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 125 毫秒
1.
The apparent molal volume φ v of boric acid has been determined in various sodium chloride solutions at 0 and 25°C from precise density measurements. Similar to its behavior in pure water, the φ v of boric acid in NaCl solutions is a linear function of the concentration. The infinite dilution φ v ° and the slope S v * of B(OH)3 are larger in NaCl solutions than in pure water. NaCl appears to be able to dehydrate B(OH)3 and cause an increase in B(OH)3-B(OH)3 interactions. The mean apparent molal volumes Φ v of the B(OH)3−NaCl solutions are predicted from pure water data using a modification of Young's rule for electrolyte-nonelectrolyte mixtures and are compared to the directly measured values. A similar treatment was carried out on the density data of acetic acid-sodium chloride solutions. The modified Young's rule was found to give a good first approximation of the mean apparent molal volumes of nonelectrolyte-electrolyte systems. The deviations from the Young's rule approximation are studied as excess volumes of mixing boric acid and NaCl solutions. Taken from a thesis submitted by Gary K. Ward in partial fulfillment of the requirements of the Master of Science degree, University of Miami, Miami, Florida 33149. Scientific Contribution Number 1731 from the University of Miami, Rosenstiel School of Marine and Atmospheric Science, Miami, Florida 33149.  相似文献   

2.
Vibrational relaxation dynamics of monomeric water molecule dissolved in d-chloroform solution were revisited using the two dimensional Infrared (2D IR) spectroscopy. The vibrational lifetime of OH bending in monomeric water shows a bi-exponential decay. The fast component (T1=(1.2±0.1) ps) is caused by the rapid population equilibration between the vibrational modes of the monomeric water molecule. The slow component (T2=(26.4±0.2) ps) is mainly caused by the vibrational population decay of OH bending mode. The reorientation of the OH bending in monomeric water is determined with a time constant of τ=(1.2±0.1) ps which is much faster than the rotational dynamics of water molecules in the bulk solution. Furthermore, we are able to reveal the direct vibrational energy transfer from OH stretching to OH bending in monomeric water dissolved in d-chloroform for the first time. The vibrational coupling and relative orientation of transition dipole moment between OH bending and stretching that effect their intra-molecular vibrational energy transfer rates are discussed in detail.  相似文献   

3.
The present work reviews a systematic spectroscopic investigation of water in the single crystal K4Fe(CN)6 · 3H2O (KFCT) and its isotopic solid solutions. On the basis of precise polarization measurements of Raman spectra and infrared studies, the positions of the OH(OD) oscillators in the ordered crystal state (T < 150 K) are determined. Overtones and combinations of water vibrations in the near infrared region are studied that allow for the evalution of the coupling and anharmonicity constants of the modes. The isotopic dilution technique is used in order to investigate the influence of intra- and intermolecular interactions on the OH(OD) vibrations. Information about the distortion of the water molecules, as well as many correlations between spectral parameters of OH(OD) stretchings and structural characteristics of H bonds in KFCT, are derived. As a result, a complete picture of the vibrational dynamics and the structure of the crystallization water in KFCT is obtained.  相似文献   

4.
To understand the effect of different vibrational and rotational modes of reactant to enhance the reactivity of the O + HO2 → OH + O2 reaction, we revisited this important atmospheric reaction. We report here a quasi-classical trajectory (QCT) study of the reaction dynamics on a recently developed full-dimensional potential energy surface (PES). Our previous work has indicated that this reaction has two pathways, the H abstraction (HA) channel and the O abstraction (OA) channel, which lead to totally different product energy distribution. In this work, we identified that the vibrational excitation of the OH stretching (v1) mode of HO2 is the switch of the HA channel at low collision energy; meanwhile, the rotational excitation can also greatly change the branching ratio of the two pathways. With the excitation of v1 mode, the original negligible HA channel controlled by the tight transition state becomes quite important. This work presents an approach to control the branching ratio via collaboration between vibrational and rotational excitation and will enrich the knowledge of the O + HO2 reaction in atmospheric chemistry and physics.  相似文献   

