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 46 毫秒
1.
A semiempirical approach is used to fix the α value for use in the extraatomic regions in multiple-scattering (MS Xα) calculations which retain the muffin-tin treatment of the potential. Such a “molecular” α value for an atom is determined by requiring the corresponding homonuclear diatomic molecule to have its minimum at the experimentally determined equilibrium separation; hence they are called α R. Molecular α R values are determined for the ground state Li2 and F2 molecules and are tested in a calculation of the ground state LiF potential curve. We find a binding energy at the calculated equilibrium separation to be within 1% of the experimental value. The LiF curve based entirely on the ordinary atomic α values is substantially inferior. The present MT Xα R approach appears to be competitive with others which are intended to improve the muffin-tin version of MS Xα calculations.  相似文献   

2.
The hydride complex K[(η5‐C5H5)Mn(CO)2H] reacted with a range of dihalo(organyl)boranes X2BR (X = Cl, Br; R = tBu,Mes, Ferrocenyl) to give the corresponding borane complexes[(η5‐C5H5)Mn(CO)2(HB(X)R)]., The presence of a hydride in bridging position between manganese and boron was deduced from 11B decoupled 1H NMR spectra. Additionally, the structure of the tert‐butyl borane complex was confirmed by single‐crystal X‐ray diffraction.  相似文献   

3.
Dicopper dicarboxylates [(R3P)mCuXCu(PR3)m] ( 5a , X = O2CCO2, R = Ph, m = 2; 5b , X = O2CCO2, R = nBu, m = 3) were prepared by treatment of [Cu2O] ( 1a ) with HO2CCO2H ( 2a ) in presence of PR3 ( 4a , R = Ph; 4b , R = nBu). A further synthesis approach to mono‐ and dicopper dicarboxylates is given using an electrolysis cell equipped with Cu electrodes and charged with acids H2X and phosphanes R3P. Without addition of a base mononuclear [(nBu3P)mCuXH] ( 6a , m = 3, XH = O2CCO2H, 6b , m = 3, XH = O2CCH2CO2H, 6c , m = 3, XH = O2CCH2CH2CO2H, 6d , m = 2, XH = O2C‐2‐C5H4N‐6‐CO2H) was formed, whereas in presence of NEt3 ( 3 ), the dicopper systems [(R3P)mCuXCu(PR3)m] ( 5a , X = O2CCO2, R = Ph, m = 2; 5b , X = O2CCO2, R = nBu, m = 3; 5c , X = O2CCH2CO2, R = nBu, m = 3; 5d , X = O2CCH2CH2CO2, R = nBu, m = 3; 5e , X = O2C‐2‐C5H4N‐6‐CO2, R = nBu, m = 3) were produced. When 6a reacted with [(tmeda)Zn(nBu)2] ( 7 ), trimetallic [(tmeda)Zn((nBu3P)3CuO2CCO2)2] ( 8 ) was accessible. In this heterobimetallic complex the Zn(tmeda) unit spans two CuO2CCO2 entities. The molecular structures of 5a , 6a and 6d in the solid state were determined by single X‐ray structure analysis. Complexes 5a and 6a are monomers, whereas 6d creates in the solid state a linear open chain coordination polymer by hydrogen bridge formation. Characteristic for 6d is the somewhat distorted trigonal bipyrimidal arrangement around the copper atom with the nBu3P ligands in axial and the C5H3NCO2H oxygen and nitrogen atoms in equatorial positions. In 5a the oxalate connectivity binds in a μ‐1,2,3,4 fashion being part of a planar Cu2(oxalate) core. TG studies of several mono‐ and dicopper dicarboxylates were carried out. Release of the PR3 ligands is recognized and the remaining Cu‐(di)carboxylate unit decomposes to afford elemental copper and CO2. The deposition of copper onto pieces of PVD‐Cu oxidized silicon wafers by applying the spin‐coating process and using 5c and 5d as precursors is discussed.  相似文献   

