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 640 毫秒
1.
This study reports the Raman (FT‐RS) and absorption infrared (FT‐IR) spectra, based on calculated wavenumbers and normal modes of vibrations, of the following compounds: L ‐Ala‐L ‐NH‐CH(Me)‐PO3H2 (alafosfalin, A1), L ‐Ala‐D ‐NH‐CH(Me)‐PO3H2 (A2), L ‐Ala‐L ‐NH‐CH(Et)‐PO3H2 (A3), D ,L ‐Ala‐D,L ‐NH‐CH(Et)‐PO3H2 (A4), L ‐Ala‐D ‐NH‐CH(iPr)‐PO3H2 (A5), L ‐Ala‐D,L ‐NH‐CH(iPr)‐PO3H2 (A6), L ‐Ala‐D,L ‐NH‐CH(tBu)‐PO3H2 (A7), L ‐Ala‐D,L ‐NH‐CH(iBu)‐PO3H2 (A8), L ‐Ala‐D,L ‐NH‐CH(cBu)‐PO3H2 (A9), L ‐Ala‐D,L ‐NH‐CH(nPA)‐PO3H2 (A10), β‐Ala‐D ‐NH‐CH(Me)‐PO3H2 (A11), and D,L ‐Ala‐NH‐C(Me,Me)‐PO3H2 (A12). The equilibrium geometries and vibrational wavenumbers are calculated using density functional theory (DFT) at the B3LYP; 6–31 + + G** level of theory using Gaussian'03, GaussSum 0.8, and GAR2PED software. We briefly compare and analyze the experimental and calculated vibrational wavenumbers in the range of 3600–400 cm−1. In addition, Raman wavenumbers are compared to those from surface‐enhanced Raman scattering (SERS) for the phosphonodipeptides of alanine (Ala) adsorbed on a colloidal silver surface. The geometry of these molecules etched on the silver surface is deduce from the observed changes in both the intensity and breadth of Raman bands in the spectra of the bound vs free species. For example, A7, A8, A1, A3, and A4 appear to adsorb onto the colloidal silver particles mainly through the phosphonate terminus, and for A3 and A4, through the  C‐NH2 and  CONH fragments. The most dominant SERS bands of A5, A6, A9, A10, and A11 are due to the amide bond vibrations, as well as to the vibrations of the  C‐NH2 group (A9 and A10) and the C C group (A6 and A11). The differences recorded for the A5, A6, A9, A10, and A11 and those of A2 and A12 are due to interactions between the amine and methyl groups with the silver surface, and they reflect vibrational characteristic of these groups. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

2.
Eritadenine, 2(R),3(R)‐dihydroxy‐4‐(9‐adenyl)‐butyric acid, is a cholesterol‐reducing compound naturally occurring in the shitake mushroom (Lentinus edodes). To identify the unknown Raman spectrum of this compound, pure synthetic eritadenine was examined and the vibrational modes were assigned by following the synthesis pathway. This was accomplished by comparing the known spectra of the starting compounds adenine and D ‐ribose with the spectra of a synthesis intermediate, methyl 5‐(6‐Aminopurin‐9H‐9‐yl)‐2,3‐O‐isopropylidene‐5‐deoxy‐β‐D ‐ribofuranoside (MAIR) and eritadenine. In the Raman spectrum of eritadenine, a distinctive vibrational mode at 773 cm−1 was detected and ascribed to vibrations in the carbon chain, ν(C C). A Raman line that arose at 1212 cm−1, both in the Raman spectrum of MAIR and eritadenine, was also assigned to ν(C C). Additional Raman lines detected at 1526 and at 1583 cm−1 in the Raman spectrum of MAIR and eritadenine were assigned to ν(N C) and a deformation of the purine ring structure. In these cases the vibrational modes are due to the linkage between adenine and the ribofuranoside moiety for MAIR, and between adenine and the carbon chain for eritadenine. This link is also the cause for the disappearance of adenine specific Raman lines in the spectrum of both MAIR and eritadenine. Several vibrations observed in the spectrum of D ‐ribose were not observed in the Raman spectrum of eritadenine due to the absence of the ribose ring structure. In the Raman spectrum of MAIR some of the D ‐ribose specific Raman lines disappeared due to the introduction of methyl and isopropylidene moieties to the ribose unit. With the approach presented in this study the so far unknown Raman spectrum of eritadenine could be successfully identified and is presented here for the first time. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