5.
Density functional method B3LYP plus the AUG‐cc‐pVDZ and AUG‐cc‐pVTZ basis sets is used to investigate ring normal modes of halogen‐substituted pyridines involved in the N ··· H? X H‐bonds with HX (X = F, Cl). The results demonstrated that the formation of hydrogen bond leads to an increase in the frequencies of the ring breathing mode v1, the N‐para‐C stretching mode v6a and the meta‐CC stretching mode v8a, whereas there is no change in the triangle mode v12 for free pyridine and a smaller blue shift for substituted pyridines. There is a strong coupling between the C? Y stretching vibration and the triangle mode (ortho‐ and para‐substituted) or the breathing mode (meta‐substituted) in substituted pyridines, which leads to the frequency decrease in the triangle or breathing modes. The natural bond orbital analysis suggests that electrostatic interaction and charge transfer caused by the intermolecular and intramolecular hyperconjugations are the origin of the frequency blue shift in the ring stretching modes. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2012  相似文献   

6.
Preparation of trans-[Pt(ox)2X2]2? (X = Cl, Br, I, SCN, OH) by Oxidative Addition to [Pt(ox)2]2? in Organic Solvents After extraction of [Pt(ox)2]2? with long-chain alkyl-ammonium ions into organic solvents the new PtIV complexes trans-[Pt(ox)2X2]2?, X = Cl, Br, I, SCN, OH, are formed directly by oxidative addition. In nonpolar solvents the bulky organic cations prevent the formation of compounds with columnar structure which by partial oxidation in aqueous solution are formed immediately. The IR and Ra spectra of the stable anhydrous (TBA) salts are assigned according to point group D2h. A characteristical dependence of the C?O, C? O, and Pt? O stretching modes in response to the oxidation state of the central ion is observed. There is vibrational fine structure in the absorption spectrum of [Pt(ox)2]2? measured at 10 K with long progressions by coupling of d—d transitions with vs(Pt? O) and vs(C?O). The characteristical feature in the UV/VIS spectra of the PtIV complexes is caused by intensive π(O, X) ← eg(Pt) CT transitions.  相似文献   

7.
Infrared spectra (IR, FIR, DRIFT, 90 and 295 K) and DSC measurements of the various polymorphs of iron oxide hydroxide, viz. goethite (α), akaganéite (β), lepidocrocite (γ), and feroxyhite (δ), and of deuterated specimens are reported. They are discussed with respect to the crystal structures proposed in the literature, the hydrogen bonds present, the energies of the OH stretching, OH bending (librational), and translational modes, and their thermal decomposition. From the two space groups proposed for β- and γ-FeO(OH), the groups I4/m and Cmc21, respectively, seem to be more reliable. The disorder of the OH ions of γ-FeO(OH) has not been confirmed in contrast to that of δ-FeO(OH). The intraionic O(H,D) distances of γ- and δ-FeO(OH) derived from neutron powder diffraction studies have to be doubted. The greater strength of the OHOH hydrogen bonds of lepidocrocite, for example, compared to that of the OHO hydrogen bonds of goethite despite the larger hydrogen bond acceptor capability of O2− is due to the strong cooperativity of the hydrogen bonds of the γ-polymorph. The extremely different strength of the hydrogen bonds of isostructural α-AlO(OH) (vOH = 2950 cm−1, 295 K), α-MnO(OH) (vOH = 2686 cm−1), and α-FeO(OH) (vOH = 3130 cm−1) is caused by the different synergetic effect of the metal ions involved, especially that of Mn3+ due to its Jahn-Teller behaviour. The decomposition temperatures and heats of the various FeO(OH) modifications as well as the halfwidths of the DSC peaks evidence a much faster decomposition rate of akaganéite than those of the other polymorphs. This is obviously due to the Cl ion impurities present in this compound.  相似文献   

