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 984 毫秒
1.
《Vibrational Spectroscopy》2009,49(2):259-262
In order to evidence the structural changes induced by CuO and V2O5 in the phosphate glass network and their modifier or former role, x(CuO·V2O5)(100  x)[P2O5·CaO] glass system was prepared and investigated using Raman spectroscopy (0  x  40 mol%).Raman spectra of the studied glasses present the specific bands of the phosphate glasses at low concentration of transition metal (TM) ions, but at higher concentration (x > 7 mol%) a strong depolymerization of the phosphate network appears; non-bridging oxygen atoms are involved in VOP and CuOP bonds and new short units are formed. For a high concentration of V2O5 (x > 10 mol%) the Raman bands of V2O5 prevail in the spectra; this fact suggests that vanadium oxide imposes its structural units in the network acting thus as a network glass former.2D correlation analysis was also applied for the concentration-dependent Raman spectra in order to verify the assignments of the vibration modes and to find correlations in the changes induced by TM ions content. 2D correlation maps indicate a good correlation between the bands at ∼705 cm−1 assigned to POP stretching vibration and at ∼1175 cm−1 assigned to PO2 groups which suggest the depolymerization of the phosphate network. The correlation between the 1270 cm−1 and 930 cm−1 bands also suggests that V2O5 oxide is responsible for PO bonds breaking and POV formation.  相似文献   

2.
KOH activation of petroleum coke (PC) was conducted with 30 vol%H2 + 70 vol% N2 as carrier gas. TG-DTG, FTIR, elemental analysis, N2 adsorption, GC and XRD techniques were used to investigate the effects of hydrogen on the activation. During the initial stage of the activation, i.e. the carbonization of the PC, additional CH and CH2 species were formed due to the chemisorption of hydrogen on the nascent sites of the PC created by the removal of the surface heteroatom groups. The formation of the CH and CH2 species increased the quantity of ‘active sites’ which is favorable to the further activation reaction, and developed the porous structure of the activated carbons. The micropore volume and BET surface areas of the activated carbon prepared under 30 vol% H2 + 70 vol% N2 and with a relatively low KOH/PC weight ratio of 2:1 have been increased from 0.78 cm3/g and 1936 m2/g to 0.97 cm3/g and 2477 m2/g, respectively, compared to that prepared in pure N2 atmosphere with the same KOH/PC ratio.  相似文献   

3.
Physical and thermal properties of polyoxyethylene glycol glycerides (Gelucire 50/13) used as sustained release matrix forming agent in pharmaceutical applications are studied by Raman spectroscopy combined with X-ray diffraction and differential scanning calorimetry methods.At first, Raman spectroscopy was used to characterize the polymorphs and liquid state of PEG 1500, with emphasis placed on the evolution of the Raman-active CC and CO stretching region (1300–1100 cm−1), along with complementary analysis of the Raman-active CH stretching modes (3000–2800 cm−1) in comparison with temperature. Unique Raman signatures were obtained for all phases, with their identity confirmed using DSC and XRD. The CC and CO stretching modes, which provided insight into the trans/gauche content, permitted polymorph discrimination due to differences in the number of modes, their relative frequencies and their full-widths at half-maximum. CH stretching generally increased with polymorph stability, indicating the dominance of methylene antisymmetric CH2 vibrations as the PEG 1500 crystal lattice became more ordered. The change in the intensities of the CH stretching bands was used to probe the order–disorder transition.The second time, Raman spectroscopy of Gelucire 50/13 was performed to characterize the contribution of its each component, with emphasis placed on the evolution of the t(CH2) and ν(CC) vibrational mode regions (1300–1200 cm−1), along with analysis of the Raman-active CH stretching modes (3000–2800 cm−1), δ(CH2) and δ(CH3) deformation region (1500–1400 cm−1), and ras(CH2) rocking region (900–800 cm−1). In comparison with temperature, the changes of the ratios of Is(CH2)]/Ias(CH2)] (I2850/I2890), Ias(CH2)]/Is(CH3)] (I2890/I2950), I[δ(CH2)]/I[δ(CH3)] (I1444/I1490), I1296/I1282 and I[ras(CH2)]/I[t(CH2)] (I845/I1282) were directly correlated with conformational changes of the Gelucire structure. Overall, Raman spectroscopy clearly demonstrated that the different functional groups studied could be characterized independently, allowing for the understanding of their role in Gelucire polymorphism.  相似文献   

