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 218 毫秒
1.
Intermediate neglect of differential overlap (INDO) method was used to study the structures and the electronic spectra of C60M12 (M=Li, Na, Be). The calculations indicate that in the minimal energy configuration of C60M12 (M=Li, Na) the C60 cage still retains Ih symmetry and the 12 Li or Na atoms are symmetrically located above the pentagons of the C60 cage, whereas the difference between the double and single bonds has been significantly reduced. In contrast, because six electrons are filled in the fivefold‐degenerated hg orbital of C60, the Cs structure of C60Be12 has illustrated the occurrence of Jahn‐Teller distortion. Based on the optimized geometries, the electronic absorption spectra were calculated and the nature of red shift was discussed. ©1999 John Wiley & Sons, Inc. Int J Quant Chem 73: 505–509, 1999  相似文献   

2.
Ab initio RHF SCF calculations are used for some small clusters MxXy, where M=Cd, Ag; X=S, I; and x, y≤7. Variation of electronic structure with size for some clusters with the bulklike tetrahedral coordination and with the lower symmetry allows one to predict their possible geometries which are compared with experimental data on the existence of the clusters. The chemical‐bonding factor (the chemical nature of bounded atoms, coordination number for metal and nonmetal atoms, hybridization, etc.) is of more importance for properties of the clusters than is the familiar quantum confinement effect of semiconductor clusters. The essential difference in regularities of small cluster formation is analyzed for CdS‐ and AgI‐based structures. ©1999 John Wiley & Sons, Inc. Int J Quant Chem 71: 337–341, 1999  相似文献   

3.
New equiatomic stannide CeRuSn was synthesized from the elements by arc‐melting. CeRuSn was investigated by X‐ray powder and single crystal diffraction: C2/m, a = 1156.1(4), b = 475.9(2) and c = 1023.3(4) pm, β = 102.89(3)°, wR2 = 0.0466, 1229 F2 values and 38 variables. CeRuSn adopts a superstructure of the monoclinic CeCoAl type through a doubling of the subcell c axis. In the superstructure two crystallographically independent cerium sites occur. Based on the interatomic distances the two sites can be assigned to trivalent Ce2 and intermediate valent Ce1. This trivalent‐intermediate valent cerium ordering is underlined by magnetic susceptibility measurements χm(T): below 150 K χm, measured with decreasing temperature, follows a Curie‐Weiss law χm = Cm/(T–θp) giving Cm = 0.38 emuK/mol as Curie constant per CeRuSn mol; a value showing that only half of the cerium atoms are trivalent in CeRuSn (Cm = 0.807 emuK/mol for one free Ce3+ ion). A remarkable feature of the CeRuSn structure are the short Ce1–Ru1 (233 pm) and Ce1–Ru2 (246 pm) distances. The crystal chemistry of CeRuSn is discussed on the basis of a group‐subgroup scheme.  相似文献   

4.
Using the symmetrized boson representation technique, concise algebraic expressions of the irreducible bases symmetry adapted to the group chain IhC5 for the fullerene molecules C20H20, C80, and C240 are derived for the most general cases and those for any specific case can be derived from them easily without a projection procedure. ©1999 John Wiley & Sons, Inc. Int J Quant Chem 73: 283–297, 1999  相似文献   

5.
The integrated molecular transform (FTm) is a unitary numerical index of structure that is capable of uniquely representing different molecular structure conformations with the exception of enantiomers. Other molecular indices have been derived from FTm as well as from the normalized molecular moment (Mn), for example, the analogous electronic and charge transforms (FTe and FTc) and moments (Me and Mc). In this study, each of these indices was calculated for up to 10 sampled conformations of each of the C1–C10 normal alkanes as they were subjected to a standard annealing process. Statistical analyses of the resulting data in the individual series and subsequent box plots, permitting facile examination of those results, indicated that the respective transform indices (FTm, FTe, FTc) are unique, that is, with no statistically significantly overlap across the series. For the Mn and Me indices, the numerical values for methane overlapped those of ethane in the first instance and both ethane and propane in the second. The Mc index values overlapped in several instances in the series. Inasmuch as the noted molecular indices are based only on parameters of structural origin, these results have profound implications for the correlation and estimation of properties derived not only from a general structure representation, but also for those properties which may be dependent on specific molecular conformations. This includes the potential for indices of molecular flexibility and conformationally dependent atomic electron densities. © 1998 John Wiley & Sons, Inc. Int J Quant Chem 70: 1185–1194, 1998  相似文献   

