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.
Certain toxic elements support the metabolism of diverse prokaryotes by serving as respiratory electron acceptors for growth. Here, we demonstrate that two anaerobes previously shown to be capable of respiring oxyanions of selenium also achieve growth by reduction of either tellurate [Te(VI)] or tellurite [Te(IV)] to elemental tellurium [Te(0)]. This reduction achieves a sizeable stable-Te-isotopic fractionation (isotopic enrichment factor [] = −0.4 to −1.0 per ml per atomic mass unit) and results in the formation of unique crystalline Te(0) nanoarchitectures as end products. The Te(0) crystals occur internally within but mainly externally from the cells, and each microorganism forms a distinctly different structure. Those formed by Bacillus selenitireducens initially are nanorods (~10-nm diameter by 200-nm length), which cluster together, forming larger (~1,000-nm) rosettes composed of numerous individual shards (~100-nm width by 1,000-nm length). In contrast, Sulfurospirillum barnesii forms extremely small, irregularly shaped nanospheres (diameter < 50 nm) that coalesce into larger composite aggregates. Energy-dispersive X-ray spectroscopy and selected area electron diffraction indicate that both biominerals are composed entirely of Te and are crystalline, while Raman spectroscopy confirms that they are in the elemental state. These Te biominerals have specific spectral signatures (UV-visible light, Raman) that also provide clues to their internal structures. The use of microorganisms to generate Te nanomaterials may be an alternative for bench-scale syntheses. Additionally, they may also generate products with unique properties unattainable by conventional physical/chemical methods.  相似文献   

2.
The self-complementary oligonucleotides [r(CGC)d(CGC)]2 and [d(CCCCGGGG)]2 in single-crystal and solution forms have been investigated by Raman spectroscopy. Comparison of the Raman spectra with results of single-crystal X-ray diffraction and with data from polynucleotides permits the identification of a number of Raman frequencies diagnostic of the A-helix structure for GC sequences. The guanine ring frequency characteristic of C3'-endo pucker and anti base orientation is assigned at 668 +/- 2 cm-1 for both dG and rG residues of the DNA/RNA hybrid [r(GCG)d(CGC)]2. The A-helix backbone of crystalline [r(GCG)d(CGC)]2 is altered slightly in the aqueous structure, consistent with the conversion of at least two residues to the C2'-endo/anti conformation. For crystalline [d(CCCCGGGG)]2, the Raman and X-ray data indicate nucleosides of alternating 2'-endo-3'-endo pucker sandwiched between terminal and penultimate pairs of C3'-endo pucker. The A-A-B-A-B-A-A-A backbone of the crystalline octamer is converted completely to a B-DNA fragment in aqueous solution with Raman markers characteristic of C2'-endo/anti-G (682 +/- 2) and the B backbone (826 +/- 2 cm-1). In the case of poly(dG).poly(dC), considerable structural variability is detected. A 4% solution of the duplex is largely A DNA, but a 2% solution is predominantly B DNA. On the other hand, an oriented fiber drawn at 75% relative humidity reveals Raman markers characteristic of both A DNA and a modified B DNA, not unlike the [d-(CCCCGGGG)]2 crystal. A comparison of Raman and CD spectra of the aqueous [d(CCCCGGGG)]2 and poly(dG).poly(dC) structures suggests the need for caution in the interpretation of CD data from G clusters in DNA.  相似文献   

3.
Certain anaerobic bacteria respire toxic selenium oxyanions and in doing so produce extracellular accumulations of elemental selenium [Se(0)]. We examined three physiologically and phylogenetically diverse species of selenate- and selenite-respiring bacteria, Sulfurospirillum barnesii, Bacillus selenitireducens, and Selenihalanaerobacter shriftii, for the occurrence of this phenomenon. When grown with selenium oxyanions as the electron acceptor, all of these organisms formed extracellular granules consisting of stable, uniform nanospheres (diameter, approximately 300 nm) of Se(0) having monoclinic crystalline structures. Intracellular packets of Se(0) were also noted. The number of intracellular Se(0) packets could be reduced by first growing cells with nitrate as the electron acceptor and then adding selenite ions to washed suspensions of the nitrate-grown cells. This resulted in the formation of primarily extracellular Se nanospheres. After harvesting and cleansing of cellular debris, we observed large differences in the optical properties (UV-visible absorption and Raman spectra) of purified extracellular nanospheres produced in this manner by the three different bacterial species. The spectral properties in turn differed substantially from those of amorphous Se(0) formed by chemical oxidation of H(2)Se and of black, vitreous Se(0) formed chemically by reduction of selenite with ascorbate. The microbial synthesis of Se(0) nanospheres results in unique, complex, compacted nanostructural arrangements of Se atoms. These arrangements probably reflect a diversity of enzymes involved in the dissimilatory reduction that are subtly different in different microbes. Remarkably, these conditions cannot be achieved by current methods of chemical synthesis.  相似文献   

