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 156 毫秒
1.
The vibrational characteristics of deuterated acetonitrile dissolved in isopropanol, dimethyl formamide (DMF), and dimethyl sulfoxide (DMSO) have been studied. Observed vibrational bands show substantial frequency shifts, the amounts of which vary almost linearly with concentration. The absorption feature in the region of 2220–2280 cm−1 was deconvoluted to the consisting absorption bands. The band at 2258 cm−1 of pure CD3CN, which is on the low frequency side of the monomer CN stretch (ν2), is attributed to the CN stretch of the dimer (ν′2). The shoulder found on the further low frequency side of the ν2 band, particularly in dilute solution, is believed to be due to ν5, and its frequency and intensity vary largely as a function of concentration along with those of other vibrational bands involved with the CD3 group. The ν5 band of pure CD3CN is believed to be active and located at about 2251 cm−1. Ab initio calculations have also been performed for the solute–solvent complexes, CD3CN–DMF and CD3CN–DMSO, at the MP2/6-31+G(2d,p) level assuming anti-parallel configurations. The calculated results show a good agreement with the observed results.  相似文献   

2.
Exposure of strong Lewis (coordinatively unsaturated metal atoms) and Bronsted (proton donor OH-groups) acid sites on solid surfaces is a prime demand for potential adsorptive and catalytic applications. In situ FTIR spectroscopy of small adsorbed base molecules, often NH3, pyridine, CH3CN, NO or CO molecules, has been well established as a powerful surface analytical technique for characterization of nature, strength and concentration of acid sites. Pyridine (Py) has been preferred as an IR probe molecule of finely divided metal oxide surfaces at room (RT) and higher temperature regimes, since it is (i) more selective and stable than NH3; (ii) much more strongly adsorbed than CO and CH3CN; and (iii) relatively more sensitive to the strength of Lewis acid sites than NO. In the present work, in situ IR spectra of Py adsorbed at ≥RT on characterized alumina, silica, silica–alumina, titania, zirconia and ceria were measured, and compared with RT-spectra of liquid and gas phase Py obtained under identical spectroscopic conditions, in order to characterize spectral consequences of mutual Py–Py interactions in the adsorbed phase. It has been concluded that the availability of Lewis acid sites can be unequivocally monitored by formation of coordinated Py molecules giving rise to IR-absorption(s) due to the ν8a mode of νCCN vibrations at 1630–1600 cm−1, where the higher the frequency assumed, the stronger the acidity of the site. Formation of pyridinium surface species (PyH+) is identifiable by (i) an ν8a-absorption at ≥1630 cm−1; (ii) an ν19b-absorption at 1550–1530 cm−1; as well as (iii) νN+---H and δN+---H absorptions occurring, respectively, near 2450 and 1580 cm−1, and, thus, the availability of Bronsted acid sites. Moreover, products and IR-characteristics of Py surface reactions at >RT have been identified, and used to imply nature of surface base sites (OHand O2−) involved in formation of acid–base site pairs.  相似文献   

3.
FTIR spectra of the four isotopically substituted 1:1 complexes of acetonitrile and boron trifluoride were recorded in Ar, N2 and Xe matrices. In Ar, previously unreported three vibrational modes of the complex were clearly observed. Several significant vibrational bands were also observed in N2 and Xe. The observed frequency shifts on complexation, Δν, were qualitatively in good agreement with the computational results, which were calculated at the B3LYP/6-311++G(d,p) level using the optimized geometry of the C3v eclipsed conformation. The observed magnitudes of Δν for most of the complex bands were larger than the calculated values. The BF3 symmetric deformation mode is an exception. The observed frequency shits for this mode were smaller than the calculated values, especially in N2. This suggests that even an inert matrix can significantly affect the vibrational spectrum of the complex.  相似文献   

4.
Poly(β-phenylpropyl -aspartate), poly(β-phenylbutyl -aspartate), and poly(β-phenylpentyl -aspartate) exhibit a reversible transformation from a right-handed -helix (R) to a left-handed ω-helix (ωL) in the solid state. During this transition, the infrared (IR) dichroism of the side-chain ester group and the birefringence change drastically, showing that the side-chain conformations are different for these two helices. In the present study, for the purpose of elucidating the preferred side-chain conformation in each helix, we performed the computational analyses. The energy contours, the directions of the IR transition moments and the anisotropies in polarizability as functions of the first two dihedral angles of the side chain, χ1 and χ2 were calculated. Then, comparing them with the experimental IR dichroism and birefringence data, we elucidated the specific side-chain conformation preferred for each R or ωL skeleton. The preferred values of (χ1, χ2) were found to be (−75, −60°) for R and (180, 45°) for ωL.  相似文献   