6.
The geometric configurations and electronic structures of the TinC2n (n=1–6) clusters were studied by using the quantum chemical ab initio density functional theory (DFT) method. Our studies showed that these TinC2n (n=1–6) could grow gradually to form cyclic clusters through the subunits TiC2 bonding to each other by C C or Ti C bond. The result could explain the existing experimental fact. The studies might also be helpful to the knowledge of the formation mechanism of the Met‐Cars. ©1999 John Wiley & Sons, Inc. Int J Quant Chem 71: 313–318, 1999  相似文献   

7.
α-Diimine Ni complexes (7, 8) were used as catalyst precursors with MAO in co- and terpolymerization of ethylene/propylene/α-olefins with OH and COOH functional groups. Trimethylaluminium was used to protect the functional group of polar monomers. The presence of 5-hexen-1-ol seems to have no effect on the polymerization rate at all for the N,N′-bis(2,6-diisopropylphenyl) derivative 8 but caused activity decreases of about fivefold in copolymerization and around two times in terpolymerization for the N,N-dimesityl derivative 7. The effect levels off at higher polar comonomer concentration. This system, (7)/MAO, also incorporates well both 10-undecen-1-ol and 10-undecen-1-oic acid. The activities obtained with these α-diimine Ni complexes in co- and terpolymerization are three to twenty times higher than those obtained with group 4 Cp based complexes especially at concentrations of polar monomer in the feed higher than 80 mM. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 2471–2480, 1999  相似文献   

8.
Metals can interact with carbon cages in the following ways: (1) stable carbon cages (i.e., fullerenes) function as electronegative olefins in their exohedral η2 bonding to transition metals; (2) endohedral metallofullerenes with a highly electropositive lanthanide (Ln) inside the carbon cage can be considered to be ionic with lanthanide cations, Ln3+, and fullerene anions; (3) fullerenes too small for independent existence can be stabilized by internal covalent bonding to an endohedral metal atom using the central carbon atoms of pentagon triplets,i.e triquinacene, units, in complexes such as M@C28 (M=Ti, Zr, Hf, and U), derived from the tetrahedral fullerene C28; (4) metal atoms can occur as vertices of binary mixed metal-carbon cages in both early transition metal complexes of the types M14C13, M8C12, and M13C22 (e.g., M=Ti) and copper-carbon cages of the types Cu2n +1C2n + (n≤10), Cu7C8 +, Cu9C10 + and Cu12C12 +. The presence of metal atoms as vertices of carbon cages changes radically their stoichiometries and thus their structures. Thus, early transition metals form cages such as Ti14C13 assumed to have titanium atoms at the vertices and face midpoints of a 3×3×3 cube and carbon atoms at the edge midpoints and center of the cube and Ti13C22 assumed to have titanium atoms at the edge midpoints and center of a 3×3×3 cube as well as C2 units and carbon atoms at the vertices and face midpoints, respectively, of the cube. Elimination of the face metal atoms from the Ti14C13 structure as well as the center carbon atom, which has been achieved experimentally by photofragmentation, leads to the Ti8C12 cluster. The structure of this cluster is based on a tetracapped tetrahedron withT d symmetry with two distinct quartets of titanium atoms, six distinct C2 pairs, and 36 direct Ti−C interactions. The copper-carbon cages of various stoichiometries are suggested to have prismatic, antiprismatic, or cuboctahedral structures in which the electronic configurations of the copper atoms approach the favored 18-electron rare gas configuration. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 5, pp. 862–869, May, 1998.  相似文献   

