首页 | 官方网站   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.
The potential of nuclear magnetic resonance (NMR) technique in probing the structure of porous systems including carbon nanostructures filled with inert gases is analysed theoretically using accurate calculations of neon (21Ne) nuclear magnetic shieldings. The CBS estimates of 21Ne NMR parameters were performed for single atom, its dimer and neon interacting with acetylene, ethylene and 1,3‐cyclopentadiene. Several levels of theory including restricted Hartree‐Fock (RHF), Møller‐Plesset perturbation theory to the second order (MP2), density functional theory (DFT) with van Voorhis and Scuseria's t‐dependent gradient‐corrected correlation functional (VSXC), coupled cluster with single and doubles excitations (CCSD), with single, doubles and triples included in a perturbative way (CCSD(T)) and single, doubles and tripes excitations (CCSDT) combined with polarization‐consistent aug‐pcS‐n series of basis sets were employed. The impact of neon confinement inside selected fullerene cages used as an NMR probe was studied at the RHF/pcS‐2 level of theory. A sensitivity of neon probe to the proximity of multiple CC bonds in C2H2, C2H4, C5H6 and inside C28, C30, C32, C34 and C60 fullerenes was predicted from 21Ne NMR parameters' changes. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

2.
High resolution NMR spectroscopy was applied to precisely determine the 83Kr nuclear magnetic dipole moment on the basis of new results available for nuclear magnetic shielding in krypton and helium‐3 atoms. Small amounts of 3He as the solutes and 83Kr as the buffer gas were observed in 3He and 83Kr NMR spectra at the constant external field, B0 = 11.7578 T. In each case, the resonance frequencies (νHe and νKr) were linearly dependent on the density of gaseous solvent. The extrapolation of experimental points to the zero density of gaseous krypton allowed for the evaluation of both resonance frequencies free from intermolecular interactions. By combining these measurements with the recommended 83Kr chemical shielding value, the nuclear magnetic moment could be determined with much better precision than ever before, μ(83Kr) = ?0.9707297(32)μN, with the improvement due to the greater accuracy of the spectral data. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

3.
tBuN = VCl2 · 1,2-Dimethoxoethane, a Precursor in the Synthesis of Binuclear Diamagnetic tert -Butylimidovanadium(IV) Compounds Syntheses of the paramagnetic tert-butylimidovanadium(IV) complexes tBuN = VCl2 · DME ( 7 ), tBuN = VCl2 · 2 L (L = 1,4-dioxane, thf, PMe3, PEt3, pyridine) and tBuN = VBr2 · DME are described; the free Lewis acids has been found by mass spectroscopy to be the binuclear compounds [(μ-NtBu)2V2Cl4] und [(μ-NtBu)2V2Br4]. 7 reacts with LiOR, LiOAr and LiNR2 forming binuclear diamagnetic tert-butylimidovanadium(IV) compounds: [(μ-NtBu)2V2Cl2(OiPr)2] ( 18 ), [(μ-NtBu)2V2(OR)4], [(μ-NtBu)2V2Cl2(OAr)2], [(μ-NtBu)2V2(OAr)4] and [(μ-NtBu)2V2Cl2(NR2)2]. In additional experiments the complexes [(μ-NtBu)2V2(CH2CMe3)2(OAr)2], [(μ-NtBu)2V2Me2(NR2)2], [(μ-NtBu)2V2Cl4] and tBuN = V(OAr)3 has been prepared. All compounds obtained are characterized by spectroscopic methods (MS; 1H, 13C, 51V NMR), [(μ-NtBu)2V2Cl2(NtBuSiMe3)2] ( 21 ) by single crystal x-ray diffraction. For 18 the presence of cis/trans isomeres was shown by NMR spectroscopy. The 51V NMR spectra of the binuclear diamagnetic vanadium (IV) compounds are discussed.  相似文献   

