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 15 毫秒
1.
In this study, the transition process of the helix‐sense inversion of poly(β‐phenethyl‐L‐aspartate) was investigated by Raman scattering and 2‐dimensional correlation spectroscopy. Temperature‐dependent Raman spectra were obtained during the helix‐sense inversion. The results of 2‐dimensional correlation analysis in the spectral regions of 1600‐1800 and 3200‐3400 cm?1 showed that the intensity changes of the side‐chain ester C═O stretching bands occurred prior to those of amide A and amide I bands in the unwinding process of αR‐helix on heating. The sequential order of the intensity changes for amide A, amide I, and the side‐chain ester C═O stretching bands during the inversion process was determined. It was found that the conformation change of the side chain occurred prior to that of the main chain for the αR‐helix on heating. Thus, we concluded that the transformation of the backbone chain from right‐handed to left‐handed is triggered by the conformational change of the side chains.  相似文献   

2.
Analysis of insulin amyloid fibrils by Raman spectroscopy   总被引:1,自引:1,他引:0  
The formation of amyloid fibrils from insulin is investigated using drop-coating-deposition-Raman (DCDR) difference spectroscopy and atomic force microscopy (AFM). Fibrils formed using various co-solvents and heating cycles are found to induce the appearance of Raman difference peaks in the amide I (approximately 1675 cm(-1)), amide III (approximately 1220 cm(-1)), and peptide backbone (approximately 1010 cm(-1)), consistent with an increase in beta-sheet content. Comparisons of results obtained from fibrils in either H2O or D2O suggest that the NH/ND stretch bands (at approximately 3300 cm(-1)/ approximately 2400 cm(-1)) are also enhanced in intensity upon fibril formation. If there is any water trapped in the core of the fibrils its OH/OD Raman intensity is too small to be detected in the presence of the stronger NH/ND bands which appear in the same region. AFM is used to confirm the formation of fibrils of about 5 nm diameter (and various lengths).  相似文献   

3.
Breast cancers are the leading cancers among females. Diagnosis by fine needle aspiration cytology (FNAC) is the gold standard. The widely practiced screening method, mammography, suffers from high false positive results and repeated exposure to harmful ionizing radiation. As with all other cancers survival rates are shown to heavily depend on stage of the cancers (Stage 0, 95%; Stage IV, 75%). Hence development of more reliable screening and diagnosis methodology is of considerable interest in breast cancer management. Raman spectra of normal, benign, and malignant breast tissue show significant differences. Spectral differences between normal and diseased breast tissues are more pronounced than between the two pathological conditions, malignant and benign tissues. Based on spectral profiles, the presence of lipids (1078, 1267, 1301, 1440, 1654, 1746 cm(-1)) is indicated in normal tissue and proteins (stronger amide I, red shifted DeltaCH2, broad and strong amide III, 1002, 1033, 1530, 1556 cm(-1)) are found in benign and malignant tissues. The major differences between benign and malignant tissue spectra are malignant tissues seem to have an excess of lipids (1082, 1301, 1440 cm(-1)) and presence of excess proteins (amide I, amide III, red shifted DeltaCH2, 1033, 1002 cm(-1)) is indicated in benign spectra. The multivariate statistical tool, principal components analysis (PCA) is employed for developing discrimination methods. A score of factor 1 provided a reasonable classification of all three tissue types. The analysis is further fine-tuned by employing Mahalanobis distance and spectral residuals as discriminating parameters. This approach is tested both retrospectively and prospectively. The limit test, which provides the most unambiguous discrimination, is also considered and this approach clearly discriminated all three tissue types. These results further support the efficacy of Raman spectroscopic methods in discriminating normal and diseased breast tissues.  相似文献   

