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.
The NMR spectra of solutions of 30%17O-enriched H2O and D2O in nitromethane display the resonances of the three isotopomers H2O, HDO, and D2O. All17O,1H and17O,2H coupling constants and the primary and secondary isotope effects onJ(17O,1H) have been determined. The primary effect is −1.0 ± 0.2 Hz and the secondary effect is −0.07 ± 0.04 Hz. Using integrated intensities in the17O NMR spectra, the equilibrium constant for the reaction H2O + D2O 2HDO is found to be 3.68 ± 0.2 at 343 K. From the relative integrated intensities of proton-coupled and -decoupled spectra the17O–{1H} NOE is estimated for the first time, resulting in values of 0.908 and 0.945 for H2O and HDO, respectively. This means that dipole–dipole interactions contribute about 2.5% to the overall17O relaxation rate in H2O dissolved in nitromethane.  相似文献   

2.
Existing evidence indicates that between 248°C and the melting point at 406°C, KOH is a rotator phase. We have shown that, as might be expected, this results in enhanced proton conductivity, and a value of 2×10−3 ohms−1 cm−1 was found at 350°C, which is the highest reported for proton conducting solid electrolytes in this temperature range. Excess protons are provided by water molecules residing on the normal OH- sites, and charge compensation is provided by CO2−3 ions in the solid solution of KOH(≈ 1 m/o K2CO3, 1.3m/o H2O). The activation energy for proton hopping between adjacent H2O and OH species probably accounts for most of the observed activation energy of 53±3 kJ mol−1. From TGA studies the isobars at 0.05 and 10 Torr were established for KOH-rich compositions in the KOH---H2O system, and it was shown that the rotator phase of KOH is stable between these vapour pressures.  相似文献   

3.
By using resonance-enhanced two-photon ionization, rotationally resolved spectra of the 610 band of 12C6D6 and (13C12C5D6 molecules have been obtained for the first time at a rotational temperature of 0.7 K in a pulsed supersonic beam. From the former, the values of B″ = 0.1573 ± 0.0008 cm−1, B′ = 0.1508 ± 0.0008 cm−1, and ξ′ = −0.412 ± 0.050 have been derived for rotational and Coriolis constants in the lower and upper levels of 12C6D6. Also, the spectra corresponding to 12C6H6 and 13C12C5H6 have been measured and the values B″ = 0.1892 ± 0.0008 cm−1, B′ = 0.1815 ± 0.0008 cm−1, and ξ′ = −0.586 ± 0.050 have been obtained for 12C6H6, in agreement with previous results. Rotational constants of 13C labeled benzene molecules have been geometrically deduced from the constants obtained. Experimental isotopic shifts of the vibronic origins of the 6a10 and 6b10 bands have been determined. There is agreement with previous 13C-benzene-h6 data. The present results are −0.91 ± 0.05 and 3.09 ± 0.05 cm−1 for 13C12C5D6 and −1.64 ± 0.05 and 2.64 ± 0.05 cm−1 for 13C12C5H6. The splittings of vibrational modes 6b and 6a in the 1B2u state are 4.00 ± 0.10 cm−1 for 13C12C5D6 and 4.28 ± 0.10 cm−1 for 13C12C5H6.  相似文献   

4.
The pure rotational spectrum of CH2F2 was recorded in the 20–100 cm−1 spectral range and analyzed to obtain rotation and centrifugal distortion constants. Analysis of the data yielded rotation constants: A = 1.6392173 ± 0.0000015, B = 0.3537342 ± 0.00000033, C = 0.3085387 ± 0.00000027, τaaaa = −(7.64 ± 0.46) × 10−5, τbbbb = −(2.076 ± 0.016) × 10−6, τcccc = −(9.29 ± 0.12) × 10−7, T1 = (4.89 ± 0.20) × 10−6, and T2 = −(1.281 ± 0.016) × 10−6cm−1.  相似文献   