4.
Three different H/D isotope effect in nine H3XH(D)YH3 (X = C, Si, or Ge, and Y = B, Al, or Ga) hydrogen‐bonded (HB) systems are classified using MP2 level of multicomponent molecular orbital method, which can take account of the nuclear quantum nature of proton and deuteron. First, in the case of H3CH(D)YH3 (Y = B, Al, or Ge) HB systems, the deuterium (D) substitution induces the usual H/D geometrical isotope effect such as the contraction of covalent R(C? H(D)) bonds and the elongation of intermolecular R(H(D)Y) and R(CY) distances. Second, in the case of H3XH(D)YH3 (X = Si or Ge, and Y = Al or Ge) HB systems, where H atom is negatively charged called as charge‐inverted hydrogen‐bonded (CIHB) systems, the D substitution leads to the contraction of intermolecular R(H(D)Y) and R(XY) distances. Finally, in the case of H3XH(D)BH3 (X = Si or Ge) HB systems, these intermolecular R(H(D)Y) and R(XY) distances also contract with the D substitution, in which the origin of the contraction is not the same as that in CIHB systems. The H/D isotope effect on interaction energies and spatial distribution of nuclear wavefunctions are also analyzed. © 2015 Wiley Periodicals, Inc.  相似文献   

5.
Zusammenfassung -Substituierte -Acylvinylphosphonate3 mitE-Konfiguration [R 2CO-CH=C(R 1)-P(O)(OR)2], werden in guten Ausbeuten durchWittig-Reaktion von Acylphosphonsäureestern1 [R 1CO-P(O)(OR)2,R 1=Alkyl oder Aryl] mit (2-Oxoalkyliden)triphenylphosphoranen2 [R 2CO-CH=PPh 3,R 2=Alkyl, O-Alkyl oder CH2 X (X=Br, OMe, CO2 Et)] erhalten.
A convenient route to -substituted dialkyl (E)-3-oxo-1-alkenylphosphonates
-Substituted dialkyl (E)--acylvinylphosphonates [R 2CO-CH=C(R 1)-P(O)(OR)2,3], are easily obtained in good yields byWittig-reaction of dialkyl acylphosphonates1 [R 1CO-P(O)(OR)2,R 1=alkyl or aryl) with 2-oxoalkylidene triphenylphosphoranes2 [R 2CO-CH=PPh 3,R 2=alkyl, O-alkyl and CH2 X (X=Br, OMe, CO2 Et)].
  相似文献   

6.
The structures of α-X-cyclopropyl and α-X-isopropyl radicals (X = H, CH3, NH2, OH, F, CN, and NC) are reported at the RHF 3-21G level of theory. The isopropyl radicals are pyramidal with out-of-plane angles varying from 12° (X = CN) to 39° (X = NH2), and barriers to inversion ranging from 0.4 kcal/mol (X = H) to 4.0 kcal/mol (X = NH2). The cyclopropyl radicals have larger out-of-plane angles, from 39.9° (X = CN) to 49.4° (X = NH2), and their barriers to inversion, which increase with the inclusion of polarization functions, vary from 5.5 kcal/mol (X = H) to 16.7 kcal/mol (X = F). In both types of radicals the amino group is the most stabilizing substituent, while the α-fluoro has little effect. The β-fluoro group is weakly destabilizing in the cyclopropyl radical. The strain energies of the cyclopropyl radicals (36–43 kcal/mol) are compared with those of similarly substituted anions, cations, and cyclopropanes.  相似文献   

7.
CCSD(T) calculations have been used for identically nucleophilic substitution reactions on N‐haloammonium cation, X? + NH3X+ (X = F, Cl, Br, and I), with comparison of classic anionic SN2 reactions, X? + CH3X. The described SN2 reactions are characterized to a double curve potential, and separated charged reactants proceed to form transition state through a stronger complexation and a charge neutralization process. For title reactions X? + NH3X+, charge distributions, geometries, energy barriers, and their correlations have been investigated. Central barriers ΔE for X? + NH3X+ are found to be lower and lie within a relatively narrow range, decreasing in the following order: Cl (21.1 kJ/mol) > F (19.7 kJ/mol) > Br (10.9 kJ/mol) > I (9.1 kJ/mol). The overall barriers ΔE relative to the reactants are negative for all halogens: ?626.0 kJ/mol (F), ?494.1 kJ/mol (Cl), ?484.9 kJ/mol (Br), and ?458.5 kJ/mol (I). Stability energies of the ion–ion complexes ΔEcomp decrease in the order F (645.6 kJ/mol) > Cl (515.2 kJ/mol) > Br (495.8 kJ/mol) > I (467.6 kJ/mol), and are found to correlate well with halogen Mulliken electronegativities (R2 = 0.972) and proton affinity of halogen anions X? (R2 = 0.996). Based on polarizable continuum model, solvent effects have investigated, which indicates solvents, especially polar and protic solvents lower the complexation energy dramatically, due to dually solvated reactant ions, and even character of double well potential in reactions X? + CH3X has disappeared. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

