首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Tetraethylammonium tetrahydroborate, Et4NBH4, in suspension in refluxing decane-dodecane mixtures has been pyrolysed at temperatures between 175 and 190 °C. Et3NBH3, which is produced by partial decomposition of Et4NBH4, reacts with Et4NBH4 to give the intermediate Et4NB3H8. Et4NBH4 and Et3NBH3 are also involved in the conversion of Et4NB3H8. to (Et4N)2B9H9, (Et4N)2B10H10, Et4NB11H14 and (Et4N)2B12H12 which are formed in varying proportions during the pyrolysis. A 1:1 Et4NBH4Et3NBH3 mixture gives the same mixture of final products in the same proportions as Et4NBH4 alone, but the reaction time is shorter.Results obtained under various conditions, for instance without solvent at 10−2 torr (50% yield), are explained by the transfer of BH3 groups occurring not only through Et3NBH3, but also by solid—solid reactions involving Et4NBH4. A more complete reaction of Et3NBH3 is obtained, giving quantitative yields, only when Et3N is evacuated from the reaction mixture. Optimum conditions for the formation of each hydroborate are examined.  相似文献   

2.
Molecular processes of the action of polycationic peptides that represent polylysine homo- and heterodendrimers on the functional activity of the biogenic amine- and peptide hormone-sensitive adenylyl cyclase signaling system (AC system) in rat myocardium and brains were studied. An intended use of these peptides is that of highly effective polymer carriers for biologically active substances. The polylysine homodendrimers of the third [(NH2)16(Lys)8(Lys)4(Lys)2Lys-Ala-NH2] (I), fourth [(NH2)32(Lys)16(Lys)8(Lys)4(Lys)2Lys-Ala-NH2] (II), and fifth [(NH2)64(Lys)32(Lys)16(Lys)8(Lys)4(Lys)2Lys-Ala-NH2] (III) generations, as well as polylysine heterodendrimers of the fifth generation, [(NH2)64(Lys-Glu)32(Lys-Glu)16(Lys-Glu)8(Lys-Glu)4(Lys-Glu)2Lys-Ala-Ala-Lys(ClAc)-Ala-NH2] (IV), [(NH2)64(Lys-Ala)32(Lys-Ala)16(Lys-Ala)8(Lys-Ala)4(Lys-Ala)2Lys-Ala-Lys(ClAc)-Ala-Ala-NH2] (V) and [(NH2)64(Lys-Gly-Gly)32(Lys-Gly-Gly)16(Lys-Gly-Gly)8(Lys-Gly-Gly)4(Lys-Gly-Gly)2Lys-Gly-Gly-Lys(ClAc)-Ala-Ala-NH2] (VI), interact with the C-terminal regions of α subunits of the heterotrimeric G proteins, preferably of the inhibitor type, and stimulate its activity in respector-independent manner. The most effective G-protein activators were homodendrimers II and III and heterodendrimer V. The polylysine dendrimers disturbed the functional coupling of receptors of biogenic amines and peptide hormones with Gi proteins and, to a lesser extent, Gs proteins. This was manifested as a decrease in the regulatory effects of the hormones on AC activity and the GTP binding of the G protein, as well as by a decrease in the affinity of receptors to agonists in the presence of polylysine dendrimers, which is a consequence of the dissociation of the receptor-G protein complex. It has also been shown that, based on their molecular mechanisms and selectivity of action on G proteins, polylysine dendrimers are similar to mastoparan and melittin, which are natural toxins of insect venom.  相似文献   