4.
An overview is given on synthesis and structures of new bidentate phosphaalkene ligands [(RMe2Si)2CP]2E (E = O, NR, N?) and (RMe2Si)2CPN(R′)PR′′2. Exceptional properties of these ligands, extending beyond predictable properties of phosphaalkenes are: (i) the NSi bond cleavage of [(iPrMe2Si)2CP]2NSiMe3 with AuI and RhI chloro complexes under mild conditions leading to binuclear complexes of the 6π-delocalised imidobisphosphaalkene anion [(iPrMe2Si)2CP]2N?, and (ii) the chlorotropic formation of molecular 1:2 PdII and PtII metallochloroylid complexes with novel ylid-type ligands [(RMe2Si)2CP(Cl)N(R)PR2]?, and the transformation of a P-platina-P-chloroylid complex into a C-platina phosphaalkene by intramolecular chlorosilane elimination. Properties of the heavier congeners [(RMe2Si)2CP]2E (E = S, Se, Te, PR, P?, As?) and (RMe2Si)2CPEPR′′2 (E = S, Se, Te) are also described.  相似文献   

5.
The molecular structure of caffeine (3,7-dihydro-1,3,7-trimethyl-1H-purine-2,6-dione) was determined by means of gas electron diffraction. The nozzle temperature was 185 °C. The results of MP2 and B3LYP calculations with the 6-31G7 basis set were used as supporting information. These calculations predicted that caffeine has only one conformer and some of the methyl groups perform low frequency internal rotation. The electron diffraction data were analyzed on this basis. The determined structural parameters (rg and ∠α) of caffeine are as follows: <r(NC)ring> = 1.382(3) Å; r(CC) = 1.382(←) Å; r(CC) = 1.446(18) Å; r(CN) = 1.297(11) Å; <r(NCmethyl)> = 1.459(13) Å; <r(CO)> = 1.206(5) Å; <r(CH)> = 1.085(11) Å; ∠N1C2N3 = 116.5(11)°; ∠N3C4C5 = 121. 5(13)°; ∠C4C5C6 = 122.9(10)°; ∠C4C5N7 = 104.7(14)°; ∠N9–C4=C5 = 111.6(10)°; <∠NCHmethyl> = 108.5(28)°. Angle brackets denote average values; parenthesized values are the estimated limits of error (3σ) referring to the last significant digit; left arrow in parentheses means that this parameter is bound to the preceding one.  相似文献   

6.
Bis(betainium) p-toluenesulfonate monohydrate (abbreviated as BBTSH) was studied at various temperatures by X-ray diffraction, differential scanning calorimetry and vibrational spectroscopy methods. DSC curves of BBTSH show a peak at about 349 K which corresponds to water escape from the crystal, and reveal the “cold crystallization” phenomenon. BBTSH crystallizes in the P21/c space group of monoclinic system. After heating above 349 K the compound dehydrates, the crystal system changes to triclinic, the monocrystalline samples become non-merohedral twins. The BBTSH crystal comprises p-toluenesulfonic anions, monoprotonated betaine dimers and water molecules. Three kinds of hydrogen bonds are present in the crystal: strong, asymmetric and almost linear OH⋯O hydrogen bond (R(O⋯O) = 2.463(2) Å), weak OwH⋯O hydrogen bonds (R(Ow⋯O) = 2.820(2)  2.822(2) Å) and weak CH⋯O hydrogen bonds (R(C⋯O) = 3.295(2)  3.416(2) Å). The νaOHO vibration of the strongest hydrogen bond in the crystal gives rise to an intense broad absorption with numbers of transmission windows in the low wavenumber region of the infrared spectra. Coupling between νCO stretching vibrations of two COO groups of the betaine dimer was detected. The process corresponding to the loss of water is accompanied by the breakage of strong OH⋯O hydrogen bonds in betaine dimers and rearrangement inside half of the betaine dimers. This rearrangement results in formation of the new betaine dimers with OH∙∙∙O hydrogen bond of similar strength as corresponding bond in the hydrated form (BBTSH).  相似文献   

