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.
13C, 17O and 57Fe NMR spectra of several carbonmonoxy hemoprotein models with varying polar and steric effects of the distal organic superstructure, constraints of the proximal side, and porphyrin ruffling are reported. Both heme models and heme proteins obey a similar excellent linear delta(13C) versus nu(C-O) relationship which is primarily due to modulation of pi-back-bonding from the Fe d(pi) to CO pi* orbital by the distal pocket polar interactions. The lack of correlation between delta(13C) and delta(17O) suggests that the two probes do not reflect a similar type of electronic and structural perturbation. delta(17O) is not primarily influenced by the local distal field interactions and does not correlate with any single structural property of the Fe-C-O unit; however, atropisomerism and deformation of the porphyrin geometry appear to play a significant role. 57Fe shieldings vary by nearly 900 ppm among various hemes and an excellent correlation was found between delta(57Fe) and the absolute crystallographic average displacement of the meso carbon atoms, /Cm/, relative to the porphyrin core mean plane. The excellent correlation between iron-57 shieldings and the average shieldings of the meso carbons of the porphyrin skeleton of TPP derivatives suggests that the two probes reflect a similar type of electronic and structural perturbation which is primarily porphyrin ruffling.  相似文献   

2.
Binding structures of metal complexes of deglyco-peplomycin (dPEP) on DNA were investigated by comparing dPEP complexes with those of bleomycin (BLM) using DNA fiber EPR spectroscopy. A low spin species of Fe(III)dPEP observed in the DNA pellet changed irreversibly to several high spin species after the fabrication of the DNA fibers. The g values of the high spin species were different from those of Fe(III)BLM. The high spin species could not be nitrosylated reductively to ON-Fe(II)dPEP, suggesting that some nitrogen atoms coordinated to the Fe(III) were displaced on the DNA fibers. On the other hand, O(2)-Co(II)dPEP remained intact on the fibers similarly to O(2)-Co(II)BLM but with an increased randomness in the orientation on the DNA. In contrast to Cu(II)BLM, a considerable amount of Cu(II)dPEP bound almost randomly on B-form DNA fibers. These results indicated that the sugar moiety in peplomycin or bleomycin is playing an important role in enhancing the stability of the metal-binding domain and in the stereospecificity of the binding on DNA.  相似文献   

3.
The DNA-binding proteins from starved cells (Dps) are a family of proteins induced in microorganisms by oxidative or nutritional stress. Escherichia coli Dps, a structural analog of the 12-subunit Listeria innocua ferritin, binds and protects DNA against oxidative damage mediated by H(2)O(2). Dps is shown to be a Fe-binding and storage protein where Fe(II) oxidation is most effectively accomplished by H(2)O(2) rather than by O(2) as in ferritins. Two Fe(2+) ions bind at each of the 12 putative dinuclear ferroxidase sites (P(Z)) in the protein according to the equation, 2Fe(2+) + P(Z) --> [(Fe(II)(2)-P](FS)(Z+2) + 2H(+). The ferroxidase site (FS) bound iron is then oxidized according to the equation, [(Fe(II)(2)-P](FS)(Z+2) + H(2)O(2) + H(2)O --> [Fe(III)(2)O(2)(OH)-P](FS)(Z-1) + 3H(+), where two Fe(II) are oxidized per H(2)O(2) reduced, thus avoiding hydroxyl radical production through Fenton chemistry. Dps acquires a ferric core of approximately 500 Fe(III) according to the mineralization equation, 2Fe(2+) + H(2)O(2) + 2H(2)O --> 2Fe(III)OOH((core)) + 4H(+), again with a 2 Fe(II)/H(2)O(2) stoichiometry. The protein forms a similar ferric core with O(2) as the oxidant, albeit at a slower rate. In the absence of H(2)O(2) and O(2), Dps forms a ferrous core of approximately 400 Fe(II) by the reaction Fe(2+) + H(2)O + Cl(-) --> Fe(II)OHCl((core)) + H(+). The ferrous core also undergoes oxidation with a stoichiometry of 2 Fe(II)/H(2)O(2). Spin trapping experiments demonstrate that Dps greatly attenuates hydroxyl radical production during Fe(II) oxidation by H(2)O(2). These results and in vitro DNA damage assays indicate that the protective effect of Dps on DNA most likely is exerted through a dual action, the physical association with DNA and the ability to nullify the toxic combination of Fe(II) and H(2)O(2). In the latter process a hydrous ferric oxide mineral core is produced within the protein, thus avoiding oxidative damage mediated by Fenton chemistry.  相似文献   