9.
The symmetry orbital tensor (SOT) method, which makes full use of symmetries in all point groups and can be applied to the self-consistent field (SCF) and post-SCF calculations, is introduced. The principal feature of this method is the definition of the symmetry orbitals (SOs). Any element in a molecular point group will transform one SO to another equivalent SO or simply to itself, and no mixture among SOs exists. Thus, although the SOs for non-Abelian point groups may adapt to reducible representations, their transformation properties are much simpler than in conventional treatments. This article also presents a general scheme to generate SOs for all point groups. The direct products of N SOs form an Nth-rank SOT group, and each matrix element between SOTs is the product of a physical factor and a geometric factor. Compared with the canonical molecular orbitals, the use of SOs can noticeably reduce the computation efforts by decreasing the number of integrals needed in the SCF calculations or the number of configurations needed in the configuration interaction (CI) calculations. The SOT-SCF and SOT-CI approaches are formulated and a preliminary SOT-SCF program is written. Pilot calculations demonstrate the value of the SOT approach, at least at the closed-shell Hartree–Fock level. ©1999 John Wiley & Sons, Inc. J Comput Chem 20: 305–321, 1999  相似文献   

10.
The ascorbate reduction reaction of the native and urea-perturbed forms, 0–8M urea, of horse heart ferricytochrome c is found to be a three-step process: a urea-dependent equilibrium step between a reducible and an irreducible form with a midconcentration of urea of 7.4M, a binding step with a binding constant of 5.9M–1, and a reduction step with a urea-independent rate constant of 2.9 ± 0.3 s–1 [J. Biol. Chem. 255 , 9666 (1980)]. The effect of adding urea, in addition to the generation of an irreducible form, is a slight lowering of the ascorbate-protein binding constant, 5.9 to 2.7M–1, which is limited to the 0–5.5M concentration range. The thermodynamics of the ureadenaturation process also yields a three-step mechanism, N? X1? X2? D, with midconcentrations of urea of 2.5–3M, 6.2M, and 7.5M, respectively, where N, D, and the Xs are the native, the 9-M-urea, and the intermediate forms. The three processes are described as the loosening of the heme crevice opening, the solvent exposure of the polypeptide backbone, and the disruption of the tryptophan–porphyrin interactions, respectively [Biochemistry 19 , 199 (1980)]. The reaction of the protein with 2,3-butanedione, a group-specific reagent for the guanidinium groups and an electron donor for this protein, is inhibited in the presence of ascorbate, but only one of the two functional groups is involved [J. Biol. Chem. 255 , 11094 (1980)]. A correlation of kinetic and thermodynamic observations led to the conclusion that the ascorbate reduction of the protein is independent of the state of the heme crevice opening and of the polypeptide organized structures; instead, it is determined by the integrity of the tryptophan indole–porphyrin interactions. This information, when taken in conjunction with the selective inhibition of the reaction of the arginine side chains by ascorbate, establishes the binding site of ascrobate as one of the two arginyl side chains, and not the opening of the crevice or its vicinity. From the three-dimensional structure of the protein, and taking into consideration the variability of the protein sequence, it is suggested that Arg-38 is the ascorbate binding site, and that the electronic interaction between the indole of Trp-59 and the porphyrin moiety must constitute, at least in part, the electron-transfer path to heme iron.  相似文献   

11.
The title compound, [KCr(C2O2)2(C6H8N2)]n, was obtained from aqueous solution and analyzed with single‐crystal X‐ray diffraction at 100 K. It crystallizes in the monoclinic space group C2/c and displays a three‐dimensional polymeric architecture built up by bimetallic oxalate‐bridged CrIII–K helical chains linked through centrosymmetric K2O2 units to yield a sheet‐like alternating P/M arrangement which looks like that of the previously described two‐dimensional [NaCr(ox)2(pyim)(H2O)]·2H2O [pyim is 2‐(pyridin‐2‐yl)imidazole; Lei et al. (2006). Inorg. Chem. Commun. 9 , 486–488]. The CrIII ions in each helix have the same chirality. The infinite neutral sheets are eclipsed with respect to each other and are held together by a hydrogen‐bonding network involving 2‐(aminomethyl)pyridine H atoms and oxalate O atoms. Each sheet gives rise to channels of Cr4K4 octanuclear rings and each resultant hole is occupied by a pair of 2‐(aminomethyl)pyridine ligands with partial overlap. The shortest Cr...Cr distance [5.593 (4) Å] is shorter than usually observed in the K–MIII–oxalate family.  相似文献   