5.
Several elementary reactions of formyl radical of combustion importance were studied using pulsed laser photolysis coupled to transient UV–Vis absorption spectroscopy: HCO → H + CO (1), HCO + HCO → products (2), and HCO + CH3 → products (3). One-pass UV absorption, multi-pass UV absorption as well as cavity ring-down spectroscopy in the red spectral region were used to monitor temporal profiles of HCO radical. Reaction (1) was studied over the buffer gas (He) pressure range 0.8–100 bar and the temperature range 498–769 K. Reactions (2a), (2b), (2c), (3a) and (3b) as well as the UV absorption spectrum of HCO, were studied at 298 and 588 K, and the buffer gas (He) pressure of 1 bar. Pulsed laser photolysis (308, 320, and 193 nm) of acetaldehyde, propionaldehyde, and acetone was used to prepare mixtures of free radicals. The second-order rate constant of reaction (1) obtained from the data at 1 bar is: k1(He) = (0.8 ± 0.4) × 10−10exp(−(66.0 ± 3.4) kJ mol−1/RT) cm3 molecule−1 s−1. The HCO dissociation rate constants measured in this work are lower than those reported in the previous direct work. The difference is a factor of 2.2 at the highest temperature of the experiments and a factor of 3.5 at the low end. The experimental data indicate pressure dependence of the rate constant of dissociation of formyl radical 1, which was attributed to the early pressure fall-off expected based on the theory of isolated resonances. The UV absorption spectrum of HCO was revised. The maximum absorption cross-section of HCO is (7.3 ± 1.2) × 10−18 cm2 molecule−1 at 230 nm (temperature independent within the experimental error). The measured rate constants for reactions (2a), (2b), (2c), (3a) and (3b) are: k2 = (3.6 ± 0.8) × 10−11 cm3 molecule−1 s−1 (298 K); k3 = (9.3 ± 2.3) × 10−11 cm3 molecule−1 s−1(298 and 588 K).  相似文献   

6.
The interaction of Na9[SbW9O33]·19.5H2O (SbW) with bovine serum albumin (BSA) is studied by spectroscopic and voltammetric methods. Absorption spectroscopy of BSA and the linear sweep voltammetry of SbW proved the formation of ground-state SbW–BSA complex. Fluorescence quenching of serum albumin by SbW is also found to be a static quenching process. The binding constant Ka is 4.13×104 L mol−1 for SbW–BSA at pH 7.40 Tris–HCl buffer at 295 K. The number of binding sites and the apparent binding constants at different temperatures are obtained from the analysis of the fluorescence quenching data. The thermodynamic parameters determined by the Van’t Hoff analysis of the binding constants (ΔH=−80.01 kJ mol−1 and ΔS=−182.85 J mol−1 K−1) clearly show that the binding is absolutely entropy driven. Hydrogen bonding and van der Waals interaction force play major role in stabilizing the complex. The effect of SbW on the conformation of BSA is analyzed using synchronous fluorescence spectroscopy.  相似文献   

7.
Yb3+-doped ceramic strontium cerate of exactly the composition SrCe0.95Yb0.05O3 − α was prepared, having a relative density of 99.0 (± 0.3%). Great care was taken to obtain homogeneous, carbonate free material. Analysis are made of the X-ray powder diffraction pattern of the as-prepared dense ceramic, resulting in the orthorhombic unit cell parameters a = 6.997(2) Å, b = 12.296(3) Å, c = 8.588(2) Å, Z = 8 and dx = 5.806(2) g cm−3. Bending strength values of the ceramic in non-proton and proton conducting state are found to be 177 and 194 MPa respectively. The ceramic kept under proton conducting conditions for 500 h at 300 °C to 800 °C in a N2 flow containing 155 mbar water vapour and 245 mbar H2, have shown to remain chemically and structurally stable. Impedance spectroscopy measurements of the bulk conductivity of the proton conducting ceramic revealed an activation energy of 53.2 kJ mol−1 and a preexponential factor of 359.1 (Ω cm)−1 K. In the non-proton conducting state the ceramic is mainly oxygen ion vacancy conducting, which indicates that charge compensation on substituting Yb+3 in SrCeO3 takes place by oxygen ion vacancies.  相似文献   

