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.
Measurements are reported on the formation of negative ions in O2, O2/Ar and O2/Ne clusters aimed at establishing the mechanisms of anion formation and the role of inelastic electron scattering by the cluster constituents on negative ion formation in clusters. In the case of pure O2 clusters the main anions we detected are of two types: O(O2) n0 and (O2) n 1– . The yields of O(O2) n showed maxima at 6.3, 8.0 and 14.0 eV and the data suggest O as their precursor; the maxima at 8 and 14 eV are due to the production of O via symmetry forbidden dissociative attachment processes in O2 at these energies which become allowed in clusters. The yields of (O2) n showed a strong maximum at near-zero energy (0.5 eV) and also at 6.3, 8 and 14 eV. With the exception of the near-zero energy resonance, the (O2) n anions at 6.3, 8 and 14 eV are attributed to nondissociative attachment of near-zero energy secondary electrons to O2 clusters. The slow secondary electrons result predominantly from scattering via the O 2 negative ion states of incident electrons with energies in their respective regions. Similar results were obtained for the mixed O2/rare gas clusters except that now a feeble and distinctly structured contribution in the yields of O(O2) n , (O2) n (and Ar(O2) n ) was observed at energies >10 eV. These anions are believed to have the lowest negative ion states of Ar* (Ne*) as their precursors.  相似文献   

2.
We have performed ab initio fourth-order Møller–Plesset perturbation theory calculations in the framework of the supermolecule approach on the vertical excitation spectra of the weakly bound van der Waals N2–He dimer. They indicate a ``T-shaped' stablest ground N2(X1g+)–He(1S) electronic state with a well depth, De, of 21.63 cm–1 at a minimum distance, Re, of 3.44 Å and zero-point vibration correction, Do, of 7.07 cm–1. They also indicate a ``T-shaped' stablest excited conformer with Re=3.25 Å, De=36.85 cm–1 and Do=17.06 cm–1 for the N2(B3g)–He(1S) triplet electronic level. In order to investigate the use of less-demanding correlation methods, test density functional theory calculations using the mPW1PW exchange–correlation functional are also presented for comparison.  相似文献   

3.
The negative ion photoelectron spectrum of7Li 2 is reported at 488 nm (2.540 eV). Three electronic bands are observed in this spectrum and are assigned to the following photodetachment transitions:7Li2,X 1 g + +e 7Li 2 ,X 2 u + ;7Li2,a 3 u + +e 7Li 2 ,X 2 u + ; and7Li2,A 1 u + +e 7Li 2 ,X 2 u + . The electron affinity of7Li2 is determined to be 0.437±0.009 eV, leading to an anion dissociation energy,D 0, of 0.865±0.022 eV for the ground state of7Li 2 . A Franck-Condon analysis of the7Li2,X 1 g + +e 7Li 2 ,X 2 u + band yields the following spectroscopic constants for the ground state of7Li 2 :B e =0.502±0.005 cm–1,r e =3.094±0.015 Å, and e =232±35 cm–1.  相似文献   

4.
We investigated the electronic structure and chemical bonding of the B3 , Al3 , and Ga3 anions, and the gas phase NaB3, NaAl3, and NaGa3 molecules. We found that the ground state of the neutral gas phase salts contains an equilateral triangular anion interacting with a Na+ cation. The B3 , Al3 , and Ga3 anions possess two delocalized electrons and are found to be aromatic. The triangular anions have been shown to be related to recently synthesized organometallic compound containing an aromatic -Ga3 2– unit, but they are differ from them by four valence electrons. The reason for earlier appearance of the -orbital in the B3 , Al3 , and Ga3 anions is discussed.  相似文献   

5.
By using pH-metric and conductometric methods it has been found that tetracycline (H3TC) forms with WO 4 2– and MoO 4 2– ions the following complex compounds: [WO3HTC]2–, [WO3(H2TC)2]2– and [MoO3(H2TC)2]2–. Stability constants log/gb 1 k =7.86 and log 1 k =7.80 for [WO3HTC]2– and [MoO3HTC]2–, respectively, have been calculated from pH-metric measurements.  相似文献   