4.
Reactions of rhodium(III) halides with multidentate N,S‐heterocycles, (LH3) 1,3,5‐tris(benzimidazolyl)benzene (L1H3; 1 ), 1,3,5‐tris(N‐methylbenzimidazolyl) benzene (L2H3; 2 ) and 1,3,5‐tris(benzothiazolyl)benzene (L3H3; 3 ), in the molar ratio 1:1 in methanol–chloroform produced mononuclear cyclometallated products of the composition [RhX2(LH2)(H2O)] (X = Cl, Br, I; LH2 = L1H2, L2H2, L3H2). When the metal to ligand ( 1–3 or 1,2,4,5‐tetrakis(benzothiazolyl)benzene [L4H2; 4 ]) molar ratio was 2:1, the reactions yielded binuclear complexes of the compositions [Rh2Cl5(LH2)(H2O)3] (LH2 = L1H2, L2H2, L3H2) and [Rh2X4(L4)(H2O)2] (X = Cl, Br, I). Elemental analysis, IR and 1H nuclear magnetic resonance (NMR) chemical shifts supported the binuclear nature of the complexes. Cyclometallation was detected by conventional 13C NMR spectra that showed a doublet around ~190 ppm. Cyclometallation was also detected by gradient‐enhanced heteronuclear multiple bond correlation (g‐HMBC) experiment that showed cross‐peaks between the cyclometallated carbon and the central benzene ring protons of 1–3 . Cyclometallation was substantiated by two‐dimensional 1H? 1H correlated experiments (gradiant‐correlation spectroscopy and rotating frame Overhauser effect spectroscopy) and 1H? 13C single bond correlated two‐dimensional NMR experiments (gradient‐enhanced heteronuclear single quantum coherence). The 1H? 15N g‐HMBC experiment suggested the coordination of the heterocycles to the metal ion via tertiary nitrogen. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

5.
Summary Six-coordinate dioxomolybdenum(VI) complexes of the type [MoO2(SCZ)2] (HSCZ=semicarbazone) were prepared and characterized by elemental analyses, molecular weight determinations and conductance and magnetic measurements. On the basis of electronic, i.r., proton magnetic resonance, and13C nuclear magnetic resonance spectral studies, an octahedral geometry withcis-MoO2 is proposed.  相似文献   

6.
The reactivity of aryl monocarboxylic acids (benzoic, 1- or 2-naphtoic, 4’-methylbiphenyl-4-carboxylic, and anthracene-9-carboxylic acids) as complexing agents for the ethoxide niobium(V) (Nb(OEt)5 precursor has been investigated. A total of eight coordination complexes were isolated with distinct niobium(V) nuclearities as well as carboxylate complexation states. The use of benzoic acid gives a tetranuclear core Nb42-O)4(L)4(OEt)8] (L=benzoate ( 1 )) with four Nb−(μ2-O)−Nb linkages in a square plane configuration. A similar tetramer, 7 , was obtained with 2-naphtoic acid by using a 55 % humid atmosphere synthetic route. Two types of dinuclear brick were identified with one central Nb−(μ2-O)−Nb linkage; they differ in their complexation state, with one bridging carboxylate ([Nb22-O)(μ2-OEt)(L)(OEt)6], with L=1-naphtoate ( 3 ) or anthracene-9-carboxylate ( 5 )) or two bridging carboxylate groups ([Nb22-O)(L)2(OEt)6], with L=4’-methylbiphenyl-4-carboxylic ( 4 ) or anthracene-9-carboxylate ( 6 )). An octanuclear moiety [Nb82-O)12(L)81-L)4−x(OEt)4+x] (with L=2-naphtoate, x=0 or 2; 8 ) was obtained by using a solvothermal route in acetonitrile; it has a cubic configuration with niobium centers at each node, linked by 12 μ2-O groups. The formation of the niobium oxo clusters was characterized by infrared and liquid 1H NMR spectroscopy in order to analyze the esterification reaction, which induces the release of water molecules that further react through oxolation with niobium atoms, in different {Nb2O}, {Nb4O4} and {Nb8O12} nuclearities.  相似文献   

7.
3He, 129Xe and 131Xe NMR measurements of resonance frequencies in the magnetic field B0 = 11.7586 T in different gas phase mixtures have been reported. Precise radiofrequency values were extrapolated to the zero gas pressure limit. These results combined with new quantum chemical values of helium and xenon nuclear magnetic shielding constants were used to determine new accurate nuclear magnetic moments of 129Xe and 131Xe in terms of that of the 3He nucleus. They are as follows: μ(129Xe) = ?0.7779607(158)μN and μ(131Xe) = +0.6918451(70)μN. By this means, the new ‘helium method’ for estimations of nuclear dipole moments was successfully tested. Gas phase NMR spectra demonstrate the weak intermolecular interactions observed on the 3He and 129Xe and 131Xe shielding in the gaseous mixtures with Xe, CO2 and SF6. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

