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.
Triprismane, a C6H6 isomer of benzene, belongs to a class of strained hydrocarbons. In this work, the energy and charge distributions in a series of aza (C6NH6–) and nitroaza (CNO2)6–N, (with =1...6) derivatives of triprismane have been analyzed using the ab initio Hartree–Fock (HF)-derived molecular electrostatic potentials and molecular electron densities. Electrostatic potential investigations have shown that the electron-rich regions around nitrogen along a series of azatriprismanes and those near oxygens of nitro group in nitroazatriprismanes become smaller on encompassing from the hexanitroaza to nitroazatriprismane. As revealed from the molecular electrostatic potential (MESP) topography for the series of azatriprismanes the MESP minimum near nitrogen become shallow with increasing nitrogen content. Heats of formation obtained from the isodesmic reaction approach in the C6N6H6– and (CNO2)6–N series correlate well with the electron density at the bond-critical point of the X— N (X = C or N) bonds of the triprismane framework.SPG is grateful to the Council of Scientific and Industrial Research (CSIR project 01(1772)/02/EMR-II), New Delhi, India. KAJ thanks CSIR for the Junior Research Fellowship. Thanks are due to C-DAC, Pune, for providing computational facilities.  相似文献   

2.
Photochemical transformations of phenothiazine (PTA) in solutions of halomethanes CHnX4–n (X = Cl, Br; n = 0, 1, 2) and in n-hexane—CHnX4–n mixtures under the irradiation with = 337 and 365 nm were studied. The rate constants of quenching of PTA fluorescence with halomethanes (k q) are 4·105—1.3·1010 L mol–1 s–1. The process occurs due to electron transfer with the C—X bond cleavage in the radical anion fragment of the primary radical ion pair. This results in the formation of the stable radical cation salt (PTA·+X). The plot of k q vs. free energy of electron transfer corresponds to the Rehm—Weller empirical equation for a one-electron process and is satisfactorily described in terms of the theory of nonradiative electron transitions in the approximation of one quantum vibration.  相似文献   

3.
Metal–ligand bonding in transition metal halide molecules and complexes with different central ions, oxidations states, and coordination numbers: CrIIIX63–, CrIVX4, CrIIX2 (X = F,Cl,Br,I), MIIICl63–(M = Mo,W), MIII(H2O)63+(M = Cr,Co) and Re2Cl82– has been studied in terms of the Extended Transition State (ETS) energy patitioning scheme within the DFT and electron density analysis (the Laplacian of the electron density and the electronic localization function). Bonding is found to be dominated by ionicity in all cases, especially so for complexes with higher coordination numbers. Covalent contributions to the metal–ligand bond are found to be mainly due to the nd-electrons and to lesser extent due to the metal (n+1)s and (n+1)p-orbitals, contributions from (n+1)s increasing when going to lower coordination numbers. Metal–ligand bonding analysis have been used in order to check some concepts emerging from ligand field theory when applied to the spectroscopy and magnetism of transition metal complexes. It is pointed out that for complexes of high symmetry (MX6, Oh, MX4, Td, and MX2, Dh) electron density analyses gain interpretative power when partitioned into contributions from occupied orbitals of different symmetry.  相似文献   

4.
Force constants were determined for the C8, C10, C12 and C14 series ofn-alkanes C n H2n + 2 using an approximate SVFF calculation and observed LAM = 1 wave numbers. In this calculation the hydrogen atoms were neglected and only the carbon backbone chain and terminal atoms were considered; this was valid since only low-frequency vibrations were under consideration. Using force constant transfer, the wavenumbers of the LAM = 1 accordion modes for the analogous -Cn H2n + 1 X and ,-Cn H2nX2 species, where X = C1, Br or I were calculated. For -chloroalkanes and ,-dichloroalkanes, them = 1 accordion modes are calculated to be in the 220–130 cm–1 and 200–120 cm–1 regions, respectively. For the bromo- and iodo-analogues them = 1 accordion modes are calculated to be in the 200–100 cm–1, 150–90 cm–1 and in the 170–100, 135–80 cm–1 regions, respectively.  相似文献   

5.
The relationship 1– m=1.062n 1/1–n , well-known in non-isothermal kinetics and applied for determination of the kinetic exponentn, is not universal. A more accurate modification of the given equation is proposed in the form of a more complex function 1– m=(n, Xm), whereX m =E/RT m is the dimensionless Arrhenius' criterion.  相似文献   

