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 62 毫秒
1.
MNDO calculations were made for triethylboroxin (EtBO)3 and triphenylboroxin (PhBO)3 using both X-ray determined and optimized geometry of these molecules. The results were compared with hypothetical “monomeric” molecules R---B=O. Calculated energies of trimerization are about −200 kJ mol−1 for both compounds and confirm the much higher stability of the “trimer”. Ebulliometric determination of molecular weight of triphenylboroxin in 2-pentanone confirms its trimeric character.  相似文献   

2.
The X-ray structure of [S4N3]Cl reveals three independent molecules, which all display π-facial interactions between the Cl and the pseudo-aromatic [S4N3]+ rings to produce a structure containing “inverse sandwich” systems.  相似文献   

3.
4.
The overall conformational order of the alkyl tail of the lyotropic lamellar phase of the dimethyldodecyl amineoxide, (CH3)2ON+(CH2)11CH3 (DDAO)---H2O system (75.7 wt.%) has been studied by using Fourier transform infrared spectroscopy. The spectral region 1000–1400 cm−1, covering the CH2 wagging modes and the methyl umbrella modes of DDAO, has been recorded in the temperature interval 3–60°C and at different mole fractions of gramicidin-D with respect to DDAO (Xg = 0.03, 0.05 and 0.1) for both DDAO-H2O and DDAO-D2O systems. It has been shown that the DDAO amphiphile molecules of the lamellar phase reorganise in a phase-like transition near 25–30°C. The DDAO-water system does not show any significant bands corresponding to a double gauche conformation at 1355 cm−1 nor to a gauche-transgauche (kinked) conformation at 1367 cm−1. These bands are probably present but hidden in the broad low-frequency side of the CH3 umbrella band at 1377 cm−1. Upon incorporation of gramicidin into the lamellar phase both head group and acyl chain spectra of the lipid change in such a way as to indicate a decreased “ordering” of the molecules, as judged by comparison with spectra of the same molecule in a micellar environment, and with increased fluidity of the acyl chains.  相似文献   

5.
Interactions between ethylselenoglycollic, selenoglycollic and ethylene-bis- selenoglycollic acids and some “soft”, “borderline” and “hard” metallic ions have been studied. The interactions of [PdCl4]aq2− with the three ligands were verified conductometrically and spectrophotometrically. The stability constants β1 and β2 for the system [PdCl4]2-- ethylene-bis-selenoglycollic acid have been determined at 25°C at ionic strength 3.0 M (NaCl). The stoichiometric ionization constant of the mentioned acid was also studied.  相似文献   

6.
The Raman spectra of F3PBH3 and F3PBD3 have been recorded (2500-10 cm−1) of the liquids (−80°C) and solids (−196°C) as well as the infrared spectra (4000-33 cm−1) of the solids. In the spectrum of the solid state many of the 10B and 11B fundamentals were clearly defined and it was also possible to assign the BH3 torsional frequency from the infrared and Raman spectra of the solids. A complete vibrational assignment is proposed and a normal coordinate calculation carried out. The force constant of 2.46 mdyn Å−1 for the P-B stretching mode is consistent with the short P-B bond; this constant is compared to the similar quantity for several other phosphorus-boron compounds. All of the E modes for the “free” molecule are shown to be split by the site symmetry which indicates that the molecules occupy Cs or C1 sites. The large number of observed lattice modes is consistent with two or more molecules per primitive cell. The torsional frequency was observed at 224 cm−1 and 167 cm−1 in hydrogen and deuterium compounds in the solid, respectively. These frequencies gave a periodic barrier of 4.15 kcal mole−1 for F3PBH3 and 4.31 kcal mole−1 for F3PBD3. CNDO/2 calculations have been carried out for F3PBH3 and the isoelectronic F3SiCH3 molecule in both the staggered and eclipsed forms and the dipole and barrier origins are discussed.  相似文献   