4.
Intravenous iron therapy is used to treat anemia associated with chronic kidney disease. The chemical structures of parenteral iron agents have not been characterized in detail, and correlations between structure, efficiency of iron delivery, and toxicity via catalysis of oxygen-derived free radical creation remain to be established. In this study, two formulations of parenteral iron have been characterized by absorption spectroscopy, X-ray diffraction analysis (XRD), transmission electron microscopy (TEM), atomic force microscopy (AFM), and elemental analysis. The samples studied were Venofer (Iron Sucrose Injection, USP) and Ferrlecit (Sodium Ferric Gluconate in Sucrose Injection). The 250-800-nm absorption spectra and the XRD patterns showed that both formulations contain a mineral core composed of iron oxyhydroxide in the beta-FeOOH mineral polymorph known as akaganeite. This was further confirmed for each formulation by imaging using TEM and AFM. The average core size for the nanoparticles, after dialysis to remove unbound or loosely bound carbohydrate, was approximately 3+/-2 nm for the iron-sucrose, and approximately 2+/-1 nm for the iron-gluconate. Each of the nanoparticles consists of a mineral core, surrounded by a layer of bound carbohydrate. The overall diameter of the average bead in the dialyzed preparations was approximately 7+/-4 nm for the iron-sucrose, and 3+/-1 nm for the iron-gluconate. Undialyzed preparations have particles with larger average sizes, depending on the extent of dilution of unbound and loosely bound carbohydrate. At a dilution corresponding to a final Fe concentration of 5 mg/mL, the average particle diameter in the iron-sucrose formulation was approximately 22+/-9 nm, whereas that of the iron-gluconate formulation was approximately 12+/-5 nm.  相似文献   

5.
Silicon nanonowires (SiNWs) were synthesized with approximately 100 nm/min by very-high-frequency plasma-enhanced chemical vapor deposition method via a vapor–liquid–solid mechanism. Pure silane gas was deposited on Au-coated silicon substrates at various radio frequency (RF) powers varied from 5 to 20 W, keeping other deposition parameters constant. Synthesized SiNWs were needle-like shape with an ultra-sharp tip diameter of about 10 nm and a length about 3 μm, which differs from the standard cylindrical nanowire produced by similar techniques. Furthermore, it was shown that the average length of nanoneedles augmented with the increase of RF power. High-resolution transition electron microscopy displayed that the nanowires were composed of a crystalline Si core with an amorphous oxide shell. Raman spectroscopy also revealed the presence of crystalline Si in the grown Si nanoneedles.  相似文献   

6.
Raman spectra of six synthetic polydeoxyribonucleotide duplexes with different base sequences have been examined in aqueous solutions with different salt or nucleotide concentrations. Detailed conformational differences have been indicated between B and Z forms of poly[d(G-C)] X poly[d(G-C)], between B forms of poly[d(G-C)] X poly[d(G-C)] and poly[d(G-m5C)] X poly[d(G-m5C)], between A and B forms of poly(dG) X poly(dC), between B and "CsF" forms of poly[d(A-T)] X poly[d(A-T)], between B forms of poly[d(A-U)] X poly[d(A-U)] and poly[d(A-T)] X poly[d(A-T)], and between low- and high-salt (CsF) forms of poly(dA) X poly(dT). The Raman spectrum of calf-thymus DNA in aqueous solution was also observed and was compared with the Raman spectra of its fibers in A, B, and C forms.  相似文献   

