ? 2 ?Br10 ( 3 ) and an extended 1‐D zigzag sheet 1 ?( 2 )3?Br8(NO3)6 ( 4 ). Single crystal X‐ray analysis showed that the dimensions of the assemblies could be stoichiometrically controlled. Resonance Raman spectra suggested the presence of an intervalence interaction, which is typically observed for quasi‐1‐D halogen‐bridged MII/MIV complexes. The intervalence interaction indicates the presence of an isolated {PtII???X? PtIV? X???PtII} moiety in the structure of 4 . On the basis of electronic spectra and polarized reflectance measurements, we conclude that 4 exhibits intervalence charge transfer (IVCT) bands. A Kramers–Kronig transformation was carried out to obtain an optical conductivity spectrum, and two sub‐bands corresponding to slightly different PtII–PtIV distances were observed.  相似文献   

11.
The title compound, {[PtIIPtIVI2(C2H8N2)4](HPO4)(H2PO4)I·3H2O}n, has a chain structure composed of square‐planar [Pt(en)2]2+ and elongated octa­hedral trans‐[PtI2(en)2]2+ cations (en is ethyl­ene­diamine) stacked alternately along the c axis and bridged by the I atoms; a three‐dimensionally valence‐ordered system exists with respect to the Pt sites. The title compound also has a unique cyclic tetra­mer structure composed of two hydrogenphosphate and two dihydrogenphosphate ions connected by strong hydrogen bonds [O⋯O = 2.522 (10), 2.567 (10) and 2.569 (11) Å]. The Pt and I atoms form a zigzag ⋯I—PtIV—I⋯PtII⋯ chain, with PtIV—I bond distances of 2.6997 (7) and 2.6921 (7) Å, inter­atomic PtII⋯I distances of 3.3239 (8) and 3.2902 (7) Å, and PtIV—I⋯PtII angles of 154.52 (3) and 163.64 (3)°. The structural parameters indicating the mixed‐valence state of platinum, expressed by δ = (PtIV—I)/(PtII—I), are 0.812 and 0.818 for the two independent I atoms.  相似文献   

12.
A series of closely related dinuclear (head-head) PtII complexes of general composition cis-[a2PtL2Pta′2]2+ with a,a′ = NH3 or CH3NH2 and L = 1-methyluracilate-N3,O4 (1-MeU) or 1-methylthyminate-N3,O4 (1-MeT) has been prepared and the solution behavior toward CeIV oxidation studied. The X-ray crystal structure of a representative example cis-[(CH3NH2)2Pt(1-MeU)2Pt(CH3NH2)2](ClO4)2 · 0.5 H2O ( 1b ), has been determined: Monoclinic, space group P21/c, a = 11.907(7) Å, b = 19.087(14) Å, c = 12.525(7) Å, β = 90.49(4)°, Z = 4. Oxidation of these diplatinum(II) complexes ([Pt2.0]2) with CeIV in aqueous solution to the corresponding diplatinum(III) species ([Pt3.0]2) proceeds via tetranuclear [Pt2.25]4 or dinuclear [Pt2.5]2 mixed-valence state compounds, depending on the nature of the a′ ligands: with a′ = NH3, blue green [Pt2.25]4 forms, whereas with a′ = CH3NH2, purple [Pt2.5]2 represents the intermediate. This difference is interpreted in terms of differences in bulk between NH3 and CH3NH2 ligands trans to the O(4) positions of the bridging nucleobases which influence the ability of dinuclear species to associate via the O(4)2 Pt a2′ faces.  相似文献   

13.
The title compound, [Pt2III(C5H10NO)2(SO4)2(C10H8N2)2]·4H2O, is the first reported example of a complex in which an amidate‐bridged Pt(bpy) dimer is stabilized in the oxidation level of PtIII (bpy is 2,2′‐bi­pyridine). The asymmetric unit consists of one half of the formula unit with a twofold axis passing through the center of the dimer. The intradimer PtIII—PtIII bond distance [2.5664 (6) Å] is comparable to those reported for α‐pyridonate‐bridged cis‐diammineplatinum(III) dimers [2.5401 (5)–2.5468 (8) Å; Hollis & Lippard (1983). Inorg. Chem. 22 , 2605–2614], in spite of the close contact between the bpy planes within the dimeric unit. The axial Pt—Osulfate distance is 2.144 (7) Å.  相似文献   

