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.
Selective methods for the synthesis of the cluster Pd23(CO)20L8, L=PEt3, have been suggested. The compound has been prepared by two routes: by the reaction of Pd10(CO)12L6 with Me3NO in the presence of HOAc with removal of CO from the gas phase, and by the reaction of Pd10(CO)12L6 with Pd(OAc)2 and CO followed by oxidation by Me3NO in an inert atmosphere.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 7, pp. 1299–1300, July, 1993.  相似文献   

2.
The reaction of Rh4(CO)12 with Pd(PBu t 3)2 yielded the high nuclearity bimetallic hexarhodium-tripalladium cluster complex Rh6(CO)16[Pd(PBu t 3)]3, 10, in 11% yield. Compound 10 was converted to the hexarhodium-tetrapalladium cluster Rh6(CO)16[Pd(PBu t 3)]4, 11, in 62% yield by reaction with an additional quantity of Pd(PBu t 3)2. Both compounds were characterized crystallographically. Structurally, both compounds consist of an octahedral cluster of six rhodium atoms with sixteen carbonyl ligands analogous to that of the known compound Rh6(CO)16. Compound 10 also contains three Pd(PBu t 3) groups that bridge three Rh–Rh bonds along edges of the Rh6 octahedron to give an overall D3 symmetry to the Rh6Pd3 cluster. Compound 11 contains four edge bridging Pd(PBu t 3) groups distributed across the Rh6 octahedron to give an overall D2d symmetry to the Rh6Pd4 cluster. Each Rh–Pd connection in both compounds contains a bridging carbonyl ligand that helps to stabilize the bond between the Pd(PBu t 3) groups and the Rh atoms. Both compounds can be regarded as Pd(PBu t 3) adducts of Rh6(CO)16.  相似文献   

3.

Structural isomerism of the Pd4(L)4(RCO2)4 (L = CO, CH2, NO; R = CC13, CF3, CH3) complexes was studied in the framework of the density functional theory (DFT). Among the Pd4(CO)4(RCO2)4 and Pd4(CH2)4(RCO2)4 complexes the most stable were the isomers with alternate coordination of pairs of carbonyl and carboxylate ligands on the sides of a planar rectangular metal core. The isomers with the pairwise coordination of NO/RCO2 on one side of the metal core are the most stable between the Pd4(NO)4(RCO2)4 complexes. The features of mutual coordination of ligands in polynuclear complexes of palladium are clarified using the obtained results.