6.
The cyclopropyl carbanions C3H4X (X=H (I), F (II), Cl (III), CF3 (IV), COCH3 (V), CN (VI), NO2 (VII), which are formed by the removal of a proton at the position, were calculated by the semiempirical MINDO/3 and MNDO methods with full optimization of the geometry. The MINDO/3 method, in agreement with the results from nonempirical methods and experimental data, leads to a nonplanar configuration at the anionic center; the angle between the CX bond and the CCC plane was 58–62°. In all probability the conclusion about a planar structure for all the investigated carbanions except (II), which emerges from the MNDO calculations, is unreliable. By calculation of the change of energies in the isodesmic reactions C3H5 + C3H5X C3H4X + C3H6 it was found that substitution of hydrogen by the polar substituent X stabilizes the carbanion. On the basis of an analysis of the changes in the length of the cyclic and extracyclic bonds in the transition from the neutral molecule to the carbanion it was established that the introduction of substituents N02, COCH3, and CN into the carbanion gives rise to predominant donation of electron density along the system through the ring to the substituent while the introduction of a halogen leads to transfer of electron density from the substituent to the anionic center.Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 21, No. 5, pp. 590–594, September–October, 1985.  相似文献   

7.
We present computational results from detailed gas-phase conformational analyses of the -substituted aldehydes, glycinal and alaninal. A synplanar conformer of glycinal and a synperiplanar conformer of alaninal in which the C=O and C–N bonds are in an eclipsing orientation are found to be lowest in energy; the two amino hydrogen atoms in these conformers are both directed over the C–C bond, i.e., in a compact arrangement. For the Group VA analogs, H2P–CH2–CHO and H2P–CH(CH3)–CHO, skew conformers in which the C–H and C–Me groups, respectively, are in an eclipsing orientation with the C=O bond are found to be lower in energy than the syn(peri)planar conformers. The results of various self-consistent reaction-field calculations suggest that the lowest-energy conformer of glycinal in 1, 2-dichloroethane is still synperiplanar, although the orientation of the amino hydrogen atoms may be different from that in the gas phase. Similar reaction-field calculations for alaninal raise the possibility that in this solvent a skew conformer, in which the C–H bond is nearly eclipsing the C=O bond, is energetically competitive with synperiplanar conformers.  相似文献   

8.
Potentially bi- and tetra-dentateSchiff bases derived from salicylaldehyde react with hydrated uranyl salts to give complexes: UO2H2 LX 2, UO2H2 LX 2 and UO2(HL)2 X 2 [H2 L=N,N-propane-1,3-diylbis(salicylideneimine), H2 L=N,N-ethylenebis(salicylideneimine) and HL=N-phenylsalicylideneimine;X =Cl, Br, I, NO3 , ClO4 , and NCS]. Because of marked spectral similrities with the structurally known Ca(H2 L) (NO3)2, theSchiff bases are coordinated through the negatively charged phenolic oxygen atoms and not the nitrogen atoms of the azomethine groups which carry the protons transferred from phenolic groups on coordination. Halide, nitrate, perchlorate and thiocyanate groups are covalently bonded to the uranyl ion, resulting a 6-coordinated uranium ion in the halo and thiocyanato complexes and 8-coordinated in nitrato and perchlorato complexes.
Komplexe von Dioxouranyl(VI) mit zwitterionischen Formen von zwei- und vierzähnigen Schiff-Basen
Zusammenfassung Von Salizylaldehyd abgeleitete zwei- und vierzähnigeSchiff-Basen reagieren mit hydratisierten Uranylsalzen zu Komplexen folgenden Typs: UO2H2 LX 2, UO2H2 LX 2 und UO2(HL)2 X 2 [H2 L=N,N-Propan-1,3-diylbis(salicylidenimin), H2 L=N,N-Ethylen-bis(salicylidenimin) und HL=N-Phenylsalicylidenimin;X =Cl, Br, I, NO3 , ClO4 und NCS]. Auf Grund eindeutiger spektraler Ähnlichkeiten mit dem bekannten Ca(H2 L) (NO3)2 wird auf Koordination über die negativ geladenen phenolischen Sauerstoffatome (und nicht über die Azomethin-Stickstoffe) geschlossen. Die AnionenX sind kovalent an das Uranyl-Ion gebunden; damit ergibt sich ein hexakoordiniertes Uranyl-Ion für die Halogen- und Thiocyanat-Komplexe und Oktakoordination für die Nitrat- und Perchlorat-Komplexe.
  相似文献   