14.
The reactions of AuIII, PtII and PdII complexes with 2-pyridinecarboxaldehyde (2CHO-py) have been examined in protic (H2O, MeOH, EtOH) and aprotic (DMF, CH2Cl2) solvents. Compounds in which the pyridine ligand is N-coordinated, either in the original aldehydic form or in a new form derived from addition of one or two protic molecules, have been isolated, namely: [Au(2CHO-py · H2O)Cl3], [Au(2CHO-py · MeOH)Cl3], [Au(2CHO-py · 2EtOH)Cl3], cis-[Pt(2CHO-py)2Cl2], trans-[Pd(2CHO-py)2Cl2], trans-[Pt(dmso)(2CHO-py)Cl2], [Pt{C5H4N-(CH2SMe)}Cl(2CHO-py)](ClO4), [Pt(terpy)(2CHOpy)](ClO4)2, [Pt(terpy)(2CHO-py · H2O)](ClO4)2 (terpy = 2,2′:6′,2′′-terpyridine). 1H-n.m.r. experiments show that the addition of the protic molecule(s) to the PtII and PdII complexes is reversible. The effects of the nature of the metal ion and the ancillary ligands as well as of the total charge of the complexes on the relative stability of the addition products are discussed. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

15.
A strategy is presented that enables the quantitative assembly of a heterobimetallic [PdPtL4]4+ cage. The presence of two different metal ions (PdII and PtII) with differing labilities enables the cage to be opened and closed selectively at one end upon treatment with suitable stimuli. Combining an inert PtII tetrapyridylaldehyde complex with a suitably substituted pyridylamine and PdII ions led to the assembly of the cage. 1H and DOSY NMR spectroscopy and ESI mass spectrometry data were consistent with the quantitative formation of the cage, and the heterobimetallic structure was confirmed using single‐crystal X‐ray crystallography. The structure of the host–guest adduct with a 2,6‐diaminoanthraquinone guest molecule was determined. Addition of N,N′‐dimethylaminopyridine (DMAP) resulted in the formation of the open‐cage [PtL4]2+ compound and [Pd(DMAP)4]2+ complex. This process could then be reversed, with the reformation of the cage, upon addition of p‐toluenesulfonic acid (TsOH).  相似文献   

16.
Treatment of [M(AMP)Cl2] (M = PtII, PdII; AMP = 2-aminomethylpyridine) with 1 mole of AgX (X = ClO4, BF4, PF6) in dmso yields [M(AMP)(dmso)Cl]X. Single crystal X-ray structure determinations of the PdII and PtII complexes indicate that dmso is S-bondedtrans to the pyridyl ring in both complexes. (2-Aminomethylpyridine)chloro(dimethylsulphoxide-S) palladium(II) tetrafluoroborate.  相似文献   

17.
Mono(nucleobase) complexes of the general composition cis‐[PtCl2(NH3)L] with L=1‐methylcytosine, 1‐MeC ( 1 a ) and L=1‐ethyl‐5‐methylcytosine, as well as trans‐[PtX2(NH3)(1‐MeC)] with X=I ( 5 a ) and X=Br ( 5 b ) have been isolated and were characterized by X‐ray crystallography. The Pt coordination occurs through the N3 atom of the cytosine in all cases. The diaqua complexes of compounds 1 a and 5 a , cis‐[Pt(H2O)2(NH3)(1‐MeC)]2+ and trans‐[Pt(H2O)2(NH3)(1‐MeC)]2+, display a rich chemistry in aqueous solution, which is dominated by extensive condensation reactions leading to μ‐OH‐ and μ‐(1‐MeC?N3,N4)‐bridged species and ready oxidation of Pt to mixed‐valence state complexes as well as diplatinum(III) compounds, one of which was characterized by X‐ray crystallography: h,t‐[{Pt(NH3)2(OH)(1‐MeC?N3,N4)}2](NO3)2 ? 2 [NH4](NO3) ? 2 H2O. A combination of 1H NMR spectroscopy and ESI mass spectrometry was applied to identify some of the various species present in solution and the gas phase, respectively. As it turned out, mass spectrometry did not permit an unambiguous assignment of the structures of +1 cations due to the possibilities of realizing multiple bridging patterns in isomeric species, the occurrence of different tautomers, and uncertainties regarding the Pt oxidation states. Additionally, compound 1 a was found to have selective and moderate antiproliferative activity for a human cervix cancer line (SISO) compared to six other human cancer cell lines.  相似文献   