  相似文献   

4.
This research was an outgrowth of previous reactions with [Pd13Ni13(CO)34]4? which produced a tetragonal crystal form of Pd23(CO)20(PEt3)10 (1) that has the same cuboctahedral-based Pd23 framework with an identical number of PEt3 ligands but two fewer CO ligands than the monoclinic crystal form of Pd23(CO)22(PEt3)10 (3) originally reported from reactions with Pd10(CO)12(PEt3)6. A subsequent investigation presented herein to establish whether the carbonyl capacity is influenced by the nature of the phosphine ligands has led to syntheses of Pd23(CO) x (PR3)10 [R3=Et3 (1), Bu n 3 (4), and Me2Ph (5)] with 20 CO ligands (x=20) from corresponding Pd10(CO)12(PR3)6 precursors either by deligation with Pd(OAc)2, CF3CO2H, Ni(1,5-COD)2, [NMe4]2[Ni6(CO)12], or HCO2H or by spontaneous enlargement; yields varied from 15 to 79%. Although attempts to obtain the original Pd23(CO)22(PEt3)10 (3) were unsuccessful, a highly significant outcome was the isolation (one time) of another monoclinic crystal form possessing the triethylphosphine Pd23(CO) x (PEt3)10 cluster with 21 COs (2). Both the compositions and atomic arrangements for each of five Pd23 clusters [1a (solvated); 1b (unsolvated); 2, 4, and 5] were unambiguously established from low-temperature single-crystal CCD X-ray crystallographic determinations in accordance with their nearly identical IR carbonyl frequencies. Solution 31P{1H} NMR spectra of 1 and 4 at room temperature displayed three distinct signals with expected integral ratios of 2/4/4 that are consistent with the solid-state structures of Pd23(CO)20(PR3)10 [R3=Et3 (1), Bu n 3 (4)] remaining intact in solution. The metal-core geometries of all of these Pd23(CO) x (PR3)10 clusters, including the thermodynamically stable ones with 20 CO ligands and the kinetic products with additional CO ligands (x=21, 22), are essentially the same. The common Pd23 core may be best described as possessing a centered hexacapped cuboctahedral Pd19 kernel (alternatively denoted as a centered ν2 Pd19 octahedron) with four edge-connected exopolyhedral wingtip Pd(exo) atoms that reduce the pseudo metal-core symmetry from Oh to D2h. The 10 PR3 ligands are linked to the six tetracapped Pd(cap) and four edge-capped wingtip Pd(exo) atoms; the latter four Pd(exo) atoms are each composed of four trigonal-planar Pd(μ2-CO)2(PR3) units. These crystallographic results provide compelling geometrical evidence for a heretofore unknown stereochemical example involving variable carbonyl ligation (x=20, 21, 22) of a close-packed nanosized Pd n (CO) x (PR3) y cluster (in this case with identical PEt3 ligands) without significant changes being induced in either the overall metal-core architecture or steric dispositions of the same number of PR3 ligands. These experimental findings have particular relevance to the long-standing Muetterties cluster/surface science analogy in showing that the different number (as well as different modes) of carbonyl ligations observed in these large metal carbonyl clusters are directly related to pressure-induced dissociative/nondissociative migratory coverages in CO chemisorptions on metal surfaces. The observed expanded capacity of CO coordination on the same Pd23 polyhedron without notable changes in geometry is no doubt a consequence of its virtually nanosized metal-core architecture; distances between outermost centrosymmetrically related pairs of Pd(cap) and Pd(exo) atoms in the Pd23 framework are 0.8 and 0.9 nm, respectively. An electrochemical (CV) study revealed that 1 undergoes one quasi-reversible two-electron reduction to 1 2? (E1/2=?0.91 V) and two consecutive quasi-reversible one-electron oxidations to 1/1 + at E1/2=0.08 V and 1 +/1 2+ at E1/2=0.32 V (THF; Ag/AgCl as reference electrode). A stereochemical/electronic analysis with the isostructural Au2Pd21(CO)20(PEt3)10 analogue (9) and resulting implications are given.  相似文献   

5.
Reductive condensation of Pd(OAc)2 in dioxane in the presence of CO and PR3 (R = Et, Bun) with addition of CF3COOH leads to the formation of decanuclear Pd103-CO)42-CO)8(PBun3)6 (I) and Pd10(CO)14(PBun3) (II) at Pd(OAc)2:PR3 molar ratios of 1:4–1:10 and 1:1.5–1:2.5, respectively. The use of CH3COOH instead of CF3COOH results in tetranuclear clusters Pd4(CO)5(PR3)4 (III) and Pd42-CO)6(PBun3) (IV). I ? III and III → IV transformations occur in organic media. The structures of I (space group P21/n, Z = λMo, 12125 independent reflections, R = 0.047) and IV (Pz:3, Z = λMo, 3254 reflections, R = 0.098) were established by X-Ray diffractions analysis. Cluster I is a 10-vertex Pd10 polyhedron, an octahedron with four unsymmetrically centered non-adjacent faces. The average PdPd distances in the octahedron are 2.825 Å, in the eight short Pdoct.Pdcap. bonds with the “equatorial” Pd atoms of the inner octahedron, bridged by the μ2-CO ligands, are 2.709 Å, and in the four elongated (without bridging CO groups) bonds with the apical Pd atoms of the octahedron are 3.300–3.422 Å. The PBun3 ligands are coordinated to the apical Pd atoms and the capping atoms (PdP 2.291–2.324 Å). Cluster IV is tetrahedral, with the CO ligands symmetrically bridged; PdPd 2.778–2.817; PdP 2.232–2.291; PdC 2.06 Å (average).  相似文献   