7.
Variable temperature (−105 to −150 °C) studies of the infrared spectra (3500–400 cm−1) of 1,1-dimethylhydrazine, (CH3)2NNH2, in liquid krypton have been carried out. No convincing spectral evidence could be found for the trans conformer which is expected to be at least 600 cm−1 less stable than the gauche form. The structural parameters, dipole moments, conformational stability, vibrational frequencies, and infrared and Raman intensities have been predicted from MP2/6-31G(d) ab initio calculations. The predicted infrared and Raman spectra are compared to the experimental ones. The adjusted r0 parameters from MP2/6-311+G(d,p) calculations are compared to those reported from an electron diffraction study. The energy differences between the gauche and trans conformers have been obtained from MP2 ab initio calculations as well as from density functional theory by the B3LYP method calculations from a variety of basis sets. All of these calculations indicate an energy difference of 650–900 cm−1 with the B3LYP calculations predicted the larger values. The potential function governing the conformational interchange has been predicting from both types of calculations and comparisons have been made. The barrier to internal rotation by the independent rotor model of the inner methyl group is predicted to have a value of 1812 cm−1 and that of the outer one of 1662 cm−1 from ab initio MP2/6-31G(d) calculations. These values agree well with the experimentally determined values of 1852±16 and 1558±12 cm−1, respectively, from a fit of the torsional transitions with the coupled rotor model. For the coupled rotor model the predicted V33 (sin 3τ0 sin 3τ1 term) value which ranged from 190 to 232 cm−1 is in reasonable agreement with the experimental value of 268±3 cm−1 but the predicted V33 (cos 3τ0 cos 3τ1 term) value of −73 to −139 cm−1 is 25% smaller and of the opposite sign of the experimental value of 333±22 cm−1. These theoretical and spectroscopy results are compared to similar quantities of some corresponding molecules.  相似文献   

8.
We present a molecular dynamics study of the solvation properties of large spherical ions S+ and S of same size, in water, chloroform and acetonitrile solutions, and at a water–chloroform interface. According to the “extrathermodynamic” TATB hypothesis, such ions have identical free energies of transfer from water to any solvent. We find that this is not the case, because S interacts better than S+ with water (by about 20 kcal mol−1), while S+ is better solvated by acetonitrile (by about 2 kcal mol−1) and chloroform (about 8 kcal mol−1) solvents. The importance of “long-range” electrostatic interactions on the charge discrimination by solvent is demonstrated by the comparison of standard vs corrected methods to calculate: (i) the electrostatic potential at the centre of the solute; (ii) the interaction energies between the ions and the solvents; and (iii) the free energies of charging the neutral sphere S0 to S+ and S, respectively. These conclusions are obtained with several solvent models and simulation conditions. The question of ion pairing for the S+S, S+Cl and SNa+ pairs is also examined in the three solvents. Finally, simulations at a liquid–liquid water–chloroform interface represented explicitly, show that S+ and S are highly surface active, although they do not possess, like classical surfactants, an amphiphilic topology. Adsorption at the interface is found with different methodologies and at different ion concentrations. These results are important in the context of the “TATB hypothesis”, and for our understanding of solvation of large hydrophobic ions in pure liquids or in heterogeneous liquid environments.  相似文献   

9.
Co-aggregation of fullerene C60 and thiophene has been studied calorimetrically in cyclohexene at 25 °C. The total aggregation heat is found to depend on initial concentration of thiophene and fall between −1.9 and −5.8 kJ mol−1. The corresponding thiophene/fullerene molar ratio (“co-aggregation number”) ranges from 7 to 12. The data are rationalized by formation of heteromolecular nanoaggregates with intermolecular contacts of both fullerene–thiophene and fullerene–fullerene types. A physical model describing interaction between fullerenes and π-donors in solution is substantiated and used to explain heterogeneity of composites containing fullerenes.  相似文献   

10.
The diffuse bands near 6100 Å in the laser-induced fluorescence spectrum of Cs2 are analyzed through quantum-mechanical spectral simulations. These bands are interpreted as bound-free emission to the vibrational continuum of the ground state from an excited state of ion-pair character. The lower region of this state, which we have labeled E′, is described approximately by the spectroscopic constants, Te = 19400 cm−1, Re = 9 Å, and ωe = 13 cm−1. Experiments with a single-mode Ar+ laser as excitation source clearly reveal fine structure in the E′ → X spectrum, which was not evident for multimode laser excitation. This fine structure confirms our analysis and supports our suggestion that extensive averaging over initial (υ′, J′) levels is responsible for the absence of fine structure in the spectra excited by a multimode laser. Various averaging mechanisms are investigated in the spectral calculations. The paper includes a brief review of other work on “structured continua” in diatomic spectra, and a semiclassical treatment of such structure, with emphasis on the distinction between “reflection” structure and “interference” structure.  相似文献   