7.
The interaction of Escherichia coli RNA polymerase with poly[d(A-T)] and poly[d-(I-C)] was studied by difference absorption spectroscopy at temperatures, from 5 to 45 degrees C in the absence and presence of Mg2+. The effect of KCl concentration, at a fixed temperature, was studied from 12.5 to 400 mM. Difference absorption experiments permitted calculation of the extent of DNA opening induced by RNA polymerase and estimation of the equilibrium constant associated with the isomerization from a closed to an open RNA polymerase-DNA complex. delta H0 and delta S0 for the closed-to-open transition with poly[d(A-T)] or poly[d(I-C)] complexed with RNA polymerase are significantly lower than the values associated with the helix-to-coil transition for the free polynucleotides. For the RNA polymerase complexes with poly[d(A-T)] and poly[d(I-C)] in 50 mM KCl, delta H0 approximately 15-16 kcal/mol (63-67 kJ/mol) and delta S0 approximately 50-57 cal/K per mol (209-239 J/K per mol). The presence of Mg2+ does not change these parameters appreciably for the RNA polymerase-poly[d(A-T)] complex, but for the RNA polymerase-poly[d(I-C)] complex in the presence of Mg2+, the delta H0 and delta S0 values are larger and temperature-dependent, with delta H0 approximately 22 kcal/mol (92 kJ/mol) and delta S0 approximately 72 cal/K per mol (approx. 300 J/K per mol) at 25 degrees C, and delta Cp0 approximately 2 kcal/K per mol (approx. 8.3 kJ/K per mol). The circular dichroism (CD) changes observed for helix opening induced by RNA polymerase are qualitatively consistent with the thermally induced changes observed for the free polynucleotides, supporting the difference absorption method. The salt-dependent studies indicate that two monovalent cations are released upon helix opening. For poly[d(A-T)], the temperature-dependence of enzyme activity correlates well with the helix opening, implying this step to be the rate-determining step. In the case of poly[d(I-C)], the same is not true, and so the rate-determining step must be a process subsequent to helix opening.  相似文献   

8.
The thermotropic properties of N-(alpha-hydroxyacyl)-sphingosine (CER[AS]) in dry and hydrated state were studied by means of X-ray powder diffraction and FT-Raman spectroscopy. The polymorphic states of the CER[AS]/water mixture (lamellar crystalline, lamellar hexagonal gel, liquid crystalline) depend on the thermal pre-treatment of the sample. Only by heating the CER[AS]/water mixture above the melting chain transition can the system be hydrated. At room temperature, both dry and hydrated states form lamellar structures, which differ in their repeat distance and packing of hydrocarbon chains. Above the melting chain transition, hydrated CER[AS] forms a liquid crystalline hexagonal phase, whereas anhydrous CER[AS] forms an isotropic liquid phase. The various phases of hydrated CER[AS] are distinguished on the basis of the corresponding Raman spectra.  相似文献   

9.
Mono Lake sediment slurries incubated with lactate and tellurite [Te(IV)] turned progressively black with time because of the precipitation of elemental tellurium [Te(0)]. An enrichment culture was established from these slurries that demonstrated Te(IV)-dependent growth. The enrichment was purified by picking isolated black colonies from lactate/Te(IV) agar plates, followed by repeated streaking and picking. The isolate, strain MLTeJB, grew in aqueous Te(IV)-medium if provided with a small amount of sterile solid phase material (e.g., agar plug; glass beads). Strain MLTeJB grew at high concentrations of Te(IV) (~8 mM) by oxidizing lactate to acetate plus formate, while reducing Te(IV) to Te(0). Other electron acceptors that were found to sustain growth were tellurate, selenate, selenite, arsenate, nitrate, nitrite, fumarate and oxygen. Notably, growth on arsenate, nitrate, nitrite and fumarate did not result in the accumulation of formate, implying that in these cases lactate was oxidized to acetate plus CO2. Strain MLTeJB is a low G + C Gram positive motile rod with pH, sodium, and temperature growth optima at 8.5–9.0, 0.5–1.5 M, and 40°C, respectively. The epithet Bacillus beveridgei strain MLTeJBT is proposed.  相似文献   