3.
The surface‐enhanced Raman scattering (SERS) of sodium alginates and their hetero‐ and homopolymeric fractions obtained from four seaweeds of the Chilean coast was studied. Alginic acid is a copolymer of β‐D ‐mannuronic acid (M) and α‐L guluronic acid (G), linked 1 → 4, forming two homopolymeric fractions (MM and GG) and a heteropolymeric fraction (MG). The SERS spectra were registered on silver colloid with the 632.8 nm line of a He Ne laser. The SERS spectra of sodium alginate and the polyguluronate fraction present various carboxylate bands which are probably due to the coexistence of different molecular conformations. SERS allows to differentiate the hetero‐ and homopolymeric fractions of alginic acid by characteristic bands. In the fingerprint region, all the poly‐D ‐mannuronate samples present a band around 946 cm−1 assigned to C O stretching, and C C H and C O H deformation vibrations, a band at 863 cm−1 assigned to deformation vibration of β‐C1 H group, and one at 799–788 cm−1 due to the contributions of various vibration modes. Poly‐L ‐guluronate spectra show three characteristic bands, at 928–913 cm−1 assigned to symmetric stretching vibration of C O C group, at 890–889 cm−1 due to C C H, skeletal C C, and C O vibrations, and at 797 cm−1 assigned to α C1 H deformation vibration. The heteropolymeric fractions present two characteristic bands in the region with the more important one being an intense band at 730 cm−1 due to ring breathing vibration mode. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

4.
In this paper, a novel approach to analyze in situ (−)‐bornyl acetate (BA) in pichtae essential oil (Siberian fir needle oil, Abies sibirica oil) by means of Raman optical activity (ROA) is reported. As part of this approach, a conformational study in the gas phase of (+)‐ and (−)‐BA has been carried out, predicting the presence of three conformers for each enantiomer at 298.15 K. The structures of these conformers were optimized with density functional theory with the Becke 3LYP functional and 6–311 + + g** basis set. Subsequently, the Raman and ROA spectra were simulated in order to compare them with the experimentally measured spectra of the neat enantiomers of BA. Finally, the combination of Raman and ROA spectroscopy as well as DFT calculations was successfully applied not only for the detection of BA but also for the determination of the specific enantiomer of BA present in the investigated pichtae essential oil samples. Thus, the ROA technique described here has the potential to be used as a fast and easy commercial method to control the quality of essential oils. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

5.
Natural resonance electronic Raman optical activity (ROA) is observed for the first time. Coincidently, the first example of vibrational ROA enhanced by low‐lying electronic transition is reported. These new phenomena were measured using the rare‐earth complex Eu(tfc)3 (+)‐tris[3‐trifluoroacetyl‐D ‐camphorato]europium(III), where electronic resonance occurs between the 532‐nm laser excitation and the 7F15D1 transition of the Eu3+ metal center. Electronic Raman spectra involve the Raman transitions terminating on the low‐lying electronic states of Eu(tfc)3. The observed vibrational ROA spectra are enhanced relative to typical ROA spectra by the proximity of vibrational states of Eu(tfc)3 to its low‐lying electronic states with significant magnetic‐dipole character, whereas the parent vibrational Raman spectra do not appear to be resonance‐enhanced since the 532‐nm vibrational Raman spectrum has similar relative intensities to the corresponding Raman spectrum measured with 1064‐nm laser excitation. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