8.
A linear relationship between the half-wave reduction potentials of α,β-unsaturated carbonyl compounds R–CHCH–COX and the Hammett σp values of R and X is proposed: E1/2=−1.341σp(X)σp(R)+1.123σp(X)+1.746σp(R)−1.694. A linear relationship is also observed for the LUMO's energy values, the absolute chemical hardness η, the chemical potential μ, the electrophilicity power ω, or the polarisation of the ethylenic double bond with the Hammett σp values of R and X.  相似文献   

9.
The short-bite ligands CH2(PR 2)2 or CH(PR 2)3 (R = Me, Ph),RN(PX 2)2 (R=H, Me, Et;X = F, OR (R= Me, Et, i-Pr, Ph), Ph),RE(CH2 ER2)2 (E = P, As;R = Me, Ph ), Ph2 P(2-C5H4N) and related species are particularly versatile for the synthesis of di- and polynuclear complexes which frequently possess metal-metal bonds. In addition to homometallic products, these ligands often permit the directed synthesis of heterometallic complexes. Selected aspects of the chemistry of these complexes are also reviewed.  相似文献   

10.
It was observed that adsorption of trace amounts of organic anions from aqueous permeants causes a logarithmic decrease in permeability of the microporous membrane until all of the available adsorption sites along the internal perimeters of the microcapillary inlets are filled. Thereafter, impedance appears to increase almost linearly with continued water flow reflecting the increase in depth of these perimeter areas saturated by uniform anion adsorption. The magnitude of this adsorption effect caused by anions of the homologous series H(CH2)nCO2 is proportional to n from n = 1 to n = 17, which is consistent with expectation on the basis of Traube's Rule. The total equivalent methylene impedance, X, caused by adsorption of other organic anions, such as R1R2CO2, is equal to the sum of its segments R1 and R2 linked in series to CO2, i.e., XT = X1 + X2. This is not true, however, for molecules such as R1AR2CO2, where A is a nucleophilic atom or group of atoms that can serve as a second adsorption site. Apparently, the segment, R2, is convoluted or held flat to the surface by the anchor points A and CO2, and only the segments R1 extend full-length above the plane of anchor points, so that XT is about equal to X1.  相似文献   

11.
Treatment of the dimeric gallium hydrazide [Me2Ga‐N(2‐Ad)‐NC5H10]2 ( 5 ) with Me2GaH resulted in the formation of an adduct 6 by Ga–N bond cleavage and coordination of the metal hydride via a Ga–N and a 3c–2e Ga–H–Ga bond. This reaction reflects the typical behavior of frustrated Lewis pairs. Reactions with heterocumulenes R–N=C=X (R = Ph, CMe3, Dipp, X = O; R = Ph, X = S) or X=C=X (X = O, S) resulted in the formation of the cyclic Ga–N insertion products Me2Ga–N(R)C(O)N(2‐Ad)‐NC5H10 ( 7a – c ), Me2GaS2C‐N(2‐Ad)‐NC5H10 ( 8 ) or Me2GaX2C‐N(2‐Ad)‐NC5H10 [X = O ( 9 ); S ( 10 )] in moderate to good yields. Three different structural motifs were observed in the solid state: Five‐membered GaNCN2 heterocycles with exocyclic C=O bonds for compounds 7a – c , four‐membered GaSCN or GaSCS heterocycles for compounds 8 and 9 (chelating coordination of the Ga atoms by SCN and CS2 ligands) and an eight‐membered (GaOCO)2 heterocycle for the dimeric CO2 insertion product 10 . Treatment of 5 with PhCN or Ph2CO resulted in a completely different reaction and afforded a dimeric Ga imide 11a or an alcoholate 11b . These reactions may start by retro‐hydrogallation with the formation of H10C5N–N=C(C9H14) and Me2GaH and proceed by addition of the metal hydride to the polar multiple bonds of the nitrile or ketone.  相似文献   

12.
A series of twelve 5-trihalomethyl-3-arylisoxazoles was synthesized and screened for antibacterial and antifungal activities. The compounds were synthesized from the cyclondensation of 1,1,1-trihalo-4-alkoxy-3-alken-2-ones [CX 3C(O)C(R 2)=C(R 1)OR, where X = Cl and F; R=Me; R 2=H; R 1=H, Me, F, Cl, Br, and NO2] with hydroxylamine hydrochloride through a rapid one-pot reaction via microwave irradiation. Some of the 5-trihalomethyl-3-arylisoxazoles exhibited good in vitro anti-Cryptococcus activity. Correspondence: Marcos A. P. Martins, Núcleo de Química de Heterociclos (NUQUIMHE), Centro de Ciências Naturais e Exatas, Departamento de Química, Universidade Federal de Santa Maria, 97.105-900 Santa Maria, RS, Brazil.  相似文献   