4.
Listeria innocua Dps (DNA binding protein from starved cells) affords protection to DNA against oxidative damage and can accumulate about 500 iron atoms within its central cavity through a process facilitated by a ferroxidase center. The chemistry of iron binding and oxidation in Listeria Dps (LiDps, formerly described as a ferritin) using H(2)O(2) as oxidant was studied to further define the mechanism of iron deposition inside the protein and the role of LiDps in protecting DNA from oxidative damage. The relatively strong binding of 12 Fe(2+) to the apoprotein (K(D) approximately 0.023 microM) was demonstrated by isothermal titration calorimetry, fluorescence quenching, and pH stat experiments. Hydrogen peroxide was found to be a more efficient oxidant for the protein-bound Fe(2+) than O(2). Iron(II) oxidation by H(2)O(2) occurs with a stoichiometry of 2 Fe(2+)/H(2)O(2) in both the protein-based ferroxidation and subsequent mineralization reactions, indicating complete reduction of H(2)O(2) to H(2)O. Electron paramagnetic resonance (EPR) spin-trapping experiments demonstrated that LiDps attenuates the production of hydroxyl radical by Fenton chemistry. DNA cleavage assays showed that the protein, while not binding to DNA itself, protects it against the deleterious combination of Fe(2+) and H(2)O(2). The overall process of iron deposition and detoxification by LiDps is described by the following equations. For ferroxidation, Fe(2+) + Dps(Z)--> [(Fe(2+))-Dps](Z+1) + H(+) (Fe(2+) binding) and [(Fe(2+))-Dps](Z+1) + Fe(2+) + H(2)O(2) --> [(Fe(3+))(2)(O)(2)-Dps](Z+1) + 2H(+) (Fe(2+) oxidation/hydrolysis). For mineralization, 2Fe(2+) + H(2)O(2) + 2H(2)O --> 2Fe(O)OH((core)) + 4H(+) (Fe(2+) oxidation/hydrolysis). These reactions occur in place of undesirable odd-electron redox processes that produce hydroxyl radical.  相似文献   

5.
Treatment of aqueous suspensions of different asbestos fibers (amosite, anthophyllite, chrysotile, and crocidolite) at 0-4°C and pH 7.2 with H O results in the consumption of H O with concomitant release of iron and production of O. During incubations, [H O] decreased in proportion to the mass of the suspended fiber, the duration of incubation, and the initial [H O]. The consumption of H O, production of O and release of iron all vary synergistically with the structure of the asbestos fiber. Release of silicon during the incubation was small relative to the decrement in [H O], reflecting a lack of dissolution of the fiber. The data are consistent with a redox process for the release of surface bound iron and it is significant that iron release occurs in the absence of a Fe(II) or Fe(III) chelator. The implications of iron release from the asbestos surface may be important in inflammatory disorders in which both silicate bound iron and H O accumulate. © Rapid Science 1998.  相似文献   

6.
The origin of previously observed variations in stoichiometry of iron oxidation during the oxidative deposition of iron in ferritin has been poorly understood. Knowledge of the stoichiometry of Fe(II) oxidation by O2 is essential to establishing the mechanism of iron core formation. In the present work, the amount of Fe(II) oxidized was measured by M?ssbauer spectrometry and the O2 consumed by mass spectrometry. The number of protons produced in the reaction was measured by "pH stat" titration and hydrogen peroxide production by the effect of the enzyme catalase on the measured stoichiometry. For protein samples containing low levels of iron (24 Fe(II)/protein) the stoichiometry was found to be 1.95 +/- 0.18 Fe(II)/O2 with H2O2 being a product, viz. Equation 1. 2Fe2+ + O2 + 4H2O----2FeOOH + H2O2 + 4H+ (1) EPR spin trapping experiments showed no evidence of superoxide radical formation. The stoichiometry markedly increased with additional iron (240-960 Fe/protein), to a value of 4 Fe(II)/O2 as in Equation 2. 4Fe2+ + O2 + 6H2O----4FeOOH + 8H+ (2) As the iron core is progressively laid down, the mechanism of iron oxidation changes from a protein dominated process with H2O2 being the primary product of O2 reduction to a mineral surface dominated process where H2O is the primary product. These results emphasize the importance of the apoferritin shell in facilitating iron oxidation in the early stage of iron deposition prior to significant development of the polynuclear iron core.  相似文献   

