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.
The thermodynamic behavior of representative short (UQ2), middle (UQ4 and UQ6) and long-chain (UQ10) ubiquinones (UQ) mixed with dipalmitoyl-phosphatidylcholine (DPPC) was studied in monolayers at the air-water interface. The influence of isoprenoid chain-length of UQ on miscibility of both lipids was investigated by analysis of surface pressure-area isotherms and using fluorescence microscopy. Analysis of excess areas (A(ex)) and free energies of mixing (DeltaGm), calculated from compression isotherms in the full range of ubiquinones concentrations, has given evidences for UQ-rich constant-size (UQ6, UQ10) or less growth limited (UQ2, UQ4) microdomains formation within mixed films. Fluorescence microscopy observation revealed that ubiquinones are preferentially soluble in the expanded phase. When lateral pressure increased, concomitant evolutions of A(ex) and DeltaGm parameters, and composition dependence of collapse surface pressures, argue for an evolution towards a total segregation, never reached due to expulsion of ubiquinones from the film. The possible significance of these observations is discussed in relation to ubiquinones organization and similar chain length effects in membranes.  相似文献   

2.
Sucrose isomerase (SI) genes from Pantoea dispersa UQ68J, Klebsiella planticola UQ14S, and Erwinia rhapontici WAC2928 were cloned and expressed in Escherichia coli. The predicted products of the UQ14S and WAC2928 genes were similar to known SIs. The UQ68J SI differed substantially, and it showed the highest isomaltulose-producing efficiency in E. coli cells. The purified recombinant WAC2928 SI was unstable, whereas purified UQ68J and UQ14S SIs were very stable. UQ68J SI activity was optimal at pH 5 and 30 to 35°C, and it produced a high ratio of isomaltulose to trehalulose (>22:1) across its pH and temperature ranges for activity (pH 4 to 7 and 20 to 50°C). In contrast, UQ14S SI showed optimal activity at pH 6 and 35°C and produced a lower ratio of isomaltulose to trehalulose (<8:1) across its pH and temperature ranges for activity. UQ68J SI had much higher catalytic efficiency; the Km was 39.9 mM, the Vmax was 638 U mg−1, and the Kcat/Km was 1.79 × 104 M−1 s−1, compared to a Km of 76.0 mM, a Vmax of 423 U mg−1, and a Kcat/Km of 0.62 × 104 M−1 s−1 for UQ14S SI. UQ68J SI also showed no apparent reverse reaction producing glucose, fructose, or trehalulose from isomaltulose. These properties of the P. dispersa UQ68J enzyme are exceptional among purified SIs, and they indicate likely differences in the mechanism at the enzyme active site. They may favor the production of isomaltulose as an inhibitor of competing microbes in high-sucrose environments, and they are likely to be highly beneficial for industrial production of isomaltulose.  相似文献   

3.
Quinone and inhibitor binding to Rhodopseudomonas sphaeroides (R-26 and GA) reaction centers were studied using spectroscopic methods and by direct adsorption of reaction centers onto anion exchange filters in the presence of 14C-labelled quinone or inhibitor. These measurements show that as secondary acceptor, QB, ubiquinone (UQ) is tightly bound in the semiquinone form and loosely bound in the quinone and quinol forms. The quinol is probably more loosely bound than the quinone. o-Phenanthroline and terbutryn, a triazine inhibitor, compete with UQ and with each other for binding to the reaction center. Inhibition by o-phenanthroline of electron transfer from the primary to the secondary quinone acceptor (QA to QB) occurs via displacement of UQ from the QB binding site. Displacement of UQ by terbutryn is apparently accessory to the inhibition of electron transfer. Terbutryn binding is lowered by reduction of QB to Q?B but is practically unaffected by reduction of QA to Q?A in the absence of QB. UQ-9 and UQ-10 have a 5- to 6-fold higher binding affinity to the QB site than does UQ-1, indicating that the long isoprenoid chain facilitates the binding to the QB site.  相似文献   