8.
The novel ternary polyphosphides M4P21I (M = K, Rb) have been synthesized from the elements in single crystalline form, representing further examples for the formation of mixed crystals between simple salts and binary phosphides. They form as ruby‐red platelets and dark‐red prisms, respectively, and are only slightly sensitive to moisture and oxygen. The compounds are isotypic (Ccmm (no 63); Z = 4; oP104; K4P21I: a = 12.853Å; b = 21.795Å; c = 9.748Å; 1168 hkl, R = 0.033; Rb4P21I: a = 13.281Å; b = 21.868Å; c = 9.771Å; 777 hkl, R = 0.053) and feature corrugated 2D networks formed from two different types of polymerized P7 units. The networks form large cavities filled by M+ and I ions. Zigzag chains of condensed trigonal M6 prisms, centered by the I anions, separate the polyphosphide nets. The mean homoatomic P‐P bond length (d = 2.216Å) corresponds to a P‐P single bond. However, the individual P‐P distances vary with position and function (2.126 ‐ 2.247Å) and these are compared with those of the isolated P21‐3 anion.  相似文献   

9.
Nuclear magnetic resonance is defined as a quantitative spectroscopic tool that enables a precise determination of the number of substances in liquids as well as in solids. There is few report demonstrating the application of NMR in the quantification of avermectin B1a (AVB1a); here, a proton nuclear magnetic resonance spectroscopy (1H NMR) using benzene [1‐methoxy‐4‐(2‐nitroethyl) (PMN)] as an internal standard and deuterochloroform as an NMR solvent was tested for the quantitative determination of AVB1a. The integrated signal of AVB1a at 5.56 ppm and the signal of PMN at 8.14 ppm in the 1H NMR spectrum were used for quantification purposes. Parameters of specificity, linearity, accuracy, precision, intermediate precision, range, limit of detection (LOD), limit of quantification (LOQ), stability and robustness were validated. The established method was accurate and precise with good recovery (98.86%) and relative standard deviation (RSD) of assay (0.34%) within the linearity of the calibration curve ranging from 5.08 to 13.58 mg/ml (R2 = 0.9999). The LOD and LOQ were 0.009 and 0.029 mg/ml, which indicated the excellent sensitivity of the method. The stability of the method was testified by a calculated RSD of 0.11%. The robustness was testified by modification of four different parameters, and the differences among each parameter were all less than 0.1%. Comparing with the assay described by the manufacturer of avermectin tablets, there was no significant difference between the assay obtained by HPLC and quantitative NMR (qNMR), which indicated qNMR was a simple and efficient method for the determination of AVB1a in commercial formulation products. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

10.
The miscibility of poly(hydroxyether of bisphenol A) (phenoxy) and poly(N-vinyl pyrrolidone) (PVP) was investigated by differential scanning calorimetry (DSC) and high-resolution solid-state nuclear magnetic resonance (NMR) techniques. The DSC studies showed that the phenoxy/PVP blends have a single, composition-dependent glass transition temperature (Tg). The S-shaped Tg-composition curve of the phenoxy/PVP blends was reported, which is indicative of the strong intermolecular hydrogen-bonding interactions. To examine the miscibility of the system at molecular level, high-resolution solid-state 13C nuclear magnetic resonance (NMR) technique was employed. Upon adding phenoxy to system, the chemical shift of carbonyl carbon resonance of PVP was observed to shift downfield by 1.6 ppm in the 13C cross-polarization (CP)/magic angle spinning (MAS) together with the high-power dipolar decoupling (DD) spectra when the concentration of phenoxy is 90 wt %. The observation was responsible for the formation of intermolecular hydrogen bonding. The proton spin-lattice relaxation time T1(H) and the proton spin-lattice relaxation time in the rotating frame T(H) were measured as a function of the blend composition. The T1(H) result was in good agreement with the thermal analysis, i.e., the blends are completely homogeneous on the scale of 20 ∼ 30 nm. The six results of T(H) further indicated that the blends were homogeneous on the scale of 40 ∼ 50Å. © 1998 John Wiley & Sons, Inc. J. Polym. Sci. B Polym. Phys. 36: 2291–2300, 1998  相似文献   