6.
UV resonance Raman spectra of the antimicrobial peptide (AMP) Anoplin (L ‐Anoplin‐NH2) and two of its derivatives (enantiomer D ‐Anoplin‐NH2 and C‐terminus deamidated L ‐Anoplin‐OH) were measured in aqueous buffer solution and in membrane‐mimetic environments including 2,2,2‐trifluoro ethanol (TFE), zwitterionic lipid dipalmitoylglycerophosphocholine (DPPC) and anionic lipid dipalmitoylglycerophosphoglycerol (DPPG) vesicle solutions. All three peptides were found to adopt random‐coil/β turn‐like conformation in aqueous solution over the temperature range of 1–60 °C. The conformation was found to become more α‐helical in membrane‐mimetic solutions such as TFE and DPPG but not in DPPC for all Anoplin derivatives. The data demonstrate that Anoplin preferentially binds to the anionic over the zwitterionic model cell membranes. Results also showed that deamidation does not change the conformation of L ‐Ano‐NH2 very significantly, but does alter membrane rupturing and antimicrobial activities thus confirming that it is the physicochemical properties rather than the peptide conformation that define the mechanism of AMP action. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

7.
This work used infrared absorption and Raman spectroscopy to determine the structure of seven modified fragments (residues 6–14 of the polypeptide chain) of bombesin (BN6–14). The peptides studied are cyclo[D ‐Phe6, His7, Leu14]BN6–14, [D ‐Phe6, Leu‐NHEt13, des‐Met14]BN6–14, [D ‐Phe6, Leu13‐®‐p‐chloro‐Phe14]BN6–14, [D ‐Phe6, β‐Ala11, Phe13, Nle14]BN6–14, [D ‐Tyr6, β‐Ala11, Phe13, Nle14]BN6–14, [D ‐Tyr6, β‐Phe11, Phe13, Nle14OH]BN6–14 and [D ‐Cys6, Asn7, D ‐Ala11, Cys14]BN6–14. These peptides are potent bombesin agonists useful in the treatment of tumors. Surface‐enhanced Raman scattering (SERS) spectroscopy was also used to examine the behavior of these molecules on an electrochemically roughened silver surface. The SERS spectra reveal that substituting native amino acids in these molecules with synthetic ones changes their adsorption state slightly on an electrochemically roughened surface of silver. The peptides [D ‐Tyr6, β‐Ala11, Phe13, Nle14]BN6–14 and [D ‐Tyr6, β‐Phe11, Phe13, Nle14OH]BN6–14 tend to adsorb strongly on this surface via C fragment (∼1400 cm−1). The observed medium enhancement of the Trp8 residue and amide bond Raman signals indicate further interactions between these fragments and the surface. [D ‐Phe6, Leu‐NHEt13, des‐Met14]BN6–14 and [D ‐Cys6, Asn7, D ‐Ala11, Cys14]BN6–14 are shown to be coordinated to the silver through  CONH , CO, and the indole ring. The strongest SERS bands (∼1506, ∼1275, ∼1149, and ∼1007 cm−1) of [D ‐Phe6, Leu13‐®‐p‐chloro‐Phe14]BN6–14 and [D ‐Phe6, β‐Ala11, Phe13, Nle14]BN6–14 suggest that these two peptides bind to the silver via Trp8 and  CONH . In the case of cyclo[D ‐Phe6, His7, Leu14]BN6–14, the formation of a peptide/Ag complex is confirmed by the strong SERS bands involving Trp8 and  CONH vibrations, which are accompanied by a SERS signal due to the CO vibrations. For these analogs, the relative potency for inhibition of binding of 125I‐[Tyr4]BN to rat pancreas acini cells was correlated with the behavior of the amide bond on the silver surface, while the contribution of the structural components to the ability to interact with the rGRP‐R was correlated with the SERS patterns. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

8.
In this work, multivariate data analysis methods were applied to the analysis and interpretation of micro‐Raman spectra, collected from a broad set of historical iron‐based ink samples, previously characterised for the content of organic acids (gallic acid, ellagic acid and protocatechuic acid). The proposed method relies on principal component analysis of the noisy spectra typically obtained on original, degraded, organic samples, where fluorescence could affect the Raman signal. The signal components could be distinguished from the noise components and then used to build a linear discriminant analysis (LDA) model, achieving separation of the spectra into three classes. Selection of pure signal factors also improved effectiveness and performances of partial least square regression (PLS) algorithms, allowing quantification of condensed tannic acid residuals. Application of multivariate methods to discriminate signal from noise removes the need for spectral data manipulation (filtering, smoothing and differentiating). The obtained classification method for discrimination of historic inks and the regression method for determination of condensed tannic acid residuals supports the use of Raman analysis of fluorescing organic materials, and may provide information to scholars on ink composition and potentially on its provenance. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