4.
Podstawka E 《Biopolymers》2008,89(11):980-992
This work presents a Fourier-transform absorption infrared, Fourier-transform Raman, and surface-enhanced Raman scattering (SERS) study of the following peptides belonging to the bombesin-like family: phyllolitorin, [Leu(8)]phyllolitorin, NMB, NMC, and PG-L. The SERS study was undertaken to understand the adsorption mechanism of bombesin-like peptides on an electrochemically roughened silver electrode surface and to show changes in the adsorption mechanism with alterations in amino acids and small tertiary structures. The SERS spectra presented here shows bands mainly associated with the Trp(8) residue vibrations. The presence of mainly pyrrole coring vibrations for phyllolitorin and [Leu(8)]phyllolitorin and mainly benzene coring modes for NMB and NMC indicated that these groups interact with the roughened silver electrode surface. Furthermore, N(1)--C(8) and C(3)--C(9) bonds of the PG-L indole ring seemed to have nearly a vertical orientation on the electrode surface. In addition, distinct vibrations of the C--S fragment were observed in the SERS spectra of [Leu(8)]phyllolitorin and PG-L. The strong enhancement of the nu(C==O) vibration in the [Leu(8)]phyllolitorin SERS spectrum yielded evidence that the intact C==O bond(s) bind strongly to the silver electrode surface, whereas NMC, phyllolitorin, and NMB were located near the silver surface. This finding was supported by the presence of the nu(C--C(==O)) mode. The amide I band observed at 1642 and 1634 cm(-1) for NMB and NMC, respectively, and the Raman amide III band seen in the 1282-1249 cm(-1) range for all peptides except PG-L, indicate that the strongly hydrogen-bonded alpha-helical conformation and random-coil structure are favored for binding to the surface. (c) 2008 Wiley Periodicals, Inc. Biopolymers 89: 980-992, 2008.This article was originally published online as an accepted preprint. The "Published Online" date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com.  相似文献   

5.
Ye M  Zhang QL  Li H  Weng YX  Wang WC  Qiu XG 《Biophysical journal》2007,93(8):2756-2766
The infrared (IR) absorption of the amide I band for the loop structure may overlap with that of the alpha-helices, which can lead to the misassignment of the protein secondary structures. A resolution-enhanced Fourier transform infrared (FTIR) spectroscopic method and temperature-jump (T-jump) time-resolved IR absorbance difference spectra were used to identify one specific loop absorption from the helical IR absorption bands of horse heart cytochrome c in D2O at a pD around 7.0. This small loop consists of residues 70-85 with Met-80 binding to the heme Fe(III). The FTIR spectra in amide I' region indicate that the loop and the helical absorption bands overlap at 1653 cm(-1) at room temperature. Thermal titration of the amide I' intensity at 1653 cm(-1) reveals that a transition in loop structural change occurs at lower temperature (Tm=45 degrees C), well before the global unfolding of the secondary structure (Tm approximately 82 degrees C). This loop structural change is assigned as being triggered by the Met-80 deligation from the heme Fe(III). T-jump time-resolved IR absorbance difference spectra reveal that a T-jump from 25 degrees C to 35 degrees C breaks the Fe-S bond between the Met-80 and the iron reversibly, which leads to a loop (1653 cm(-1), overlap with the helical absorption) to random coil (1645 cm(-1)) transition. The observed unfolding rate constant interpreted as the intrachain diffusion rate for this 16 residue loop was approximately 3.6x10(6) s(-1).  相似文献   

6.
Measurement of the elastic properties of single osteon lamellae is still one of the most demanding tasks in bone mechanics to be solved. By means of site-matched Raman microspectroscopy, acoustic microscopy and nanoindentation the structure, chemical composition and anisotropic elasticity of individual lamellae in secondary osteons were investigated. Acoustic impedance images (911-MHz) and two-dimensional Raman spectra were acquired in sections of human femoral bone. The samples were prepared with orientations at various observation angles theta relative to the femoral long axis. Nanoindentations provided local estimations of the elastic modulus and landmarks necessary for spatial fusion of the acoustic and spectral Raman images. Phosphate nu(1) (961 cm(-1)) and amide I (1665 cm(-1)) band images representing spatial distributions of mineral and collagen were fused with the acoustic images. Acoustic impedance was correlated with the indentation elastic modulus E(IT) (R(2)=0.61). Both parameters are sensitive to elastic tissue anisotropy. The lowest values were obtained in the direction perpendicular to the femoral long axis. Acoustic images exhibit a characteristic bimodal lamellar pattern of alternating high and low impedance values. Since this undulation was not associated with a variation of the phosphate nu(1)-band intensity in the Raman images, it was attributed to variations of the lamellar orientation. After threshold segmentation and conversion to elastic modulus the orientation and transverse isotropic elastic constants were derived for individual ensembles of apparent thin and thick lamellae. Our results suggest that this model represents the effective anisotropic properties of an asymmetric twisted plywood structure made of transverse isotropic fibrils. This is the first report that proves experimentally the ability of acoustic microscopy to map tissue elasticity in two dimensions with micrometer resolution. The combination with Raman microspectroscopy provides a unique way to study bone and mineral metabolism and the relation with mechanical function at the ultrastructural tissue level.  相似文献   