9.
Effective reaction order by cyanide ions, p, effective transfer coefficient ef, and exchange current i 0of the gold dissolution process are measured at a constant surface coverage by thallium adatoms Tl. The constancy of Tlis ensured by constant time tof the electrode's contact with solution after renewal of its surface and before starting the measurements. Solutions containing 2.5 × 10–6to 10–5M TlNO3, 0.1 M KCN, and 0.1 M KOH are studied. With an increase in t, quantities p, ef, and i 0increase from values typical for pure KCN solutions (0.17, 0.1, and 10–5A/dm2, respectively) to p 1.1, ef 0.4, and i 0 3 × 10–4A/dm2. The increase in pby nearly unity is interpreted as indicating a change in the limiting stage in the presence of thallium ions: in the latter case, it involves not only adsorbed cyanide ions (as in pure cyanide solutions), but also those located beyond the adsorption layer. A possible mechanism explaining the acceleration of gold dissolution in the presence of thallium adatoms is put forward.  相似文献   

10.
Densities and heat capacities of aqueous solutions of azoniaspiroalkane halides, (CH2) n N+ (CH2) n X (where X=Cl, I andn=5,6), have been measured at 25°C using a flow densimeter and a flow microcalorimeter. The limiting apparent molal volumes (ø v ) and apparent molal heat capacities (ø cp ) obtained from these data are compared with those of the azoniaspiroalkane bromides and the corresponding tetraalkylammonium halides. The concentration dependence of ø v and øcp are examined for clues on the influence of solute hydration, structure, and conformational flexibility on the excess functions of quaternary ammonium halides.  相似文献   

11.
The influence of the nature of the heteroatom on the Ni+ gas-phase binding energies of HCC–XH3 (X is C, Si, or Ge) compounds has been investigated through the use of high-level density functional theory methods. The structures of the corresponding Ni+ complexes were optimized at the B3LYP/6-311G(d,p) level of theory. Final energies were obtained in single-point B3LYP/6-311+G(2df,2p) calculations. Nonconventional complexes, in which the metal cation interacts simultaneously with the CC system and with one of the X–H bonds of the substituent XH3 group, play a significant role in the binding of Ni+ to HCC–XH3 (X is Si or Ge) derivatives. Conversely, such nonconventional complexes are not local minima of the propyne–Ni+ potential-energy surface. This establishes a clear distinction between unsaturated carbon derivatives and the Si- and Ge-containing analogues as far as bonding to transition-metal monocations is concerned. Actually, the attachment of Ni+ to HCC–XH3 (X is Si or Ge) compounds in the gas phase yields a mixture of conventional and nonconventional complexes. These agostic-type interactions can be viewed as a dative bond from the X–H bonding orbital toward the empty s orbital of the metal, and a back-donation from the valence electron pairs of the metal into the X–H antibonding orbital of the neutral species.Proceedings of the 11th International Congress of Quantum Chemistry satellite meeting in honor of Jean-Louis Rivail  相似文献   

12.
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   

13.
Wavelength-dispersive electron probe microanalysis (WDS-EPMA) of TiN/Ti diffusion couples was carried out in order to determine the nitrogen profiles across the phase bands. Thein situ TiN/Ti diffusion couples were used for phase equilibrium studies. The problem of complete overlap of the Ti L1 and N K lines was solved using chemically well-characterized single-phase titanium nitride samples (-Ti(N), -Ti2N and -TiN1–X ) as external standards. By using such homotypic standard materials it was possible to eliminate systematic errors introduced by imperfect correction algorithms. The composition of the phases in multiphase TiN/Ti diffusion couples could be determined with an accuracy of better than ±1 at% N.  相似文献   

14.
Self-assemblies of 2-hydroxy-1,4-naphthoquinone (HNQ) have been investigated using the (HNQ)n (n=1–4) series as modeled systems employing ab initio Hartree–Fock calculations. The energetics and charge distribution in these molecular systems are presented. As revealed from the electron density in the highest occupied molecular orbital of the lowest energy conformers of (HNQ)n (n=1–4) the charge percolates to the end unit of the assembly. This has been supported by the molecular electrostatic potential topography.  相似文献   