9.
Experimental vibrational spectroscopic studies and density functional theory (DFT) calculations of the di‐amino acid peptide derivatives α‐ and β‐N‐acetyl‐L‐Asp‐L‐Glu have been undertaken. Raman and infrared spectra have been recorded for samples in the solid state. DFT simulations were conducted using the B3‐LYP correlation functional and the cc‐pVDZ basis set to determine energy minimized/geometry optimized structures (based on a single isolated molecule in the gaseous state). Normal coordinate calculations have provided vibrational assignments for fundamental modes, including their potential energy distributions. Significant differences are observed between α‐ and β‐N‐acetyl‐L‐Asp‐L‐Glu both in the computed structures and in the vibrational spectra. The combination of experimental and calculated spectra provide an insight into the structural and vibrational spectroscopic properties of di‐amino acid peptide derivatives. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

10.
Investigations of the vibrational spectra of cyclo(Gly‐Gly), cyclo(L‐Ala‐L ‐Ala) and cyclo(L ‐Ala‐Gly) are reported. Raman scattering and Fourier transform infrared (FTIR) spectra of solid‐state and aqueous protonated samples, as well as their corresponding N‐deuterated isotopomers, have been examined. In addition, density functional theory (DFT) (B3‐LYP/cc‐pVDZ) calculations of molecular structures and their associated vibrational modes were carried out. In each case, the calculated structures of lowest energy for the isolated gas‐phase molecules have boat conformations. Assignments have been made for the observed Raman and FTIR vibrational bands of the cyclic di‐amino acid peptides (CDAPs) examined. Raman polarization studies of aqueous phase samples are consistent with C2 and C1 symmetries for the six‐membered rings of cyclo(L‐Ala‐L‐Ala) and cyclo(L‐Ala‐Gly), respectively. There is a good correlation between experimental and calculated vibrational bands for the three CDAPs. These data are in keeping with boat conformations for cyclo(L‐Ala‐L‐Ala) and cyclo(L‐Ala‐Gly) molecules, predicted by the ab initio calculations, in both the solid and aqueous solution states. However, Raman spectroscopic results might infer that cyclo(L‐Ala‐Gly) deviates only slightly from planarity in the solid state. The potential energy distributions of the amide I and II modes of a cis‐peptide linkage are shown to be significantly different from those of the trans‐peptides. For example, deuterium shifts have shown that the cis‐amide I vibrations found in cyclo(Gly‐Gly), cyclo(L‐Ala‐L‐Ala), and cyclo(L‐Ala‐Gly) have larger N‐H contributions compared to their trans‐amide counterparts. Compared to trans‐amide II vibrations, cis‐amide II vibrations show a considerable decrease in N H character. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

11.
Vibrational bands of L ‐tryptophan which was adsorbed on Ag nanoparticles (∼10 nm in diameter) have been investigated in the spectral range of 200–1700 cm−1 using surface‐enhanced Raman scattering (SERS) spectroscopy. Compared with the normal Raman scattering (NRS) of L ‐tryptophan in either 0.5 M aqueous solution (NRS‐AS) or solid powder (NRS‐SP), the intensified signals by SERS have made the SERS investigation at a lower molecular concentration (5 × 10−4 M ) possible. Ab initio calculations at the B3LYP/6‐311G level have been carried out to predict the optimal structure and vibrational wavenumbers for the zwitterionic form of L ‐tryptophan. Facilitated with the theoretical prediction, the observed vibrational modes of L ‐tryptophan in the NRS‐AS, NRS‐SP, and SERS spectra have been analyzed. In the spectroscopic observations, there are no significant changes for the vibrational bands of the indole ring in either NRS‐AS, NRS‐SP, or SERS. In contrast, spectral intensities involving the vibrations of carboxylate and amino groups are weak in NRS‐AS and NRS‐SP, but strong in SERS. The intensity enhancement in the SERS spectrum can reach 103–104‐fold magnification. On the basis of spectroscopic analysis, the carboxylate and amino groups of L ‐tryptophan are determined to be the preferential terminal groups to attach onto the surfaces of Ag nanoparticles in the SERS measurement. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