13.
Reactions of KI, Pr, PrI3, and Os in niobium tubes at 800° yielded black, air- and moisture-sensitive crystals of Kpr6I10Os which were characterized by single crystal X-ray diffraction (orthorhombic, Pnma, a = 15.362(3), b = 13.498(2), c = 14.128(3) Å, Z = 4, R(F)/Rw = 4.4/5.6%). Subsequent parallel experiments also gave, according to Guinier powder pattern data, the isostructural compounds CsPr6I10Fe (a = 15.312(2), b = 13.426(1), c = 14.154(1) Å), CsLa6I10Fe (a = 15.523(2), b = 13.646(2), c = 14.334(1) Å) and CsLa6I10Mn (a = 15.457(4), b = 13.737(2), c = 14.329(2) Å). The important structural feature is the presence of octahedral rare-earth-metal cluster units R6 that are centered by a transition-metal atom Z and bridged and interconnected by halide atoms. The new compounds exhibit the same general pattern of halide connectivity (R6Z)XXXX as do the triclinic compounds R6X10Z. However, the structural arrangement of the metal octahedra is significantly different; they are linked by Ii–i atoms into zigzag chains along [010] and these are interconnected into a three dimensional network by Ii–a atoms to form channels in which the alkali-metal atoms are located. The introduction of alkali-metal atoms into reactions leads to new quaternary compounds with discrete rare-earth-metal clusters centered by transition metals and more open structure frameworks. Measurements of the temperature dependencies of the magnetic susceptibilities for CsLa6I10Fe and CsLa6I10Mn are consistent with expectations for 17- and 16-electron cluster systems, respectively.  相似文献   

14.
On the Reaction of Halomethylphosphonium Halides, [R3PCYnX3–n]X, with Phosphanes, R′3P The results of the reaction of 19 different halomethylphosphonium halides, [R3PCYnX3–n]X (R = Ph, n-Bu, Me2N, Et2N; Y = H, F; X = Cl, Br, I; n = 0–2), with Ph3P, n-Bu3P, and (R2N)3P are presented. As reaction products bisphosphonium salts, [R3P? CYnX2–n? PR′3]X2, and phosphoranylphosphonium salts, [R3P=CY? PR′3]X, or reduced (halo)methyl-phosphonium salts, [R3PCHYnX2–n]X, are obtained. [Ph3PCBrF2]Br and [Bu3PCBrF2]Br react with R′3P by trans-alkylation forming [R′3PCBrF2]Br. The factors influencing the course of the reaction are discussed.  相似文献   

15.
The title compounds have been prepared in water by reaction of SbF3 with dihydrogen phosphates or arsenates and characterized by single crystal X-ray work, IR, Raman, and Mössbauer spectroscopy. They have identical layer structures. Layers of composition [(SbF)XO4] (X = P, As) were formed by sharing four corners between XO4 tetrahedra and SbFO4 pseudooctahedra. The lengths of the terminal Sb---F bond (with the lone pair in a trans-position) and the Sb---O bonds are 192 and 219 pm, respectively. The stacking of the layers and the interlayer distance depend on the cations and the number of intercalated water molecules. In Na(SbF)AsO4 the Na+ ion is coordinated by only two oxygen atoms within 300 pm. Crystal data: Na(SbF)PO4 · 5H2O, monoclinic, P21/m, A = 656.2(5), B = 654.1(5), C = 867.9(3) pm, β = 92.43(1)°, 889 reflections, 81 parameters, R = 0.044, Rw = 0.046. NH4(SbF)PO4 · H2O, tetragonal, I4/m, A = 656.6(3), C = 1439.8(5) pm, 680 reflections, 31 parameters, R = 0.023, Rw = 0.021. Na(SbF)AsO4, tetragonal, P4/ncc, A = 671.8(1), C = 1756.4(4) pm, 1056 reflections, 28 parameters, R = 0.052, Rw = 0.065. NH4(SbF)AsO4 · 3H2O, tetragonal, P4/ncc, A = 683.8(2), C = 1873.0(7) pm, 1194 reflections, 30 parameters, R = 0.042, Rw = 0.050.  相似文献   