4.
The diversity of dinucleotide sequences at the 5′ ends of vaccinia virus mRNA's was determined by a two-dimensional electrophoresis procedure. RNA labeled with S-adenosyl[methyl-3H]methionine was synthesized in vitro by enzymes present in vaccinia virus cores. The RNA, ending in m7G(5′)pppNmpN−, was β-eliminated and treated with alkaline phosphatase. After digestion with RNases T2, T1, and A, all eight possible dinucleotides containing Gm and Am were identified. They are, in decreasing order of abundance: GmpUp (22%), AmpCp (18%), GmpAp (16%), GmpCp (15%), AmpAp (11%), AmpUp (10%), AmpGp (7%), and GmpGp (2%).  相似文献   

5.
Purple, photosynthetic reaction centers from Rhodobacter sphaeroides bacteria use ubiquinone (UQ10) as both primary (QA) and secondary (QB) electron acceptors. Many quinones reconstitute QA function, while a few will act as QB. Nine quinones were tested for their ability to bind and reconstitute QA and QB functions. Only ubiquinone (UQ) reconstitutes both functions in the same protein. The affinities of the non-native quinones for the QB site were determined by a competitive inhibition assay. The affinities of benzoquinones, naphthoquinone (NQ), and 2-methyl-NQ for the QB site are 7 ± 3 times weaker than that at QA site. However, di-ortho-substituted NQs and anthraquinone bind tightly to the QA site (K d ≤ 200 nM), and ≥1,000 times more weakly to the QB site, perhaps setting a limit on the size of the site. With a low-potential electron donor, 2-methyl, 3-dimethylamino-1,4-NQ, (Me-diMeAm-NQ) at QA, QB reduction is 260 meV, more favorable than with UQ as QA. Electron transfer from Me-diMeAm-NQ at the QA site to NQ at the QB site can be detected. In the QB site, the NQ semiquinone is estimated to be ≈60–100 meV higher in energy than the UQ semiquinone, while in the QA site, the semiquinone energy level is similar or lower with NQ than with UQ. Thus, the NQ semiquinone is more stable in the QA than in the QB site. In contrast, the native UQ semiquinone is ≈60 meV lower in energy in the QB than in the QA site, stabilizing forward electron transfer from QA to QB.  相似文献   

6.
The action of phospholipase C (Bacillus cereus) toward mixed micelles of phosphatidylcholine and the nonionic surfactant Triton X-100 is analyzed according to the “surfaceas-cofactor” kinetic scheme recently proposed for characterizing the action of lipolytic enzymes [Deems, R. A., Eaton, B. R., and Dennis, E. A. (1975) J. Biol. Chem.250, 9013–9020]. According to this scheme, the enzyme first associates with the surface or mixed micelles, where the dissociation constant is KsA. The enzyme, now part of the mixed micelle surface, then binds the substrate phospholipid molecule in its active site and this binding is related to the Michaelis constant, KmB. The surface, or mixed micelles in this scheme, behaves kinetically as a cofactor in that, under initial rate conditions, the surface properties of the mixed micelles are virtually unchanged after catalysis. For phospholipase C with egg phosphatidylcholine as substrate, it was found that at pH 6.4 (the pH optimum for the enzyme) and 40 °C, V is about 2 × 103 μmol min?1 (mg of protein)?1. KsA is about 2 mm and KmB is 1 to 2 × 10?10 mol cm?2. The kinetic constants for phospholipase C are compared with those previously reported for phospholipase A2 and the membrane-bound enzyme phosphatidylserine decarboxylase determined under similar conditions.  相似文献   

7.
Mixed monolayers of the ganglioside GM1 and the lipid dipalmitoylphosphatidlycholine (DPPC) at air-water and solid-air interfaces were investigated using various biophysical techniques to ascertain the location and phase behavior of the ganglioside molecules in a mixed membrane. The effects induced by GM1 on the mean molecular area of the binary mixtures and the phase behavior of DPPC were followed for GM1 concentrations ranging from 5 to 70 mol %. Surface pressure isotherms and fluorescence microscopy imaging of domain formation indicate that at low concentrations of GM1 (<25 mol %), the monolayer becomes continually more condensed than DPPC upon further addition of ganglioside. At higher GM1 concentrations (>25 mol %), the mixed monolayer becomes more expanded or fluid-like. After deposition onto a solid substrate, atomic force microscopy imaging of these lipid monolayers showed that GM1 and DPPC pack cooperatively in the condensed phase domain to form geometrically packed complexes that are more ordered than either individual component as evidenced by a more extended total height of the complex arising from a well-packed hydrocarbon tail region. Grazing incidence x-ray diffraction on the DPPC/GM1 binary mixture provides evidence that ordering can emerge when two otherwise fluid components are mixed together. The addition of GM1 to DPPC gives rise to a unit cell that differs from that of a pure DPPC monolayer. To determine the region of the GM1 molecule that interacts with the DPPC molecule and causes condensation and subsequent expansion of the monolayer, surface pressure isotherms were obtained with molecules modeling the backbone or headgroup portions of the GM1 molecule. The observed concentration-dependent condensing and fluidizing effects are specific to the rigid, sugar headgroup portion of the GM1 molecule.  相似文献   