12.
不同的不饱和脂肪酸各自具有其不同的生理功能,但常见的不饱和脂肪酸产品大部分为几种脂肪酸的混合物,故在应用前对不纯的脂肪酸产品进行组成分析是必须的。测量了不饱和脂肪酸产品组分中最常见的油酸和亚油酸的拉曼光谱,确定了各拉曼谱线的振动模归属,分析了其分子的构象特征。该结果为研究长链不饱和脂肪酸的振动能级结构及能级间跃迁等做了基础工作,丰富了有机物分子的价键数据和性质。同时详细分析比较了油酸和亚油酸拉曼光谱的差异,为定性鉴别脂肪酸产品的成分提供了一种简便有效的方法,对拉曼光谱在地沟油检测方面的应用具有重要的指导意义。  相似文献   

13.
The vibrational spectra of trans‐p‐coumaric acid (pCA) and syringic acid (SA) are discussed. The spectral fingerprints of the organic acids observed in the infrared and Raman spectra are assigned to fundamental vibrational wavenumbers supported by quantum chemical computations. The average surface‐enhanced Raman scattering spectra of both acids have been obtained on silver colloidal solutions and the interpretation of the spectra is presented based on complementary Raman spectra and computational results for the silver salts. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

14.
Four L ‐valine (L ‐Val) phosphonate dipeptides that are potent inhibitors of zinc metalloproteases, namely, L ‐Val‐C(Me)2‐PO3H2 (V1), L ‐Val‐CH(iP)‐PO3H2 (V2), L ‐Val‐CH(iB)‐PO3H2 (V3), and L ‐Val‐C(Me)(iP)‐PO3H2 (V4), are studied by Fourier‐transform infrared (FT‐IR) spectroscopy, Fourier‐transform Raman spectroscopy (FT‐RS), and surface‐enhanced Raman scattering (SERS). The band assignment (wavenumbers and intensities) is made based on (B3LYP/6‐311 + + G**) calculations. Comparison of theoretical FT‐IR and FT‐RS spectra with those of SERS allows to obtain information on the orientation of these dipeptides as well as specific‐competitive interactions of their functionalities with the silver substrate. More specifically, V1 and V4 appear to interact with the silver substrate mainly via a  CsgCH3 moiety localized at the  NamideCsg(CH3)P molecular fragment. In addition, the  POH and isopropyl units of V4 assist in the adsorption process of this molecule. In contrast, the  CαNH2 and  PO3H groups of V2 and V3 interact with the silver nanoparticles, whereas their isopropyl and isobutyl fragments seem to be repelled by the silver substrate (except for the  CH2  of V3), similar to the  Cβ(CH3)2 fragment of L ‐Val for all L ‐Val phosphonate dipeptides investigated in this work. The adsorption mechanism of these molecules onto the colloidal silver surface is also affected by amide bond behavior. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

15.
Chemical imaging method of vibrational spectroscopy, which provides both spectral and spatial information, creates a three‐dimensional (3D) dataset with a huge amount of data. When the components of the sample are unknown or their reference spectra are not available, the classical least squares (CLS) method cannot be applied to create visualized distribution maps. Raman image datasets can be evaluated even in such cases using multivariate (chemometric) methods for extracting the needed hidden information. The capability of chemometrics‐assisted Raman mapping is evaluated through the analysis of pharmaceutical tablets (considered as unknown) with the aim of estimating the pure component spectra based on the collected Raman image. Six chemometric methods, namely, principal component analysis (PCA), maximum autocorrelation factors (MAF), sample–sample 2D correlation spectroscopy (SS2D), self‐modeling mixture analysis (SMMA), multivariate curve resolution–alternating least squares (MCR‐ALS), and positive matrix factorization (PMF), were compared. SMMA was found to be the best choice to determine the number of components. MCR‐ALS and PMF provided the pure component spectra with the highest quality. MCR‐ALS was found to be superior to PMF in the estimation of Raman scores (which correspond to the concentrations) and yielded almost the same results as CLS (using the real reference spectra). Thus, the combination of Raman mapping and chemometrics could be successfully used to characterize unknown pharmaceuticals, identify their ingredients, and obtain information about their structures. This may be useful in the struggles against illegal and counterfeit products and also in the field of pharmaceutical industry when contaminants are to be identified. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