11.
A Doppler-based velocity selection technique has been used to measure the relative velocity dependence of the cross sections σji,Δr) for rotationally inelastic collisions from level ji to ji + Δν1 = 8,22,42) in 7Li*2 A 1Σ+u)—Xe. The σjν±2r) are strongly attenuated at a smaller νr by “torque averaging” due to molecular rotation; in contrast, for large |Δ|, σj = νrn (1 n 2). An empirical intermolecular potential which reproduces these types of behavior for 3-D classical trajectories is exhibited.  相似文献   

12.
The photochemical reaction of azide derivatives induced by ultraviolet (UV) laser in matrix-assisted laser desorption/ionization mass spectrometry (MALDI) is reported. A novel synthesized class of azide aromatic derivatives, spin-labeled photoaffinity non-nucleoside adenosine triphosphate (ATP) analogs which are useful probes in study of muscle contraction mechanism, is used in this investigation. In the negative ion MALDI spectra of these ATP analogs, “fingerprint” peaks corresponding to [M − 10 − 1], [M − 12 − 1], [M − 16 − 1], [M − 26 − 1], [M − 28 − 1], [M − 41 − 1], and [M − 42 − 1] were observed with relative intensities depending on the MALDI matrix. Only the [M − 16 − 1] is present in the similar mass spectra of the analog in which the azido group is replaced by a hydrogen. A model is suggested for the photochemical reactions of azide derivatives under UV laser irradiation. The photoreaction fingerprint information is diagnostically useful in characterization of azido compounds, especially for spin-labeled photoaffinity non-nucleoside ATP analogs.  相似文献   

13.
The Gibbs energy of formation of IrO2(s) has been measured by means of oxygen dissociation pressure measurements, and by EMF measurements using ZrO2 (+ CaO) as the solid electrolyte. In addition, high-temperature enthalpy increments of IrO2 have ben measured from 416 to 940 K using a drop calorimeter. A “third law” evaluation of the experimental results and data from literature has been made. For the enthalpy of formation of IrO2(s) the value ΔH°f (298.15 K) - −(59.60 ± 0.03) kcal mole−1 has been selected. The thermodynamic functions of IrO2(s) have been calculated in the temperature range 298–1200 K.  相似文献   

14.
The phase speciation of thorium and consequences for the residence times of colloids have been examined in seawater of the Middle Atlantic Bight (MAB) and the Gulf of Mexico. Two fractions of colloidal organic matter (COM), 0.2 μm > COM1 > 1 kD and 0.2 μm > COM10 > 10 kD, were sampled using cross-flow ultrafiltration techniques and measured for their 234Th activity and organic carbon concentration. The ratios of mass concentrations of COM1 to those of suspended particulate matter were as high as 10 in the MAB and 6–34 in the Gulf of Mexico. Higher concentrations of colloids may be of great importance in the biogeochemical cycling of many particle-reactive nuclides or trace elements owing to their high specific surface area and complexation capacity. A significant fraction of 234Th in the traditionally defined “dissolved” pool was found to be associated with colloids. On average, about 10% of “dissolved” 234Th was in the colloidal fraction of sizes between 10 kDa and 0.2 μm, and 50% was in the 1 kDa-0.2 μm fraction. Values of the partition coefficients [Kc: (0.5−4) × 106 ml g−1 for Kc1 and (0.5−7) × 106 ml g−1 for Kc10] of 234Th between truly dissolved (<1 kDa) and colloidal fractions approximated those for Th-particle interactions [Kp: (0.3−10) × 106 ml g−1], indicating that colloid and suspended particle surface sites are similar. The distribution of 234Th between dissolved, colloidal, and particulate phases was broadly similar to that of organic carbon in these oceanic environments. Thus, thorium isotopes might be used as tracers of marine organic carbon cycling. Residence times of colloids derived from 234Th:238U disequilibria were consistently short, ranging from 1 to 14 days for COM10 and from 5 to 65 days for COM1, suggesting that marine colloids are highly reactive in marine biogeochemical processes. The discrepancy between apparent turnover times of colloids (1 kDa) derived from Th scavenging and 14C measurements suggest that 234Th and 14C may trace different geochemical pathways of colloids in the ocean.  相似文献   