3.
Interaction between the sodium salt of a DNA extracted from salmon sperm (41% GC) with [Pt(NH3)4]Cl2, [Pt(NH2? (CH2)2? NH? (CH2)2? NH2Cl]Cl, cis-Pt(NH2? (CH2)2? NH2)Cl2, cis-Pt(NH3)2Cl2, trans-Pt(NH3)2Cl2, K[Pt(C2H4)Cl3], and K2[PtCl4) indicates at least three types of complexation. A correlation is found between the change of pH and the number of platinum atoms fixed per (AT + GC) unit. The first binding site is located on the G-C pairs (guanine–cytosine), most likely the N-7(G) site, as it was shown in a previous study of the guanosine-platinum salts. The fixation of the second platinum atom by the pair (AT + GC) takes place with liberation of protons. In the case of the complexes cis-Pt(NH2? (CH2)2? NH2)Cl2, cis-Pt(NH3)2Cl2, and trans-Pt(NH3)2Cl2 the second interaction seems to involve simultaneously the N-7(A) and the N-1(G) and N-3(C) sites. This latter intercrosslink between guanine and cytosine obviously liberates protons and the decrease of pH is related in this case to the trans effect of the platinum compounds. The first two platinum atoms in the reaction of K2PtCl4] or the Zeise salt, K[Pt(C2H4)Cl3] with DNA are fixed on the G-C pairs. A maximum of six platinum atoms per (AT + GC) unit were fixed in this case. Preliminary experiments with a DNA extracted from bacteria Micrococcus lysodeikticus (72% GC) give similar results.  相似文献   

4.
《Inorganica chimica acta》1986,121(2):213-217
Treatment of [IrCl(C2H4)4] with K(C9H7) (C9H7 =indenyl) gives [Ir(C2H4)2(η-C9H7)]. This compound is converted quantitatively into [Ir(CO)2(η-C9H7)] by treatment with carbon monoxide. By reacting together these two iridium complexes [Ir2(μ-CO)(CO)2(ηC9H7)2] has been obtained. The compound [Ir(CO)2(η-C9H7)] reacts with [Pt(C2H4)2{P(cyclo-C6H11)3}] to give the complex [Ir2Pt(CO)3{P(cyclo-C6H11)3}(η-C9H7)2]. Protonation of the latter affords the salt [Ir2Pt(μ-H)(CO)3{P(cyclo-C6H11)3}(μ-C9H7)2] [BF4]. The main features of the molecular structure of these complexes have been established by IR and NMR spectroscopy.  相似文献   

5.
The molybdenum hydride complexes Mo(PMe3)5H2 and Mo(PMe3)4H4 are capable of cleaving the C-S bonds of thiophene, benzothiophene and dibenzothiophene. For example, Mo(PMe3)5H2 reacts with thiophene to give the η5-thiophene and butadiene-thiolate complexes, (η5-C4H4S)Mo(PMe3)3 and (η5-C4H5S)Mo(PMe3)22-CH2PMe2). These complexes are also obtained from the reaction between Mo(PMe3)4H4 and thiophene under photochemical conditions, whereas at elevated temperatures thiophene is desulfurized to liberate but-1-ene. Similarly, Mo(PMe3)4H4 desulfurizes benzothiophene at elevated temperatures to liberate ethylbenzene, while the arylthiolate complex Mo(PMe3)4(SC6H4Et)H3 is obtained photochemically. Furthermore, Mo(PMe3)4H4 cleaves the C-S bond of dibenzothiophene to give [η61-C6H5C6H4S]Mo(PMe3)2H.  相似文献   

6.
Large-conductance calcium and voltage-dependent potassium (BKCa) channel is an important determinant of vascular tone. It is activated by hydrogen peroxide (H2O2) which occurs in various physiological and pathological processes. However, the regulation mechanism is not fully understood. In the present study, the mSlo in the presence or absence of hβ1 were cotransfected with the PTENwt, PTENC124S, PTENG129E in HEK 293 cells. Typical BKCa channel currents could be recorded in cell-attached configurations. We found that PTENwt reduced the H2O2-induced BKCa channel activation during the initial 10 min treatment. In contrast, coexpression with catalytically inactive PTENC124S/PTENG129E mutants that lack lipid phosphatase activity produced no regulation on the H2O2-induced BKCa channel activation. These results demonstrated that PTEN regulated the H2O2-induced BKCa channel activation through phosphatidylinositol 3-phosphatse. However, the inhibitory effect of PTEN on the H2O2-induced BKCa channel activation was attenuated when cells were treated with H2O2 at concentrations higher than 100 μM or at 100 μM for long-term treatment. In addition, the p-AKT expression level in PTENwt overexpressing cells was lower than that in control cells, and the increase of cytoplasmic free calcium concentration ([Ca2+]i) induced by H2O2 was also inhibited. These findings may elucidate a new mechanism for H2O2-induced BKCa channel activation and provide some evidences for the role of PTEN on vasodilation induced by H2O2.  相似文献   