7.
This study investigates the solution thermodynamics of the iron complexes of dexrazoxane (ICRF-187, (+)-1,2-bis(3,5-dioxopiperazinyl-1-yl)propane), [Fe(ADR-925)](+/0), and its desmethyl derivative ICRF-154, [Fe(ICRF-247)H2O](+/0). The solid state structure of [Fe(ICRF-247)H2O]+ is also reported. [Fe(ICRF-247)H2O]Br x 0.5NaBr x H2O crystallizes in the P42(1)2 space group with Z = 4, a = 14.9851(8), b = 14.9851(8), c = 8.0825(9) A and R = 0.03(2) for 1839 reflections and exhibits a pentagonal bipyramidal geometry with a labile water molecule occupying the seventh coordination site. Potentiometric titrations (FeL = 8.5 mM, 0.1 M NaNO3, 25 degrees C) reveal stable monomeric complexes (log Kf = 18.2 +/- 0.1, [Fe(ADR-925)]+, and 17.4 +/- 0.1, [Fe(ICRF-247)H2O]+) exist in solution at relatively low pH. Upon addition of base, the iron-bound water is deprotonated; the pKa values for [Fe(ICRF-247)H2O]+ and [Fe(ADR-925)]+ are 5.63 +/- 0.07 and 5.84 +/- 0.07, respectively. At higher pH both complexes undergo mu-oxo dimerization characterized by log Kd values of 2.68 +/- 0.07 for [Fe(ICRF-247)H2O]+ and 2.23 +/- 0.07 for [Fe(ADR-925)]+. In the presence of an oxidant and reductant, both [Fe(ICRF-247)H2O]+ and [Fe(ADR-925)]+ produce hydroxyl radicals that cleave pBR322 plasmid DNA at pH 7 in a metal complex concentration-dependent manner. At low metal complex concentrations (approximately 10(-5) M) where the monomeric form predominates, cleavage by both FeICRF complexes is efficient while at higher concentrations (approximately 5 x 10(-4) M) DNA cleavage is hindered. This change in reactivity is in part accounted for by dimer formation.  相似文献   

8.
Radiotherapy is the major treatment for nasopharyngeal carcinoma, a malignant tumor of epithelial origin. In this process, a tracer with high sensitivity is pivotal for diagnostic imaging in radiotherapy. Here, we designed a novel multifunctional magnetic silver nanocomposite, Fe(3)O(4)/Ag conjugated to an epidermal growth factor receptor-specific antibody (C225), which can be potentially used for synchronous cancer therapy and diagnosis via magnetic resonance imaging. Characteristics of Fe(3)O(4)/Ag/C225 were determined by transmission electron microscopy, energy dispersive X-ray spectroscopy, ultraviolet spectra, and dynamic light scattering. The results demonstrated that Fe(3)O(4)/Ag/C225 nanoparticles were spherical and dispersed well in water. The activity of C225 was preserved ~80% in the Fe(3)O(4)/Ag/C225 nanoparticles. Futhermore, we tested the cytotoxicity and radiosensitivity of the nanocomposite for human nasopharyngeal carcinoma cell lines (CNEs) in vitro. MTT analysis revealed that Fe(3)O(4)/Ag/C225 could inhibit the proliferation of CNEs in a dose- and time-dependent manner. The clonogenic assay indicated that Fe(3)O(4)/Ag/C225 combined with X-ray treatment could increase the sensitivity of CNEs to irradiation. In a summary, the novel multifunctional nanocomposite Fe(3)O(4)/Ag/C225 might be a potential radiosensitizer for treating malign tumors in the clinic.  相似文献   

