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 62 毫秒
1.
The iron-containing B2 subunit of ribonucleotide reductase from Escherichia coli has been investigated by Raman spectroscopy. Both the tyrosyl radical-containing native protein and the radical-free protein exhibit a resonance-enhanced Raman band at 500 cm?1. This band is assigned to an Fe-O vibrational mode arising from an oxygen-containing ligand. The failure to observe any tyrosinate ring modes makes it unlikely that ribonucleotide reductase is an iron-tyrosinate protein and rules out tyrosinate oxygen as a ligand. It is proposed that the 500 cm?1 band in ribonucleotide reductase is analogous to the 510 cm?1 Fe-O vibrational mode of methemerythrin and arises from an oxo- or carboxylate-bridge between the antiferromagnetically-coupled Fe(III) ions.  相似文献   

2.
Raman spectra were measured for poly(L -histidine) in H2O, poly(L -histidine-d2 and -d3) in D2O, L -histidine in H2O, L -histidine-d3 (and d4) in D2O, and 4-methylimidazole in H2O with various pH (or pD) values. The Raman scattering peaks observed for these samples were ascribed to the neutral and positively charged imidazole groups on the basis of the spectral changes due to the pH variation and to the deuterium substitution of the imino protons. The vibrational modes of these peaks were deduced from the normal coordinate analysis made on the positively charged and neutral 4-ethylimidazoles. The Raman scattering peaks from the imidazole groups in the neutral form clearly indicate that these imidazole groups exist in the equilibrium between the two tautomeric forms, the 1-N protonated from (tautomer I) and the 3-N protonated one (tautomer II). For example, the breathing vibration of the 1-N protonated form is observed at 1282 cm?1 for L -histidine and at 1304 cm?1 for 4-methylimidazole, while the breathing vibration of the 3-N protonated form is observed at 1260 cm?1 for L -histidine and 4-methylimidazole. From the temperature dependence of the relative intensities of the tautomer I peak to that of the tautomer II, it was concluded that the tautomer I is energetically more stable than the tautomer II, and the ΔH value is 1.0 ± 0.3 kcal/mol for L -histidine and 0.4 ± 0.1 kcal/mol for 4-methylimidazole. Poly(L -histidine) with the neutral imidazole side chains shows the amide I peak at 1672 cm?1, indicating that the sample assumes the antiparallel pleated-sheet structure. Poly(L -Ala75L -His25) and poly(L -Ala50L -His50) were found to take the α-helical and β-form conformations, respectively.  相似文献   

3.
The reduction of methemerythrin (Hr+) by dithionite produces deoxyhemerythrin (Hro) in multi, possibly three, stages. The kinetics were examined at pH 8·2 and 25 °C. The first stage is reduction of methemerythrin to an intermediate A by SO2- (k = 1.3 × 105m?1s?1). The much slower second and third stages have rates independent of dithionite concentrations. Reaction is completed after about 10 h. The kinetics of reactions of A with N3-, H2O2, and O2 were examined, as well as the conversion of A to intermediate B (k = 4·4 × 10?4s?1). It is concluded that A is an (Fe(II)Fe(III))8 species, and that in B the unit (Fe(II)Fe(II))8 is well developed, judging by its unreactivity towards N3?, its reaction with H2O2, and its reversible uptake of O2 (85–90% of the final product). There is little effect of adjusting the pH to 6·3 on the rates of the processes examined.  相似文献   

4.
Coherent anti-Stokes Raman scattering spectra, in resonance with the isoalloxazine visible electronic transition, have been obtained down to 300 cm?1 for flavin adenine dinucleotide, riboflavin binding protein and glucose oxidase, in H2O and D2O. Several isoalloxazine vibrational modes can be identified by analogy with those of uracil. Of particular interest is a band at ~1255 cm?1 in H2O, which is replaced by another at ~1295 cm?1, in D2O. The H2O band appears to be a sensitive monitor of H-bonding of the N3 isoalloxazine proton to a protein acceptor group. It shifts down by 10 cm?1 in riboflavin binding protein, and disappears altogether in glucose oxidase. Other band shifts, of 3–5 cm?1, are similar for the two flavoproteins, and may reflect environmental changes between aqueous solution and the protein binding pockets.  相似文献   