7.
A critical point in the V1 sector and entire V1VO complex is the interaction of stalk subunits G (Vma10p) and E (Vma4p). Previous work, using precipitation assays, has shown that both subunits form a complex. In this work, we have analysed the N-terminal segment of subunit G (G1–59) of the V1VO ATPase from Saccharomyces cerevisiae by using nuclear magnetic resonance (NMR) spectroscopy. Analyses of 1H-15N heteronuclear single quantum coherence (HSQC) spectra of G1–59 in the absence and presence of the N-terminal peptides E1–18 and E18–38 as well as the produced and purified C-terminal segment (E39–233) shows specific interactions only with the peptide fragment E18–38. The binding of this peptide occurs via the residues M1, V2, S3, and K5 as well for V22, S23, K24, A25 and R26 of G1–59. The specific E18–38/G1–59 binding has been confirmed by fluorescence correlation spectroscopy data. The E18–38 peptide has been studied by CD spectroscopy and NMR. The 3D structure of this peptide adopts a stable helix-hinge-helix formation in solution. A model structure of the E18–38/G1–59 complex reveals the orientation of E18–38 relative to G1–59 via salt-bridges of the polar residues and van der Waals forces at the very N-terminus of both segments.  相似文献   

8.
《Inorganica chimica acta》1986,119(2):149-163
Kinetics of the base-induced decomposition of five 2-alkoxyethyl(aquo)cobaloximes, ROCH2CH2- Co(D2H2)OH2 (R = C6H5, CF3CH2, CH3, CH3CH2, (CH3)2CH), have been studied manometrically in aqueous base, ionic strength 1.0 M (KC1) at 25.0± 0.1 °C under an argon atmosphere. For the complexes with good leaving group alkoxide substituents (R = C6H5 and CF3CH2) the reactions are first- order in cobaloxime and first-order in hydroxide ion and produce stoichiometric amounts of ethylene and leaving group alcohol (ROH). NMR observation of decomposing solutions and workup of cobalt chelate products show that the reaction is initiated by hydroxide ion attack on an equatorial quaternary carbon leading to formation of an altered cobal- oxime product in which one of the Schiff's base linkages has become hydrated. For the remainer of the complexes the yield of ethylene is less than stoichiometric and pH-dependent, and the ethylene evolving reaction is second-order in hydroxide ion activity. The yield-limiting side reaction is shown to be base-catalyzed formation of a base-stable but photolabile alkoxyethylcobaloxime analog in which a Schiff's base linkage of the chelate has become hydrated, β-Elimination to form alkyl vinyl ethers was not observed for any of the alkoxyethylcobal- oximes. The second-order dependence of ethylene formation on hydroxide ion activity for R = CH3, CH2CH3, and CH(CH3)2 is discussed at some length, but is not well understood at present.  相似文献   

9.
Conditions for assay of molybdenum cofactor in barley shoot extracts in the presence of molybdate (25 mM N2MoO4) and the sulphydryl-group protector, reduced glutathione (5 mM) were optimized. Both total Mo-cofactor (assayed after heat-treatment of cell-free extracts) and ‘free’ Mo-cofactor (assayed in untreated cell-free extracts) were assayed. Compared to control plants grown in the absence of an exogenous nitrogen source total Mo-cofactor levels increased around 70 % when plants were grown for 4 days in the presence of either 15 mM KNO3 or 15 mM NH4NO3. Growth in the presence of 15 mM (NH4)2SO4 did not affect the Mo-cofactor level. Very similar results were seen when plants were transferred to these nitrogen sources for 24 hr after previous growth in the absence of an exogenous nitrogen source. In contrast ‘free’ Mo-cofactor levels of both KNO3 and NH4NO3-treated plants were increased 2-3-fold over untreated controls. Growth in the presence of (NH4)2SO4 did not affect the ‘free’ Mo-cofactor level.  相似文献   