8.
Rotational analyses have been performed on the emission spectra of the 0-0, 1-1, 2-2, and 3-3 bands of the β system (c1Φ - a1Δ) of the TiO molecule, excited in a microwave discharge through a mixture of helium, oxygen and TiCl4 vapor. Rotational constants were obtained for all the bands from which the following equilibrium constants were derived. Be=0.52301±0.00008 cm, αe=0.00313±0.0006 cm, re=1.6391±0.0001 AHigher order constants, Dv and Hv, were calculated for the various vibrational levels.  相似文献   

9.
Using a high-resolution Fourier transform spectrum of hydrogen selenide in natural abundance, about 600 intensities of lines belonging to the ν1, ν3, and 2ν2 bands of H280Se were measured. A least-squares fit of these intensities was performed, allowing determination of the vibrational transition moments of these bands and their rotational corrections. Finally, the first derivatives of the dipole moment with respect to the normal coordinates q1 and q3 were found to be ∂μχ/∂q1 = (−0.5938 ± 0.010) × 10−1 and ∂μz/∂q3 = (0.5683 ± 0.010) × 10−1 Debye, respectively.  相似文献   

10.
Ro-vibrational spectra of HNCS and DNCS have been obtained in the spectral range 300–4000 cm−1 with a practical resolution limit of 0.06 cm−1 in the region 350–1200 cm−1 and 0.15 cm−1 in the region 1200–4000 cm−1. The observed fine structure permitted definitive assignments for some of the PQK, QQK, and RQK branches in both molecules, and yielded sets of rotational constants in substantial agreement with those obtained from recent microwave and far-infrared studies. Precise estimates of the band origins have been obtained and there is evidence of second-order Coriolis coupling between the three bending modes in each molecule. The isolation of the out-of-plane bending modes has lead to a re-assignment of ν3, ν4, ν5, and ν6 for each molecule. The band origins, uncorrected for Coriolis interaction, are for HNCS and DNCS, respectively. v1:3538.6 ±0.3, 2644.5±0.5cm−1;v2:1989.0 ±0.3, 1944.3±0.5cm−1;v3:857.0 ±0.6, 851.0±0.1cm−1;v4:615.0 ±0.5, 549.1±0.2cm−1;v5:469.2 ±0.1, 365.8 ±0.2cm−1;v6:539.2 ±0.5, 481.0±0.1cm−1;  相似文献   

11.
The overtone band 2ν08 of CH3CN around 720 cm−1 has been measured on a Bruker Fourier transform spectrometer at a resolution of 0.003 cm−1. Only the parallel band was observed, but due to the l(2, 2) resonance, ΔK = −2 lines leading to the v8 = 2, l8 = −2 levels with K = 1-3 could be seen. More information for the l8 = ±2 component of the vibrational state v8 = 2 was evaluated from the hot band 2ν±28 - ν±18. Altogether more than 1000 lines were assigned. In the fit pure rotational lines from literature were also combined. Among the results the anomalous A0 - A′ values 4.6722(13) × 10−3 cm−1 for the 2ν08 band and 7.0324(32) × 10−3 cm−1 for the 2ν±28 band are striking.  相似文献   