7.
《Vibrational Spectroscopy》2007,43(1):104-110
The Raman spectra of serine [α-amino-β-hydroxypropionic acid; HOCH2CH(NH3)+COO] and 3,3-dideutero-serine [HOCD2CH(NH3)+COO] in aqueous solution were studied in the range 4000–300 cm−1. The data obtained for the deuterated compound are novel and provide compelling evidence that previously reported assignments for the undeuterated amino acid should be revised.  相似文献   

8.
The structure of the complex of dimethylphenyl betaine (DMPB) with dichloroacetic acid (DCA) (1) has been investigated by X-ray diffraction, FTIR and Raman spectroscopy, and B3LYP/6-311 + + G(d,p) calculations. The crystal is monoclinic, space group P21. The acid is connected with betaine through the OH⋯O hydrogen bond of 2.480(2) Å. In the optimized structure the short, asymmetric O⋯O distance is 2.491 Å. FTIR spectrum shows a broad absorption in the 1500–400 cm−1 region characteristic of very short OH⋯O hydrogen bond caused by Fermi resonance between νOH and overtones of δOH and γOH. In the Raman spectrum this broad absorption is not observed. The potential energy distributions (PED) were used for the assignments of IR and Raman frequencies in the experimental and calculated spectra. The FTIR and Raman spectra of the crystal complex are consistent with the X-ray results.  相似文献   

9.
Continuous gradient temperature Raman spectroscopy (GTRS) applies the temperature gradients utilized in differential scanning calorimetry (DSC) to Raman spectroscopy, providing a straightforward technique to identify molecular rearrangements that occur near phase transitions. Herein we apply GTRS and DSC to the solid dipeptides Ala-Pro, Pro-Ala, and the mixture Ala-Pro/Pro-Ala 2:1. A simple change in residue order resulted in dramatic changes in thermal stability and properties. Characteristic Pro vibrations were observed at ∼75 °C higher temperature in Pro-Ala than Ala-Pro. The appearance/disappearance of characteristic vibrational modes with increasing temperature showed that a double peak in the Ala-Pro major phase transition (174–184 °C) was due to a gauche to anti 165° rotation of H3CC*NH3 about C*. CH2 rocking and wagging frequencies present in Pro-Ala were not observed in Ala-Pro. For Ala-Pro, the Ala +NH3, and Pro COO sites were flexible whereas the Pro ring moiety was not; since the OCN (C)2 amide bond is planar the CNC moiety keeps the Pro ring rigid. For Pro-Ala, CH2 sites in the Pro ring were flexible and the OCNH amide bond is perpendicular to the Pro ring. Since the mass of the Pro ring is significantly larger than the mass of the flexible Ala +NH3 moiety, Pro-Ala absorbs more thermal energy, corresponding to a higher phase transition temperature (240–260 °C). Ala-Pro, Pro-Ala, and Ala-Pro/Pro-Ala 2:1 exhibited α-helix, β-sheet, α-helix secondary structure conformations, respectively.  相似文献   

10.
The 1:1 complex of piperidine-4-carboxylic acid (isonipecotic acid, P4C) with 2,6-dichloro-4-nitrophenol (DCNP), has been investigated by single-crystal X-ray analysis, Raman and FTIR spectroscopy and theoretical calculations. The hydrogen-bonded-ion-pair complex is observed in the crystalline state with the O⋯H⋯OOC hydrogen bond of 2.453(16) Å. FTIR spectrum shows a broad absorption in the 1600–400 cm−1 region characteristic of very short OHO hydrogen bond, broken by the Evans holes. The complexes are joined through NH⋯O into a H-bonding network. The NH⋯O mode appears as a broad band in the range of 3100–2000 cm−1. In the structure optimized at the B3LYP/6–311 + +G(d,p) level of theory the proton is transferred from DCNP to P4C, and molecules are joined through the O⋯HOOC hydrogen bond of 2.640 Å. The experimental and theoretical infrared spectra are discussed. Detail interpretation of the vibrational spectra has been carried out with the use of computed Potential Energy Distribution (PED).  相似文献   