8.
Because of their surface localization, G protein-coupled receptors (GPCRs) are often pharmaceutical targets as they respond to a variety of extracellular stimuli (e.g., light, hormones, small molecules) that may activate or inhibit a downstream signaling response. The adenosine A2A receptor (A2AR) is a well-characterized GPCR that is expressed widely throughout the human body, with over 10 crystal structures determined. Truncation of the A2AR C-terminus is necessary for crystallization as this portion of the receptor is long and unstructured; however, previous work suggests shortening of the A2AR C-terminus from 412 to 316 amino acids (A2AΔ316R) ablates downstream signaling, as measured by cAMP production, to below that of constitutive full-length A2AR levels. As cAMP production is downstream of the first activation event—coupling of G protein to its receptor—investigating that first step in activation is important in understanding how the truncation effects native GPCR function. Here, using purified receptor and Gαs proteins, we characterize the association of A2AR and A2AΔ316R to Gαs with and without GDP or GTPγs using surface plasmon resonance (SPR). Gαs affinity for A2AR was greatest for apo-Gαs, moderately affected in the presence of GDP and nearly completely ablated by the addition of GTPγs. Truncation of the A2AR C-terminus (A2AΔ316R) decreased the affinity of the unliganded receptor for Gαs by ~20%, suggesting small changes to binding can greatly impact downstream signaling.  相似文献   

9.
All tRNAs have numerous modifications, lack of which often results in growth defects in the budding yeast Saccharomyces cerevisiae and neurological or other disorders in humans. In S. cerevisiae, lack of tRNA body modifications can lead to impaired tRNA stability and decay of a subset of the hypomodified tRNAs. Mutants lacking 7-methylguanosine at G46 (m7G46), N2,N2-dimethylguanosine (m2,2G26), or 4-acetylcytidine (ac4C12), in combination with other body modification mutants, target certain mature hypomodified tRNAs to the rapid tRNA decay (RTD) pathway, catalyzed by 5’-3’ exonucleases Xrn1 and Rat1, and regulated by Met22. The RTD pathway is conserved in the phylogenetically distant fission yeast Schizosaccharomyces pombe for mutants lacking m7G46. In contrast, S. cerevisiae trm6/gcd10 mutants with reduced 1-methyladenosine (m1A58) specifically target pre-tRNAiMet(CAU) to the nuclear surveillance pathway for 3’-5’ exonucleolytic decay by the TRAMP complex and nuclear exosome. We show here that the RTD pathway has an unexpected major role in the biology of m1A58 and tRNAiMet(CAU) in both S. pombe and S. cerevisiae. We find that S. pombe trm6Δ mutants lacking m1A58 are temperature sensitive due to decay of tRNAiMet(CAU) by the RTD pathway. Thus, trm6Δ mutants had reduced levels of tRNAiMet(CAU) and not of eight other tested tRNAs, overexpression of tRNAiMet(CAU) restored growth, and spontaneous suppressors that restored tRNAiMet(CAU) levels had mutations in dhp1/RAT1 or tol1/MET22. In addition, deletion of cid14/TRF4 in the nuclear surveillance pathway did not restore growth. Furthermore, re-examination of S. cerevisiae trm6 mutants revealed a major role of the RTD pathway in maintaining tRNAiMet(CAU) levels, in addition to the known role of the nuclear surveillance pathway. These findings provide evidence for the importance of m1A58 in the biology of tRNAiMet(CAU) throughout eukaryotes, and fuel speculation that the RTD pathway has a major role in quality control of body modification mutants throughout fungi and other eukaryotes.  相似文献   