15.
Complexes having the formula [C5Me5Fe(CO)-η4-diene]+ BF4 have been prepared by reaction of [C5Me5Fe(CO)2THF]+BF4 with a range of dienes. A Brönsted plot derived from the rates of addition of amine nucleophiles suggests an essentially “soft” nature for the complexes diene.  相似文献   

16.
Molar excess enthalpies HmE, isobaric heat capacities CP,mE, volumes VmE and isothermal compressibilities κTE for the 1,3-dioxane(3DX) + cyclohexane mixture were measured at 298.15 K, in order to compare to those of the 1,4-dioxane(4DX) + cyclohexane mixture. HmE is endothermic and the maximum value about 1.5 kJ mol−1 at x ≈ 0.45, and lower than that of the 4DX mixture by about 80 J mol−1. VmE is positive over the whole concentration and the maximum value is about 0.85 cm3 mol−1 at x ≈ 0.45, and lower than that of the 4DX mixture. The above results suggest the energetic unstabilization, resulting in the volume expansion in the mixture. CP,mE shows the characteristic W-shaped concentration dependence, which has maximum at x ≈ 0.45 and two minima at x ≈ 0.1 and 0.9. The maximum CP,mE value for 3DX mixture shifts toward the positive side, compared to that of 4DX mixture. κTE were estimated from speeds of sound, densities, thermal expansion coefficients and isobaric heat capacities of the pure component liquids and the mixtures. The κTE result shows the positive concentration dependence over the whole composition range. The 3DX mixture has the similar thermodynamic properties to the 4DX mixture, despite that 4DX is the nonpolar solvent and 3DX is the dipolar liquid. this means that there exists the local dipolar interaction between 4DX molecules, and the prevalence of “microheterogeneity” in the both mixtures.  相似文献   

17.
Pérez-Bustamante JA 《Talanta》1974,21(12):1291-1295
The preparation and spectrophotometric properties of a new type of complex compound of arsenazo I with Pu(IV) in the presence of H2O2 are described. The new compound has a blue colour, derived from a wide absorption band with a maximum at 610 nm. and a corresponding molar absorptivity of 4 × 104 l. mole−1.cm−1. From 2 hr after its preparation this curious new compound undergoes for several days a steady decomposition accompanied by decolorization. The formation of similar peroxy Pu(IV) complexes has not so far been shown to take place with arsenazo III or with any other “arsenazo-type” reagent.  相似文献   

18.
The generality of a two-electron reduction process involving an mechanism has been established for M3(CO)12 and M3(CO)12n(PPh3)n (M = Ru, Os) clusters in all solvents. Detailed coulometric and spectral studies in CH2Cl2 provide strong evidence for the formation of an ‘opened’ M3(CO)122− species the triangulo radical anions M3(CO)12−· having a half-life of < 10−6 s in CH2Cl2. However, the electrochemical response is sensitive to the presence of water and is concentration dependent. An electrochemical response for “opened” M3(CO)122− is only detected at low concentrations < 5 × 10−4 mol dm−3 and under drybox conditions. The electroactive species ground at higher concentrations and in the presence of water M3(CO)112− and M6(CO)182− were confirmed by a study of the electrochemistry of these anions in CH2Cl2; HM3(CO)11 is not a product. The couple [M6(CO)18]−/2− is chemically reversible under certain conditions but oxidation of HM3(CO)11 is chemically irreversible. Different electrochemical behaviour for Ru3(CO)12 is found when [PPN][X] (X = OAc, Cl) salts are supporting electrolytes. In these solutions formation of the ultimate electroactive species [μ-C(O)XRu3(CO)10] at the electrode is stopped under CO or at low temperatures but Ru3(CO)12−· is still trapped by reversible attack by X presumably as [η1-C(O)XRu3(CO)11]. It is shown that electrode-initiated electron catalysed substitution of M3(CO)12 only takes place on the electrochemical timescale when M = Ru, but it is slow, inefficient and non-selective, whereas BPK-initiated nucleophilic substitution of Ru3(CO)12 is only specific and fast in ether solvents particulary THF. Metal---metal bond cleavage is the most important influence on the rate and specificity of catalytic substitution by electron or [PPN]-initiation. The redox chemistry of M3(CO)12 clusters (M = Fe, Ru, Os) is a consequence of the relative rates of metal---metal bond dissociation, metal-metal bond strength and ligand dissociation and in many aspects resembles their photochemistry.  相似文献   