5.
The vibrational Raman optical activity (ROA) spectra of di- and tri-L -alanine in the range 650–1750 cm?1 have been measured in H2O and D2O solution at high, neutral, and low pH and pD. Corresponding ROA spectra for tetra- and penta-L -alanine have also been obtained, but over a more restricted set of pH and pD conditions. There are similarities with the ROA spectrum of L -alanine below ~ 1200 cm?1, but the spectra are very different above this wavenumber due to the influence of the vibrational coordinates of the peptide group. The similar overall appearance of the di-, tri-, and tetrapeptide ROA under selected conditions of pH and pD, and of all four peptide ROA spectra in DCl and HCl solutions, in the backbone skeletal stretch region ~ 1050–1200 cm?1 and the extended amide III region ~ 1250–1350 cm?1, suggests that the backbone conformation is approximately the same in all four structures. One difference, however, is a shift of a large positive ROA band in H2O at ~ 1341 cm?1 in the dipeptide, assigned to Cα–H and in-plane N–H deformations, down to ~ 1331 cm ?1 in the tripeptide and to ~ 1315 cm?1 in the tetrapeptide and pentapeptide (the last in HCl due to insufficient solubility in H2O), which indicates increasing delocalization of the corresponding normal mode with increasing chain length. Our results do not support the suggestion that stabilizing interactions of the zwitterionic end groups in tri-L -alanine at neutral pH leads to a different solution structure to that at high pH. © 1994 John Wiley & Sons, Inc.  相似文献   

6.
The pH dependence for the interconversion of the acid and base forms of methemerythrin from Themiste dyscritum was investigated by difference spectroscopy. A new technique was designed to be able to study mixtures without knowledge of extinction coefficients or exact protein concentrations. The resultant pKa value of 8.4 proved that T. dyscritum hemerythrin crystals used for previous X-ray crystallographic studies at pH less than or equal to 6.5 were in the acid form. Since this material contains a 5-coordinate iron atom with no evidence of a ligated water molecule, it is more appropriately referred to as methemerythrin than aquomethemerythrin. The presence of an iron-bound hydroxide in the base form of methemerythrin was verified by resonance Raman spectroscopy for both T. dyscritum and Phascolopsis gouldii. At pH greater than 9, the protein from either species exhibited a new feature at 490 cm-1 that shifted to 518 cm-1 in D2O and was assigned to a coupled Fe-OH stretching and O-H bending vibration. Thus, hydroxomethemerythrin is the correct designation for the base form of the protein. The other resonance-enhanced vibration, the Fe-O-Fe symmetric stretch, was observed at 506 cm-1 in hydroxomethemerythrin and at 511 cm-1 in methemerythrin and was unaffected by deuteration. Addition of perchlorate to methemerythrin had no effect on the Raman spectrum, despite its known role in stabilizing the met form relative to the hydroxomet form.  相似文献   

7.
Thermally induced unfolding of ribonuclease at low pH produces a large increase of Raman scattering intensity at ~ 2930cm?1 (CH3-stretching region). Comparisons of the CH3-stretching spectra of various model compounds in the presence or absence of H2O (2H2O), indicate that the changes observed with ribonuclease arise from the insertion of previously buried, aliphatic amino acid residues into water.  相似文献   

8.
Raman spectra are presented for sarcoplasmic reticulum membranes. Interpretation of the 1000–1130 cm?1 region of the spectrum indicates that the sarcoplasmic reticulum membrane may be more fluid than erythrocyte membranes that have been examined by the same technique. The fluidity of the membrane also manifests itself in the amide I portion of the membrane spectrum with a strong 1658 cm?1 band characteristic of CC stretching in hydrocarbon side chains exhibiting cis conformation. This band is unaltered in intensity and position in H2O and in 2H2O thus obscuring amide I protein conformation. Of particular interest is the appearance of strong, resonantly enhanced bands at 1160 and 1527 cm?1 attributable to membrane-associated carotenoids.  相似文献   