12.
Bis-alkynylated oligoethyleneglycol (OEG) and a monopropargyl-functionalized perfluorinated ethylene glycol (FEG) were clicked to azide-functionalized gold surface (Au–N3) at room temperature via the well known 1,3 cycloaddition click chemical reaction. The Au–N3 substrate was obtained by nucleophilic attack of NaN3 on gold substrates modified by the electrochemical reduction of the , +N2–C6H4–CH2Br diazonium salt. This electrochemical process yields aryl layer-modified gold of the type Au–C6H4–CH2Br (hereafter Au–Br). The untreated and modified gold plates were examined by XPS, PMIRRAS and contact angle measurements. XPS brought evidence for electrografting aryl layers by the detection of Br3d; azide functionalization by the increase of the N/Br atomic ratio; and click reaction of OEG with Au–N3 by the increase of O/N ratio. In addition, the perfluorinated plate (Au-FEG) exhibited F1s and characteristic C1s peaks from -(CF2)7- chain and terminal CF3. Infra red spectroscopy (PMIRRAS) evidenced (i) grafting N3 to Au–Br; (ii) characteristic stretching bands, from ethylene glycol units, C–O–C (1100–1300 cm−1); CF2 (1000–1100 cm−1) and CF3 (1100–1350 cm−1) from FEG grafts; and (iii) suppression of alkynyl bands from OEG and FEG after surface click chemistry. More importantly, PMIRRAS results support an important bridging of the bispropargyl oligoethylene glycol at the gold surface. Water drop contact angles were found to be 48.7° and 83.0° for Au-OEG and Au-FEG, respectively, therefore highlighting the control over the hydrophilic/hydrophobic character of the clicked substrate.This work shows that clicking macromolecules to grafted, diazonium salt-derived aryl layers is a novel, simple and valuable approach for designing robust, functional surface organic coatings.  相似文献   

13.
Amorphous, nanocrystalline, and bulk AlO(OH) · xH2O crystals have six fundamental modes (FM) of vibration in a nonlinear AlO(OH) molecular structure. Most of them appear in groups of four IR and Raman bands. Their positions and relative intensities differ significantly in three specimens. The nanocrystals (monoclinic structure with z=8 molecules per unit cell) have four OH stretching bands at values enhanced by up to 360 cm−1 at 3120, 3450, 3560 cm−1 in comparison to those in bulk crystals or amorphous specimens. The first two bands are broad, bandwidth Δν1/2200 to 350 cm−1, while the other two are sharp, Δν1/290 cm−1. The sharp bands shift to 3525 and 3595 cm−1 after heating the sample at 100°C. They no longer appear after heating at 300 or 500°C for 2 h (the specimen decomposes to Al2O3), leaving behind only two bands at 3100 and 3400 cm−1. A Δν1/2 value of 500 cm−1 appears in the 3400 cm−1 in a delocalized distribution of H atoms. Two bands also occur at 3098 and 3300 cm−1 in bulk crystals (orthorhombic structure with z=4) or at 2990 and 3515 cm−1 in an amorphous sample. More than one bands appear in a FM vibration in occurrence of sample in more than one conformers. The amorphous sample has approximately the same conformer structure as the bulk crystals. An amorphous surface structure exists in nanocrystals with a group of three bands at 1420, 1510 and 1635 cm−1 in an interconnected network structure. It encapsulates the nanocrystals in an amorphous shell. Its volume fraction, 33% estimated from the integrated intensity in three bands, determines 2.2 nm thickness in the shell in spherical shape of nanocrystals in 35 nm diameter.  相似文献   

14.
Studies on the acid-base properties and solubility of a polyammonium polyelectrolyte (chitosan) with different molecular weights (MW 310 and 50 kDa), were performed at T = 25 °C, in the pH range 2.5–7. The protonation of chitosan was investigated by potentiometry ([H+]-glass electrode) in NaCl, NaNO3 and mixed NaNO3 + Na2SO4 ionic media, at different ionic strengths. Protonation constants were calculated as a function of dissociation degree α by means of two different models, namely, a simple linear model and the modified Henderson–Hasselbalch equation. Experimental data were also fitted using a model independent of α (Diprotic-like model), according to which the acid-base properties can be simply described by two protonation constants in all the acidic pH range. The dependence on ionic strength of protonation constants in NaCl aqueous solution was modelled by Specific ion Interaction Theory (SIT). The ion pair formation between protonated chitosan and Cl, NO3 and SO42− was also considered, and the relative formation constants are reported.Solubility investigations were performed in NaCl aqueous solutions in a wide range of ionic strength (0.1 < I/mol L− 1 < 3.0), with the aim to determine the activity coefficients of neutral species and the Setschenow coefficient of chitosan 310 kDa.  相似文献   