9.
Heme oxygenase converts heme to biliverdin, iron, and CO in a reaction with two established intermediates, alpha-meso-hydroxyheme and verdoheme. Transient kinetic studies show that the conversion of Fe(3+)-heme to Fe(3+)-verdoheme is biphasic. Electron transfer to the heme (0.11 s(-1) at 4 degrees C and 0.49 s(-1) at 25 degrees C) followed by rapid O(2) binding yields the ferrous dioxy complex. Transfer of an electron (0.056 s(-1) at 4 degrees C and 0.21 s(-1) at 25 degrees C) to this complex triggers the formation of alpha-meso-hydroxyheme and its subsequent O(2)-dependent fragmentation to Fe(3+)-verdoheme. The conversion of Fe(3+)-verdoheme to Fe(3+)-biliverdin is also biphasic. Thus, reduction of Fe(3+) to Fe(2+)-verdoheme (0.15 s(-1) at 4 degrees C and 0.55 s(-1) at 25 degrees C) followed by O(2) binding and an electron transfer produces Fe(3+)-biliverdin (0.025 s(-1) at 4 degrees C and 0.10 s(-1) at 25 degrees C). The conversion of Fe(3+)-biliverdin to free biliverdin is triphasic. Reduction of Fe(3+)-biliverdin (0.035 s(-1) at 4 degrees C and 0.15 s(-1) at 25 degrees C), followed by rapid release of Fe(2+) (0.19 s(-1) at 4 degrees C and 0.39 s(-1) at 25 degrees C), yields the biliverdin-enzyme complex from which biliverdin slowly dissociates (0.007 s(-1) at 4 degrees C and 0.03 s(-1) at 25 degrees C). The rate of Fe(2+) release agrees with the rate of Fe(3+)-biliverdin reduction. Fe(2+) release clearly precedes biliverdin dissociation. In the absence of biliverdin reductase, biliverdin release is the rate-limiting step, but in its presence biliverdin release is accelerated and the overall rate of heme degradation is limited by the conversion of Fe(2+)-verdoheme to the Fe(3+)-biliverdin.  相似文献   

10.
The reaction of horse spleen ferritin (HoSF) with Fe2+ at pH 6.5 and 7.5 using O2, H2O2 and 1:1 a mixture of both showed that the iron deposition reaction using H2O2 is approximately 20- to 50-fold faster than the reaction with O2 alone. When H2O2 was added during the iron deposition reaction initiated with O2 as oxidant, Fe2+ was preferentially oxidized by H2O2, consistent with the above kinetic measurements. Both the O2 and H2O2 reactions were well defined from 15 to 40 degrees C from which activation parameters were determined. The iron deposition reaction was also studied using O2 as oxidant in the presence and absence of catalase using both stopped-flow and pumped-flow measurements. The presence of catalase decreased the rate of iron deposition by approximately 1.5-fold, and gave slightly smaller absorbance changes than in its absence. From the rate constants for the O2 (0.044 s(-1)) and H2O2 (0.67 s(-1)) iron-deposition reactions at pH 7.5, simulations of steady-state H2O2 concentrations were computed to be 0.45 microM. This low value and reported Fe2+/O2 values of 2.0-2.5 are consistent with H2O2 rapidly reacting by an alternate but unidentified pathway involving a system component such as the protein shell or the mineral core as previously postulated [Biochemistry 22 (1983) 876; Biochemistry 40 (2001) 10832].  相似文献   

11.
Ferritin catalyzes the oxidation of Fe2+ by O2 to form a reconstituted Fe3+ oxy-hydroxide mineral core, but extensive studies have shown that the Fe2+ to O2 stoichiometry changes with experimental conditions. At Fe2+ to horse spleen ferritin (HoSF) ratios greater than 200, an upper limit of Fe2+ to O2 of 4 is typically measured, indicating O2 is reduced to 2H2O. In contrast, a lower limit of Fe2+ to O2 of approximately 2 is measured at low Fe2+ to HoSF ratios, implicating H2O2 as a product of Fe2+ deposition. Stoichiometric amounts of H2O2 have not been measured, and H2O2 is proposed to react with an unknown system component. Evidence is presented that identifies this component as amine buffers, including 3-N-morpholinopropanesulfonic acid (MOPS), which is widely used in ferritin studies. In the presence of non-amine buffers, the Fe2+ to O2 stoichiometry was approximately 4.0, but at high concentrations of amine buffers (0.10 M) the Fe2+ to O2 stoichiometry is approximately 2.5 for iron loadings of eight to 30 Fe2+ per HoSF. Decreasing the concentration of amine buffer to zero resulted in an Fe2+ to O2 stoichiometry of approximately 4. Direct evidence for amine buffer modification during Fe2+ deposition was obtained by comparing authentic and modified buffers using mass spectrometry, NMR, and thin layer chromatography. Tris(hydroxymethyl)aminomethane, MOPS, and N-methylmorpholine (a MOPS analog) were all rapidly chemically modified during Fe2+ deposition to form N-oxides. Under identical conditions no modification was detected when amine buffer, H2O2, and O2 were combined with Fe2+ or ferritin separately. Thus, a short-lived ferritin intermediate is required for buffer modification by H2O2. Variation of the Fe2+ to O2 stoichiometry versus the Fe2+ to HoSF ratio and the amine buffer concentration are consistent with buffer modification.  相似文献   