8.
Monomeric and Dimeric Chromium(III) Phthalocyanines: Synthesis and Properties of Hydroxopyridinophthalocyaninatochromium(III) and μ-Oxodi(pyridinophthalocyaninatochromium(III)) Heating of ?[Cr(OH)Pc2?]”? in pyridine (Py) gives the paramagnetic (T = 273 K) complexes [Cr(OH)(Py)Pc2?] (μCr = 3.84 μB) and [(Cr(Py)Pc2?)2O] (μCr = 1.24 μB) by consecutive substitution and condensation reactions. The UV-VIS spectra are characterized by the typical B, Q, and N regions of the Pc2? ligand being shifted hypsochromically for the dimer with respect to the monomer due to excitonic coupling (1.5 kK). Regions of weak absorbance between 8 and 13 resp. 19 kK are assigned to trip-quartet transitions for both complexes. A weak band at 870 cm?1 in the FIR/MIR spectra is assigned to vas(Cr? O? Cr). In the resonance Raman(RR) spectra v(Cr? O) at 514 cm?1 resp. vs(Cr? O? Cr) at 426 cm?1 is selectively enhanced. Further strong RR-lines of the μ-Oxo dimer at 110 and 631 cm?1 are assigned to a (Py? Cr? O)- resp. internal pyridine deformation of a1g symmetry. An assignment as 2vas(Cr? O? Cr) is proposed for the remarkable RR line at 1740 cm?1.  相似文献   

9.
The melting point, T f of water in a pore decreases as the surface area to pore volume ratio of the pore decreases. Analysis of water absorbed in the pores of silica gels using differential scanning calorimetry (DSC) and dielectric relaxation spectroscopy (DRS) shows that the thickness of the bound, non-freezing water layer adjacent to the pore surface increases as its temperature increases, but that it is independent of the surface silanol concentration, [SisOH]. In contrast, the thickness decreases as the cylindrical pore radius r H decreases. Thus, the increase in the bound water thickness from 0.45 nm for gels with r H =1.2 nm to 1.2 nm for gels with r H =7.5 nm is due to the increase from –53°C to –7°C of the temperature (e.g., the melting point T f ) at which the bound water thickness was measured, and not due to the increase in t H or the decrease in [SisOH]. The T f of bulk water measured in a DSC was –0.3°C. The boiling point T v of bulk water measured in a DSC was 81.3°C. T v increased to 94°C in 7.5 nm pores and to 109°C K in 1.2 nm pores.  相似文献   

10.
NiH3IO6 · 6 H2O — Crystal Structures and Vibrational Spectra The crystal structure of NiH3IO6 · 6 H2O has been determined by X-ray single-crystal diffraction (Pc, Z = 2, a = 516.74(9), b = 981.5(2), c = 1052.5(2) pm, β = 116.496(8)°) on the basis of 4169 unique reflections (R = 1.96%). The structure is built up of distorted Ni(H2O)62+ and H3IO62? octahedra linked by hydrogen bonding. IR and Raman spectra of both the title compound and isostructural MgH3IO6 · 6 H2O as well as of deuterated specimens are given. There are up to 14 different OH(OD) modes in the spectra of isotopically dilute samples due to the 15 different hydrogen positions of the structure. The internal modes of the meridional H3IO62? ions (pseudo C2v symmetry) are discussed with respect to that double T-shaped entity, which gives rise to only two instead of 3I? O, I? O(H), and OH stretches in the IR and Raman spectra, i.e. the same as for facial (C3v) structured ions.  相似文献   

11.
The intermolecular interactions of the ?non-planar deformation modes”? in arbonates The v2 coupling constants of some carbonates of the aragonite structure, and of structures with similar configuration of the nearest anion neighbours, are investigated by means of 14C isotopic substitution. A relationship is found between the coupling constants and the fifth power of the anion distances. This relation shows that short-range repulsion forces and not dipole-dipole interaction governs the v2 coupling of the carbonates investigated in these studies.  相似文献   