11.
In this paper, we report the gas phase infrared spectra of fluorene and its methylated derivatives using a heated multipass cell and argon as a carrier gas. The observed spectra in the 4000–400 cm−1 range have been fitted using the modified scaled quantum mechanical force field (SQMFF) calculation with the 6-311G** basis. The advantage of using the modified SQMFF method is that it scales the force constants to find the best fit to the observed spectral lines by minimizing the fitting error. In this way we are able to assign all the observed fundamental bands in the spectra. With consecutive methyl substitutions two sets of bands are found to shift in a systematic way. The set of four aromatic CH stretching vibrations around 3000 cm−1 shifts toward lower frequencies while the single most intense aromatic CH out-of-plane bending mode around 750 cm−1 shifts toward higher frequencies. The reason for shifting of aromatic CH stretching frequency toward lower wave numbers with gradual methyl substitution has been attributed to the lengthening of the CH bonds due to the +I effect of the methyl groups to the ring current as revealed from the calculations. While the unexpected shifting of the aromatic CH out-of-plane bend toward higher wave numbers with increasing methyl substitution is ascribed to the lowering of the number of adjacent aromatic CH bonds on the plane of the benzene ring with gradual methyl substitutions.  相似文献   

12.
Reaction of pentadienyl radicals (C5H7) with O2 has been studied by a combination of pulsed laser photolysis and photoionization mass spectrometry. These radicals could be generated either by the photolysis of 1,3-pentadiene or by the two-step reaction of carbon tetrachloride photolysis followed by the H-atom abstraction reaction of Cl atom with 1,4-pentadiene. The equilibrium between pentadienyl radicals, O2 and pentadienylperoxy radicals could be observed over the range 268–308 K. An analysis of the temporal signal of pentadienyl radicals was used to evaluate the equilibrium constant. Third-law analysis was used to evaluate the enthalpy change for the reaction C5H7 + O2 ⇌ C5H7O2. The observed CO bond energy in the C5H7O2 adduct was found to be 56.0 ± 2.2 kJ·mol–1, which is lower than the values of peroxy radicals formed with allyl and cyclohexenyl radicals which have an allylic resonance structure.  相似文献   

13.
Raman experiments of aluminum chloride and formamide (FA) solutions in different compositions and temperatures were carried out. Spectral changes provoked by the increase of the salt concentration were observed in different regions. The νCO and νCN modes of FA upon complexation were upshifted and suggest that the CONH hybrid (II) is stabilized by Al(III). Bands at 547 and 295 cm−1, which are assigned to the νAlO and νAlN vibrations, respectively, evidence coordination through both O and N atoms of FA. The quantitative analysis performed at the carbonyl stretching region found 5 FA molecules around this cation, resulting in the formation of the [Al(FA)5]Cl3 complex. Its stability is maintained by whole studied concentration range and up to around 100 °C. At higher temperatures, distortions in the FA shell begin occurring and a new component at 356 cm−1 is then observed and assigned to the [AlCl4] complex.  相似文献   