11.
Previously reported ferromagnetic triangles (NnBu4)2[Cu3(μ3-Cl)2(μ-4-NO2-pz)3Cl3] ( 1 ), (PPN)2[Cu3(μ3-Cl)2(μ-pz)3Cl3] ( 2 ), (bmim)2[Cu3(μ3-Cl)2(μ-pz)3Cl3] ( 3 ) and newly reported (PPh4)2[Cu3(μ3-Cl)2(μ-4-Ph-pz)3Cl3] ( 4 ) were studied by magnetic susceptometry, electron paramagnetic resonance (EPR) spectroscopy and ab initio calculations to assess the origins of their ferromagnetism and of the magnetic anisotropy of their ground S=3/2 state (PPN+=bis(triphenylphosphine)iminium, bmim+=1-butyl-3-methylbenzimidazolium, pz=pyrazolate). Ab initio studies revealed the d character of the magnetic orbitals of the compressed trigonal bipyramidal copper(II) ions. Ferromagnetic interactions were attributed to weak orbital overlap via the pyrazolate bridges. From the wavefunctions expansions, the ratios of the magnetic couplings were determined, which were indeterminate by magnetic susceptometry. Single-crystal EPR studies of 1 were carried out to extend the spin Hamiltonian with terms which induce zero-field splitting (zfs), namely dipolar interactions, anisotropic exchange and Dzyaloshinskii–Moriya interactions (DMI). The data were treated through both a giant-spin model and through a multispin exchange-coupled model. The latter indicated that ≈62 % of the zfs is due to anisotropic and ≈38 % due to dipolar interactions. The powder EPR data of all complexes were fitted to a simplified form of the multispin model and the anisotropic and dipolar contributions to the ground state zfs were estimated.  相似文献   

12.
The measurement of the magnitude and sign of 2J(C,H) couplings offers a reliable way to determine the absolute configuration at a carbon center in a fixed cyclic system. A decrease of the dihedral angle ? in the O—CA—CB—H fragment always leads to a change of the 2J(CA,HB) coupling to more negative values, independent of the type and position of substituents at the two carbon centers. The orientations of the two substituents at C‐3 of the epimeric pair 1 and 2 were determined unambiguously through the measurement of the geminal coupling constants between C‐3 and the hydrogen atoms at C‐2 and C‐4. In particular, 2J(C‐3,H‐2ax) with ?1.5 Hz, ? = 174° in 1 and ?6.6 Hz, ? = 47° in 2 , and 2J(C‐3,H‐4) with +1.5 Hz, ? = 175° in 1 and ?4.7 Hz, ? = 49° in 2 showed the greatest differences between the two epimers. Both couplings therefore allow the determination of the absolute configuration at C‐3. It should be noted, however, that the size of the coupling constants can be different for dihedral angles of nearly identical size, when there are different numbers of electronegative substituents on the two coupling pathways, i.e. no O‐substituent at C‐2, but one axial O‐substituent at C‐4. It becomes clear that it is not sufficient to measure the magnitude of 2J coupling constants only, but that the sign of the geminal coupling is needed to identify the absolute configuration at a chiral center. The coupling of C‐3 with H‐2eq is not useful for the determination of the configuration at C‐3, as the similarity of the dihedral angles ? (O—C‐3—C‐2—H‐2eq) (57° in 1 and 70° in 2 ) leads to identical coupling constants (?6.1 Hz) for both epimers. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

13.
To understand the effects of picolinamide-like ligands on reaction equilibrium, a series of picolinamide derivatives were synthesized and the interactions between the diperoxovanadate complex [OV(O2)2L]? (L = D2O or HOD, abbr. bpV) and picolinamide ligands in solution were explored using multinuclear (1H, 13C, and 51V) magnetic resonance, COSY, and HSQC in 0.15 mol L?1 NaCl ionic medium for mimicking physiological conditions. Formation constants among the picolinamide-like ligands are N-(1-hydroxypropan-2-yl)-picolinamide ≈N-(2-hydroxypropyl)-picolinamide>N-(3-hydroxypropyl)-picolinamide>N-propyl-picolinamide. The substituting group influences the equilibrium by electronic effects. The interactions result in a series of new seven-coordinate diperoxovanadate species [OV(O2)2L′]? (L′ = picolinamide-like ligands).  相似文献   