12.
Microbial nitrate-dependent Fe(II) oxidation is known to contribute to iron biogeochemical cycling; however, the microorganisms responsible are virtually unknown. In an effort to elucidate this microbial metabolic process in the context of an environmental system, a 14-cm sediment core was collected from a freshwater lake and geochemically characterized concurrently with the enumeration of the nitrate-dependent Fe(II)-oxidizing microbial community and subsequent isolation of a nitrate-dependent Fe(II)-oxidizing microorganism. Throughout the sediment core, ambient concentrations of Fe(II) and nitrate were observed to coexist. Concomitant most probable number enumeration revealed the presence of an abundant nitrate-dependent Fe(II)-oxidizing microbial community (2.4 x 10(3) to 1.5 x 10(4) cells g(-1) wet sediment) from which a novel anaerobic, lithoautotrophic, Fe(II)-oxidizing bacterium, strain 2002, was isolated. Analysis of the complete 16S rRNA gene sequence revealed that strain 2002 was a member of the beta subclass of the proteobacteria with 94.8% similarity to Chromobacterium violaceum, a bacterium not previously recognized for the ability to oxidize nitrate-dependent Fe(II). Under nongrowth conditions, both strain 2002 and C. violaceum incompletely reduced nitrate to nitrite with Fe(II) as the electron donor, while under growth conditions nitrate was reduced to gaseous end products (N2 and N2O). Lithoautotrophic metabolism under nitrate-dependent Fe(II)-oxidizing conditions was verified by the requirement of CO2 for growth as well as the assimilation of 14C-labeled CO2 into biomass. The isolation of strain 2002 represents the first example of an anaerobic, mesophilic, neutrophilic Fe(II)-oxidizing lithoautotroph isolated from freshwater samples. Our studies further demonstrate the abundance of nitrate-dependent Fe(II) oxidizers in freshwater lake sediments and provide further evidence for the potential of microbially mediated Fe(II) oxidation in anoxic environments.  相似文献   

13.
A comparison of an undecairon(III) complex with the ferritin iron core   总被引:1,自引:0,他引:1  
The iron core of ferritin is comprised of up to 4,500 Fe(III) atoms as Fe2O3.nH2O, which is maintained in solution by a surrounding, spherical coat of protein. Organisms as diverse as bacteria and man use the ferritin iron-protein complex as a reservoir of stored iron for other essential proteins. To extend studies of the steps in polynuclear iron core formation, a recently characterized undecairon(III) oxo-hydroxo aggregate [Fe11 complex] (Gorun et al., J. Am. Chem. Soc. 109, 3337 [1987]) was examined by x-ray absorption spectroscopy as a model for an intermediate. The results, which are comparable to the previous x-ray diffraction studies, show near neighbors (Fe-O) at 1.90 A that are distinct from those in ferritin and a longer distance of 2.02 A. However, contributions from neighbors (Fe-C) known to exist at ca. 2.7 A were obscured by a highly ordered Fe-Fe interaction and were not detectable in the Fe11 complex in contrast to a previously characterized Fe(III) cluster bound to the protein coat. Of the two Fe-Fe interactions detectable in the Fe11 complex, the shortest, at 3.0 A is particularly interesting, occurring at the same distance as a full shell (CN = 6) in ferritin, but having fewer Fe neighbors (CN = 2-3) characteristic of an intermediate in core formation. The incomplete Fe-Fe shell is much more ordered than in ferritin, suggesting that the disorder in ferritin cores may be associated with the later steps of the core growth. Differences between the Fe11 complex and the full core of ferritin indicate the possibility of intermediates in ferritin iron formation that might be like Fe11.  相似文献   