16.
A comparative study of molecular structures of five L ‐proline (L ‐Pro) phosphonodipeptides: L ‐Pro‐NH‐C(Me,Me)‐PO3H2 (P1), L ‐Pro‐NH‐C(Me,iPr)‐PO3H2 (P2), L ‐Pro‐L ‐NH‐CH(iBu)‐PO3H2 (P3), L ‐Pro‐L ‐NH‐CH(PA)‐PO3H2 (P4) and L ‐Pro‐L ‐NH‐CH(BA)‐PO3H2 (P5) has been carried out using Raman and absorption infrared techniques of molecular spectroscopy. The interpretation of the obtained spectra has been supported by density functional theory calculations (DFT) at the B3LYP; 6–31 + + G** level using Gaussian 2003 software. The surface‐enhanced Raman scattering (SERS) on Ag‐sol in aqueous solutions of these phosphonopeptides has also been investigated. The surface geometry of these molecules on a silver colloidal surface has been determined by observing the position and relative intensity changes of the Pro ring, amide, phosphonate and so‐called spacer (−R) groups vibrations of the enhanced bands in their SERS spectra. Results show that P4 and P5 adsorb onto the silver as anionic molecules mainly via the amide bond (∼1630, ∼1533, ∼1248, ∼800 and ∼565 cm−1), Pro ring (∼956, ∼907 and ∼876 cm−1) and carboxylate group (∼1395 and ∼909 cm−1). Coadsorption of the imine nitrogen atom and PO group with the silver surface, possibly by formation of a weaker interaction with the metal, is also suggested by the enhancement of the bands at 1158 and 1248 cm−1. P1, P2 and P3 show two orientations of their main chain on the silver surface resulting from different interactions of the  C CH3,  NH and  CONH fragments with this surface. Bonding to the Ag surface occurs mainly through the imino atom (1166 cm−1) for P2, while for P1 and P3 it occurs via the methyl group(s) (1194–1208 cm−1). The amide group functionality (CONH) is practically not involved in the adsorption process for P1 and P2, whereas the Cs P bonds do assist in the adsorption. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

17.
Solid‐state protonated and N,O‐deuterated Fourier transform infrared (IR) and Raman scattering spectra together with the protonated and deuterated Raman spectra in aqueous solution of the cyclic di‐amino acid peptide cyclo(L ‐Asp‐L ‐Asp) are reported. Vibrational band assignments have been made on the basis of comparisons with previously cited literature values for diketopiperazine (DKP) derivatives and normal coordinate analyses for both the protonated and deuterated species based upon DFT calculations at the B3‐LYP/cc‐pVDZ level of the isolated molecule in the gas phase. The calculated minimum energy structure for cyclo(L ‐Asp‐L ‐Asp), assuming C2 symmetry, predicts a boat conformation for the DKP ring with both the two L ‐aspartyl side chains being folded slightly above the ring. The CO stretching vibrations have been assigned for the side‐chain carboxylic acid group (e.g. at 1693 and 1670 cm−1 in the Raman spectrum) and the cis amide I bands (e.g. at 1660 cm−1 in the Raman spectrum). The presence of two bands for the carboxylic acid CO stretching modes in the solid‐state Raman spectrum can be accounted for by factor group splitting of the two nonequivalent molecules in a crystallographic unit cell. The cis amide II band is observed at 1489 cm−1 in the solid‐state Raman spectrum, which is in agreement with results for cyclic di‐amino acid peptide molecules examined previously in the solid state, where the DKP ring adopts a boat conformation. Additionally, it also appears that as the molecular mass of the substituent on the Cα atom is increased, the amide II band wavenumber decreases to below 1500 cm−1; this may be a consequence of increased strain on the DKP ring. The cis amide II Raman band is characterized by its relatively small deuterium shift (29 cm−1), which indicates that this band has a smaller N H bending contribution than the trans amide II vibrational band observed for linear peptides. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