12.
Ab initio calculations are reported for dimerization-induced changes, Δk, in the harmonic force constant k of the H-bonded OH in water dimer. Two dimer geometries are considered. Δk is obtained by considering the perturbation of a given monomer OH potential by the interaction energy in the dimer in question. The interaction energy is partitioned to identify the role of the various contributions to Δk. The sensitivity of Δk to the choice of the one-electron basis set is studied by using five different basis sets, some of which have a set of bond functions in the H? O bond. At the correlated level, correction for basis set superposition error is found to be essential. A comparison is made of the correlation contribution to Δk as given by the CEPA1, MP2, MP3, and MP4 methods. Of these, MP2 gives exaggerated results. Nevertheless, for economical and reasonably accurate calculations on large systems the MP2 approach in the ESPB basis set is advocated. The most accurate calculations yield a shift Δv0-;1 of – 121 cm?1 for the uncoupled donor O-H vibrational frequency in water dimer.  相似文献   

13.
The integrated intensity change by H-bonds are measured for CH3OH solved in different solvents of fundamental, 1. and 2. overtone OH stretching bands. A function A=f(ν) for the strong intensity change by H-bonds of the fundamental band is given, it shows a kink between pure van der Waals solvents and H-bond acceptors. - The contrary behavior of fundamental and 1. overtone bands for the T-dependence of pure CH3OH and its LiClO4-solutions could be canceled if the fundamental spectra are intensity corrected by A=f(ν). It is shown that the discussions between species and continuum models of water could become unique taking into account the function f(ν) and its kink, different for fundamental and overtone bands.  相似文献   

14.
The generation of metastable O2(1Σg+) and O2(1Δg) in the H + O2 system of reactions was studied by the flow discharge chemiluminescence detection method. In addition to the O2(1Σg+) and O2(1Δg) emissions, strong OH(v = 2) → OH(v = 0), OH(v = 3) → OH(v = 1), HO2(2A000) → HO2(2A000), HO2(2A001) → HO2(2A000), and H O2(2A200) → HO2(2A000) emissions were detected in the H + O2 system. The rate constants for the quenching of O2(1Σg+) by H and H2 were determined to be (5.1 ± 1.4) × 10?13 and (7.1 ± 0.1) × 10?13 cm3 s?1, respectively. An upper limit for the branching ratio to produce O2(1Σg+) by the H + HO2 reaction was calculated to be 2.1%. The contributions from other reactions producing singlet oxygen were investigated.  相似文献   

15.
By applying the projection operator method it is shown that the complicated Hamiltonian of a E ? ? JT system with nonlinear coupling coefficients can be written in terms of two Hamiltonians which are simple to handle and transform according to irreducible representations E, A1, and A2 of C3v point group. A variational approach is then used to calculate the ground state energy, using the Hamiltonian that transforms according to E, as an explicit function of the linear and nonlinear coupling parameters. The energies calculated in the strong coupling limit are finally compared with the corresponding previously calculated energies.  相似文献   

16.
Vibrational Spectra and Force Constants of Symmetric Tops. XXI. Rovibrational Spectrum of CF379Br and CF381Br in the v2 Region The rovibrational spectrum of CF379Br and of natural CF3Br has been recorded in the region of the fundamental v2 near 750 cm?1 in the i.r. with a resolution of 0.04 cm?1. From the analysis of the spectra the values of v0, the anharmonicity constants x23, x25 and x26 as well as B″ and B′ from the J structure have been obtained both for CF379Br and CF381Br. v2 displays Fermi resonance with 2v3 which is located near 700 cm?1.  相似文献   