14.
On the limited ability of superoxide to release iron from ferritin   总被引:3,自引:0,他引:3  
Reductive release of iron from ferritin may catalyze cytotoxic radical reactions like the Haber-Weiss reaction. The ability of .O2- to mobilize Fe(II) from ferritin was studied by using the xanthine/xanthine oxidase reaction, with and without superoxide dismutase, and with bathophenanthroline sulphonate as the chelator. Not more than one or two Fe(II)/ferritin molecules could be released by an .O2(-)-dependent mechanism, even after repeated exposures of ferritin to bursts of .O2-. The amount of releaseable iron depended on the size and the age of the iron core, but not on the iron content of the protein shell of ferritin which was manipulated by chelators and addition of FeCl3. The kinetic characteristics of the .O2(-)-mediated iron release indicated the presence of a small pool of readily available iron at the surface of the core. The very limited .O2(-)-dependent release of iron from ferritin is compatible with a protective role of ferritin against toxic iron-catalyzed reactions.  相似文献   

15.
The so-called Fe/Mn-oxidizing bacteria have long been recognized for their potential to form extracellular iron hydroxide or manganese oxide structures in aquatic environments. Bacterial species belonging to the genus Gallionella, one type of such bacteria, oxidize iron and produce uniquely twisted extracellular stalks consisting of iron oxide-encrusted inorganic/organic fibers. This paper describes the ultrastructure of Gallionella cells and stalks and the visualized structural and spatial localization of constitutive elements within the stalks. Electron microscopy with energy-dispersive X-ray microanalysis showed the export site of the stalk fibers from the cell and the uniform distribution of iron, silicon, and phosphorous in the stalks. Electron energy-loss spectroscopy revealed that the stalk fibers had a central carbon core of bacterial exopolymers and that aquatic iron interacted with oxygen at the surface of the carbon core, resulting in deposition of iron oxides at the surface. This new knowledge of the structural and spatial associations of iron with oxygen and carbon provides deeper insights into the unique inorganic/organic hybrid structure of the stalks.  相似文献   

16.
The structure of the core oligosaccharide moiety of the lipopolysaccharide (LPS) of Plesiomonas shigelloides O54 (strain CNCTC 113/92) has been investigated by (1)H and (13)C NMR, fast atom bombardment mass spectrometry (MS)/MS, matrix-assisted laser-desorption/ionization time-of-flight MS, monosaccharide and methylation analysis, and immunological methods. It was concluded that the main core oligosaccharide of this strain is composed of a decasaccharide with the following structure: (see text) in which l-alpha-D-Hepp is l-glycero-alpha-D-manno-heptopyranose. The nonasaccharide variant of the core oligosaccharide ( approximately 10%), devoid of beta-D-Glcp substituting the alpha-D-GlcpN at C-6, was also identified. The core oligosaccharide substituted at C-4 of the outer core beta-D-Glcp residue with the single O-polysaccharide repeating unit was also isolated yielding a hexadecasaccharide structure. The determination of the monosaccharides involved in the linkage between the O-specific polysaccharide part and the core, as well as the presence of -->3)-D-beta-D-Hepp-(1--> instead of -->3,4)-D-beta-D-Hepp-(1--> in the repeating unit, revealed the structure of the biological repeating unit of the O-antigen. The core oligosaccharides are not substituted by phosphate residues and represent novel core type of bacterial LPS that is characteristic for the Plesiomonas shigelloides serotype O54. Serological screening of 69 different O-serotypes of P. shigelloides suggests that epitopes similar to the core oligosaccharide of serotype O54 (strain CNCTC 113/92) might also be present in the core region of the serotypes O24 (strain CNCTC 92/89), O37 (strain CNCTC 39/89) and O96 (strain CNCTC 5133) LPS.  相似文献   

17.
Serological studies using SDS-PAGE and immunoblotting revealed that from five strains that are ascribed to Citrobacter serogroup O2, four strains, PCM 1494, PCM 1495, PCM 1496 and PCM 1507, are reactive with specific anti-Citrobacter O2 serum. In contrast, strain PCM 1573 did not react with anti-Citrobacter O2 serum and, hence, does not belong to serogroup O2. The LPS of Citrobacter youngae O2a,1b (strain PCM 1507) was degraded under mild acidic conditions and the O-specific polysaccharide (OPS) released was isolated by gel chromatography. Sugar and methylation analyses along with (1)H- and (13)C-NMR spectroscopy, including two-dimensional (1)H,(1)H COSY, TOCSY, NOESY and (1)H,(13)C HSQC experiments, showed that the repeating unit of the OPS has the following structure: [structure: see text]. NMR spectroscopic studies demonstrated that Citrobacter werkmanii O20 and C. youngae O25 have the same OPS structure as C. youngae O2. Sugar and methylation analyses of the core oligosaccharide fractions demonstrated structural differences in the lipopolysaccharide core regions of these strains, which may substantiate their classification in different serogroups.  相似文献   