18.
The asymmetric unit of the title complex, [PtCl2(C14H38B10P2)]·0.5CH2Cl2 or cis‐[PtCl2{1,2‐(PiPr2)2‐1,2‐C2B10H10}]·0.5CH2Cl2, contains one disordered solvent mol­ecule and two mol­ecules of the complex, in which each PtII atom displays slightly distorted square‐planar coordination geometry. The P atoms connected to the cage C atoms are coordinated to the PtII atom. The Pt—P distances vary slightly [2.215 (3) and 2.235 (4) Å] and the Pt—Cl distances are equal [2.348 (3) and 2.353 (5) Å].  相似文献   

19.
Of the numerous ways in which two adenine and two guanines (N9 positions blocked in each) can be cross‐linked by three linear metal moieties such as trans‐a2PtII (with a=NH3 or MeNH2) to produce open metalated purine quartets with exclusive metal coordination through N1 and N7 sites, one linkage isomer was studied in detail. The isomer trans,trans,trans‐[{Pt(NH3)2(N7‐9‐EtA‐N1)2}{Pt(MeNH2)2(N7‐9‐MeGH)}2][(ClO4)6] ? 3H2O ( 1 ) (with 9‐EtA=9‐ethyladenine and 9‐MeGH=9‐methylguanine) was crystallized from water and found to adopt a flat Z‐shape in the solid state as far as the trinuclear cation is concerned. In the presence of excess 9‐MeGH, a meander‐like construct, trans,trans,trans‐[{Pt(NH3)2(N7‐9‐EtA‐N1)2}{Pt(MeNH2)2(N79‐MeGH)2}][(ClO4)6] ? [(9‐MeGH)2] ? 7 H2O ( 2 ) is formed, in which the two extra 9‐MeGH nucleobases are hydrogen bonded to the two terminal platinated guanine ligands of 1 . Compound 1 , and likewise the analogous complex 1 a (with NH3 ligands only), undergo loss of an ammonia ligand and formation of NH4+ when dissolved in [D6]DMSO. From the analogy between the behavior of 1 and 1 a it is concluded that a NH3 ligand from the central Pt atom is lost. Addition of 1‐methylcytosine (1‐MeC) to such a DMSO solution reveals coordination of 1‐MeC to the central Pt. In an analogous manner, 9‐MeGH can coordinate to the central Pt in [D6]DMSO. It is proposed that the proton responsible for formation of NH4+ is from one of the exocyclic amino groups of the two adenine bases, and furthermore, that this process is accompanied by a conformational change of the cation from Z‐form to U‐form. DFT calculations confirm the proposed mechanism and shed light on possible pathways of this process. Calculations show that rotational isomerism is not kinetically hindered and that it would preferably occur previous to the displacement of NH3 by DMSO. This displacement is the most energetically costly step, but it is compensated by the proton transfer to NH3 and formation of U(?H+) species, which exhibits an intramolecular hydrogen bond between the deprotonated N6H? of one adenine and the N6H2 group of the other adenine. Finally the question is examined, how metal cross‐linking patterns in closed metallacyclic quartets containing two adenine and two guanine nucleobases influence the overall shape (square, rectangle, trapezoid) and the planarity of a metalated purine quartet.  相似文献   

20.
The reaction of cis-[Pt(15NH3)2(H2O) 2] 2+ (3) with N-acetylcysteine [H3accys] was investigated in aqueous solution. In this reaction, the ammine in the platinum complex formed was liberated. A mono-dentate sulfur-boundplatinum(II) product cis-[Pt(15NH3)2(H2O)(H2accys-S)]+ (7) and six-membered che-late ring complex cis-[Pt(15NH3)2 (Haccys-S,O)] (8) were formed in solution. The dinuclear sulfur-bridged complex 9, giving a broad peak in 15N NMR, was also observed, but only present in very tiny amounts. The mass spectrometry (ES-MS) was undertaken from this re action, and the product detected was only the dinuclear sulfur bridged platinum species and species related to it by ammine loss.  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