16.
The isotypical crystal structures of the mixed valent trihalides PtCl3 and PtBr3 were redetermined by single crystal methods (space group R3¯; trigonal setting; PtCl3: a = 21.213Å, c = 8.600Å, c/a = 0.4054; Z = 36; 1719 hkl; R = 0.035; PtBr3: a = 22.318Å, c = 9.034Å; c/a = 0.4048; Z = 36; 1606 hkl; R = 0.027). A cubic closest packing of X anions forms the basis of an optimized arrangement of cuboctahedrally [Pt6X12] cluster molecules with PtII and enantiomers of helical chains of edge‐condensed [PtX2X4/2] octahedra with PtIV in cis‐Δ‐ and cis‐Λ‐configuration, respectively. The bond lengths vary with the function of the X ligands (d¯(PtII—X) = 2.315 and 2.445Å; d¯(PtII—PtII) = 3.336 and 3.492Å; d(PtIV—X) = 2.286 — 2.417Å and 2.437 — 2.563Å). The PtII atoms are shifted outwards the X12 cuboctahedra by 0.045Å and 0.024Å, respectively. The symmetry governed Periodic Nodal Surface, PNS, perfectly separates the regions of different valencies. Quantum chemical calculations exclude the possible additional interactions between PtII and one of the exo‐ligands of PtIV.  相似文献   

17.
18.
In this study, the optimum extraction conditions for maximum recovery of the content of total phenolics (TPC) and total antioxidant abilities were analyzed for Malus baccata (Linn.) Borkh. using response surface methodology. The effects of ethanol percentage (X1,%), ultrasonic power (X2, W) and extraction temperature (X3, °C) on the total phenolic content (Y1) and antioxidant ability (Y2) were evaluated. A second‐order polynomial model produced a satisfactory fitting of the experimental data with regard to total phenolic content (R2 = 0.9942, P < 0.0001) and antioxidant ability (R2 = 0.9966, P < 0.0001). The optimized conditions were ethanol concentration of 61.0%, ultrasonic power of 308.6 W, extraction temperature of 51.1°C for TPC and 60.5%, 311.4 W, 51.6°C for antioxidant ability, the predicted values agreed well with the experimental values. Results implied that the major phenolic compounds in obtained extracts as chlorogenic acid, quercetin‐3‐gal/glu, quercetin‐3‐xyl/ara, phloretin‐2‐xyloside, quercetin‐3‐ rhamnoside, and phloridzin.  相似文献   

19.
Abstract

IR spectra and molecular structures of ZnL2X2 (L = 2-NH2-4-R-thiazole, R = H (at), CH3; 2-NH2-6-R1 -benzothiazole, R1 = H (abt), CH3, OCH3, OC2H5 or 2-NH2-tetrahydrobenzothiazole; X = Cl, Br, I) have been studied. It was found that the complex character of the far IR spectra and difficulties in v(NH2) interpretation make conclusions regarding structure based only on IR data ambiguous, and in some cases discrepant. Single crystal X-ray data for the complexes with L = at, X = Br, I and L = abt, X = Br show that structures are built up of discrete tetrahedral ZnL2X2 molecules with monodentate ligands coordinated via endo-N atoms. It was found that in the coordination tetrahedra MN2Cl2 (M = Co2+, Zn2+) the central atom redistributes electron density between the thiazole ligands and the terminal Cl atoms.  相似文献   

20.
A simple, sensitive and accurate analytical method was optimized and developed for the determination of deoxynivalenol and aflatoxins in cereals intended for human consumption using high‐performance liquid chromatography with diode array and fluorescence detection and a photochemical reactor for enhanced detection. A response surface methodology, using a fractional central composite design, was carried out for optimization of the water percentage at the beginning of the run (X1, 80–90%), the level of acetonitrile at the end of gradient system (X2, 10–20%) with the water percentage fixed at 60%, and the flow rate (X3, 0.8–1.2 mL/min). The studied responses were the chromatographic peak area, the resolution factor and the time of analysis. Optimal chromatographic conditions were: X1 = 80%, X2 = 10%, and X3 = 1 mL/min. Following a double sample extraction with water and a mixture of methanol/water, mycotoxins were rapidly purified by an optimized solid‐phase extraction protocol. The optimized method was further validated with respect to linearity (R2>0.9991), sensitivity, precision, and recovery (90–112%). The application to 23 commercial cereal samples from Greece showed contamination levels below the legally set limits, except for one maize sample. The main advantages of the developed method are the simplicity of operation and the low cost.  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