15.
The structures and energies of axial and equatorial conformers and rotamers of 4-substituted tetrahydro-2H-thiopyran-1,1-dioxides (tetrahydrothiopyran-1,1-dioxides, thiacyclohexane-1,1-dioxides, thiane-1,1-dioxides, and 1,1-dioxothianes; CH3, CH2OH, CHO, COCH3, CN, F, Cl, Br, and OCOCH3) were calculated using the hybrid density functionals B3LYP, B3P86, and B3PW91, as well as MP2 and the 6-31G(d), 6-31G(2d), 6-31G(3d), 6-31G(d,p), and 6-31+G(d) basis sets. MP2/6-31+G(d)/ /HF/6-31+G(d) [–G° = 1.73 kcal/mol], B3P86/6-31G(d) [–G° = 1.75 kcal/mol], and B3PW91/6-31G(d) [–G° = 1.85 kcal/mol] gave conformational free energy (G°) values at 180 K for 4-methyltetrahydro-2H-thiopyran-1,1-dioxide which were similar to the reported experimental values for methylcyclohexane (–G° = 1.80 kcal/mol), 4-methyltetrahydro-2H-thiopyran (–G° = 1.80 kcal/mol), and other 4-methyl-substituted heterocycles. All levels of theory showed that the conformational preferences of the 4-methanoyl (4-formyl), 4-ethanoyl (4-acetyl), and 4-cyano substituents were small. The HF calculations gave conformational free energy (G°) values for 4-chlorotetrahydro-2H-thiopyran-1,1dioxide which were closer to the experimental value than the MP2 and density functional methods. The best agreement with available experimental data for 4-bromotetrahydro-2H-thiopyran-1,1-dioxide was obtained from the HF/6-31G(2d), HF/6-31G(3d), and B3LYP/6-31G(2d) calculations, and, for 4-acetoxytetrahydro-2H-thiopyran-1,1-dioxide, from the HF/6–31G(3d) calculations. The conformational free energies (G°) and relative energies (E) of the conformers and rotamers have been compared with the correspondingly substituted cyclohexanes and tetrahydro-2H-thiopyrans and are discussed in terms of dipole–dipole (electrostatic) interactions and repulsive nonbonded interactions (steric) in the most stable axial and equatorial conformers. The axial S=O bond lengths are shorter than the equatorial S=O bond lengths and the C2–C3 bond lengths in the substituents with carbon-bonded to the ring are shorter than the C3–C4 and C4–C-5 bond lengths. In contrast, the C2–C3 bond lengths in the 4-halogen and 4-acetoxy substituents are longer than the C3–C4 and C4–C-5 bond lengths.  相似文献   

16.
The electronic structures of the Cl3 , Br3 , I3 , BrCl2 , BrI2 , BrICl, and BrII anions were calculated using the discrete variational X method. These calculations showed that the adiabatic electron affinity (EA) of the corresponding trihalogen molecules is close to the vertical EA and first vertical ionization potential (IP) of the corresponding singly charged anions. The calculated first IP of all these anions are rather similar (3.3–3.8 eV). Thus, these trihalides may be considered weak superhalogens.Branch of the Institute of Chemical Physics, Academy of Sciences of the USSR. Translated from Zhurnal Strukturnoi Khimii, Vol. 30, No. 5, pp. 41–47, September–October, 1989.  相似文献   

17.
With replacement of N atoms by CH groups in the most stable chain isomer of N8H8, 34 possible isomers of Nn(CH)8−nH8 (n = 0–7) have been designed and optimized at the B3LYP/6-311++G** level of theory. The natural bond orbital (NBO) and atoms in molecules (AIM) analysis are carried out to study the bonding nature and relative stabilities of these conformers. G3MP2 method is applied to calculate energies and heats of formation. The results indicate that the hyperconjugation effect from lone pairs of nitrogen atoms to germinal C–N bonds is the major factor which caused the change of the C–N bond length. With the more replacement of nitrogen atoms by CH groups, the heats of formation of the isomers of Nn(CH)8−nH8 (n = 0–7) decrease gradually, but the energies increase linearly.  相似文献   

18.
Zusammenfassung Rhenium(VII) wird mit Tetraphenylarsoniumchlorid durch Verteilen zwischen Wasser und Chloroform vom 100 000fachen Überschuß an Wolfram, 100 00fachen an Molybdän sowie von Kupfer und Nickel abgetrennt. Zur photometrischen Bestimmung wird Re(VII) in der Chloroformphase reduziert und mit -Furildioxim umgesetzt, 530nm= 29 750 l · Mol–1 · cm–1.
Photometric determination of small amounts of rhenium in the organic layer after separation of tungsten and molybdenum by liquid-liquid extraction
Rhenium(VII) is separated by extraction of tetraphenylarsoniumperrhenate from a 100 000-fold excess of W, 100 00-fold excess of Mo, and from Cu and Ni. For photometric determination Re is reduced in the organic layer and the colour formed by addition of -furildioxime is used for the measurement at 530 nm; =297501 · Mol–1 · cm–1.