10.
Redox thermodynamic data provide a detailed insight into control of the reduction potential E degrees' of the [Fe(S-Cys)4] site in rubredoxin. Mutant forms were studied in which specific structural changes were made in both the primary and secondary coordination spheres. Those changes have been probed by resonance Raman spectroscopy. The decrease of approximately 200 mV in E degrees' observed for the [Fe(S-Cys)3(O-Ser)]-/2- couples in the surface ligand mutants C9S and C42S is essentially enthalpic in origin and associated with the substitution of ligand thiolate by ligand olate. However, the pH dependence of the potentials below characteristic pKa(red) approximately equals 7 is an entropic contribution, plausibly associated with increased conformational flexibility induced by a longer Fe(II)-O(H)-Ser bond in the reduced form. The presence of a second surface Ser ligand in the new double mutant protein C9S/C42S affects the enthalpic term primarily for pH>pKa(red) > or = 9.3, but for pHpKa approximately 9: [Fe(III)(S-Cys)3(OH)]- + e- --> [Fe(II)(S-Cys)3(OH)]2-. pH [Fe(II)(S-Cys)3(OH2)]-.  相似文献   

11.
BackgroundBiosynthesis of Te nanoparticles may occur in higher plants exposed to Te, as reported on microorganisms. However, unambiguous observations of the biogenic nanoparticles (BgNPs) of Te in plants are lacking. Hence, in this study, we investigated the formation of insoluble BgNPs of Te in garlic (Allium sativum) as a model plant.MethodWe performed elemental analysis based on inductively coupled plasma-mass spectrometry (ICP-MS) technique, and obtained Te concentration and distribution in various parts of garlic. In addition, insoluble Te particles were detected by fast time-resolved ICP-MS. Direct observation of the insoluble Te particle was also conducted by scanning electron microscope (SEM) and transmission electron microscope (TEM).ResultsA part of the roots and clove from Te-exposed garlic showed black coloration. Te concentrations in the black-colored parts were significantly increased compared with the non-colored parts. Transient signals of Te unique to nanoparticles were detected from the insoluble fractions of the black-colored parts. Finally, rod-shaped biogenic Te nanoparticles consisting of highly crystalline elemental Te was observed by SEM and TEM.ConclusionOur data provide new insights to the metabolic pathway of Te in higher plants for the formation of insoluble biogenic nanoparticles, which is extremely important for the detoxification of Te.  相似文献   

12.
Two gold(III) compounds [Au(TACN)Cl(2)]Cl (1) and [Au(TACN)Cl(2)][AuCl(4)] (2) (where TACN=1,4,7-triazacyclononane), have been synthesized and characterized by electrospray ionization mass spectrometry (ESI-MS), (1)H NMR spectroscopy and elemental analyses. The structure of compound 2 was determined by X-ray crystallography, in which TACN coordinates to the gold(III) center in a bidentate mode and the unbound amine group forms a very short intramolecular Au-H(-N) contact (1.91A). Biological activity data showed that compound 1 is more cytotoxic than cisplatin against A-549 and HCT-116 tumor cell lines. The interactions of compound 1 with CT-DNA were studied by UV-Vis, fluorescence and CD spectroscopy, which suggests that compound 1 can induce the distortion of DNA double helix.  相似文献   