10.
The dynamics of the opening-closing of the constituent base-pairs as well as of the exchange kinetics of the base-paired imino and amino protons with water in a DNA-RNA hybrid, [5′r(G1A2U3U4G5A6A7)3′]:5′p[d(T8C9A10A11T12C13)]3′-Pzn] duplex (I), are reported here in details for the first time. The exchange kinetics of amino and imino protons in the DNA-RNA hybrid (duplex I) have been compared with identical studies on the following B-DNA duplexes: d(C1G2T3A4C5G6)2 (II), d[p(5′T1G2T3T4T5G6 G7C8)3′]:d[p(5′C9C10A11A12A13C14A15)3′] (III), d(C5G6C7G8A9A10T11T12C13G14C15G16)2 (IV) and d(C1G2C3G4C5G6C7G8A9A10T11T12C13G14C15G16C17G18C19G20)2 (V). This comparative study shows that the life-times τo of various base-pairs in the DNA-RNA hybrid (I) varies in the range of ∼ 1 ms, and they are quite comparable to those of the shorter B-DNA duplexes (II) and (III), but very different from the τo of the larger duplexes (IV) and (V): the τo for the base pair of T11 and T12 residues in the 20-mer (duplex V) are 2.9 ± 2.3 ms and 23.2 ± 8.9 ms, respectively, while the corresponding τo in the 12-mer (duplex IV) are 2.8 ± 2.2 ms and 17.4 ± 5.4 ms. It has also been shown that the total energy of activation (Ea) assessed from the exchange rates of both imino and amino protons, representing energetic contributions from both base-pair and helix opening-closing as well as from the exchange process of the imino protons from the open state with the bound water, is close to the Ea of the short B-DNA duplex (Ea ≈ 28–47 kcal/mol).  相似文献   

11.
The aim of this study was to prepare aptamer-modified liposomes loaded with gadolinium (Gd) to enhance the effective diagnosis for tumor by MRI. A modified GBI-10 (GBI-10m) was used to prepare targeted liposomes (GmLs). Liposomes with GBI-10 aptamer (GLs) and without aptamer (non-targeted liposomes (NLs)) were also prepared as controls. The particle size and zeta potential of GmLs, GLs, and NLs were all assayed. A clinical 3.0 T MR scanner was employed to assess the imaging efficiency and measure the longitudinal relaxivity (r 1) of the above liposomes. Confocal laser scanning microscopy and flow cytometry were used to analyze and compare the targeting effects of GmLs, GLs, and NLs to MDA-MB-435s cells at 37°C. The particle size of the prepared liposomes was scattered at 100–200 nm, and their values of r 1 were ~4 mM?1 s?1. The images of confocal laser scanning microscopy showed that GmLs in the cytoplasm were significantly more than GLs and both GmLs and GLs were more than NLs. The fluorescence intensity of GmLs was increased by about two times than that of GLs and three times than that of NLs by flow cytometry. Therefore, the GmLs in this initial study were suggested to be a potential MRI contrast agent at 37°C for diagnosing tumors with the protein of tenascin-C over-expressed.  相似文献   

12.
Gd1a, Gd1b and Gt1b gangliosides were dispersed in the following membrane-mimicking systems: (a) homogeneous micelles; (b) mixed micelles with Gm1 ganglioside (which is resistant to the enzyme action), Triton X-100 or bovine serum albumin; (c) small unilamellar vesicles of egg phosphatidylcholine. The effect of dispersion on sialic acid release by Vibrio cholerae sialidase was studied. As reference substrates freely interacting with the enzyme the lipid-free carbohydrates of Gd1a and 3′-sialosyl-lactose were employed. The apparent Vmax. of the enzyme was, with all the gangliosides, dependent on the type of ganglioside dispersion. It was lowest for homogeneous micelles and mixed micelles with ganglioside Gm1, and increased about 6-fold for ganglioside/bovine serum albumin lipoprotein micelles, 15-fold for mixed-ganglioside/Triton X-100 micelles (optimal molar ratio 1:7.5) and 30-fold for phosphatidylcholine vesicles containing 2.5 mol% ganglioside (this proportion was optimal for enzyme activity on the vesicles). For ganglioside Gd1a, the activity on Triton X-100 mixed micelles and on mixed vesicles was even greater (3- and 6-fold respectively) than that displayed on Gd1a lipid-free carbohydrate. With each of the used gangliosides the apparent Km values were very similar values for homogeneous micelles and vesicular dispersions, but showed marked increases for Triton X-100 mixed micelles, approaching the values exhibited by reference oligosaccharides. Triton X-100 micelles and phosphatidylcholine vesicles did not appreciably alter the kinetics of sialidase action on 3′-sialosyl-lactose and on Gd1a lipid-free carbohydrate, indicating that the above effects are dependent on the intrinsic characteristics of the membrane-like systems containing gangliosides.  相似文献   