7.
The filamentous bacteriophage PH75, which infects the thermophile T. thermophilus, assembles in vivo at 70 degrees C and is stable to at least 90 degrees C. Although a high-resolution structure of PH75 is not available, the virion is known to comprise a closed single-stranded (ss) DNA circle of 6500 nucleotides sheathed by a capsid comprising 2700 copies of a 46-residue subunit (pVIII). Here, we employ Raman and UV-resonance Raman (UVRR) spectroscopy to identify structural details of the pVIII and DNA constituents of PH75 that may be related to the high thermostability of the native virion assembly. Analysis of the Raman amide I and amide III signatures reveals that the capsid subunit secondary structure is predominantly (87%) alpha-helical but contains a significant number of residues (6 +/- 1 or 13 +/- 3%) differing from the canonical alpha-helix. This minor structural component is not apparent in capsid subunits of the mesophilic filamentous phages, fd, Pf1, and Pf3, previously examined at similar spectral resolution. The Raman signature of PH75 also differs from those of fd, Pf1, and Pf3 by virtue of an unusual alanine marker (898 cm(-)(1) band), which is attributed to C(alpha)-H hydrogen-bond donation by subunit Ala residues. Because alanines of the PH75 subunit occur primarily within sXXXs motifs (where s is a small side chain, e.g. Gly, Ala, Ser), and because the occurrence of such motifs in alpha-helices is believed to thermostabilize interhelix associations via C(alpha)-H...O interactions [G. Kleiger et al. (2002) Biochemistry 41, 5990-5997], we propose that such hydrogen bonding may explain both the alanyl and amide I/III markers of PH75 capsid subunits and that C(alpha)-H...O interactions may serve as a significant source of virion thermostabilization. Raman and UVRR signatures of PH75 are also distinguished from those of fd, Pf1, and Pf3 by several marker bands that are indicative of hydrophilic Trp and Tyr environments, including hydrogen bonding interactions of aromatic ring substituents. These interactions are likewise proposed as contributors to the high thermostability of PH75 vis-a-vis fd, Pf1, and Pf3. Finally, PH75 is the only filamentous phage exhibiting UVRR markers diagnostic of a highly base-stacked ssDNA genome incorporating the low energy C2'-endo/anti deoxynucleoside conformation. The present results suggest that both intersubunit interactions and genome organization contribute to the enhanced thermostability of PH75 relative to mesophilic filamentous bacteriophages.  相似文献   

8.
Monti P  Taddei P  Freddi G  Ohgo K  Asakura T 《Biopolymers》2003,72(5):329-338
This study focuses on the conformational characterization of poly(alanine-glycine) II (pAG II) as a model for a Bombyx mori fibroin silk I structure. Raman, IR, and 13C-cross-polarization/magic angle spinning NMR spectra of pAG II are discussed in comparison with those of the crystalline fraction of B. mori silk fibroin (chymotryptic precipitate, Cp) with a silk I (silk I-Cp) structure. The spectral data give evidence that silk I-Cp and the synthetic copolypeptide pAG II have similar conformations. Moreover, the spectral findings reveal that silk I-Cp is more crystalline than pAG II; consequently, the latter contains a larger amount of the random coil conformation. Differential scanning calorimetry measurements confirm this result. N-Deuteration experiments on pAG II allow us to attribute the Raman component at 1320 cm(-1) to the amide III mode of a beta-turn type II conformation, thus confirming the results of those who propose a repeated beta-turn type II structure for silk I. The analysis of the Raman spectra in the nuNH region confirms that the silk I structure is characterized by the presence of different types of H-bonding arrangements, in agreement with the above model.  相似文献   