5.
The vibrational spectra of σ-(C3H5)Mn(CO)5 are reported. Assignment of bands is made and carbonyl force constants are calculated. The results indicate that the Mn(CO)5 moiety has C symmetry. The calculated angle between the axial and equatorial carbonyl groups is approximately 95°. The bonding in this compound is very similar to that in (CH3)Mn(CO)5.

In the far-infrared region, seven bands are expected in C symmetry (3A1 + 4E), and all are observed.  相似文献   


6.
This work uses cavity ring-down spectroscopy to measure two high vibrational overtones (4ν9 and 5ν9) of the O–H stretch in sulfuric acid. The frequencies, bandwidths, and intensities are obtained for these previously unobserved transitions. The atmospheric J-values for the overtone-induced photodissociation are calculated using the experimental cross-sections. Accurate J-values are essential for understanding the formation of the springtime polar sulfate layer by overtone-induced dehydration of H2SO4. The results are compared to previous experimental and theoretical studies of sulfuric acid.  相似文献   

7.
The molecular structures of n-hexane were determined by RHF/4-21G ab initio geometry optimization at 30° grid points in its three-dimensional τ1(C11–C8–C5–C1), τ2(C14–C11–C8–C5), τ3(C17–C14–C11–C8) conformational space. Of the resulting 12×12×12=1728 grid structures, 468 are symmetrically non-equivalent and were optimized constraining the torsions τ1, τ2, and τ3 to the respective grid points, while all other structural parameters were relaxed without any constraints. From the results, complete parameter surfaces were constructed using natural cubic spline functions, which make it possible to calculate parameter gradients, |P|=[(∂P/∂τi)2+(∂P/∂τj)2]1/2, where P is a C–C bond length or C–C–C angle. The parameter gradients provide an effective measure of the torsional sensitivity of the system and indicate that dynamic activities in one part of the molecule can significantly affect the density of states, and thus the contributions to vibrational entropy, in another part. This opens the possibility of dynamic entropic conformational steering in complex molecules; i.e. the generation of free energy contributions from dynamic effects of one part of a molecule on another. When the conformational trends in the calculated C–C bond lengths and C–C–C angles are compared with average parameters taken from some 900 crystallographic structures containing n-hexyl fragments or longer C–C bond sequences, some correlation between calculated and experimental trends in angles is found, in contrast to the bond lengths for which the two sets of data are in complete disagreement. The results confirm experiences often made in crystallography. That is, effects of temperature, crystal structure and packing, and molecular volume effects are manifested more clearly in bond lengths than bond angles which depend mainly on intramolecular properties. Frequency analyses of the τ1, τ2 and τ3 torsional angles in the crystal structures show conformational steering in the sense that, if τ1 is trans peri-planar (170°≤τ1≤180°; −180°≤τ1≤−170°), the values of τ2 and τ3 are clustered closely around the ideal gauche (±60°) and trans (±180°) positions. In contrast, when τ1 is in the region (50°≤τ1≤70°), there is a definite increase in the populations of τ2 and τ3 at −90 and −150°.  相似文献   

8.
Basis set dependence of the vibrational wavenumbers of out-of-plane modes calculated at the MP2 level of ab initio molecular orbital theory is examined for benzene, p-benzoquinodimethane, p-benzoquinone, furan, and thiophene. Various polarization functions up to (3df,p) are used in combination with the 6-31G and 6-311G basis sets. It is shown that, especially in the case of normal modes with alternate out-of-plane motions of the carbon atoms (such as the ν4 (b2g) mode of benzene), the calculated wavenumbers depend strongly on the exponents (d) of the d functions on the carbon atoms. It is therefore necessary to include d functions with an optimum exponent (d0.4) on the carbon atoms to obtain reasonable out-of-plane vibrational force fields. In a few cases (such as the ν16 (a2) mode of furan), inclusion of a set of f functions on the carbon atoms has some effects on the calculated wavenumbers of out-of-plane modes. However, unless the basis set contains an optimum set of d functions, inclusion of a set of f functions does not improve the agreement between the observed and calculated vibrational wavenumbers. As a case with an exaggerated effect of basis set, it is shown that the wavenumber of the ν41 (b2g) mode of the planar optimized structure of p-benzoquinodimethane is calculated to be imaginary by using the 6-311G(d,p) or 6-311G(df,p) basis set at the MP2 level. For all the molecules treated in the present study, reasonable out-of-plane vibrational force fields are obtained by using the 6-31G(2df,p) and 6-311G(2df,p) basis sets.  相似文献   