14.
Blends of poly(vinyl chloride) (PVC) with Poly(N‐vinyl pyrrolidone) (PVP) were investigated by Fourier infrared spectroscopy (FTIR) and high‐resolution solid‐state 13C cross‐polarization/magic angle spinning (CP/MAS) nuclear magnetic resonance (NMR) spectroscopy. The intermolecular interactions between PVP and PVC are weaker than the self‐association of PVP and the inclusion of the miscible PVC results in the decreased self‐association of PVP chains, which was evidenced by the observation of high‐frequency shift of amide stretching vibration bands of PVP with inclusion of PVC. This result was further substantiated by the study of 13C CP/MAS spectra, in which the chemical shift of carbonyl resonance of PVP was observed to shift to a high field with inclusion of PVC, indicating that the magnetic shielding of the carbonyl carbon nucleus is increased. The proton spin‐lattice relaxation time in the laboratory frame (T1 (H)) and the proton spin‐lattice relaxation time in the rotating frame (T(H)) were measured as a function of the blend composition to give the information about phase structure. It is concluded that the PVC and PVP chains are intimately mixed on the scale of 20–30Å. © 1999 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 37: 2412–2419, 1999  相似文献   

15.
The development of nuclear spins hyperpolarization, and the search for molecules that can be efficiently hyperpolarized is an active area in nuclear magnetic resonance. In this work we present a detailed study of SABRE SHEATH (signal amplification by reversible exchange in shield enabled alignment transfer to heteronuclei) experiments on 15N2-azobenzene. In SABRE SHEATH experiments the nuclear spins of the target are hyperpolarized through transfer of spin polarization from parahydrogen at ultralow fields during a reversible chemical process. Azobenzene exists in two isomers, trans and cis. We show that all nuclear spins in cis-azobenzene can be efficiently hyperpolarized by SABRE at suitable magnetic fields. Enhancement factors (relative to 9.4 T) reach up to 3000 for 15N spins and up to 30 for the 1H spins. We compare two approaches to observe either hyperpolarized magnetization of 15N/1H spins, or hyperpolarized singlet order of the 15N spin pair. The results presented here will be useful for further experiments in which hyperpolarized cis-15N2-azobenzene is switched by light to trans-15N2-azobenzene for storing the produced hyperpolarization in the long-lived spin state of the 15N pair of trans-15N2-azobenzene.  相似文献   

16.
The reactivity of the complexing anthracene-9-carboxylate ligand has been investigated with a niobium(IV) tetrachloride precursor (NbCl4 ⋅ 2THF) in isopropanol solvent. This resulted in the crystallization of a molecular assembly containing two distinct {Nb12O21} cores surrounded by multiple isopropanolate and anthracenoate ligands. The compound is formulated [Nb12(3-O)3(μ-O)18(C15H9O2)8(OiPr)10] ⋅ [Nb123-O)2(μ-O)19(C15H9O2)8(OiPr)10] illustrating the two different dodecameric oxo-clusters, for which the niobium(IV) precursor was oxidized in the niobium(V) state during the reactional process. The two distinct {Nb12O21} units mainly differs by the environment of the niobium centers, which exhibits unexpected five-fold coordination (square pyramid) for some of them, together with the classical six-fold coordination (octahedron) as usually found for niobium(V). In the crystallization process, the. IR spectroscopy was used to analyze the esterification reaction occurring between the anthracene acid an isopropanolate ligands responsible of the production of water used in the oxo-condensation of the niobium centers. 93Nb Solid state NMR was tentatively used to assess the occurrence of the different niobium environments.  相似文献   