9.
Inelastic neutron-scattering (INS) spectra of three isotopic derivatives of polyglycine I (-COCH2NH-)n, (-COCD2NH-)n, and (-COCH2ND-)n at 20 K are presented from 30 to 4000 cm(-1). The band frequencies are compared to those observed in the infrared and Raman. Assignments in terms of group vibrations are proposed. These mostly resemble previous assignment schemes, except for the amide bands. The INS intensities reveal that the proton dynamics for the (N)H proton are totally different from those proposed previously. They are independent of the molecular frame and the valence bond approach is not consistent with observation. A phenomenological approach is proposed in terms of localized modes. The calculated intensities reveal that the (N)H stretching mode has two components at approximately 1377 and 1553 cm(-1). This is a dramatic change compared to all former assignments at approximately 3280 cm(-1) based on infrared and Raman data. These proton-dynamics are associated with a weakening of the NH bond due to the ionic character of the hydrogen bond (N(delta-)...H+...O(delta'-)) and proton transfer. The infrared and Raman spectra are re-examined and a new assignment scheme is proposed for the amide bands; the amide A and B bands are re-assigned to the overtones of the stretching modes. A symmetric double-minimum potential for the proton is consistent with all the observations.  相似文献   

10.
In this study we present the electrochemically induced Fourier transform infrared (FTIR) difference spectra of the Cu(A) center derived from the ba(3)-type cytochrome c oxidase of Thermus thermophilus in the spectral range from 1800 to 500 cm(-1). The mid infrared is dominated by the nu(C[double bond]O) vibrations of the amide I modes at 1688, 1660, and 1635 cm(-1), reflecting the redox-induced perturbation of the predominantly beta-sheet type structure. The corresponding amide II signal is found at 1528 cm(-1). In the lower frequency range below 800 cm(-1), modes from amino acids liganding the Cu(A) center are expected. On the basis of the absorbance spectrum of the isolated amino acids, methionine is identified as an important residue, displaying C-S vibrations at these frequencies. This spectral range was previously disregarded by protein IR spectroscopists, mainly due to the strong absorbance of the solvent, H(2)O. With an optimized setup, however, IR is found suitable for structure/function studies on proteins.  相似文献   

11.
Changes in the amide bands in Fourier transform infrared spectra of proteins are generally attributed to alterations in protein secondary structure. In this study spectra of five different globular proteins were compared in the solid and solution states recorded with several sampling techniques. Spectral differences for each protein were observed between the various sampling techniques and physical states, which could not all be explained by a change in protein secondary structure. For example, lyophilization in the absence of lyoprotectants caused spectral changes that could (partially) have been caused by the removal of hydrating water molecules rather than secondary structural changes. Moreover, attenuated total reflectance spectra of proteins in H2O were not directly comparable to transmission spectra due to the anomalous dispersion effect. Our study also revealed that the amide I, II, and III bands differ in their sensitivities to changes in protein conformation: For example, strong bands in the region 1620-1630 and 1685-1695 cm(-1) were seen in the amide I region of aggregated protein spectra. Surprisingly, absorbance of such magnitudes was not observed in the amide II and III region. It appears, therefore, that only the amide I can be used to distinguish between intra- and intermolecular beta-sheet formation. Considering the differing sensitivity of the different amide modes to structural changes, it is advisable to utilize not only the amide I band, but also the amide II and III bands, to determine changes in protein secondary structure. Finally, it is important to realize that changes in these bands may not always correspond to secondary structural changes of the proteins.  相似文献   