9.
Depolarization ratios of Raman bands, excited at 488.0 nm, of guanosine-5′-monophosphoric 4 acid, cytidine-5′-monophosphoric acid, adenosine-5′-monophosphoric acid, thymidine-5′-monophosphoric acid, and uridine-5′-monophosphoric acid have been measured in their H2O and D2O solutions in the spectral region from 300 to 1800 cm?1. For comparison, the disodium salt of 2′-deoxyadenosine-5′-monophosphoric acid was also subjected to the depolarization measurement in its H2O solution. The results have been correlated with possible orientations of the principal axes of the Raman scattering tensors as well as with the relative magnitudes of the tensor components. Results should be useful for future polarized Raman studies of synthetic and natural DNA. © 1993 John Wiley & Sons, Inc.  相似文献   

10.
The complex of sulfide and methemerythrin has been characterized by resonance Raman spectroscopy. At pH 8.0 the complex contains two irons and one S2- at the active site. The resonance Raman spectrum of the sulfidomethemerythrin complex contains only one vibration, at 444 cm-1. This vibration is assigned to an iron-sulfide stretch. The possibility that sulfidomethemerythrin contains a mu-sulfido bridge. FeIII-S2-FeIII, analogous to the proposed mu-oxo bridge in azidomethemerythrin is discussed.  相似文献   

11.
Raman studies of nucleic acids. VII. Poly A-poly U and poly G-poly C   总被引:6,自引:0,他引:6  
L Lafleur  J Rice  G J Thomas 《Biopolymers》1972,11(12):2423-2437
Laser-excited Raman spectra of the double-helical complexes poly A·poly U and poly G·poly C are reported for 2H2O and H2O solutions. The spectra are discussed in relation to their use as quantitative reference spectra for determining the dependence of the Raman scattering of RNA on secondary structure. The Raman line at 815 cm?1, due to the phosphodiester group, exhibits the same intrinsic intensity in spectra of poly A·poly U and poly G·poly C and is thus dependent only upon the amount of ordering of the helix and not on the kinds of nucleotides involved. The hypochromic Raman lines in spectra of poly A·poly U are identified and their intensity changes are determined quantitatively over the temperature range 32–85°C. Comparison of the spectra in the 1500–1750 cm?1 region reveals that the Raman lines from carbonyl group vibrations of uracil are about sevenfold more intense than those of guanine and cytosine for both paired and unpaired states and will thus dominate the spectra of RNA. The Raman frequencies in this region are also compared with previously reported infrared frequencies and give evidence of being strongly perturbed by base-stacking interactions in the helices.  相似文献   

12.
《Carbohydrate research》1987,163(1):9-14
Laser-Raman spectra of Na+ kappa-carrageenan, Na+ neocarrabiose 4-sulphate, and neocarrabiose in the region 700–1500 cm−1 are reported for solutions in H2O and D2O. The C-1-H-1α vibration, coupled with COH related modes, is assigned to a band at 840 cm−1, close to the maximum of the symmetrical COS stretching (∼850 cm−1). The symmetrical SO stretch is proposed to occur near 1040 cm−1 and is probably coupled with COH vibrations which give rise to strong bands in the region 1000–1100 cm−1. The intense band in the region 730–740 cm−1 is ascribed to a complex ring vibration.  相似文献   

13.
Abstract

The polarized Raman spectra of a single crystal of the barium salt of inosine monophosphoric acid hexahydratc (Ba-IMP.6H2O) have been observed with 488.0 nm excitation. For each Raman band, the relative intensities of aa, bb, cc, ab and bc tensor components have been determined. The tensor quotients from the crystal were augmented with measured depolarization ratios in solution. From these experimental data, the shape and orientation of the localized Raman scattering tensor were deduced for each of the normal modes of the hypoxanthine residue, phosphate moiety and ribose portion. The hypoxanthine residue gives a strong Raman band around 1553 cm?1, which shows rather large depolarization ratio, p = 0.32, in aqueous solution, and shows a great scattering anisotropy in the single crystal of IMP. The shape and orientation of the Raman tensor associated to this 1553 cm?1 vibration have been determined: one of its principal axes (y-axis) is directed along the long axis (N1-N7) of the hypoxanthine residue and the relative magnitudes of its components are given as r 1 = αxxzz = ?1, r 2 = αyyzz = 12. Next, a general relation has been shown between the orientation angles (θ and χ) of such a local Raman tensor in a uniaxial biological fiber and the anisotropy of Raman scattering intensities from the fiber. By the use of this relation, a discussion has been made of the orientation of the hypoxanthine residue in a poly(r1). poly(rC) duplex fiber.