17.
Ionic conduction in highly designable and porous metal–organic frameworks has been explored through the introduction of various ionic species (H+, OH, Li+, etc.) using post-synthetic modification such as acid, salt, or ionic liquid incorporation. Here, we report on high ionic conductivity (σ>10−2 S cm−1) in a two-dimensionally (2D)-layered Ti-dobdc (Ti2(Hdobdc)2(H2dobdc), H4dobdc: 2,5-dihydroxyterephthalic acid) via LiX (X=Cl, Br, I) intercalation using mechanical mixing. The anionic species in lithium halide strongly affect the ionic conductivity and durability of conductivity. Solid-state pulsed-field gradient nuclear magnetic resonance (PFG NMR ) verified the high mobility of H+ and Li+ ions in the temperature range of 300–400 K. In particular, the insertion of Li salts improved the H+ mobility above 373 K owing to strong binding with H2O. Furthermore, the continuous increase in Li+ mobility with temperature contributed to the retention of the overall high ionic conductivity at high temperatures.  相似文献   

18.
Copolymers of 3,3-bis-hydroxymethyloxetane, BHMO, 3-metyl-3-hydroxymethyloxetane, MHMO, or with 3-ethyl-3-hydroxymethyloxetane, EHMO, monomer units were characterized by x-ray fiber diffraction, differential scanning calorimetry and 13C solid-state nuclear magnetic resonance (NMR). The copolymers are statistically random and crystalline throughout the range of compositions. Both P(BHMO) and P(MHMO) appear to crystallize in the same crystal form. The fiber repeat indicates a planar zigzag backbone conformation, c(fiber axis) = 4.77 ± 0.03 Å. Similarities in the x-ray fiber diagrams as well as a linear dependence of Tm with composition of copolymer with no change in fiber diagrams indicates isomorphism, a phenomenon in which the random substitution of MHMO monomeric units into the crystalline lattice of P(BHMO) occurs without hindering crystallization of the resulting copolymer. © 1994 John Wiley & Sons, Inc.  相似文献   

19.
A survey of the use of 187Os satellite subspectra in 1H and 31P{1H} spectra of triosmium carbonyl clusters is reported. By varying evolution delays in HMQC spectra of [Os3(µ‐H)2(CO)10] we have selectively extracted the values for 1J(Os,H) and 2J(Os,H), respectively. An analysis of the principal modes of phosphine coordination in triosmium clusters demonstrates that 31P{1H}187Os satellite subspectra are diagnostic for equatorial coordination [1J(Os,P) = 211–223 Hz] or for axial coordination (perpendicular to the plane of the cluster) [1J(Os,P) ≈ 147 Hz]. Chelating and bridging diphosphines yield 187Os satellite subspectra which are the sum of A2X and AA′X spin systems. If significant P–P coupling is present, the AA′X component requires simulation. All observed 2J(Os,P) trans‐equatorial couplings fall in the range 38–65 Hz. Copyright © 2001 John Wiley & Sons, Ltd.  相似文献   

20.
205Tl longitudinal relaxation rate measurements were performed on several thallium(III) complexes with the composition Tl(OH)n(H2O)6?n(3?n)+ (n = 1,2), Tl(Cl)n(H2O)m?n(3?n)+, Tl(Br)n(H2O)m?n(3?n)+ (m = 6 for n = 1–2, m = 5 for n = 3, m = 4 for n = 4), Tl(CN)n(H2O)m?n(3?n)+ (m = 6 for n = 1–2, m = 4 for n = 3–4) in aqueous solution, at different magnetic fields and temperatures. 13C and 2D isotopic labelling and 1H decoupling experiments showed that the contribution of the dipolar relaxation path is negligible. The less symmetric lower complexes (n < 4) had faster relaxation rate dominantly via chemical shift anisotropy contribution which depended on the applied magnetic field: T1 values are between 20 and 100 ms at 9.4 T and the shift anisotropy is Δσ = 1000–2000 ppm. The tetrahedral complexes, n = 4, relax slower; their T1 is longer than 1 s and the spin–rotation mechanism is probably the dominant relaxation path as showed by a temperature dependence study. In the case of the TlCl4? complex, presumably a trace amount of TlCl52? causes a large CSA contribution, 300 ppm. Since the geometry and the bond length for the complexes in solution are known from EXAFS data, it was possible to establish a correlation between the CSA parameter and the symmetry of the complexes. The relaxation behaviour of the Tl–bromo complexes is not in accordance with any known relaxation mechanism. Copyright © 2002 John Wiley & Sons, Ltd.  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-23

京公网安备 11010802026262号