12.
Formation, crystal structure, polymorphism, and transition between polymorphs are reported for M(thd)3, (M = Al, Cr, Mn, Fe, Co, Ga, and In) [(thd) = anion of H(thd) = C11H20O2 = 2, 2, 6, 6‐tetramethylheptane‐3, 5‐dione]. Fresh crystal‐structure data are provided for monoclinic polymorphs of Al(thd)3, Ga(thd)3, and In(thd)3. Apart from adjustment of the M–Ok bond length, the structural characteristics of M(thd)3 complexes remain essentially unaffected by change of M. Analysis of the M–Ok, Ok–Ck, and Ck–Ck distances support the notion that the M–Ok–Ck–Ck–Ck–Ok– ring forms a heterocyclic unit with σ and π contributions to the bonds. Tentative assessments according to the bond‐valence or bond‐order scheme suggest that the strengths of the σ bonds are approximately equal for the M–Ok, Ok–Ck, and Ck–Ck bonds, whereas the π component of the M–Ok bonds is small compared with those for the Ok–Ck, and Ck–Ck bonds. The contours of a pattern for the occurrence of M(thd)3 polymorphs suggest that polymorphs with structures of orthorhombic or higher symmetry are favored on crystallization from the vapor phase (viz. sublimation). Monoclinic polymorphs prefer crystallization from solution at temperatures closer to ambient. Each of the M(thd)3 complexes subject to this study exhibits three or more polymorphs (further variants are likely to emerge consequent on systematic exploration of the crystallization conditions). High‐temperature powder X‐ray diffraction shows that the monoclinic polymorphs convert irreversibly to the corresponding rotational disordered orthorhombic variant above some 100–150 °C (depending on M). The orthorhombic variant is in turn transformed into polymorphs of tetragonal and cubic symmetry before entering the molten state. These findings are discussed in light of the current conceptions of rotational disorder in molecular crystals.  相似文献   

13.
The bis‐sulfonamide m‐C6H4(SO2NHPh)2 efficiently promotes the ring‐opening polymerization of lactide when combined with tertiary amines, such as N,N‐dimethylaminopyridine. Polylactides of controlled molecular weights (Mn up to 17,700 g mol?1) and very narrow molecular weight distributions (Mw/Mn < 1.11) are obtained under mild conditions and in a living fashion. The reaction takes place through a bifunctional mechanism involving activation of both the alcohol and the monomer. Modulation of the sulfonamide component supports cooperative dual hydrogen‐bonding of lactide involving the two (SO2NHAr) moieties. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 959–965, 2010  相似文献   

14.
The two title compounds, [M(C6H14O2PS2)2(C12H8N2)], where M = CdII and FeII, are isomorphous. Each compound has a crystallographic twofold axis of symmetry through the metal atom and the 1,10-phenanthroline mol­ecule. The central metal atom is coordinated to four S atoms from the two dithiophos­phate groups and two N atoms from the 1,10-phenanthroline ligand. The environment of the metal atom is a distorted octahedron.  相似文献   

15.
In the title compound, C8H22Cl2N2Si3, the central Si atom is tetrahedrally coordinated by two Cl and two N atoms in a molecule that has crystallographically imposed C2 symmetry. Comparison is made with the isomorphous structure having titanium instead of silicon at the central position in the diazacyclopentane ring [Tinkler, Deeth, Duncalf & McCamley (1996). Chem. Commun. pp. 2623–2624].  相似文献   