9.
The thermodynamic interactions in aqueous solutions of uncharged polymers were studied. Using a gel-deswelling method, the water activities (chemical potentials) in binary and ternary (two polymers in one solvent) solutions of methylcellulose (MC), polyvinyl alcohol (PVA) and polyvinyl pirrolidone (PVP), respectively were determined at various polymer volume fractions (1.0 × 10−2 < v2 < 1.0 × 10−1). On the theoretical basis of the Flory–Huggins approximations, the relevant solvent–segment (χ12 or χ13) and segment–segment pair interaction parameters (χ23) have been calculated.

The solvent activity curves (ln a1 versus polymer volume fraction) can be well described by a polynomial of third-degree in both the binary and the ternary solutions of the polymers. The solvent–segment interaction parameters exhibit a slight dependence on the polymer concentration. For each binary solution, the χ12v2 function can be fitted by a straight line wich has a small positive slope. In the mixtures of two polymers, the values of the segment–segment (χ23) interaction parameters were close to zero or sligthly negative (χ23 0 ± 0.03), indicating that under the studied conditions, the polymers in the ternary solutions are compatible.  相似文献   


10.
The paper presents a new method for predicting the frequency of the b1 mode, which is infrared-inactive, in complexes of the type LM(CO)5 belonging to C4V point group. The method was based on the relation λ3=λ4+[(1−δ/δ)](λ1λ2), where δ=(λ1λ2)/(λ1λ2+λ3λ4), λ1, λ2, λ3 and λ4 are the λ parameters of the , , b1 and e modes, respectively. For a large numbers of complexes of the type LM(CO)5 the average value of δ was found to be 0.80, with a standard deviation of 0.02. With the use of average value of δ, the frequencies of b1 mode were estimated. The result obtained indicated that there exists a rather good fit between observed and calculated frequencies, with a mean error of 2.7 cm−1. In addition, it was shown that the δ parameter can be used as a criterion of the correct band assignment for the complexes understudy.  相似文献   

11.
2,3,4-triphenyl-1-oxa-4-azabutadine (C20H15NO) has been studied by X-ray analysis and AM1 molecular orbital methods. It crystallises in the triclinic space group P-1 with a=9.414(3), b=10.479(3), c=8.385(2) Å, =103.31(3)°, β=97.10(3)°, γ=74.09(1)°, V=772.5(4) Å3, Z=2, Dc=1.227 gcm−3, and μ(MoK)=0.075 mm−1 and F000=300. The structure was solved by direct methods and refined to R=0.043 for 2672 reflections [I>2σ(I)]. The conformational analysis of the title compound were investigated by semi-empirical quantum mechanical AM1 calculations. The minimum conformation energies were calculated as a function of the three torsion angles θ1(O(1)C(7)C(8)N(1)), θ2(C(8)N(1)C(15)C(16)) and θ3(C(14)C(9)C(8)N(1)). The results are compared with the X-ray results. C=O and C=N groups are twisted about each other by 95.5(2)°.  相似文献   

12.
The SCF-RPA scheme is applied to the n → π* absorption band of cyclopentanone in its (C2) ground-state nuclear configuration. Using the Cederbaum—Domcke algorithm, the effects of the (ν3) C=O stretching, (ν18) ring puckering and (ν25) C=O out- of-plane vibrations are investigated. The calculated spectrum shows both allowed and vibronic components, in agreement with the experimental observations of Howard-Lock and King.  相似文献   

13.
合成了一个新的非线性光学(NLO)有机材料1-二茂铁基-3-[(9-乙基)咔唑-3-基]丙烯酮(FCAK),并通过NMR、IR、MS和元素分析等技术手段进行了表征。 采用粉末Nd∶YAG激光技术测定了标题化合物的三阶非线性光学性质并确定了相关参数。 激光脉冲为4 ns时,非线性折射率n2=-3.5×10-18 m2/W,非线性吸收系数β=-2.7×10-11 m/W,三阶非线性极化率χ(3)=2.04×10-12 esu,三阶非线性分子超极化率γ=1.1×10-30 esu。 激光脉冲为21 ps时,n2=0.55×10-18 m2/W,β=-0.6×10-11 m/W,χ(3)=3.4×10-13 esu,γ=0.13×10-30 esu。  相似文献   