6.
The photochemical reactions of Co2(CO)6(PBu3)2 with various quinones and phenothiazine were studied by ESR. The results indicate that the photolysis of Co2(CO)6(PBu3)2 in the presence of o-quinones led to the observation of an ESR spectrum, showing a broad 8 or 12 groups of hyperfine lines due to interaction with cobalt (Co, 1 = 7/2). The results show that with o-quinones the cobalt radical adducts are formed via metal chelation to the carbonyl oxygens. However, when Co2(CO)6(PBu3)2 was irradiated in the presence of p-quinones, only the para-semiquinone radicals were observed. The photolysis of Co2(CO)6(PBu3)2 with phenothiazine in toluene yields phenothiazine radical. The sixteen and eighteen electrons rule have been used to account for these observations.  相似文献   

7.
The reaction of Na2[Fe(CO)4] with Br2CF2 in n‐pentane generates a mixture of the compounds (CO)3Fe(μ‐CO)3–n(μ‐CF2)nFe(CO)3 ( 2 , n = 2; 3 , n = 1) in low yields with 3 as the main product. 3 is obtained free from 2 by reacting Br2CF2 with Na2[Fe2(CO)8]. The non‐isolable monomeric complex (CO)4Fe=CF2 ( 1 ) can probably considered as the precursor for 2 . 3 reacts with PPh3 with replacement of two CO ligands to form Fe2(CO)6(μ‐CF2)(PPh3)2 ( 4 ). The complexes 2 – 4 were characterized by single crystal X‐ray diffraction. While the structure of 2 is strictly similar to that of Fe2(CO)9, the structure of 3 can better be described as a resulting from superposition of the two enantiomers 3 a and 3 b with two semibridging CO groups. Quantum chemical DFT calculations for the series (CO)3Fe(μCO)3–n(μ‐CF2)nFe(CO)3 (n = 0, 1, 2, 3) as well as for the corresponding (μ‐CH2) derivatives indicate that the progressively larger σ donor and π acceptor properties for the bridging ligands, in the order CO < CF2 < CH2, favor a stronger Fe–Fe bond.  相似文献   

8.
The electron impact induced mass spectra of [CF3SMn(CO)4]2, [CF3SeMn(CO)4]2, [CF3SFe(CO)3]2, [CF3SeFe(CO)3]2, CF3SeFe(CO)2C5H5 and CF3SCr(NO)2C5H5 are reported. These compounds exhibit weak molecular ion peaks and undergo preferential loss of CO or NO groups. The CO or NO free fragments suffer typical loss of ECF2(E = S, Se) with the simultaneous shift of F from carbon to metal. The ions [FFeC5H5]+ and [FCrC5H5]+ in the spectra of the cyclopentadienyl compounds prefer expulsion of π-cyclopentadienyls. The pyrolysis effects on the spectra of the compounds have been studied. An increase in temperature eases the expulsion of ECF2 groups from all the compounds and favors the formation of [Fe(C5H5)2]+ and [Cr(C5H5)2]+ in the cyclopentadienyl compounds.  相似文献   

9.
The kinetics of protonation of tungsten hydrides WH(CO)2(NO)L2 (1, L = PMe3, PEt3, P(OPri)3, PPh3) by weak OH-acids (PhOH, (CF3)2CHOH, (CF3)3COH) in hexane was studied by IR spectroscopy. The study of the reactions of compounds 1 with OH-acids at 190–270 K revealed that the first step involves the formation of dihydrogen-bonded W(CO)2(NO)L2(H)...HOR complexes. When the temperature increases to ambient, the proton transfer and evolution of molecular hydrogen occur, affording the final products: organyloxy derivatives W(OR)(CO)2(NO)L2. The study of the kinetics at 298 K found that the proton transfer is the rate-determining step. The rate constants k app are 2.2·10−5–6.3·10−4 s−1, and the free activation energies are ΔG 298K = 22–23 kcal mol−1. The rate constants depend on the proton-accepting properties of the hydride and the acidic properties of the OH-proton donor and increase in the same order as the enthalpy of hydrogen bond formation. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 5, pp. 837–841, May, 2007.  相似文献   