10.
L S Kan  J C Barrett  P S Miller  P O Ts'o 《Biopolymers》1973,12(10):2225-2240
PMR investigations on the diastereomeric phosphate methyl protons of the dinucleoside ethyl phosphotriesters Tp(C2H5)T, dA, and dIp(C2H5)dI have been used to study the conformational changes of these dimersin solution. In D2O (273°K), the diastereomeric phosphate-methly groups of Tp(C2H5)T appear as a triplet. The methyl resonances of dIp(C2H5)dI and dAp(C2H5)dA appear as two sets of triplets and their chemical shift differences (δ1 ? δ2), decrease with increasing temperature, finally becoming zero at 292°K and 333°K, respectively. The same phenomenon is observed for dAp(C2H5)dA in CD3OD; in this detacking solvent, the difference (δ1 ? δ2) diminishes to zero at a lower temperature (261°K). At room temperature in D2O, the chemical shift of the phosphate methyl of Tp(C2H5)T appears at lower field than those of dIp(C2H5)dI or dAp(C2H5)dA. The differences between the chemical shifts of these groups (δI ? δT or δA ? δT) increase with increasing temperature, and reach maximal values at 301°K and 333°K, respectively. The results suggest that at low temperature the largest fraction of the dimer population exists in a stacked state, with the phosphate-ethyl groups outside the stack. Increasing temperature causes an oscillation of the bases and a shift in the dimer population away from the stacked state. Finally at high temperature, the planar bases rorate with respect to one another and in the case of dIp(C2H5)dI and dAp(C2H5)dA, the ethyl groups experience shielding by the anisotropic ring current of the five-membered ring of the bases. Thus, the current pmr studies and those reported earlier from our laboratory support an “oscillation-rotation model” for the unstacking process of the dimers. The relationship of this model and the “two-state model” is discussed.  相似文献   

11.
Twenty metallic compounds were assayed for their genotoxic mutagenic activity by the bioluminescence test restoration of the luminescence of dark mutant of the luminous bacterium Photobacterium fischeri). The activity of the metals was tested in a liquid medium as well as on a solid medium. K2Cr2O7, MnCl2, BeCl2, KH2AsO4, ZnCl2 and Na2WO4 showed strong activity in liquid medium while AgNO3, Cd(OOCCH3)2, CoCl2, CuCl2, HgCl2, Na2SeO3 and Pb(NO3)2 were more active in the solid medium test. BaCl2, Na2MoO4, NaAsO2, NiSO4, Na2SeO4, RbCl, and SnCl2 were not active in the bioluminescence test. The correlation between the genotoxic activity of the tested metallic compounds in the bioluminescence test and other bacterial tests for genotoxic agents as well as the correlation between these results and the carcinogenicity of these compounds is discussed.  相似文献   

12.
Large ribonucleoprotein subparticles were recovered upon ribonuclease digestion of the 50 S ribosomal subunits of Escherichia coli, partially deproteinized by LiCl. Both their RNA and their protein compositions were analysed. The subunits, treated with LiCl at a concentration of 5.5 m, released an homogeneous subparticle containing proteins L3, L4, L13, L17, L22 and L29, about 70% of the 13 S fragment of 23 S RNA and about 50% of the 18 S one. Slightly larger species of subparticles were obtained from 50 S subunits treated with LiCl at concentrations between 3 m and 5 m; they contained in addition proteins L20, L21 and L23 or L2, L14, L20, L21 and L23 and a few small 23 S RNA fragments. No large subparticle was recovered from the 6 m-LiCl-treated 50 S subunits which contain only proteins L3, L13 and L17. These LiCl subparticles were compared with those obtained from intact, unfolded and sodium doecyl sulphatetreated 50 S subunits.These studies reveal that in the presence of 0.10 m-magnesium acetate there is a very compact area within 50 S subunits consisting of proteins L3, L4, L13, L17, L22 and L29 and of about 60% of 23 S RNA; this area probably has an essential structural role. The results also show that 23 S RNA has a more folded conformation when within the 50 S subunit than when isolated, this conformation being stabilized by some of the 50 S proteins, in particular proteins L4, L22, L20 and L21. Finally these data permit a more definite localization of the primary and/or secondary binding sites of proteins L2, L3, L4, L14, L17, L20, L21 and L22 on 23 S RNA.  相似文献   