6.
Water exchange on Mn centers in proteins has been modeled with density functional theory using the B3LYP functional. The reaction barrier for dissociative water exchange on [MnIV(H2O)2(OH)4] is only 9.6 kcal mol–1, corresponding to a rate of 6×105 s–1. It has also been investigated how modifications of the model complex change the exchange rate. Three cases of water exchange on Mn dimers have been modeled. The reaction barrier for dissociative exchange of a terminal water ligand on [(H2O)2(OH)2MnIV(-O)2MnIV(H2O)2(OH)2] is 8.6 kcal mol–1, while the bridging oxo group exchange with a ring-opening mechanism has a barrier of 19.2 kcal mol–1. These results are intended for interpretations of measurements of water exchange for the oxygen evolving complex of photosystem II. Finally, a tautomerization mechanism for exchange of a terminal oxyl radical has been modeled for the synthetic O2 catalyst [(terpy)(H2O)MnIV(-O)2MnIV(O)(terpy)]3+ (terpy=2,2:6,2-terpyridine). The calculated reaction barrier is 14.7 kcal mol–1.Contribution to the Björn Roos Honorary Issue  相似文献   

7.
A vibrational–rotational spectrum of the ν = 2 transitions of a high-temperature molecule AlF was observed between 1490 and 1586 cm−1 with a diode laser spectrometer. Measurements were made on the ν = 3–1, 4–2, 5–3 and 8–6 bands at a temperature of 900 °C. Measured spectral lines were fitted to effective band constants ν0, Bν and Dν for each band. Present measurements were made with only one Pb-salt laser diode. Physical significance of the effective band constants is discussed.  相似文献   

8.
Electron energy distribution functions (EDF) in SiH4, SiH4–H2 radiofrequency discharges have been calculated by solving the time-dependent Boltzmann equation in the presence of a sinusoidal field. Particular emphasis is given to the modulation of EDF as a function of the applied frequency (·106/p 0 ·108 sec–1 torr–1) and of gas composition. The results show that at /p 0 = ·106 sec–1 torr–1 EDF follows in a quasistationary mode the variation of the field with the exception of a small range of electric field near to the zero crossing. Still, at the higher considered frequency (/p 0 =·108 sec–1 torr–1), we observe some modulation of EDF. The necessity of using a time-dependent approach is tested by comparing the present results with the corresponding ones obtained by using the effective field approximation (i.e., the approximation which solves instead of the time-dependent Boltzmann equation the corresponding stationary one at the effective values of the rf field). The two sets of results can differ by orders of magnitude in the tail of EDF, the differences decreasing with increasing molar fraction of H2 and increasing field frequency. The role of excited states (second-kind collisions) is studied by inserting in the Boltzmann equation given concentrations of vibrational and electronic states. The results show that second-kind collisions strongly affect EDF especially in pure silane. Finally a satisfactory agreement has been found between theoretical and experimental results concerning the modulation of electrons of given energy in pure silane discharges.  相似文献   

9.
The phosphorescence spectra of ReBr 6 2– doped A2SnX6 (A = K, Rb, Cs; X = Cl, Br) have been measured at 10 K. The spectra consist of a weighted sum of progressions associated with the local modes of the ReBr 6 2– center. By a fit to a generalized Lorentzian line shape function the totally symmetric distortion of the 7(2 T 2g ) excited state relative to the 8(4 A 2g ) ground state has been determined.Dedicated to Professor Dr. H.-H. Schmidtke on the occasion of his 50th birthday.  相似文献   

10.
The energetics and reaction mechanism of the migratory insertion of carbon monoxide and methyl isocyanide into the zirconium–carbon and titanium–carbon bonds in [calix[4](OMe)2(O)2M–Me2], (M=Zr, Ti), have been investigated by combining static and dynamic density functional calculations. Two steps have been characterized: the coordination of the incoming nucleophilic moiety leading to relatively stable facial adducts; its subsequent insertion into the M–C bond, leading to 2-bound acyl or iminoacyl complexes, providing a rationale for the different behavior of CO and MeNC towards both insertion and deinsertion reactions. Our results indicate that the rate-determining step for the overall MeNC insertion into the M–C bond is its coordination to the electron-deficient metal center, with the titanium system featuring a higher energy barrier (12.7 versus 5.5 kcal mol–1). Ab initio molecular dynamics simulations have been performed on the Zr system by means of the Car–Parrinello method, to study the hitherto inaccessible mechanistic features of the insertion reactions.Contribution to the Björn Roos Honorary Issue  相似文献   