10.
Perfluoromethyl-Element-Ligands. XVIII. Preparation and Spectroscopic Investigation of M(CO)5L and M(CO)4L2 Complexes [L = MenP(CF3)3?n; n = 0–3; M = Cr, Mo, W] M(CO)5L and cis-M(CO)4L2 complexes, respectively [M = Cr, Mo, W; L = MenP(CF3)3?n; n = 0–3] are prepared reacting M(CO)5 · THF or M(CO)4norbor with L at room temperature. The cis-compounds isomerize above 50°C yielding the trans-complexes; the rate of isomerization increases with increasing number of CF3 groups. Thermal reaction of M(CO)6 (M = Cr, Mo, W) with P(CF3)3 yields M(CO)5P(CF3)3 and trans-M(CO)4[P(CF3)3]2. Introduction of three P(CF3)3 ligands by reaction with M(CO)3(cycloheptatriene) (M = Cr, Mo) proves unsuccessful; besides little M(CO)5P(CF3)3 trans-M(CO)4[P(CF3)3]2 is formed. The new compounds are characterized by analytical and spectroscopic (n.m.r., i.r., MS) methods.  相似文献   

11.
In this study selected bidentate (L2) and tridentate (L3) ligands were coordinated to the Re(I) or Tc(I) core [M(CO)2(NO)]2+ resulting in complexes of the general formula fac-[MX(L2)(CO)2(NO)] and fac-[M(L3)(CO)2(NO)] (M = Re or Tc; X = Br or Cl). The complexes were obtained directly from the reaction of [M(CO)2(NO)]2+ with the ligand or indirectly by first reacting the ligand with [M(CO)3]+ and subsequent nitrosylation with [NO][BF4] or [NO][HSO4]. Most of the reactions were performed with cold rhenium on a macroscopic level before the conditions were adapted to the n.c.a. level with technetium (99mTc). Chloride, bromide and nitrate were used as monodentate ligands, picolinic acid (PIC) as a bidentate ligand and histidine (HIS), iminodiacetic acid (IDA) and nitrilotriacetic acid (NTA) as tridentate ligands. We synthesised and describe the dinuclear complex [ReCl(μ-Cl)(CO)2(NO)]2 and the mononuclear complexes [NEt4][ReCl3(CO)2(NO)], [NEt4][ReBr3(CO)2(NO)], [ReBr(PIC)(CO)2(NO)], [NMe4][Re(NO3)3(CO)2(NO)], [Re(HIS)(CO)2(NO)][BF4], [99Tc(HIS)(CO)2(NO)][BF4], [99mTc(IDA)(CO)2 (NO)] and [99mTc(NTA)(CO)2(NO)]. The chemical and physical characteristics of the Re and Tc-dicarbonyl-nitrosyl complexes differ significantly from those of the corresponding tricarbonyl compounds.  相似文献   

12.
Transition Metal Phosphido Complexes. XIII. P-functional Phosphido-Bridged Heterobimetallic Complexes with and without a Metal-Metal Bond; P(SiMe3)2-Bridged cp(CO)xFe Derivatives cp(CO)2FeP(SiMe3)2 1 reacts with the carbonyl nitrosyl complexes Co(CO)3(NO), Fe(CO)2(NO)2,Mn(CO)(NO)3 substituting a CO ligand and with the THF complexes M′(CO)5THF(M′ = Cr, Mo, W), Mncp(CO)2THF MnMecp(CO)2 which can be obtained in solution substituting the THF ligand to give the phosphido-bridged bimetallic complexes cp(CO)2Fe[μ-P(SiMe3)2]M′Lm 2 (M′Lm = Co(CO)2(NO) b , Fe(CO)(NO)2 c , Mn(NO)3 d , Cr(CO)5 f , Mo(CO)5 g , W(CO)5 h , Mncp(CO)2 i , MnMecp(CO)2 j ). Solutions of Li(Me3Si)2PM′Lm 4e–l (M′Lm = Fe(CO)4 e , Crcp(CO)(NO) k , Vcp(CO)3 l ) are available by a selective cleavage reaction of a Si? P bond in the complexes (Me3Si)3PM′Lm 3e–l using n-BuLi. Reactions of cp(CO)2FeBr with 4e–l give the bimetallic complexes 2e–l . The open-chain complexes 2c, 2f, 2h–k undergo a photochemical decarbonylation reaction to form the phosphido-bridged bimetallic complexes cp(CO)Fe[μ-CO, μ-P(SiMe3)2]M′Lm?1(Fe-M′) 5 (M′Lm?1 = Fe(NO)2 c , Cr(CO)4 f , W(CO)4 h , Mncp(CO) i , MnMecp(CO) j , Crcp(NO) k ) containing a metal-metal bond. Equilibria between various isomers can partially be observed in solutions of the complexes 5. I.R., N.M.R., and mass spectral data are reported.  相似文献   