19.
We have combined the high sensitivity of the ICLAS technique with the rotational cooling effect of a slit jet expansion in order to observe and to understand the visible and near infrared NO2 spectrum. By this way, an equivalent absorption pathlength of several kilometers through rotationally cooled molecules has been achieved. Due to the vibronic interaction between the two lowest electronic states, 2A1 and à 2B2, this spectrum is vibronically dense and complex. Moreover, the dense room temperature rotational structure is perturbed by additional rovibronic interactions. In contrast, the rotational analysis of our jet cooled spectrum is straightforward. The NO2 absorption spectrum is vanishing to the IR but, owing to the high sensitivity of the ICLAS technique, we have been able to record the NO2 spectrum down to 11200 cm−1 with a new Ti:sapphire ICLAS spectrometer. As a result 249 2B2 vibronic bands have been observed (175 cold bands and 74 hot bands) in the 11200–16150 cm−1 energy range. Due to the cooling effect of the slit jet we have reduced the rotational temperature down to about 12 K and at this temperature the K = 0 subbands are dominant. Consequently, we have analysed only the K = 0 manifold for N 7 of each vibronic band. The dynamical range of the band intensities is about one thousand. Due to the strong vibronic interaction between the 2A1 and à 2B2 electronic states, we observed not only the a1 vibrational levels of the à 2B2 state but also the b2 vibrational levels of the 2A1 state interacting with the previous ones. By comparison with the calculated density of states, we conclude that we have observed about 65% of the total number of 2B2 vibronic levels located in the studied range. However, there are more missing levels in the IR because of the weakness of the spectrum in this range. The correlation properties of this set of vibronic levels have been analysed calculating the power spectrum of the absorption stick spectrum which displays periodic motions: the dominant period, at 714 ± 20 cm−1, corresponds to the bending motion of the à 2B2 state. The other observed periods remain unassigned. In contrast the next neighbor spacing distribution (NNSD) shows a strong level repulsion, i.e. a manifestation of quantum chaos. These two observations, apparently contradictory, can be rationalized as follows: the short time dynamics, for t < 10−12 s, is “regular” while for longer times the dynamics becomes “chaotic”. We suggest that this behavior may be observed directly with a pump and probe fs laser experiment.  相似文献   

20.
The paper reports results of a study on the specific adsorption of F, Cl, Br, I, ClO3, BrO3, IO3 and IO4 on hydrous γ-Al2O3. The isotherms of the anion adsorption and the adsorption dependencies on pH and the ionic strength of the solution have been determined under the equilibrium conditions. According to the degree of affinity to γ-Al2O3, the anions can be ordered as: I3334−. It has been established that the sorption of IO4 and F involves the formation of surface complexes in the inner co-ordination sphere, whereas that of Cl, Br, I, ClO3, BrO3 and IO3 takes place through formation of ion pair complexes in the outer co-ordination sphere. In the dynamic system, the exchange isoplanes and elution curves have been determined for selected anions on columns filled with Al2O3. It has been shown that γ-Al2O3 can be used for isolation and concentration of IO3 from natural waters in order to decrease the limit of the ions determination to 2 μg l−1. Using differential pulse voltammetry (DPV), after isolation and concentration on γ-Al2O3, the content of iodates has been determined in mineral, marine and tap water doped with these ions.  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