11.
Interactions of a series of amphiphilic cationic polyelectrolytes with various kinds of organic counteranions have been investigated in water by one- and two-dimensional 1H NMR spectroscopy at 20 °C. The cationic polyelectrolytes were prepared by micellar homopolymerization of tail-type cationic surface-active monomers with a cationic charge with -end, (ST–Cm–AB, m=5, 7, and 9, where ST is a styrenic group, Cm, an alkylene chain at the 4-position of styrene, and AB, alkyltrimethylammonium bromide). Aliphatic monosodium salt of maleic acid (MAS) and its stereoisomer, fumaric acid (FAS), sodium benzoate (NaB), potassium hydrogen phthalate (PHK), and sodium salicylate (NaSal) were added to a salt-free aqueous solution of the polyelectrolytes and 1H NMR measurements were carried out. Amphiphilic P(ST–Cm–AB) polyelectrolytes act as efficient hosts to strongly capture the hydrophobic counteranions B, PH, and Sal, but not MA and FA. The 1H NMR signals of these hydrophobic counteranions remarkably shift upfield and broaden in water in the presence of the amphiphilic polyelectrolytes. The nuclear Overhauser effect (NOE) signals between the cationic group of the polymer and aromatic benzoate counteranion protons are clearly observed to imply cation– interaction. The capturing of hydrophobic counterions by the polyelectrolytes is likely due to electrostatic, hydrophobic, and cation– interactions between them. The reduced viscosity, sp/Cp, for the solution at [PHK]/[P(ST–C7–AB)]=1.0 steeply increases with increasing polymer concentration (Cp) above ca. 0.9 g/dL to show pronounced viscoelasticity.  相似文献   

12.
The electronic structure of the Ca2 molecule has been investigated by use of a two-valence-electron semiempirical pseudopotential and applying the internally contracted multireference configuration interaction method with complete-active-space self-consistent-field reference wave functions. Core–valence correlation effects have been accounted for by adding a core-polarization potential to the Hamiltonian. The ground-state properties of the Ca2 and Ca2+ dimers have also been studied at the single-reference coupled-cluster level with single and double excitations including a perturbative treatment of triple excitations. Good agreement with experiment has been obtained for the ground-state potential curve and the only experimentally known A1u+ excited state of Ca2. The spectroscopic parameters De and Re deduced from the calculated potential curves for other states are also reported. In addition, spin–orbit coupling between the singlet and triplet molecular states correlating, respectively, with the (4p)1P and (4p)3P Ca terms has been investigated using a semi-empirical two-electron spin–orbit pseudopotential. Acknowledgement.This work was supported by grant 5 P03B 082 21 from the Polish State Committee for Scientific Research (KBN).  相似文献   

13.
The main objective of this study was to develop a thermodynamic model for predicting Cr(III) behavior in concentrated NaOH and in mixed NaOH–NaNO3 solutions for application to developing effective caustic leaching strategies for high-level nuclear waste sludges. To meet this objective, the solubility of Cr(OH)3(am) was measured in 0.003 to 10.5 m NaOH, 3.0 m NaOH with NaNO3 varying from 0.1 to 7.5 m, and 4.6 m NaNO3 with NaOH varying from 0.1 to 3.5 m at room temperature (22 ± 2°C). A combination of techniques, X-ray absorption spectroscopy (XAS) and absorptive stripping voltammetry analyses, were used to determine the oxidation state and nature of aqueous Cr. A thermodynamic model, based on the Pitzer equations, was developed from the solubility measurements to account for dramatic increases in aqueous Cr with increases in NaOH concentration. The model includes only two aqueous Cr species, Cr(OH) 4 and Cr2O2(OH) 4 (although the possible presence of a small percentage of higher oligomers at >5.0 m NaOH cannot be discounted) and their ion–interaction parameters with Na+. The logarithms of the equilibrium constants for the reactions involving Cr(OH) 4 [Cr(OH)3(am) + OH Cr(OH) 4 ] and Cr2O2(OH) 4 2– [2Cr(OH)3(am) + 2OH Cr2O2(OH) 4 2– + 2H2O] were determined to be –4.36 ± 0.24 and –5.24 ± 0.24, respectively. This model was further tested and provided close agreement between the observed Cr concentrations in equilibrium with Cr(OH)3(am) in mixed NaOH–NaNO3 solutions and with high-level tank sludges leached with and primarily containing NaOH as the major electrolyte.  相似文献   