  相似文献   

14.
This paper reports the first Raman spectroscopic study of the potassium complex of the cation-specific antibiotic valinomycin. Complete Raman spectra (140 to 3600 cm?1) of crystalline valinomycin-KSCN and its CCl4, CHCl3 and C2H5OH solutions are presented and used to probe the structure of the complex in these environments. In all cases a single, narrow peak is observed in the ester CO stretch region (1750 to 1775 cm?1) which contrasts strongly with the broad bands observed in solutions of uncomplexed valinomycin. This is consistent with the presence of a single conformation in which all six ester CO groups co-ordinate an enclosed potassium ion. We find that although the ester CO stretch frequencies of the complex are similar in the solid state and in non-polar solution (~1770 cm?1) they are considerably different in the presence of polar solvents (~1756 cm?1); this may indicate that the complexed potassium ion is still free to interact with nearby solvent ions (and possibly its counterion) through gaps in the hydrophobic “shield” provided by the hydrocarbon residues of valinomycin. In contrast the amide CO frequencies of the complex (~1650 cm?1) are solvent-independent. These groups are apparently strongly hydrogen-bonded to provide a rather rigid, compact framework for the complex conformation.  相似文献   

15.
The Raman spectra of guanylyl (3′-5′) guanosine (GpG) in solution in H2O and D2O at pH 3–7 have been recorded at various temperatures between 0 and 80°C. The results are consistent with the existence in the lower temperature range of stable aggregates formed by the stacking of GpG tetramers. The aggregates melt cooperatively near 60°C, which results in important changes in the spectra. Among these, a large increase in intensity of some of the bands assigned to the guanine residues shows that unstacking of the bases occurs at the melting. Also apparent in the spectra are changes in the intensity and frequency of band attributable to molecular groups involved in intermolecular hydrogen bonding between adjacent molecules in the complex. The melting temperature of GpG decreases by approximately 15°C upon lowering the concentration from 5 × 10?2 to 5 × 10?4M, as shown by Raman, calorimetric, CD, and uv measurements. The experimentally determined ΔH and ΔS for the melting transition are 9 Kcal/mol and 28 e.u./mol, respectively. The aggregation of GpG in 1.5 × 10?3M solutions was found to be very slow. The half-time of the process, which roughly follows first-order kinetics, is approximately 3 min at 10°C and 21 min at 35°C. The negative energy of activation associated with this reaction (?143 Kcal) indicated that the process involves intermediates whose concentrations decrease the temperatures raised, thus slowing down the overall process. The rate of disaggregation of GpG upon dilution to very low concentration is also extremely slow, indicating that the GpG aggregates, once formed, are very stable.  相似文献   

16.
The efficacy of various chemicals as alternatives to antibiotics for the control of fireblight (Erwinia amylovora) on pear trees was tested. The chemicals were applied in two ways. In 1999 and 2000, preselected pear twigs (80–90% bloom stage) were sprayed once either preventively 1 day before inoculation or curatively one or three days after artificial inoculation with pathogen concentrations of 105 and 107cfu ml?1. In 2000 and 2001, whole trees were sprayed 2 and 4 days before artificial inoculation of the flowers. From the incidence of diseased flowers it appeared that Bion (50% benzothiadiazole) at 0.2 g litre?1 H2O and Aliette (80% fosetyl‐Al) at 2.5 g litre?1 H2O showed considerable preventive action by eliciting systemic acquired resistance mostly when they were applied in the whole trees. However the best control was achieved with the antibiotic Agrept (20% streptomycin) at 0.5 g litre?1 H2O. This showed both preventive and curative action. Kocide (77% copper hydroxide) at 0.9 g litre?1 H2O, Dentamet (citric acid in chelate) at 1.5 ml litre?1 H2O, Bactosan (an extract from the plant Pongamia pinnatd) at 3.0 g litre?1 H2O and Bion at 0.1 g litre?1 H2O, showed preventive action, but only when the inoculum concentration was low.  相似文献   

17.
Resonance Raman spectra have been obtained for Compound II of horseradish peroxidase. Its prophyrin vibrational frequencies are consistent with a planar low-spin heme containing Fe(IV). The oxidation-state marker band is found at the unprecedentedly high value of 1382 cm?1. This band was also observed in solutions of myoglobin and cytochrome c peroxidase to which H2O2 had been added. No evidence was found for an actual FeO double bond in Compound II.  相似文献   