14.
Two carbon-rich starburst gold(I) acetylide complexes [TEE][Au(PCy3)]4 (3, [TEE]H4=tetraethynylethene) and [TEB][Au(PCy3)]3 (6, [TEB]H3=1,3,5-triethynylbenzene) were prepared and their UV–vis absorption, emission and excitation spectra have been recorded. In fluid CH2Cl2 solutions, 3 exhibits prompt 1(ππ*) fluorescence with λ0–0 and λmax at 413 and 428 nm, respectively, while 6 displays 3(ππ*) phosphorescence with λ0–0 and λmax at 446 and 479 nm, respectively. The crystal structure of 3·CH2Cl2 has been determined.  相似文献   

15.
Polarised IR and Raman spectra for KH2PO3 single crystal samples were measured at room temperature. Additionally, the IR spectra for the Xb(Z) sample were also measured at low temperatures (300–14 K). The spectra are discussed on the basis of oriented gas model and group theory. The stretching νOH vibrations of the hydrogen bonds with the OO distances of 2.547 and 2.529 Å give characteristic broad ABC-type bands in the IR (polarised parallel to the X and to the b(Z) directions) and Raman (xx, xz and yx) spectra. The Davydov-type (correlation field or factor group) splitting is not observed for the νOH modes. The presence of two independent hydrogen bonds in the crystal is manifested by splitting of the C band into two (C′, C″) components and by the different frequencies of the out-of-plane bending γOH vibrations. The in-plane bending modes δOH are strongly mixed/coupled with the stretching vibrations of the PO3 groups.

The C bands (C′ and C″) change into quite sharp bands on lowering of the temperature. Various simplified models for internal vibrations of the phosphite anions are applied for finding a correlation between the crystal structure and polarised vibrational spectra. The stretching vibrations of the νPH groups manifest their unequivalence in two symmetry-independent hydrogenphosphite anions.  相似文献   


16.
The tridecameric aluminum polymer [AlO4Al12(OH)24(H2O)12]7+ was prepared by forced hydrolysis of Al3+ up to an OH/Al molar ratio of 2.2. Upon addition of sulfate, the tridecamer crystallized as the monoclinic basic aluminum sulfate Na0.1[AlO4Al12(OH)24(H2O)12](SO4)3.55. The dehydroxylation of the basic aluminum sulfate has been studied by Fourier transform in-situ infrared emission spectroscopy over a temperature range of 200° to 750°C at 50°C intervals. The spectrum is characterized by the sulfate ν1 (1024 cm−1), ν3 doublet (1117 and 1168 cm−1) and the ν4 doublet (568 and 611 cm−1) modes. Furthermore, minor bands assigned to nitrate are observed. Upon heating from ≈350° to 400°C major changes are observed, especially in the bandwidth and band intensities. The bands in the hydroxyl stretching region due to the Al13 group disappear, whereas the bands around 1050 cm−1 display various changes in bandwidths, intensities and positions associated with the dehydration and dehydroxylation of the basic sulfate and the changing of the structure into an aluminum oxosulfate. The nitrate bands diminish upon heating.  相似文献   