18.
In this work, lipase produced from an isolated strain Burkholderia sp. C20 was immobilized on magnetic nanoparticles to catalyze biodiesel synthesis. Core-shell nanoparticles were synthesized by coating Fe(3)O(4) core with silica shell. The nanoparticles treated with dimethyl octadecyl [3-(trimethoxysilyl) propyl] ammonium chloride were used as immobilization supporters. The Burkholderia lipase was then bound to the synthesized nanoparticles for immobilization. The protein binding efficiency on alkyl-functionalized Fe(3)O(4)-SiO(2) was estimated as 97%, while the efficiency was only 76% on non-modified Fe(3)O(4)-SiO(2). Maximum adsorption capacity of lipase on alkyl-functionalized Fe(3)O(4)-SiO(2) was estimated as 29.45 mg g(-1) based on Langmuir isotherm. The hydrolytic kinetics (using olive oil as substrate) of the lipase immobilized on alkyl-grafted Fe(3)O(4)-SiO(2) followed Michaelis-Menten model with a maximum reaction rate and a Michaelis constant of 6251 Ug(-1) and 3.65 mM, respectively. Physical and chemical properties of the nanoparticles and the immobilized lipase were characterized by Brunauer-Emmett-Teller (BET) analysis, scanning electron microscope (SEM), and Fourier transform infrared spectroscopy (FT-IR). Moreover, the immobilized lipase was used to catalyze the transesterification of olive oil with methanol to produce fatty acid methyl esters (FAMEs), attaining a FAMEs conversion of over 90% within 30 h in batch operation when 11 wt% immobilized lipase was employed. The immobilized lipase could be used for ten cycles without significant loss in its transesterification activity.  相似文献   

19.
The mechanism of mitomycin C-induced lipid peroxidation has been studied at pH 7.5, using systems containing phospholipid membranes (liposomes) and an Fe3+-ADP complex with purified NADPH-cytochrome P-450 reductase. Both O2- and H2O2 are generated during the aerobic enzyme-catalyzed reaction in the presence of mitomycin C. Hydroxyl radical is formed in the reaction by the reduction of H2O2. This is catalyzed by the Fe2+-ADP complex in a phosphate buffer or to a lesser extent when in a Tris-HCl buffer. The reduction of Fe3+-ADP to Fe2+-ADP is mainly achieved by O2-. The resulting Fe2+-ADP in the presence of O2 forms a perferryl ion complex which is a powerful stimulator of lipid peroxidation. However, the formation of such an iron-oxygen complex is strongly inhibited by phosphate ions, which do not interfere with the generation of OH radicals. These findings suggest that, since lipid peroxidation occurs in a Tris-HCl buffer (but not in a phosphate buffer), the OH radical is unlikely to be involved in the observed lipid peroxidation process.  相似文献   

20.
Escherichia coli strains of group O111 were characterized with respect to sensitivity to complement killing, amount of lipopolysaccharide and O antigen-containing capsule, and distribution of O antigen. All wild-type E. coli O111 strains were resistant to complement killing in the absence of specific antibody. Presensitization of strains with antibody to whole cells (OK antibody), followed by incubation in 50% pooled normal human serum as a source of complement, subdivided wild-type strains into three types: completely resistant, partially resistant, and sensitive. Completely and partially resistant mutants were isolated by cycles of serum killing, starting with one sensitive strain. Completely resistant mutants had no O antigen-containing capsule, but had 50% more lipopolysaccharide than did the parent, and this lipopolysaccharide had 30% fewer lipid A core molecules devoid of O antigen. Partially resistant mutants still had O antigen-containing capsule, but contained 40% more lipopolysaccharide than did the parent; the extent of coverage of lipid A core with O antigen remained unchanged. No correlations were found between outer membrane protein composition and the degree of serum resistance. Since the terminal membrane attack complex (C5b-9) must stably insert into a hydrophobic membrane site to effect killing, we conclude that both increased lipid A core and increased coverage of lipid A core with O antigen preclude access of C5b-9 to lethal sites on the cell surface.  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