13.
Nitrosyl derivatives of polyoxomolybdates have been synthesized and characterized by X-ray diffraction. Most of them contain the MoII(NO)3+ unit and their structures are related to the following structural types: Lindqvist, Keggin and decatungstate [W10O32]4–. Reductive nitrosylation of (NBu4)4[-Mo8O26] by hydroxylamine in methanol yields (NBu4)2[Mo5O13(OMe)4(NO){Na(MeOH)}]. 3MeOH, which is a versatile reagent yielding a variety of derivatives (i) by the transformation of [Mo5O13(OMe)4(NO)]3– into [Mo6O18(NO)]3– in acetonitrile, (ii) by the formation of [PMo12O39(NO)]4– by reaction of [Mo5O13(OMe)4(NO)]3– with [PMO12O40]3– in basic condition and (iii) by the formation of mixed valence MoVI/MoV/MoII decamolybdates [Mo10O24(OMe)7(NO)]2–, [Mo10O25(OMe)6(NO)] and [Mo10O20(OMe)9(NO)3]2– by chemical reduction of [Mo5O13(OMe)4(NO)]3–; MoII is localized while MoV are delocalized in the first two species but localized in the third. The unique ligating properties of [Mo5O13(OMe)4(NO)]3– have been documented: this species acts as a tetradentate ligand in [Ce{Mo5O13(OMe)4(NO)}2]2–, a symmetrically tetraligating ligand in [Rh2Cp*2(-Br){-Mo5O13(OMe)4(NO)}] and a bidentate ligand in [Mo5O13(OMe)4(NO){RhCp*(H2O)}]. Some polyoxomolybdates of the type [Mo5(NO)2O12{RC(NH2)NHO}2{RC(NH)NO}2]2–, which contain the Mo0(NO) 2 2+ unit, have also been characterized.  相似文献   

14.
The endogenous gibberellins of dwarf mutants of lettuce   总被引:1,自引:1,他引:0       下载免费PDF全文
The gibberellin (GA) content of E-1, a tall genotype of early flowering lettuce (Lactuca sativa L.), and of three selected GA-responsive dwarfs, dwf1, dwf2, and dwf21, has been determined using 13C-labeled internal standards and gas chromatographymass spectrometry (GC-MS). In the shoots of the E-1 parent, GA1, 3-epi-GA1, GA3, GA5, GA8, GA19, GA20, GA29, and GA53 were identified by full scan GC-MS and Kovats retention indices. Purification by immunoaffinity chromatography selective for 13-hydroxy GAs, was necessary for GA identification. Relative to the parent E-1, the concentrations of GA1, GA8, GA20, and GA29 in the shoots of dwf2 plants were reduced to about 10% and in shoots of dwf21 plants to less than 50%. In dwf1 the levels of GA1, GA8, and GA29 were also reduced to less than 50% of the parent E-1, but the level of GA20 was fivefold higher than in E-1. Plant height was correlated with the endogenous levels of GA1 and GA8.  相似文献   

15.
Reactions of alkanolamines [R1R2NXOH; R1 = H, CH3, C2H5; R2 = H, CH3, C2H5 and X = -CH2CH2-, -CH2CH2CH2-, -CH2CHCH3, -C6H4CH2CH2-] with aluminium isopropoxide in different molar ratios (1 to 3) yield compounds of the type Al(OPri)3?n(OXNR1R2)n, where ‘n’ can be 1, 2 and 3. Most of the derivatives are distillable liquids, soluble in common organic solvents and susceptible to hydrolysis even by atmospheric moisture. The new derivatives are characterized by elemental analysis, IR and 1H NMR spectra. Molecular weight measurements of Al(OPri)3?n(OXNR1R2)n reveal them to be tetrameric in nature.  相似文献   

16.
Piyush Jha 《Luminescence》2016,31(7):1302-1305
This paper reports the luminescence behavior of Sr0.097Al2O4:Eu0.01,Dy0.02 phosphors under UV‐irradiation. The effect of UV‐irradiation on afterglow (AG), thermoluminescence (TL) and mechanoluminescence (ML) of Sr0.097Al2O4:Eu0.01,Dy0.02 phosphors is investigated. The space group of Sr0.097Al2O4:Eu0.01,Dy0.02 phosphors is monoclinic P21. The prepared phosphors exhibit a long AG, intense TL and ML. It is found that the AG, ML intensity and TL increase with increasing duration of irradiation time. The ML intensity decreases with successive impact of the load onto the phosphors, whereby the diminished ML intensity can be recovered by UV‐irradiation. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