17.
Differences in the system constants of the solvation parameter model, discontinuities in retention factor plots (log k against volume fraction of organic solvent) and retention factor correlation plots are used to study the retention mechanism on XTerra MS C18, XBridge C18 and XBridge Shield RP18 stationary phases with acetonitrile–water and methanol–water mobile phases containing from 10 to 70% (v/v) organic solvent. Wetting of XBridge C18 at 10 and 20% (v/v) acetonitrile is incomplete and is responsible for small changes in the retention mechanism. The intermolecular interactions responsible for retention on XTerra MS C18 and XBridge C18 are similar with minor differences in cavity formation and hydrogen-bonding interactions responsible for small selectivity differences. On the other hand, for bulky solutes there are large changes in retention at low volume fractions of organic solvent (<40% v/v) associated with steric repulsion on the XTerra MS C18 stationary phases that is absent for XBridge C18. Selectivity differences are more apparent for XBridge C18 and XBridge Shield RP18. For acetonitrile-water mobile phases cavity formation in the solvated XBridge Shield RP18 is slightly more difficult and hydrogen-bond acid and base interactions are more important than for XBridge C18. With methanol–water mobile phases it becomes slightly easier to form a cavity in the solvated XBridge RP18 compared with XBridge C18. In addition, the methanol-water solvated XBridge RP18 is a stronger hydrogen-bond base and more dipolar/polarizable than XBridge C18. These variations in selectivity justify the use of XBridge C18 and XBridge Shield RP18 as complementary stationary phases for method development.  相似文献   

18.
Spindly Mg5(CO3)4(OH)2·4H2O nanorods with porous and hollow nanostructures were synthesized in a water-in-oil reverse microemulsion system of cetyltrimethylammonium bromide (CTAB)/water/cyclohexane/n-pentanol. These nanorods were characterized by a number of physical techniques, including powder X-ray diffraction (XRD) and transmission electron microscopy (TEM). Comparison experiments showed that fundamental experimental parameters, such as the molar ratio of H2O to CTAB and the concentration of reactants, were important in the morphological control of Mg5(CO3)4(OH)2·4H2O nanostructures. The potential applications of the as-synthesized Mg5(CO3)4(OH)2·4H2O nanorods in water treatment were also investigated.  相似文献   

19.
李权  蔡静  陈俊蓉  赵可清 《中国化学》2008,26(2):255-259
使用密度泛函理论B3LYP方法和6-311++G**基函数对4-羟甲基吡啶与水形成的1:1和1:2(摩尔比)氢键复合物进行了理论计算研究,分别得到稳定的4-羟甲基吡啶-H2O和4-羟甲基吡啶-(H2O)2氢键复合物3个和8个。经基组重叠误差和零点振动能校正后,最稳定的1:1和1:2氢键复合物的相互作用能分别为-20.536和-44.246 kJ/mol。振动分析显示O-H···N(O)氢键的形成使复合物中O-H键对称伸缩振动频率红移(减小)。自然键轨道分析表明,4-羟甲基吡啶与水形成最稳定的1:1和1:2氢键复合物时,分子间电荷转移分别为0.02642 e 和0.03813 e 。含时密度泛函理论TD-B3LYP/ 6-311++G**计算显示,相对于4-羟甲基吡啶单体分子,氢键H-OH···N和H-OH···OH的形成分别使最大吸收光谱波长兰移8~16纳米和红移4~11纳米。  相似文献   

20.
Styrene microemulsion polymerizations with different short‐chain alcohols [n‐CiH2i+1OH (CiOH), where i = 4, 5, or 6] as the cosurfactant were investigated. Sodium dodecyl sulfate and sodium persulfate (SPS) were used as the surfactant and initiator, respectively. The desorption of free radicals out of latex particles played an important role in the polymerization kinetics. An Arrhenius expression for the radical desorption rate coefficient was obtained from the polymerizations at temperatures of 50–70 °C. The polymerization kinetics were not very sensitive to the alkyl chain length of alcohols compared with the temperature effect. The maximal polymerization rate in decreasing order was C6OH > C4OH > C5OH. This was related to the differences in the water solubility of CiOH and the structure of the oil–water interface. The feasibility of using a water‐insoluble dye to study the particle nucleation mechanisms was also evaluated. The parameters chosen for the study of the particle nucleation mechanisms include the cosurfactant type (CiOH), the SPS concentration, and the initiator type (oil‐soluble 2,2′‐azobisisobutyronitrile versus water‐soluble SPS). © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 39: 3199–3210, 2001  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