18.
We report an ab initio simulation study of the ultrafast broad bandwidth ultraviolet stimulated resonance Raman spectra (SRRS) of l ‐tyrosine, l ‐tryptophan, and trans‐l ‐tryptophan‐l ‐tyrosine (WY) dipeptide. Two‐pulse one‐dimensional SRRS and three‐pulse two‐dimensional SRRS that reveal inter‐residue and intra‐residue vibrational correlations are simulated using electronically resonant or pre‐resonant pulse configurations that select the Raman signal and discriminate against excited state pathways. Multimode effects are incorporated via the cumulant expansion. The two‐dimensional SRRS technique is more sensitive to residue couplings than spontaneous Raman. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

19.
B3‐LYP/cc‐pVDZ calculations of the gas‐phase structure and vibrational spectra of the isolated molecule cyclo(L ‐Ser‐L ‐Ser), a cyclic di‐amino acid peptide (CDAP), were carried out by assuming C2 symmetry. It is predicted that the minimum‐energy structure is a boat conformation for the diketopiperazine (DKP) ring with both L ‐seryl side chains being folded slightly above the ring. An additional structure of higher energy (15.16 kJ mol−1) has been calculated for a DKP ring with a planar geometry, although in this case two fundamental vibrations have been calculated with imaginary wavenumbers. The reported X‐ray crystallographic structure of cyclo(L ‐Ser‐L ‐Ser), shows that the DKP ring displays a near‐planar conformation, with both the two L ‐seryl side chains being folded above the ring. It is hypothesized that the crystal packing forces constrain the DKP ring in a planar conformation and it is probable that the lower energy boat conformation may prevail in the aqueous environment. Raman scattering and Fourier‐transform infrared (FT‐IR) spectra of solid state and aqueous solution samples of cyclo(L ‐Ser‐L ‐Ser) are reported and discussed. Vibrational band assignments have been made on the basis of comparisons with the calculated vibrational spectra and band wavenumber shifts upon deuteration of labile protons. The experimental Raman and IR results for solid‐state samples show characteristic amide I vibrations which are split (Raman: 1661 and 1687 cm−1, IR: 1666 and 1680 cm−1), possibly due to interactions between molecules in a crystallographic unit cell. The cis amide I band is differentiated by its deuterium shift of ∼30 cm−1, which is larger than that previously reported for trans amide I deuterium shifts. A cis amide II mode has been assigned to a Raman band located at 1520 cm−1. The occurrence of this cis amide II mode at a wavenumber above 1500 cm−1 concurs with results of previously examined CDAP molecules with low molecular weight substituents on the Cα atoms, and is also indicative of a relatively unstrained DKP ring. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

20.
Raman spectroscopy, in principle, is an excellent technique for the study of molecular species developed on metal surfaces during electrochemical investigations. However, the use of the more common laser wavelengths such as the 514.5‐nm line results in spectra of less than optimal intensity, particularly for iron oxide compounds. In the present work, near‐resonance enhancement of the Raman spectra was investigated for the iron oxide and iron oxyhydroxide compounds previously reported to be present in the passive film on iron, using a tuneable dye laser producing excitation wavelengths between 560 and 637 nm. These compounds were hematite (α‐Fe2O3), maghemite (γ‐Fe2O3), magnetite (Fe3O4), goethite (α‐FeOOH), akaganeite (β‐FeOOH), lepidocrocite (γ‐FeOOH) and feroxyhyte (δ‐FeOOH). Optimum enhancement, when compared to that with the 514.5‐nm line, was obtained for all the iron oxide and oxyhydroxide standard samples in the low wavenumber region (<1000 cm−1) using an excitation wavelength of 636.4 nm. Particularly significant enhancement was obtained for lepidocrocite, hematite and goethite. Copyright © 2010 John Wiley & Sons, Ltd.  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