17.
Benzophenone imine [M(η1-NHCPh2)(CO)nP5-n]BPh4 [M = Mn, Re; n = 2, 3; P = P(OEt)3, PPh(OEt)2, PPh2OEt, PPh3] complexes were prepared by allowing triflate M(κ1-OTf)(CO)nP5-n compounds to react with an excess of the imine. Hydride-imine [MH(η1-NHCPh2)P4]BPh4 (M = Ru, Os), triflate-imine [Os(κ1-OTf)(η1-NHCPh2)P4]BPh4 and bis(imine) [Ru(η1-NHCPh2)2P4](BPh4)2 [P = P(OEt)3] derivatives were also prepared. The complexes were characterized spectroscopically (IR, 1H, 31P, 13C NMR) and a geometry in solution was also established. Hydride-benzophenone imine [IrHCl(η1-NHCPh2)L(PPh3)2]BPh4and [IrHCl(η1-NHCPh2)L(AsPh3)2]BPh4 [L = P(OEt)3 and PPh(OEt)2] complexes were prepared by reacting hydride IrHCl2L(PPh3)2 and IrHCl2L(AsPh3)2 precursors with an excess of imine. Dihydride IrH21-NHCPh2)(PPh3)3 complex was also obtained and a geometry in solution was proposed.  相似文献   

18.
Production and localization of endogenous hydrogen peroxide (H2O2) were investigated in strains of Xanthomonas by histochemical analysis under electron microscopy. Even though the levels of endogenous H2O2 production were different among various strains, the produced H2O2 was localized in the cell wall of all Xanthomonas strains tested. The impairment of the level of endogenous H2O2 accumulation resulted in a significantly decreased growth rate of bacteria, regardless if the difference of the H2O2 level is originally present between wild type strains or caused by mutation of the ahpC gene of Xanthomonas. The endogenous accumulation of H2O2 positively correlates with the cell division. Interestingly, the accumulated H2O2 was also localized in the mesosome-like structure and nucleoids during the cell division cycle. Furthermore, results revealed quantitative and dimensional changes of H2O2 accumulation in the two additional locations. These findings indicated that the additional locations of the accumulated H2O2 were closely associated with the process of cell division. Together, these results suggest that the endogenous H2O2 production plays an important role in cell proliferation of Xanthomonas.  相似文献   

19.
A series of experiments is presented investigating short term and long term changes of the nature of the response of rate of CO2 assimilation to intercellular p(CO2). The relationships between CO2 assimilation rate and biochemical components of leaf photosynthesis, such as ribulose-bisphosphate (RuP2) carboxylase-oxygenase activity and electron transport capacity are examined and related to current theory of CO2 assimilation in leaves of C3 species. It was found that the response of the rate of CO2 assimilation to irradiance, partial pressure of O2, p(O2), and temperature was different at low and high intercellular p(CO2), suggesting that CO2 assimilation rate is governed by different processes at low and high intercellular p(CO2). In longer term changes in CO2 assimilation rate, induced by different growth conditions, the initial slope of the response of CO2 assimilation rate to intercellular p(CO2) could be correlated to in vitro measurements of RuP2 carboxylase activity. Also, CO2 assimilation rate at high p(CO2) could be correlated to in vitro measurements of electron transport rate. These results are consistent with the hypothesis that CO2 assimilation rate is limited by the RuP2 saturated rate of the RuP2 carboxylase-oxygenase at low intercellular p(CO2) and by the rate allowed by RuP2 regeneration capacity at high intercellular p(CO2).  相似文献   

20.
The reaction of [Pd(OAc)2(py)2] with [Li((OEt2)2.5)][B(C6F5)4] was conducted with intent to generate the cationic palladium complex [Pd(OAc)(py)3][B(C6F5)4], (2, py = pyridine). A single crystal structure of this material, however, reveals a 1-D polymer structure formed by the self-assembly of alternating dicationic ([Pd(py)4]2+) and neutral ([Pd(OAc)2(py)2]) palladium units bridged by acetato linkages to give [Pd(py)4][Pd(OAc)2(py)2][B(C6F5)4]2 (3). These two palladium sites are produced by disproportionation of the pyridine ligands in [Pd(OAc)(py)3][B(C6F5)4]. Proton NMR studies confirm the existence of a solvent dependent equilibrium between [Pd(py)4]2+, [Pd(OAc)2(py)2] and [Pd(OAc)(py)3]+.  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