13.
A microprobe system has been developed that can record Raman spectra from as little as 2 microL of solution containing only micrograms of biological pigments. The apparatus consists of a liquid nitrogen (l-N2)-cooled cold stage, an epi-illumination microscope, and a substractive-dispersion, double spectrograph coupled to a l-N2-cooled CCD detector. Experiments were performed on native bovine rhodopsin, rhodopsin expressed in COS cells, and four rhodopsin mutants: Glu134 replaced by Gln (E134Q), Glu122 replaced by Gln (E122Q), and Glu113 replaced by Gln (E113Q) or Ala (E113A). Resonance Raman spectra of photostationary steady-state mixtures of 11-cis-rhodopsin, 9-cis-isorhodopsin, and all-trans-bathorhodopsin at 77 K were recorded. The Raman spectra of E134Q and the wild-type are the same, indicating that Glu134 is not located near the chromophore. Substitution at Glu122 also does not affect the C = NH stretching vibration of the chromophore. The fingerprint and Schiff base regions of the Raman spectra of the 380-nm, pH 7 forms of E113Q and E113A are characteristic of unprotonated retinal Schiff bases. The C = NH modes of the approximately 500-nm, pH 5 forms of E113Q and E113A in H2O (D2O) are found at 1648 (1629) and 1645 (1630) cm-1, respectively. These frequencies indicate that the protonated Schiff base interacts more weakly with its protein counterion in the Glu113 mutants than it does in the native pigment. Furthermore, perturbations of the unique bathorhodopsin hydrogen out-of-plane (HOOP) vibrations in E113Q and E113A indicate that the strength of the protein perturbation near C12 is weakened compared to that in native bathorhodopsin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
G H Shimer  A R Wolfe  T Meehan 《Biochemistry》1988,27(20):7960-7966
We have investigated the equilibrium binding of racemic 7r,8t,9t,10c-tetrahydroxy-7,8,9,10-tetrahydrobenzo[a]pyrene to the double-stranded, synthetic polynucleotides poly[d(A-T)], poly[d(G-C)], and poly[d(G-m5C)] at low binding ratios. Difference absorption spectroscopy shows a 10-nm red shift for binding to poly[d(A-T)] and an 11-nm red shift for binding to either poly[d(G-C)] or poly[d(G-m5C)]. The value of delta epsilon for binding is approximately the same for all three hydrocarbon-polynucleotide complexes. Binding of this neutral polycyclic aromatic hydrocarbon derivative to these polynucleotides is dependent upon ionic strength and temperature. Analysis of complex formation employing polyelectrolyte theory shows a greater release of counterions associated with binding to poly[d(A-T)] than with the other two polynucleotides (0.5 and ca. 0.36, respectively). Thus, sequence-selective binding of this hydrocarbon in DNA would be expected to change depending on salt concentration. The temperature dependence of binding was studied at 100 mM Na+ where the equilibrium binding constants for poly[d(A-T)] and poly[d(G-m5C)] are roughly equivalent and 6-fold greater than the binding affinity for poly[d(G-C)]. The binding to poly[d(A-T)] and poly[d(G-C)] is characterized by a delta H omicron = -7.0 kcal/mol, and the large difference in affinity constants arises from differences in negative entropic contributions. Formation of hydrocarbon-poly[d(G-m5C)] complexes is accompanied by a delta H = -9.1 kcal/mol. However, the affinity for poly[d-(G-m5C)] is the same as that for poly[d(A-T)] due to the much more negative entropy associated with binding to poly[d(G-m5C)].  相似文献   