13.
The reaction of tetranuclear Pd4(μ-COOCH3)4(μ-CO)4 cluster (1a) with p- and o-chloronitrosobenzenes was found to give dinuclear nitrosoamide complexes, Pd2(OAc)2(p-ClC6H4N[p-ClC6H3NO])2 (4) and Pd2(OAc)2(o-ClC6H4N[o-ClC6H3NO])2 (5), respectively. The formation of complexes 4 and 5 is accompanied by evolution of CO2, resulting from oxidation of CO coordinated in cluster 1. Complexes 4 and 5 were characterized by elemental analysis and IR and 1H NMR spectroscopy; their structures were studied by EXAFS. The reactions of dinuclear complex 4 with molecular hydrogen and CO were studied. The major products of reduction of 4 with hydrogen include metallic palladium, acetic acid, cyclohexanone, and molecular nitrogen. Treatment of complex 4 with CO under mild conditions (1 atm, 20 °C) affords p-chlorophenyl isocyanate.  相似文献   

14.
The reaction of Ru3(CO)12 with tetramethyltrifluoromethylcyclopentadiene at various ratios of the reagents was studied. Refluxing of Ru3(CO)12 with a sixfold excess of tetramethyltrifluoromethylcyclopentadiene in octane in an inert atmosphere gave a complex, which is, according to X-ray diffraction data, a dimer,trans-[Ru(η5-C5Me4CF3)(CO)2]2. The reaction under the same conditions but starting from Ru3(CO)12 and C5Me4CF3H in 2∶1 molar ratio gave a hexaruthenium cluster [Ru63-H)(η24-CO)2(μ-CO)(Co)125-C5Me4CF2)], which was characterized by IR as well as1H,13C, and19F NMR spectroscopy. According to X-ray diffraction data, an Ru4 tetrahedron, in which two edges are bound by additional “briding” Ru atoms, constitutes the frame of this compound. This complex has one (η5-C5Me4CF3) ligand, as well as one (μ3-H) and two (η24-CO) groups. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 3, pp. 507–512, March, 1998.  相似文献   

15.
The sole and unexpected products from the reactions of a variety of lead (II) and lead (IV) compounds with [Co2(CO)6(L)2] complexes (L = tertiary arsine, phosphine, or phosphite) in refluxing benzene solution are the blue, air-stable percobaltoplumbanes [Pb{Co(CO)3(L)}4]. These have also been obtained from the reaction of Na[Co(CO)3(L)] (L  PBu3n) with lead (II) acetate which with Na[Fe(CO)2(NO)(L)] forms the isoelectronic [Pb{Fe(CO)2(NO)(L)}4] [L  P(OPh)3]. The IR spectra of the complexes in the v(CO) and v(NO) regions are consistent with tetrahedral PbCo4 or PbFe4 fragments, trigonal bipyramidal coordination about the cobalt or iron atoms and linear PbCoAs, PbCoP, or PbFeP systems. Unlike [Pb{Co(CO)4}4], our complexes do not dissociate to [Co(CO)3(L)]? or [Fe(CO)2(NO)(L)]? ions when dissolved in donor solvents.  相似文献   

16.
A new ruthenium-rhodium mixed-metal cluster HRuRh3(CO)12 and its derivatives HRuRh3(CO)10(PPh3)2 and HRuCo3(CO)10(PPh3)2 have been synthesized and characterized. The following crystal and molecular structures are reported: HRuRh3(CO)12: monoclinic, space group P21/c, a 9.230(4), b 11.790(5), c 17.124(9) Å, β 91.29(4)°, Z = 4; HRuRh3(CO)10(PPh3)2·C6H14: triclinic, space group P1, a 11.777(2), b 14.079(2), c 17.010(2) Å, α 86.99(1), β 76.91(1), γ 72.49(1)°, Z = 2; HRuCo3(CO)10(PPh3)2·CH2Cl2: triclinic, space group P1, a 11.577(7), b 13.729(7), c 16.777(10) Å, α 81.39(4), β 77.84(5), γ 65.56°, Z = 2. The reaction between Rh(CO)4? and (Ru(CO)3Cl2)2 tetrahydrofuran followed by acid treatment yields HRuRh3(CO)12 in high yield. Its structural analysis was complicated by a 80–20% packing disorder. More detailed structural data were obtained from the fully ordered structure of HRuRh3(CO)10(PPh3)2, which is closely related to HRuCo3(CO)10(PPh3)2 and HFeCo3(CO)10(PPh3)2. The phosphines are axially coordinated.  相似文献   