12.
The conformational changes associated with the redox transition of plastocyanin (PC) were investigated by absorption and reaction-induced infrared spectroscopy. In addition to spectral features readily ascribed to beta and turn protein secondary structures, the amide I band shows a major component band at 1647 cm(-1) in both redox states of the protein. The sensitivity of this component to deuteration and increasing temperature suggests that PC adopts an unusual secondary structure in solution, which differs from those described for other type I copper proteins, such as azurin and halocyanin. The conformations of oxidized and reduced PC are different, as evidenced (1) by analysis of their amide I band contour and the electrochemically induced oxidized-minus-reduced difference spectrum and (2) by their different thermal stability. The redox-induced difference spectrum exhibits a number of difference bands within the conformationally sensitive amide I band that could be assigned to peptide C=O modes, in light of their small shift upon deuteration, and to signals attributable to side chain vibrational modes of Tyr residues. Lowering the pH to 4.8 induces destabilization of both redox states of the protein, more pronounced for reduced PC, without significantly affecting their secondary structure. Besides the conformational differences obtained at neutral pH, the oxidized-minus-reduced difference spectrum shows two broad and strong negative bands at 1405 and 1571 cm(-1), assigned to COO(-) vibrations, and a broad positive band at 1710 cm(-1), attributed to the C=O vibration of a COOH group(s). These bands are indicative of a protonation of (an) Asp or Glu side chain(s) upon plastocyanin oxidation at acidic pH.  相似文献   

13.
Atherosclerosis is an inflammatory process occurring in arterial tissue, involving the subintimal accumulation of LDL. Measurement of the rate at which LDL and other lipoproteins, such as HDL and VLDL, enter and exit the tissue can provide insight into the mechanisms involved in the development of atherosclerotic lesions. Permeation of VLDL, LDL, HDL, and glucose was measured for both normal and atherosclerotic human carotid endarterectomy tissues (CEA) at 20°C and 37°C using optical coherence tomography (OCT). The rates for LDL permeation through normal CEA tissue were (3.16 ± 0.37) × 10(-5) cm/s at 20°C and (4.77 ± 0.48) × 10(-5) cm/s at 37°C, significantly greater (P < 0.05) than the rates for atherosclerotic CEA tissue at these temperatures [(1.97 ± 0.34) × 10(-5) cm/s at 20°C and (2.01 ± 0.23) × 10(-5) cm/s at 37°C]. This study effectively used OCT to measure the rates at which naturally occurring lipoproteins enter both normal and diseased carotid intimal tissue.  相似文献   

14.
Water is an integral part of collagen's triple helical and higher order structure. Studies of model triple helical peptides have revealed the presence of repetitive intrachain, interchain, and intermolecular water bridges (Bella et al., Structure 1995, 15, 893-906). In addition, an extended cylinder of hydration is thought to be responsible for collagen fiber assembly. Confocal Raman spectroscopy and dynamic vapor sorption (DVS) measurements of human Type I collagen and pigskin dermis were performed to probe relative humidity (RH)-dependent differences in the nature and level of collagen hydration. Raman spectra were also acquired as a function of time for both Type I collagen and pigskin dermis samples upon exchange of a 100% RH H(2) O to deuterium oxide (D(2) O) environment. Alterations in Amide I and III modes were consistent with anticipated changes in hydrogen bonding strength as RH increased and upon H → D exchange. Of note is the identification of a Raman spectral marker (band at 938 cm(-1) ) which appears to be sensitive to alterations in collagen-bound water. Analysis of DVS isotherms provided a quantitative measure of adsorbed and absorbed water vapor consistent with the Raman results. The development of a Raman spectral marker of collagen hydration in intact tissue is relevant to diverse fields of study ranging from the evaluation of therapeutics for wound healing to hydration of aging skin.  相似文献   