14.
《Vibrational Spectroscopy》2007,43(2):387-394
The metal ion distributions at the two metal sites (hexaformate-coordinated Me1 sites and mixed-coordinated Me2 sites) in the title mixed crystals as determined by single crystal X-ray diffraction and double matrix infrared spectroscopic methods are presented and discussed. The mixed formates are isostructural with the end compounds (space group P21/c). The local metal ion concentrations as a function of the total metal ion concentrations exhibit a clear preference of Zn2+ ions to Me1 sites and the Mg2+ ions to Me2 sites.The analysis of the infrared spectra reveals that the spectral regions 2300–2500 cm−1 (νOD of matrix-isolated HDO molecules) and 1300–1400 cm−1 (symmetric COO stretching (ν2) and bending CH (ν5) modes) are mostly sensitive to the metal ion environment. The inclusion of Mg2+ and Zn2+ in the structures of Zn(HCOO)2·2H2O and Mg(HCOO)2·2H2O, respectively, leads to an appearance of new infrared bands corresponding to νOD of HDO molecules bonded to the incorporated ions (i.e. new hydrogen bonding systems MgOH2⋯OCHOZn and ZnOH2⋯OCHOMg are formed in the mixed formates). The respective new bands are observed at small concentrations of included Mg2+ ions (about 5 mol%, x = 0.05) and at considerably higher concentrations of included Zn2+ ions (about 30 mol%, x = 0.7). Contrarily, the ν2 and ν5 modes caused by the incorporated cations bonded to formate ions occur at x  0.3 and x  0.85 (Mg2+ ions in Zn(HCOO)2·2H2O and Zn2+ ions in Mg(HCOO)2·2H2O, respectively). Thus, the infrared spectroscopy experiments confirm the single crystal X-ray measurements that the Mg2+ ions are localized predominantly at Me2 sites and the Zn2+ ions at Me1 sites in the title mixed crystals. The pronounced preference of the Mg2+ ions to Me2 sites is owing to the strong affinity of these ions to water molecules.  相似文献   

15.
FTIR spectra of propionic acid (PA), 2-propanol (PROH) and its binary mixtures with varying molefraction of the PA were recorded in the region 500–3500 cm?1, to investigate the formation of hydrogen bonded complexes in a mixed system. The observed features in ν(CO), ν(CO) and δ(COH) of PA, ν(CO) of PROH and δ(COH) of PA + PROH have been explained in terms of the hydrogen bonding interactions between PROH and PA and dipole–dipole interaction. The dipole moment derivative for the above mentioned vibrational modes have also been predicted from the integrated absorbance. The intrinsic linewidth for the vibrational modes ν(CO) and δ(COH) of PA has been elucidated using Bondarev and Mardaeva model.  相似文献   

16.
DFT calculations with B3LYP and PBE1PBE functionals and 6–311++G(d,p) basis set have been performed in order to obtain molecular geometries, binding energies and vibrational properties of the RCN?HF H-bonded complexes with R = NH2, CH3O, CH3, OH, SH, H, Cl, F, CF3, CN and NO2. As expected, it has been verified as a red-shift of the HF stretching frequency (νHF), in conformity with the elongation of the bond after complexation. On the other hand, the CN stretching frequency (νCN) is blue-shifted and corresponds to a shortening of the bond. The binding energies (ΔEc), including BSSE and ZPVE corrections, show a linear correlation with several structural, electronic and vibrational properties. In particular, an important linear dependence between the binding energy and the calculated dipole moment of the free RCN molecule (μRCN) has been found. This result suggests that μRCN can be a useful quantity in order to predict the ability of this fragment to form a hydrogen-bond. The IR intensities of stretching and bending modes of complexed HF acid fragment are adequately interpreted through the atomic polar tensor of the hydrogen atom in HF using the modified CCFO model for infrared intensities. The new vibrational modes arising from complexation show several interesting features.  相似文献   

17.
From thermal analyses and X-ray diffraction the phase diagram of the BiSnTe and SnTeBi2Te3 sections was determined. The local environment of Sn and Te atoms was studied by 119Sn and 125Te Mössbauer spectroscopy. The BiSnTe section showed a eutectic reaction at 267 °C and 20 % mole SnTe–80 % mole Bi. No intermediate compound was detected. The SnTeBi2Te3 section is characterized by a eutectic reaction at 585 °C and 40 % mole SnTe–60% mole Bi2Te3 and a peritectic reaction at 600 °C and 50 % mole SnTe–50% mole Bi2Te3. It corresponds to the compound SnBi2Te4, which has a rhombohedral layered structure with unit cell parameters a=4.3954(4) Å and c=41.606(1) Å. © 2000 Académie des sciences / Éditions scientifiques et médicales Elsevier SASSnTe / Bi / Bi2Te3 / phase diagram / Mössbauer  相似文献   