17.
Reaction of the carbonyl Ru3(CO)12 with water leads to the formation of polynuclear hydrides α-H4Ru4(CO)12, α-H2Ru4(CO)13; the corresponding reaction with Os3(CO)12 yields the complexes (H)(OH)Os3(CO)10, H2Os4(CO)13, H4Os4(CO)12, H2Os5(CO)16, H2Os5(CO)15, H2Os6(CO)18 and H2Os7(CO)19C.  相似文献   

18.
The reaction of H2Os3(CO)10 with CF3CN in hexane at 80°C leads to two isomeric products. The isomer constituting the major product contains a 1,1,1-tri-fluoroethylidenimido ligand which bridges one edge of the Os3 triangle via the nitrogen, atom and may be formulated as (μ-H)Os3(CO)10(μ-NC(H)CF3) (I). The minor product, formulated as (μ-H)Os3(CO)10(μ-η2-HNCCF3) (II), contains a 1,1,1-trifluoroacetimidoyl ligand which is also edge-bridging, being N-bonded to one Os atom and C-bonded to the other. Thermolysis of I and II in solution results in loss of a CO group in each case to give (μ-H)Os3(CO)9?32-NC(H)CF3) (III) and (μ-H)Os3(CO)932-HNCCF3) (IV), respectively, which, it is proposed, are structurally related to I and II, but with the CN group coordinated also to the third Os atom in place of a CO group. In the case of IV this proposal has been confirmed by an X-ray crystallographic analysis. The compound crystallises in space group C2/c with a = 14.258(7), b = 13.486(10), c = 18.193(8) Å, β = 92.68(4)°, and Z = 8. The structure was solved by a combination of direct methods and Fourier difference techniques, and refined by full-matrix least squares to R = 0.054 for 2489 unique observed diffractometer data. Reaction of I with Et3P gives a 1 : 2 adduct which is formulated as (μ-H)Os3(CO)10[μ-N?C(H)(CF3)PEt3] (V) on the basis of NMR evidence.  相似文献   

19.
Preparation and Properties of (CF3)2EMn(CO)5 (E ? P, As) The complexes (CF3)2EMn(CO)5 (E ? P, As) are formed by the reaction of E2(CF3)4 with HMn(CO)5. They can be converted quantitatively to the binuclear compounds [Mn(CO)4E(CF3)2]2 in a thermal (E ? P) or photochemical (E ? P, As) process. u. v. irradiation of a 1:1 mixture gives the mixed derivative Mn2(CO)8As(CF3)2P(CF3)2 together with the symmetrical systems. The Mn? E bond is less reactive with HBr and Me3SnBr than the M? E bond in derivatives of the type Me3ME(CF3)2 (M ? Si, Ge, Sn; Me ? CH3). The terminal (CF3)2E groups are found to be strong π-acceptor ligands.  相似文献   

20.
Abstract

The reaction of (CF3)2P-P(CF3)2 with [Ru3(CO)12] yielded compounds : [Ru14(CO)13{μ-P(CF3)2)2] (1), [Ru4(CO)14{μ-P(CF3)2}2] (2), and [Ru4(CO)11{μ-P(CF3)2}4] (3); reaction with [μ-H)4Ru4(CO)12] yielded (1) and [(μ-H)3Ru4(CO)12{μ-P(CF3)2}] (4). The reaction of (CF3)2PH with [Ru3(CO)12] yielded compounds (1) and (4) and compounds (1) and (2) using cluster : ligand ratios of 1:1 and 1:2 respectively. All the compounds have been characterised by X-ray crystallography; a schematic diagram of their structures is shown in Figure 1. The fluxional behaviour of the hydrides in (4) was studied using variable temperature 1H NMR spectroscopy (see Figure 2). The result of this study was used in the assignment of hydride positions of (4) in the solid state.  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