15.
The spin–lattice relaxation times and spin–spin relaxation times for 1H and M in M5H3(SO4)4·H2O (M=Na, K, Rb, and Cs) single crystals grown using the slow-evaporation method were measured as functions of temperature. Two kinds of protons were identified in the M5H3(SO4)4·H2O structure: acid protons and water protons. Our experimental results show that the acid and water protons in Cs5H3(SO4)4·H2O are involved in phase transitions of this crystal, whereas neither type of proton is involved in the phase transitions of the other three crystal type (M5H3(SO4)4·H2O; M=Na, K, and Rb). Moreover, the relaxation times for the M (=Na, K, and Rb) nuclei in these crystals were found to decrease with increasing temperature and can be described with (k=2). The T1 results for M (=Na, K, and Rb) in M5H3(SO4)4·H2O crystals can be explained in terms of a relaxation mechanism in which the lattice vibrations are coupled to the nuclear electric quadrupole moments.  相似文献   

16.
The electron impact behavior of CO adsorbed on was investigated. The desorption products observed were neutral CO, CO+, and O+. After massive electron impact residual carbon, C/W = 0.15, but not oxygen was also found, suggesting that energetic neutral O, not detected in a mass analyzer must also have been formed. Formation of β-CO, i.e., dissociated CO with C and O on the surface was not seen. The total disappearance cross section varies only slightly with coverage, ranging from 9 × 10 −18 cm2 at low to 5 × 10−18 cm2 at saturation (CO/W = 0.75). The cross section for CO+ formation varies from 4 × 10−22 cm2 at satura to 2 × 10−21 cm2 at low coverage. That for O+ formation is 1.4 × 10−22 cm2 at saturation and 2 × 10−21 cm2 Threshold energies are similar to those found previously [J.C. Lin and R. Gomer, Surf. Sci. 218 (1989) 406] for and CO/Cu1/W(110) which suggests similar mechanisms for product formation, with the exception of β-CO on clean W(110). It is argued that the absence or presence of β-CO in ESD hinges on its formation or absence in thermal desorption, since electron impact is likely to present the surface with vibrationally and rotationally activated CO in all cases; β-CO formation only occurs on surfaces which can dissociate such CO. It was also found that ESD of CO led to a work function increase of the remaining Pd1/W(110) surface of 500 meV, which could be annealed out only at 900 K. This is attributed to surface roughness, caused by recoil momentum of energetic desorbing entities.  相似文献   

17.
The hydrolysis of VO2+ and the complex with sulfate were studied potentiometrically, spectrophotometrically and calorimetrically, in NaCl aqueous solution (0 < I ≤ 1 mol L− 1) and at t = 25 °C. The formation of two hydrolytic species VO(OH)+ and VO2(OH)22+ and one complex with sulfate was found, with log β = − 5.65 for the reaction VO2+ + H2O = VO(OH)+ + H+, log β = − 7.02 for the reaction 2VO2+ + 2H2O = (VO)2(OH)22+ + 2H+ and log K = 1.73 for VOSO40 species (at I = 0.1 mol L− 1 and t = 25 °C). For these species, using calorimetric data, ΔH and TΔS values were also obtained. By using the above values, interactions of VO2+ with acetate (ac), malonate (mal), succinate (suc), 1,2,3-propanetricarboxylate (tca) and 1,2,3,4-butanetetracarboxylate (btc) ligands were studied potentiometrically and spectrophotometrically. The formation of ML+, ML20 and MLOH0 for ac; ML0, MLH+, ML22− and ML2H for mal; ML0, MLH+ and MLOH for suc; ML and MLH0 for tca and ML2−, MLH and MLH20 for btc were found. Formation constants are reported at I = 0.1 mol L− 1, together with SIT parameters for the dependence on ionic strength. By visible spectrophotometric measurements, λmax and εmax values for the relevant species in solution were determined.  相似文献   