17.
Whipped foams (10%, w/v protein, pH 7.0) were prepared from commercially available samples of whey protein isolate (WPI) and egg white protein (EWP), and subsequently compared based on yield stress (τ0), overrun and drainage stability. Adsorption rates and interfacial rheological measurements at a model air/water interface were quantified via pendant drop tensiometry to better understand foaming differences among the ingredients. The highest τ0 and resistance to drainage were observed for standard EWP, followed by EWP with added 0.1% (w/w) sodium lauryl sulfate, and then WPI. Addition of 25% (w/w) sucrose increased τ0 and drainage resistance of the EWP-based ingredients, whereas it decreased τ0 of WPI foams and minimally affected their drainage rates. These differing sugar effects were reflected in the interfacial rheological measurements, as sucrose addition increased the dilatational elasticity for both EWP-based ingredients, while decreasing this parameter for WPI. Previously observed relationships between τ0 and interfacial rheology did not hold across the protein types; however, these measurements did effectively differentiate foaming behaviors within EWP-based ingredients and within WPI. Interfacial data was also collected for purified β-lactoglobulin (β-lg) and ovalbumin, the primary proteins of WPI and EWP, respectively. The addition of 25% (w/w) sucrose increased the dilatational elasticity for adsorbed layers of β-lg, while minimally affecting the interfacial rheology of adsorbed ovalbumin, in contrast to the response of WPI and EWP ingredients. These experiments underscore the importance of utilizing the same materials for interfacial measurements as used for foaming experiments, if one is to properly infer interfacial information/mechanisms and relate this information to bulk foaming measurements. The effects of protein concentration and measurement time on interfacial rheology were also considered as they relate to bulk foam properties. This data should be of practical assistance to those designing aerated food products, as it has not been previously reported that sucrose addition improves the foaming characteristics of EWP-based ingredients while negatively affecting the foaming behavior of WPI, as these types of protein isolates are common to the food industry.  相似文献   

18.
二茂铁是合成新颖有机功能材料的基本单元之一。 本文设计并合成了两个基于二茂铁的同分异构查尔酮衍生物:1-二茂铁基-3-(噻吩-2-基)丙烯酮(a)和1-二茂铁基-3-(噻吩-3-基)丙烯酮(b)。 采用超快激光Z-扫描技术(脉宽180 fs,波长532 nm)测定了化合物a和b的三阶非线性光学性质。 结果表明,化合物a吸收系数β=-2.1×10-12 m/W,折射率n2=1.9×10-19 m2/W,分子超极化率γ=5.37×10-32 esu;化合物b:β=-1.2×10-13 m/W,n2=2.0×10-19 m2/W,γ=4.48×10-32 esu。 说明在飞秒激光激发下,电荷转移能够在化合物a和b分子内部快速进行,二者均具有优异的超快三阶非线性光学响应。 在B3LYP/6-311+G(d,p)理论水平下,计算了化合物a和b分子轨道能量、极化率和各基团在前线分子轨道中的占有率。 理论计算结果显示,二茂铁基团在化合物a和b前线分子轨道中占有率分别为97%和98%,对两化合物的非线性光学性能起主导作用。  相似文献   

19.
孙金鱼  石玉芳  王桂林  赵明根 《应用化学》2015,32(10):1134-1138
合成了一种新的具有潜在应用价值的非线性光学(NLO)有机材料1-(芘-1-基)-3-(4-二甲氨基苯基)丙烯酮(PMAK),并通过 NMR、IR、MS和元素分析等技术手段进行了表征。 采用溶液Nd:YAG激光技术测定了PMAK的三阶非线性光学性质并确定了相关参数。 纳秒实验结果:折射率n2=-3.5×10-17 m2/W,吸收系数β=7.0×10-10 m/W,极化率χ(3)=2.54×10-11 esu,分子超极化率γ=3.44×10-30 esu;皮秒实验结果:n2=-2.8×10-18 m2/W,β=8.3×10-11 m/W,χ(3)=2.49×10-12 esu,γ=3.33×10-31 esu。  相似文献   

20.
The crystal structure of N-(2-hydroxy-5-chlorophenyl) salicylaldimine (C13H10NO2Cl) was determined by X-ray analysis. It crystallizes orthorhombic space group P212121 with a=12.967(2) Å, b=14.438(3) Å, c=6.231(3) Å, V=1166.5(6) Å3, Z=4, Dc=1.41 g cm−3 and μ(MoK)=0.315 mm−1. The title compound is thermochromic and the molecule is nearly planar. Both tautomeric forms (keto and enol forms in 68(3) and 32(3)%, respectively) are present in the solid state. The molecules contain strong intramolecular hydrogen bonds, N1–H1O1/O2 (2.515(1) and 2.581(2) Å) for the keto form and O1–H01N1 for the enol one. There is also strong intermolecular O2–HO1 hydrogen bonding (2.599(2) Å) between neighbouring molecules. Minimum energy conformations AM1 were calculated as a function of the three torsion angles, θ1(N1–C7–C6–C5), θ2(C8–N1–C7–C6) and θ3(C9–C8–N1–C7), varied every 10°. Although the molecule is nearly planar, the AM1 optimized geometry of the title compound is not planar. The non-planar conformation of the title compound corresponding to the optimized X-ray structure is the most stable conformation in all calculations.  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