15.
The Raman spectrum of chemically denatured lysozyme was studied. The denaturants studied included dimethyl sulfoxide, LiBr, guanidine · HCl, sodium dodecyl sulfate, and urea. Previous studies have shown that the amide I and amide III regions of the Raman spectrum are sensitive to the nature of the hydrogen bond involving the amide group. The intensity of the amide III band at 1260 cm?1 (assigned to strongly hydrogen-bonded α-helix structure) relative to the intensity of the amide III band near 1240 cm?1 (assigned to less strongly hydrogen-bonded groups) is used as a parameter for comparison with other physical parameters used to assess denaturation. The correlation between this Raman parameter and denaturation as evidenced by enzyme activity and viscosity measurements is good, leading to the conclusion that the amide III Raman spectrum is useful for assessing the degree of denaturation. The Raman spectrum clearly depends on the type of denaturant employed, suggesting that there is not one unique denatured state for lysozyme. The data, as interpreted, place constraints on the possible models for lysozyme denaturation. One of these is that the simple two-state model does not seem consistent with the observed Raman spectral changes.  相似文献   

16.
Time-resolved Fourier transform infrared difference spectra of the phosphoenzyme conversion and Ca(2+) release reaction (Ca(2)E(1)-P --> E(2)-P) of the sarcoplasmic reticulum Ca(2+)-ATPase were recorded at pH 7 and 1 degrees C in H(2)O and (2)H(2)O. In the amide I spectral region, the spectra indicate backbone conformational changes preserving conformational changes of the preceding phosphorylation step. beta-sheet or turn structures (band at 1685 cm(-1)) and alpha-helical structures (band at 1653 cm(-1)) seem to be involved. Spectra of the model compound EDTA for Ca(2+) chelation indicate the assignment of bands at 1570, 1554, 1411 and 1399 cm(-1) to Ca(2+) chelating Asp and Glu carboxylate groups partially shielded from the aqueous environment. In addition, an E(2)-P band at 1638 cm(-1) has been tentatively assigned to a carboxylate group in a special environment. A Tyr residue seems to be involved in the reaction (band at 1517 cm(-1) in H(2)O and 1515 cm(-1) in (2)H(2)O). A band at 1192 cm(-1) was shown by isotopic replacement in the gamma-phosphate of ATP to originate from the E(2)-P phosphate group. This is a clear indication that the immediate environment of the phosphoenzyme phosphate group changes in the conversion reaction, altering phosphate geometry and/or electron distribution.  相似文献   

17.
Laser Raman studies of conformational variations of poly-L-lysine   总被引:4,自引:0,他引:4  
The frequencies and intensities of the laser Raman spectra of poly-L -lysine (PLL) have been observed in the following studies: (1) the thermally induced α-to-β transition which occurs with increasing temperature at high pH; (2) the ionized form to α transition at 10°C by increasing pH; and (3) the ionized form to α transition by ionic strength at low pH. The frequency-dependent bands which have been observed are the amide I (in H2O), amide I′ (in D2O), amide III, and C–C stretch. It has been found possible to assign an unique set of frequencies and intensities to each conformation of PLL of α, β, and ionized form. In this way the nature of the conformations intermediate in the transitions can be determined. The frequencies of the amide III and amide III′ are very weak in the α-helix and somewhat higher than usual in the β form. Hence it appears the amide III and amide III′ bands may differ from one type of polypeptide to another with the same backbone conformation.  相似文献   