18.
The 14N nuclear relaxation times T1 and T2 in egg yolk phosphatidylcholine have been observed in single bilayer vesicles dispersed in the media of different viscosities, 1H2O and 2H2O. The lateral diffusion coefficient of lipid molecule D has been calculated according to the method reported earlier: D = 2.2 × 10?8cm2s?1 in 1H2O and 2.1 × 10?8cm2s?1 in 2H2O at 20°C. They are in excellent agreement. This result gives a strong basis of usefulness of 14N NMR method in the evaluation of D without introducing any system perturbation.  相似文献   

19.
Kinetic studies of the reduction of ferrioxamine B (Fe(Hdesf)+) by Cr(H2O)62+, V(H2O)62+, and dithionite have been performed. For Cr(H2O)62+ and V(H2O)62+, the rate is ?d[Fe(Hdesf)+]/dt = k[Fe(Hdesf)+][M2+]. For Cr(H2O)62+, k = 1.19 × 104 M?1 sec?1 at 25°C and μ = 0.4 M, and k is independent of pH from 2.6 to 3.5. For V(H2O)62+, k = 6.30 × 102 M?1 sec?1 at 25°C, μ = 1.0 M, and pH = 2.2. The rate is nearly independent of pH from 2.2 to 4.0. For Cr(H2O)62+ and V(H2O)62+, the activation parameters are ΔH = 8.2 kcal mol?1, ΔS ?12 eu and ΔH = 1.7 kcal mol?1, ΔS = ?40 eu (at pH 2.2) respectively. Reduction by Cr(H2O)62+ is inner-sphere, while reduction by V(H2O)62+ is outer-sphere. Reduction by dithionite follows the rate law ?d[Fe(Hdesf)+]/dt =kK12[Fe(Hdesf)+][S2O42?]12 where K is the equilibrium constant for dissociation of S2O42? into SO2? radicals. The value of k at 25°C and μ = 0.5 is 2.7 × 103 M?1 sec?1 at pH 5.8, 3.5 × 103 M?1 sec?1 at pH 6.8, and 4.6 × 103 M?1 sec?1 at pH 7.8, and ΔH = 6.8 kcal mol?1 and ΔS = ?19 eu at pH 7.8.  相似文献   

20.
Infrared and Raman spectra of the Mg2+ salt of poly(U) in D2O were recorded in the 1600-1800 cm?1 region and between 1 and 20C. The ir spectra showed a melting curve similar to the uv melting curves with a temperature of transition of about 6.5°C. This spectral change is assumed to be associated with the formation of the secondary structure of Mg2+-poly(U) in D2O at this temperature. Three double-helical and two triple-helical structures were used as inputs to compute the normal modes of vibration. A double-helical structure was found to give the best agreement with the observations. Knowledge of the C=0 eigenvectors, and of the expression for transition probability from quantum mechanics, was used to explain the so far unanswered question of H. T. Miles [(1964) Proc. Natl. Acad. Sci. USA 51, 1104–1109; (1980) Biomolecular Structure, Conformation, Function and Evolution, Pergamon, Oxford, pp. 251–264] as to why there is an increase in the ir vibrational wave number of a carbonyl band when that group is H-bonded to another polynucleotide chain in a helix. Such considerations also explain why a predicted band at about 1648 cm?1 is not to be seen in the ir spectra but is present in the Raman spectra. The model incorporating the C?O transition dipole-dipole coupling interaction is able to explain also the observed higher intensity of the higher wave-number ir band. The experimental results demonstrate that the complete picture of vibrational dynamics of Mg2+-poly(U) in D2O is obtained only by looking simultaneously at ir and Raman spectra and not at only one of them. Weak ir bands were found to be as useful as the strong ones in understanding structure and vibrational dynamics. On the bases of our ir and Raman spectra, of the normal-mode analyses, and of the literature data, it is concluded that Mg2+-poly(U) in D2O is present in a double-helical structure at temperatures below the temperature of transition, whereby the uracil residues are paired according to arrangement (a) (see Fig. 1). This structure is rodlike and arises by refolding of one poly(U) chain. The computations show that no normal mode is associated with a single C?O group vibration; all C?O group vibrations are heavily mixed motions of various C?O groups.  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