15.
By using pulsed and continuous wave laser irradiation in the 350-450-nm region, we have characterized Raman scattering from horseradish peroxidase (HRP) compounds I and II and from iron porphyrin pi-cation radical model compounds. For compound II we support the suggestion [Terner, J., Sitter, A. J., & Reczek, C. M. (1985) Biochim. Biophys. Acta 828, 73-80; Proniewicz, L. M., Bajdor, K., & Nakamoto, K. (1986) J. Phys. Chem. 90, 1760-1766] that resonance enhancement of the FeIV = O vibration proceeds by way of a charge-transfer state. Our excitation profile data locate this state at approximately 400 nm. Compound I was prepared at neutral pH by rapid mixing of the resting enzyme with hydrogen peroxide. Each sample aliquot was excited by a single, 10-ns laser pulse to generate the Raman spectrum; optical spectroscopy following the Raman measurement confirmed that HRP-I was the principal product during the time scale of the measurement. The Raman spectrum of this species, however, is not characteristic of that which we observe from metalloporphyrin pi-cation radicals [Oertling, W. A., Salehi, A., Chung, Y., Leroi, G. E., Chang, C. K., & Babcock, G. T. (1987) J. Phys. Chem. 91, 5887-5898], including the iron porphyrin cation radicals reported here. Instead, the spectrum recorded for HRP-I at neutral pH is suggestive of an oxoferryl heme with the same geometric and electronic structure as that of HRP-II at high pH.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
Poly-β-benzyl-L -aspartate (poly[Asp(OBzl)]) forms either a lefthanded α-helix, β-sheet, ω-helix, or random coil under appropriate conditions. In this paper the Raman spectra of the above poly[Asp(OBzl)] conformations are compared. The Raman active amide I line shifts from 1663 cm?1 to 1679 cm?1 upon thermal conversion of poly[Asp(OBzl)] from the α-helical to β-sheet conformation while an intense line appearing at 890 cm?1 in the spectrum of the α-helix decreases in intensity. The 890 cm?1 line also displays weak intensity when the polymer is dissolved in chloroform–dichloroacetic acid solution and therefore is converted to the random coil. This line probably arises from a skeletal vibration and is expected to be conformationally sensitive. Similar behavior in the intensity of skeletal vibrations is discussed for other polypeptides undergoing conformational transitions. The Raman spectra of two cross-β-sheet copolypeptides, poly(Ala-Gly) and poly(Ser-Gly), are examined. These sequential polypeptides are model compounds for the crystalline regions of Bombyx mori silk fibroin which forms an extensive β-sheet structure. The amide I, III, and skeletal vibrations appeared in the Raman spectra of these polypeptides at the frequencies and intensities associated with β-sheet homopolypeptides. Since the sequential copolypeptides are intermediate in complexity between the homopolypeptides and the proteins, these results indicate that Raman structure–frequency correlations obtained from homopolypeptide studies can now be applied to protein spectra with greater confidence. The perturbation scheme developed by Krimm and Abe for explaining the frequency splitting of the amide I vibrations in β-sheet polyglycine is applied to poly(L -valine), poly-(Ala-Gly), poly(Ser-Gly), and poly[Asp(OBzl)]. The value of the “unperturbed” frequency, V0, for poly[Asp(OBzl)] was significantly greater than the corresponding values for the other polypeptides. A structural origin for this difference may be displacement of adjacent hydrogen-bonded chains relative to the standard β-sheet conformation.  相似文献   

17.
Raman spectroscopic study of left-handed Z-RNA   总被引:3,自引:0,他引:3  
The solvent conditions that induce the formation of a left-handed Z form of poly[r(G-C)] have been extended to include 6.5 M NaBr at 35 degrees C and 3.8 M MgCl2 at room temperature. The analysis of the A----Z transition in RNA by circular dichroism (CD), 1H and 31P NMR, and Raman spectroscopy shows that two distinct forms of left-handed RNA exist. The ZR-RNA structure forms in high concentrations of NaBr and NaClO4 and exhibits a unique CD signature. ZD-RNA is found in concentrated MgCl2 and has a CD signature similar to the Z form of poly[d(G-C)]. The loss of Raman intensity of the 813-cm-1 A-form marker band in both the A----ZR-RNA and A----ZD-RNA transitions parallels the loss of intensity at 835 cm-1 in the B----Z transition of DNA. A guanine vibration that is sensitive to the glycosyl torsion angle shifts from 671 cm-1 in A-RNA to 641 cm-1 in both ZD- and ZR-RNA, similar to the B----Z transition in DNA in which this band shifts from 682 to 625 cm-1. Significant differences in the glycosyl angle and sugar pucker between Z-DNA and Z-RNA are suggested by the 16-cm-1 difference in the position of this band. The Raman evidence for structural difference between ZD- and ZR-RNA comes from two groups of bands: First, Raman intensities between 1180 and 1600 cm-1 of ZD-RNA differ from those for ZR-RNA, corroborating the CD evidence for differences in base-stacking geometry. Second, the phosphodiester stretching bands near 815 cm-1 provide evidence of differences in backbone geometry between ZD- and ZR-RNA.  相似文献   