18.
The formation of closed icosahedral capsids from a single species of coat protein subunit requires that the subunits assume different conformations at different lattice positions. In the double-stranded DNA bacteriophage P22, formation of correctly dimensioned capsids is mediated by interaction between coat protein subunits and scaffolding protein. Raman spectroscopy has been employed to compare the conformations of coat protein subunits which have been polymerized to form capsids in the presence and absence of the of scaffolding protein display a Raman spectrum characterized by a broad amide I band centered at 1665 cm-1 with a discernible shoulder near 1653 cm-1, and a broad amide III profile centered at 1238 cm-1 but asymmetrically skewed to higher frequency. These spectral features indicate that the protein conformation in procapsid shells is rich in beta-sheet secondary structure but contains also a significant distribution of alpha-helix. When biologically active, purified subunits assemble in the absence of scaffolding protein, they form polydisperse multimers lacking the proper dimensions of procapsid closed shells. We designate these multimers as "associated subunits" (AS). The Raman spectrum of associated subunits indicates a narrower distribution of secondary structure. The associated subunits are characterized by a sharper and more intense Raman amide I band at 1666 cm-1, with no prominent amide I shoulder of lower frequency. An analogous narrowing of the Raman amide III profile is also observed for AS particles, with an accompanying shift of the amide III band center to 1235 cm-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
The kinetic and thermodynamic parameters of wheat β-amylase (WBA) were characterized and various additives were evaluated for enhancing its activity and thermostability. WBA activity was examined by neocuproine method using soluble starch as substrate. The Michaelis constant (K(m)) and molecular activity (k(cat)) were determined to be 1.0±0.1% (w/v) and 94±3s(-1), respectively, at pH 5.4 and at 25°C. The optimum reaction temperature (T(opt)) for WBA activity was 55°C and the temperature (T(50)) at which it loses half of the activity after 30-min incubation was 50±1°C. Modifications of the solvent with 182mM glycine and 0.18% (w/v) gelatin have increased the T(50) by 5°C. Glycerol, ethylene glycol, dimethylformamide (DMF) and dimethyl sulfoxide have also slightly enhanced the thermostability plausibly through weakening the water structure and decreasing the water shell around the WBA protein. Ethanol and DMF activated WBA by up to 24% at 25°C probably by inducing favorable conformation for the active site or changing the substrate structure by weakening the hydrogen bonding. Its half-life in the inactivation at 55°C was improved from 23 to 48min by 182mM glycine. The thermodynamic parameters indicate that WBA is thermo-labile and sufficient stabilization was achieved through solvent modification with additives and that the heat inactivation of WBA is entropic-driven. It is suggested that WBA could be applied more widely in starch-saccharification industries with employing suitable additives.  相似文献   

20.
Reactivity of the dicopper(I) and dicopper(II) complexes supported by novel polyamine ligands L1 (1,11-bis(6-methylpyridin-2-yl)-2,6,10-triaza-2,6,10-tribenzylundecane) and L2 (5-benzyl-1,9-bis(6-methylpyridin-2-yl)-2,8-bis(6-methylpyridin-2-ylmethyl)-2,5,8-triazanonane) towards O(2) and H(2)O(2), respectively, has been investigated in order to shed light on the ligand effects on Cu(2)/O(2) chemistry. The dicopper(I) complex of L1 (1a) readily reacted with O(2) in a 2:1 ratio at a low temperature (-94 degrees C) in acetone to afford a mixture of (mu-eta2.eta2-peroxo)dicopper(II) and bis(mu-oxo)dicopper(III) complexes. The formation of these species has been confirmed by the electron spin resonance (ESR) silence of the solution as well as their characteristic absorption bands in the UV-visible region (gammamax= 350 and 510 nm due to the peroxo complex and approximately 400 nm due to the bis(mu-oxo) complex] and the resonance Raman bands at 729 cm(-1) [Deltanu (16(O2)-18(O2)) = 38 cm(-1)] due to the peroxo complex and at 611 and 571 cm(-1) [Deltanu(16(O2)-18(O2)) = 22 and 7 cm(-1), respectively] due to the bis(mu-oxo) complex. The peroxo and bis(mu-oxo) complexes were unstable even at the low temperature, leading to oxidative N-dealkylation at the ligand framework. The dicopper(I) complex of L2 (2a) also reacted with O(2) to give (mu-hydroxo)dicopper(II) complex (2b(OH)) as the product. In this case, however, no active oxygen intermediate was detected even at the low temperature (-94 degrees C). With respect to the copper(II) complexes, treatment of the (mu-hydroxo)dicopper(II) complex of L1 (1b(OH)) with an equimolar amount of H(2)O(2) in acetone at -80 degrees C efficiently gave a (mu-1,1-hydroperoxo)dicopper(II) complex, the formation of which has been supported by its ESR-silence as well as UV-vis (370 and 650 nm) and resonance Raman spectra [881 cm(-1); [Deltanu (16(O2)-18(O2)) = 49 cm(-1)]. The (mu-1,1-hydroperoxo)dicopper(II) intermediate of L1 also decomposed slowly at the low temperature to give similar oxidative N-dealkylation products. Kinetic studies on the oxidative N-dealkylation reactions have been performed to provide insight into the reactivity of the active oxygen intermediates.  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