18.
《Physics letters. [Part B]》2008,660(5):466-470
A partial-wave analysis of the reaction πpηηπp at 18 GeV/c has been performed on a data sample of approximately 4000 events obtained by Brookhaven experiment E852. The JPC=0−+π(1800) state is observed in the a0(980)η and f0(1500)π decay modes. It has a mass of 1876±18±16 MeV/c2 and a width of 221±26±38 MeV/c2. The JPC=2−+π2(1880) meson is observed decaying through a2(1320)η. It has a mass of 1929±24±18 MeV/c2 and a width of 323±87±43 MeV/c2. Both states are potential candidates for non-exotic hybrid mesons.  相似文献   

19.
The Ag2O–TiO2–SiO2 glasses were prepared by Ag+/Na+ ion-exchange method from Na2O–TiO2–SiO2 glasses at 380–450 °C below their glass transition temperatures (Tg), and their electrical conductivities were investigated as functions of TiO2 content and the ion-exchange ratio (Ag/(Ag+Na)). In a series of glasses 20R2xTiO2·(80−x)SiO2 with x=10, 20, 30 and 40 in mol%, the electrical conductivities at 200 °C of the fully ion-exchanged glasses of R=Ag were in the order of 10−5 or 10−4 S cm−1 and were 1 or 2 orders of magnitude higher than those of the initial glasses of R=Na. The glass of x=30 exhibited the highest increase of conductivity from 3.8×10−7 to 1.3×10−4 S cm−1 at 200 °C by Ag+/Na+ ion exchange among them. When the ion-exchange ratio was changed in 20R2O·30TiO2·50SiO2 system, the electrical conductivity at 200 °C exhibited a minimum value of 7.6×10−8 S cm−1 around Ag/(Ag+Na)=0.3 and increased steeply in the region of Ag/(Ag+Na)=0.5–1.0. When the ion-exchange temperature was changed from 450 to 400 °C, the conductivity of the ion-exchanged glass of x=30 decreased. The infrared spectroscopy measurement revealed that the ion-exchange temperature of 450 °C induced a structural change in the glass of x=30. The Tg of the fully ion-exchanged glass of x=30 was 498 °C. It was suggested that the incorporated silver ions changed the average coordination number of titanium ions to form higher ion-conducting pathway and resulted in high conductivity in the titanosilicate glasses.  相似文献   

20.
Michael A. Henderson 《Surface science》2010,604(13-14):1197-1201
Temperature programmed desorption (TPD), electron energy loss spectroscopy (ELS) and low energy electron diffraction (LEED) were used to study the interaction of molecular oxygen with the (2 × 1) reconstructed surface of hematite α-Fe2O3(011­2) under UHV conditions. The (2 × 1) surface is formed from vacuum annealing of the ‘ideal’ (1 × 1) surface and possesses Fe2+ surface sites based on ELS. While O2 does not stick to the (1 × 1) surface at 120 K, the amount of O2 that can be reversibly adsorbed at 120 K on the (2 × 1) surface was estimated to be ~ 0.5 ML (where 1 ML is defined as the Fe3+ surface coverage on the ideal (1 × 1) surface), with additional O2 that is irreversibly adsorbed based on subsequent H2O TPD. Molecularly and dissociatively adsorbed O2 modifies the surface chemistry of H2O both in terms of enhanced OH stability (relative to either the (1 × 1) or (2 × 1) surfaces) and in the blocking of H2O adsorption sites. While O2 adsorption at 120 to 300 K does not transform the (2 × 1) surface into the (1 × 1) surface, the influence of O2 on the (2 × 1) surface involves both charge transfer from surface Fe2+ sites and formation of an ordered c(2 × 2) structure resulting from O2 dissociation.  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