18.
The peroxidase activity of carboxymethylated cytochrome c (Cmcytc) has been investigated by spectroscopic and kinetic techniques to examine the effect of carboxymethylation on the peroxidase activity of native cytochrome c (cytc). The optical spectrum suggests that the reaction of Cmcytc with H(2)O(2) proceeds through only one intermediate, compound I. The apparent rate constant (k(app)) for the reaction was found to be 17, 72 and 210 M(-1) s(-1) at pH 7.0, 5.0 and 3.5 respectively. These values are about 60 times larger than those reported for native cytc (0.236 M(-1) s(-1) at pH 7.0), and about five orders of magnitude lower than those for classical peroxidases. Cmcytc was found to catalyse oxidation of organic and inorganic substrates. The second order rate constant for the oxidation of 2,2'-azino-bis(3-ethylbenzthiazoline-6-sulphonic acid) (ABTS) by Cmcytc (205 [H(2)O(2)] s(-1)) is found to be larger than the corresponding value for native cytc (50 [H(2)O(2)] s(-1)) at pH 6.0. The carboxymethylation of cytc ruptures the Fe-S (Met 80) bond and increases the rate of its reaction with H(2)O(2), and its catalytic activity. The specific activity of Cmcytc was measured spectrophotometrically by the reported method using ABTS as substrate, and was found to be 288, 473 and 872 microM min(-1) mg(-1) at pH 7.0, 5.0 and 3.5 respectively. Resonance Raman studies indicated the presence of a bis-histidine coordinated form of Cmcytc at neutral pH, and the existence of a population distribution of different ligation states such as bis-histidine (HH), histidine-water (HW) and five coordinate (5C) forms at lower pH. The relative population of different species in Cmcytc was found to be HH (approximately 100%, approximately 50%, approximately 44%), HW (approximately 0%, approximately 44%, 41%) and 5C (approximately 0%, approximately 6%, 15%) at pH 7.0, 4.7 and 3.1 respectively. We have attempted to correlate the pH dependence of the reaction of Cmcytc with hydrogen peroxide and its peroxidase activity with the haem stereochemical structures observed for Cmcytc. Steady-state and time-resolved tryptophan fluorescence studies on Cmcytc were done to probe the conformational changes around the haem pocket of Cmcytc.  相似文献   

19.
Virions of the Ff group of bacteriophages (fd, f1, M13) are morphologically identical filaments (approximately 6-nm diameter x approximately 880-nm length) in which a covalently closed, single-stranded DNA genome is sheathed by approximately 2700 copies of a 50-residue alpha-helical subunit (pVIII). Orientations of pVIII tyrosines (Tyr21 and Tyr24) with respect to the filament axis have been determined by Raman linear intensity difference (RLID) spectroscopy of flow-oriented mutant virions in which the tyrosines were independently mutated to methionine. The results show that the twofold axis of the phenolic ring (C1-C4 line) of Tyr21 is inclined at 39.5 +/- 1.4 degrees from the virion axis, and that of Tyr24 is inclined at 43.7 +/- 0.6 degrees. The orientation determined for the Tyr21 phenol ring is close to that of a structural model previously proposed on the basis of fiber x-ray diffraction results (Protein Data Bank, identification code 1IFJ). On the other hand, the orientation determined for the Tyr24 phenol ring differs from the diffraction-based model by a 40 degrees rotation about the Calpha-Cbeta bond. The RLID results also indicate that each tyrosine mutation does not greatly affect the orientation of either the remaining tyrosine or single tryptophan (Trp26) of pVIII. On the basis of these results, a refined model is proposed for the coat protein structure in Ff.  相似文献   

20.
The conformation of DNA that originates from association of complementary single-stranded circles (form V DNA) is investigated in solution at low salt concentration. It is shown that circular dichroism extended to the far ultraviolet region (down to 165 nm) represents a powerful tool for determination of the handedness of double helical DNAs in solution. The positive intense band at 186 nm followed by a strong negative band around 170 nm is characteristic of all right-handed helical forms (B,A) of DNA, whereas the circular dichroism spectrum of the Z form of poly[d(G-C)] of opposite helical sense represents a quasi inversion of these far ultraviolet bands. Thus, form V DNA is found to represent a co-existence of left-handed Z-type and right-handed B double helical stretches in addition to negative superturns. The Raman spectrum of form V DNA provides further support for the contribution of a left-handed double helical conformation, as shown by comparison to the high resolution Raman spectra of poly[d(G-C)] in the Z and B forms.The analysis of present spectroscopic data and the analysis of occurrence of alternating [d(G-C)] purine-pyrimidine sequences in the form V DNA used strongly suggest that in DNA of natural sequence, topological constraint may generate left-handed double helices, a conformation thought so far to be limited to the alternating [d(G-C)] sequences. Such structure could play a role in recognition and regulation of gene expression.  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