16.
In a combined experimental and computational study, the molecular and electronic structures of the divalent bis(m-terphenyl)element cations [(2,6-Mes2C6H3)2E]+ of group 13 ( 1 , E=B; 2 , E=Al; 3 , E=Ga; 4 , E=In; 5 , E=Tl) were investigated. The preparation and characterization of 2 , 3 and 5 were previously reported by Wehmschulte's (Organometallics 2004 , 23, 1965–1967; J. Am. Chem. Soc. 2003 , 125, 1470–1471) and our groups (Organometallics 2009 , 28, 6893–6901). The indinium ion 4 was prepared and fully characterized for the first time. Attempts to prepare the borinium ion 1 by fluoride or hydride abstraction were unsuccessful. The electronic structures of 1 – 5 and the stabilization by the bulky m-terphenyl substituents were analyzed using quantum chemical calculations and compared to the divalent bis(m-terphenyl)pnictogenium ions [(2,6-Mes2C6H3)2E]+ of group 15 ( 6 , E=P; 7 , E=As; 8 , E=Sb; 9 , E=Bi) previously investigated by our group (Angew. Chem. Int. Ed. 2018 , 57, 10080–10084). The calculated fluoride ion affinities (FIA) of 1–9 are higher than that of SbF5, which classifies them as Lewis superacids.  相似文献   

17.
We have studied the polymerization of methyl methacrylate with cesium and tetramethylammonium salts of diethyl 2‐ethylmalonate carbanion. For polymerizations initiated at room temperature, no effort was made to control the exotherm. The polymerizations proceeded with high conversions and produced polymers characterized by broad polydispersities (Mw/Mn = 2–3). We consistently observed Mn (exptl) < Mn(calcd) for target = 50,000–300,000 g/mol; we observed an apparent upper limit for Mn of 60,000–70,000 g/mol. Chain transfer from impurities in reagents was eliminated as the source of molecular weight lowering. Oligomeric samples were analyzed by mass spectrometry; results were best explained by methoxide initiation of the polymerization. No evidence for malonate end groups was observed in the mass spectra. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 615–620, 1999  相似文献   

18.
A series of new AB-type poly(etherimide)s having bisphenol-type moiety was prepared by the one-pot polyimidization using triphenylphosphite(TPP) in N-methyl-2-pyrrolidone(NMP)/pyridine solution at 150°C. Complete cyclodehydration was observed in the polymerizations as well as in model reactions. Polymers were obtained with inherent viscosities in the 0.27–0.49 dL/g range. The Mn and Mw/Mn of poly[4-(1,4-phenyleneoxy-1,4-phenylenehexafluoro-isopropylidene-1,4-phenylene)oxyphthalimide] (4d) with ηinh = 0.49 dL/g were 73,400 g/mol and 1.5, respectively. Most polymers could readily be dissolved in common organic solvents such as DMAc, NMP, and m-cresol. The polymer 4d was soluble even in chloroform. These polymers had glass transition temperatures between 205 and 235°C, and 5% weight loss temperatures in the range of 511–532°C in nitrogen. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 3530–3536, 1999  相似文献   

19.
Acidic bismuth salts, such as BiCl3, BiBr3, BiJ3, and Bi‐triflate catalyzed the ring‐opening polymerization of 2‐methoxazoline (MOZ) in bulk at 100 °C, whereas less acidic salts such as Bi2O3 or Bi(III)acetate did not. Bi‐triflate‐catalyzed polymerizations of 2‐ethyloxazoline (EtOZ) were performed with variation of the monomer–catalyst ratio (M/C). It was found that the molecular weights were independent of the M/C ratio. The formation of cationic chain ends and the absence of cycles was proven by reactions of virgin polymerization products with N,N‐dimethyl‐4‐aminopyridine or triphenylphosphine. The resulting polymers having modified cationic chain ends were characterized by 1H NMR spectroscopy and MALDI‐TOF mass spectrometry. The polymerization mechanism including chain‐transfer reactions is discussed. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 4777–4784, 2008  相似文献   

20.
Density Functional Theory (DFT) calculations at the B3LYP/6-31G* level have been performed on four bowl-shaped polyaromatic hydrocarbons of C30H12 molecular formula ( 1 – 4 ) showing C3 ( 1 ), C2v ( 2 and 4 ), and C2h ( 3 ) symmetries. The geometrical and electronic properties of the compounds studied have been analyzed to explain their relative stability. NMR chemical shifts parameters for the atoms and Nucleus Independent Chemical Shifts (NICSs) for the rings were calculated using the GIAO method. The 13C and 1H chemical shifts calculated are in very good agreement with the experimental data. ©1999 John Wiley & Sons, Inc. J Comput Chem 20: 1412–1421, 1999  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