18.
The natural vermiculites from different localities (Bulgaria, Brazil, and South Africa) after acid treatment were used for this study. Differently acidified vermiculite samples were prepared from the natural vermiculite sample using different concentrations of hydrochloric acid (0.5 M and 1 M) and different reaction time (2 h and 4 h) at 80 °C. Natural vermiculites and acid treated vermiculites were analyzed by elemental analysis, X-ray diffraction (XRD) analysis and studied using Fourier transform infrared (FTIR) spectroscopy and dispersive Raman spectroscopy. According to the XRD analysis vermiculites are interstratified structures created in the different two-one-zero sheet hydrated phases. Ratio of intensities of spectrally deconvoluted bands at 1075 cm−1 and 1000 cm−1 (stretching vibration of SiO bonds of vermiculites and stretching vibration of SiO bonds of amorphous silica, respectively) was used to determine the content of amorphous silica in acid treated vermiculite samples. Study of the infrared and Raman spectra of the acidified vermiculites enable a comparison of these two spectroscopic data that have not yet been performed.  相似文献   

19.
The RuC bond of the bis(iminophosphorano)methandiide-based ruthenium(II) carbene complexes [Ru(η6-p-cymene)(κ2-C,N-C[P{NP(O)(OR)2}Ph2]2)] (R = Et (1), Ph (2)) undergoes a C–C coupling process with isocyanides to afford ketenimine derivatives [Ru(η6-p-cymene)(κ3-C,C,N-C(CNR′)[P{NP(O)(OR)2}Ph2]2)] (R = Et, R′ = Bz (3a), 2,6-C6H3Me2 (3b), Cy (3c); R = Ph, R′ = Bz (4a), 2,6-C6H3Me2 (4b), Cy (4c)). Compounds 34ac represent the first examples of ketenimine–ruthenium complexes reported to date. Protonation of 34a with HBF4 · Et2O takes place selectively at the ketenimine nitrogen atom yielding the cationic derivatives [Ru(η6-p-cymene)(κ3-C,C,N-C(CNHBz)[P{NP(O)(OR)2}Ph2]2)][BF4] (R = Et (5a), Ph (6a)).  相似文献   

20.
The reaction of RuTp(COD)Cl (1) with PR3 (PR3 = PPh2iPr, PiPr3, PPh3) and propargylic alcohols HCCCPh2OH, HCCCFc2OH (Fc = ferrocenyl), and HCCC(Ph)MeOH has been studied.In the case of PR3 = PPh2iPr, PiPr3 and HCCCPh2OH, the 3-hydroxyvinylidene complexes RuTp(PPh2iPr)(CCHC(Ph)2OH)Cl (2a) and RuTp(PiPr3)(CCHC(Ph2)OH)Cl (2b) were isolated.With PR3 = PPh2iPr and HCCCFc2OH as well as with PR3 = PPh3 and HCCCPh2OH dehydration takes place affording the allenylidene complexes RuTp(PPh2iPr)(CCCFc2)Cl (3b) and RuTp(PPh3)(CCCPh2)Cl (3c).Similarly, with PPh2iPr and HCCC(Ph)MeOH rapid elimination of water results in the formation of the vinylvinylidene complex RuTp(PPh2iPr)(CCHC(Ph)CH2)Cl (4).In contrast to the reactions of the RuTp(PR3)Cl fragment with propargylic alcohols, with HCC(CH2)nOH (n = 2, 3, 4, 5) six-, and seven-membered cyclic oxycarbene complexes RuTp(PR3)(C4H6O)Cl (5), RuTp(PR3)(C5H8O)Cl (6), and RuTp(PR3)(C6H10O)Cl (7) are obtained. On the other hand, with 1-ethynylcyclohexanol the vinylvinylidene complex RuTp(PPh2iPr)(CCHC6H9)Cl (8) is formed. The reaction of the allenylidene complexes 3ac with acid has been investigated. Addition of CF3COOH to a solution of 3ac resulted in the reversible formation of the novel RuTp vinylcarbyne complexes [RuTp(PPh2iPr)(C–CHCPh2)Cl]+ (9a), [RuTp(PPh2iPr)(C–CHCFc2)Cl]+ (9b), and [RuTp(PPh3)(C–CHCPh2)Cl]+ (9c). The structures of 3a, 3b, and 5b have been determined by X-ray crystallography.  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