Wir danken der Deutschen Forschungsgemeinschaft und dem Verband der chemischen Industrie (Fonds der Chemie) für Beihilfen, sowie Frl. M. Schriever für ihre zuverlässige Mitarbeit.  相似文献   

19.
The geometrical structure and conformation of dimethyloxalate, CH3OC(O)–C(O)OCH3, have been studied by gas electron diffraction (GED) and quantum-chemical calculations (MP2 and B3LYP methods with 6-31G* and cc-pVTZ basis sets). The GED analysis with a dynamic model (T = 323 K) results in a mixture of two planar conformers, anti (C2h symmetry) and syn (C2v symmetry) orientation of the two C=O bonds. The energy difference between these conformers is 0.02(0.18) kcal/mol and barrier to internal rotation around the C–C bond is 0.44(0.41) kcal/mol. The CH3 groups occupy synperiplanar positions with respect to the C=O bonds. The following main geometrical parameters for the anti conformer (Å and degrees) have been derived: rg(C–C) = 1.532(3), rg(C=O) = 1.203(2), rg(Csp3–O) = 1.436(3), rg(Csp2–O) = 1.333(3), (Csp2–Csp2–O) = 111.9(1.9), (Csp2–O–Csp3) = 116.3(1.6), (O–C= O) = 127.0(1.8).This paper is devoted to the 75th anniversary of gas electron diffraction method.  相似文献   

20.
Zusammenfassung Der lineare thermische Ausdehnungskoeffizient von linearem Polyäthylen hoher Dichte wurde im Temperaturbereich –20 °C bis + 40°C bestimmt. Bei isotropen Proben besteht eine lineare Beziehung zwischen Dichte bzw. Kristallisationsgrad v und. Die gemessenen Koeffizienten liegen fürT 0 = 20 °C im Bereich = 110 ... 130 · 10–6 K–1.Kalt verstreckte Proben mit Verstreckungsgraden = 8 ... 15 haben beiT 0 = 20 °C in Verstreckrichtung den Koeffizienten l = –24 · 10–6 K–1. Der negative Zahlenwert von tl ist unabhängig von und weiteren Herstellungsparametern. Seine Ursache ist einerseits die Orientierung derc-Achsen der Kristallite in Verstreckrichtung mit c = –12 · 10–6 K–1 und andererseits der negative Koeffizient am * –50 · 10–6 K–1 der verspannten amorphen Phase, der auf dem gummielastischen Verhalten der tie-molecules beruht.Beim Tempern oberhalb von +40 °C schrumpfen die Proben irreversibel, wobei | ansteigt und nach dem Aufschmelzen der Proben wieder den Wert des isotropen Materials annimmt. Nach dem Tempern wenig unterhalb der Schmelztemperatur der Kristallite wurden überhöhte Koeffizienten | gemessen, die eine Rotation der Kristallite um dieb-Achsen erkennen lassen.
Summary The linear thermal coefficient of expansion,, of high density linear polyethylene has been determined in the temperature range of –20 ° ... + 40 °C. For isotropic samples a linear relationship between density or crystallinity v and is valid. Measured values of forT 0 = 20 °C amount to = 110 ... 130 · 10–6 K–1.Cold drawn samples of draw ratios = 8 ... 15 yield | = –24 · 10–6 K–1 atT 0 = 20 °C parallel to the draw axis. The negative value of | does not depend on draw ratio or other parameters of sample processing. It is caused byc-axis orientation of the crystallites in draw direction with c = –12 · 10–6 K–1 and by a negative coefficient am * = –50 · 10–6 K–1 of the stressed amorphous phase, which is due to rubber elastic behaviour of the tie molecules.When annealed above 40 °C the samples shrink irreversibly and | is augmented. After melting the samples the value of isotropic material is restored. Annealing the samples little below the melting temperature of the crystallites leads to superelevated values all which reflect| rotation of the crystallites around theb-axis.
  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