14.
Studies on the magnetic properties of the molecular antiferromagnetic material {N(n-C5H11)4[MnIIFeIII(ox)3]}, carried out by various physical techniques (AC/DC magnetic susceptibility, magnetization, heat capacity measurements and Mössbauer spectroscopy) at low temperatures, have been presented. Different experimental observations complement each other and provide a clue for the observation of an uncompensated magnetization below the Néel temperature and short-range correlations persisting high above TN. It is understood that the honeycomb layered structure of the compound contains non-equivalent magnetic sub-lattices, (MnII–ox–FeIIIA–...) and (MnII–ox–FeIIIB–...), where different responses of the FeIIIA and FeIIIB spin sites towards an external magnetic field might be responsible for the observation of the uncompensated magnetization in this compound at T < TN. The present magnetic system is an S = 5/2 2-D Heisenberg antiferromagnet system with the intralayer exchange parameter J/kB = −3.29 K. A very weak interlayer exchange interaction was anticipated from the spin wave modeling of the magnetic heat capacity for T < 0.5TN. The positive sign of the coupling between the layers has been concluded from the Mössbauer spectrum in the applied magnetic field. Frustration in the magnetic interactions gives rise to the uncompensated magnetic moment in this compound at low temperatures.  相似文献   

15.
Raman spectra of aqueous Zn(II)–perchlorate solutions were measured over broad concentration (0.50–3.54 mol-L–1) and temperature (25–120°C) ranges. The weak polarized band at 390 cm–1 and two depolarized modes at 270 and 214 cm–1 have been assigned to 1(a 1g), 2(e g), and 5(f 2g) of the zinc–hexaaqua ion. The infrared-active mode at 365 cm–1 has been assigned to 3(f 1u). The vibrational analysis of the species [Zn(OH2) 2 + ] was done on the basis of O h symmetry (OH2 as point mass). The polarized mode 1(a 1g)-ZnO6 has been followed over the full temperature range and band parameters (band maximum, full width at half height, and intensity) have been examined. The position of the 1(a 1g)-ZnO6 mode shifts only about 4 cm–1 to lower frequencies and broadens by about 32 cm–1 for a 95°C temperature increase. The Raman spectroscopic data suggest that the hexaaqua–Zn(II) ion is thermodynamically stable in perchlorate solution over the temperature and concentration range measured. These findings are in contrast to ZnSO4 solutions, recently measured by one of us, where sulfate replaces a water molecule of the first hydration sphere. Ab initio geometry optimizations and frequency calculations of [Zn(OH2) 2 + ] were carried out at the Hartree–Fock and second-order Møller–Plesset levels of theory, using various basis sets up to 6-31 + G*. The global minimum structure of the hexaaqua–Zn(II) species corresponds with symmetry T h. The unscaled vibrational frequencies of the [Zn(OH2) 2 + ] are reported. The unscaled vibrational frequencies of the ZnO6, unit are lower than the experimental frequencies (ca. 15%), but scaling the frequencies reproduces the measured frequencies. The theoretical binding enthalpy for [Zn(OH2) 2 + ] was calculated and accounts for ca. 66% of the experimental single-ion hydration enthalpy for Zn(II).Ab initio geometry optimizations and frequency calculations are also reported for a [Zn(OH2) 2 18 ] (Zn[6 + 12]) cluster with 6 water molecules in the first sphere and 12 in the second sphere. The global minimum corresponds with T symmetry. Calculated frequencies of the zinc [6 + 12] cluster correspond well with the observed frequencies in solution. The 1-ZnO6 (unscaled) mode occurs at 388 cm–1 almost in perfect correspondence to the experimental value. The theoretical binding enthalpy for [Zn(OH2) 2 18 ] was calculated and is very close to the experimental single ion-hydration enthalpy for Zn(II). The water molecules of the first sphere form strong hydrogen bonds with water molecules in the second hydration shell because of the strong polarizing effect of the Zn(II) ion. The importance of the second hydration sphere is discussed.  相似文献   

16.
For the 102 atoms from He to Lr in their ground states, the average interelectronic angles <12> nl, n'l' between an electron in a subshellnl and another electron in a subshell n'l' are examined, where n and l are the principal and azimuthal quantum numbers, respectively. Theoretical study clarifies that <12> nl,n'l' are 90° precisely if ll' are even, while they are larger than 90° if ll' are odd. Numerical analysis of 3,275 subshell pairs with odd ll' of the 102 atoms shows that the increases in the total average interelectronic angles <12> from 90° are attributed predominantly to subshell pairs with n=n' and ll'=1.  相似文献   

17.
The particularity of metalloid clusters as a special kind of metal atom cluster is described. For the first time such metalloid clusters are investigated in the gas phase by means of FT/ICR–mass spectrometry, the results of which show that metalloid clusters represent a bridge between the bulk metal and metal compounds that can be found in solution after oxidation of the bulk metal. The metalloid clusters presented herein are [Ga19R6] (R=C(SiMe3)3), and SiAl14Cp*6 and the precursor Al4Cp*4 (Cp*= 5-C5Me5).  相似文献   

18.
17O-NMR spin-lattice relaxation timesT 1 of D2O molecules were measured at 5–85°C in D2O solutions of alkali metal halides (LiClCsCl, KBr, and KI), DCl, KOD, Ph4PCl, NaPh4B, and tetraalkylammonium bromides (Me4NBrAm4NBr) in the concentration range 0.1–1.4 mol-kg–1 TheB-coefficients of the electrolytes obtained from the concentration dependence of relaxation ratesR 1=1/T1 were divided into the ionicB-coefficients by three methods: (i) the assumption ofB (K+)=B(Cl), (ii) the assumption ofB(Ph4P+)=B(Ph4B), and (iii) the use ofB(Br) obtained from a series ofB(R4NBr). It was found that Methods (ii) and (iii) resulted in an abnormal temperature dependence of theB-coefficients of alkali metal ions and a negative values of rotational correlation times c at lower temperatures for hydroxide and halide ions. These results suggest that the methods based on the van der Waals volume are not adequate for the ionic separation of NMRB-coefficients. From the analysis using the assumption ofB(K+)=B(Cl), it was found that D3O+, OD, and Me4N+ ions are the intermediates between structure makers and breakers, and that the hydrophobicity of phenyl groups is weaker than that of alkyl groups due to the interactions between water molecules and -electrons in phenyl groups.  相似文献   

19.
Two modes of reactivity of N-silylphosphoranimines have been utilized to prepare the title compounds containing either B–N=P or Si–N=P–N–B linkages. First, silicon-nitrogen bond cleavage reactions of the N-silylphosphoranimines, Me3SiN=PMe(R)OCH2CF3 (1: R=Me, 2: R=Ph), with various chloroboranes gave the new N-borylphosphoranimines, Ph(Me2N)B–N=PMe2OCH2CF3 (2) and [(Me3Si)2N](Cl)B–N=PMe2OCH2CF3 (10). In other cases, however, the expected B–N=P products were unstable and cyclic phosphazenes [Me(R)P=N]3,4 were obtained. Second, deprotonation-substitution reactions of the aminophosphoranimines, Me3SiN=P(R)Me–N(R)H, were used to prepare a series of novel (borylamino)-phosphoranimines, Me3SiN=P(R)(Me)–N(R)–B(NMe2)2 (18: R=Me, R=t-Bu; 19: R=R=Me; 20: R=Ph, R=t-Bu; 21: R=Ph, R=Me) and Me3SiN=PMe2–N(t-Bu)–B(Ph)X (22: X=NMe2, 23: X=OCH2CF3). All of the new boron–nitrogen–phosphorus products were fully characterized by multinuclear NMR (1H, 13C, and 31P) spectroscopy and elemental analysis.  相似文献   

20.
Complete-active-space self-consistent-field calculation of the reorganisation energy, , corresponding to the strongly allowed HOMOLUMO transition in planar polyenes in the trans form (C 2 h symmetry), gives >0.5 eV. This large depends on the fact that the short and long bond lengths of the excited 1B u (or 3B u ) state compared to the 1A g ground state are almost cancelled. The emission redshift (Stokes shift) in molecules with the same type of system is quite small, however, which suggests that the Stokes shift may be dynamic, owing to the presence of another excited state at lower or about the same energy. Acknowledgement.We congratulate Björn on his birthday and at the same time thank him for the CASSCF method and for many years of collaboration and help from him and his collaborators to make this wonderful method work in our laboratory.Contribution to the Björn Roos Honorary Issue  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