13.
Pentane extraction of lyophilized mitochondria with depletion of up to 92% of endogenous ubiquinone (UQ) does not affect ubiquinol oxidase activity in terms of Km; in certain preparations the V is decreased probably because the extraction is harmful to the membrane integrity. In such case dl-α-tocopherol is able to maintain enzymatic activity up to the normal values found in control mitochondria. On the other hand, NADH-UQ-1 reductase activity is greatly affected by pentane extraction with a large decrease in V but no change in Km, but this activity is protected by addition of dl-α-tocopherol to the extraction medium. The same conclusions can be drawn for succinate-UQ-1 reductase activity. In conclusion it appears that endogenous UQ does not mediate the interaction of exogenous UQ-1 with the redox sites for UQ in the respiratory chain.  相似文献   

14.
To obtain information concerning the effects of ionophores on biological membranes, the thermotropic behavior of ionophores such as gramicidin A′ and valinomycin in monolayers was investigated by measuring the surface pressure–area (πA) and the surface viscosity-area (ηsA) isotherms. Gramicidin A’ had an isotherm having the transition from a liquid-expanded through an intermediate to a condensed state, while valinomycin had a concave isotherm. The πA isotherms for two ionophores had a decremental shift with increasing temperatures, depending upon a variety of their molecular structures. A distinct difference between the two ionophores in ηsA isotherms was observed. In addition, the interaction of dimyristoylphosphatidylcholine (DMPC) with the two ionophores in mixed monolayers was investigated. When valinomycin was mixed with DMPC, no deviation from the additivity rule occurred below and above the phase transition temperature Tc of DMPC. However, when gramicidin A′ was mixed with DMPC, a considerable negative deviation from ideal mixing occurred below Tc, suggesting the formation of an irregular ripple structure.  相似文献   

15.
《Process Biochemistry》2007,42(2):285-288
In this paper, an adsorbent with dimethylamine ligand for endotoxin removal was prepared and used to study the adsorption isotherms of endotoxin and BSA. The results showed that as the introducing of endotoxin, the maximum adsorption capacity qm of BSA increased from 7.24 to 7.74 mg/mL and the apparent association constant KA of BSA decreased from 14.06 to 11.48 mL/mg. The adsorption isotherms of BSA changed from Langmuir model to Multilayer model. All these gave the direct evidences to construct the adsorption process. By comparing adsorption isotherms of endotoxin with BSA, it was found that the apparent association constant KA of endotoxin was much higher than that of BSA.  相似文献   

16.
The aim of this study was to isolate and evaluate the withanolides in inducing apoptotic like death in Leishmania donovani in vitro. Withanolides were fractionated and isolated from the leaves of Withania somnifera and LC-MS/MS analysis of two fractions namely, F5 and F6 of ethanolic extracts, obtained through column chromatography with silica gel, was performed. The antileishmanial effect of withanolides on L. donovani promastigotes was assessed in vitro using PI dye exclusion test. The effect of withanolides on promastigote morphology was determined by scanning electron microscopy. To understand their mode of action against L. donovani, DNA fragmentation, quantification of parasites at sub G0/G1 phase, determination of phosphatidylserine externalization, measurement of reactive oxygen species (ROS) and mitochondrial membrane potential (Ψm) were done. Results showed that LC–MS/MS analysis confirmed the presence of withanolides in isolated fractions. Treatment with withanolides resulted in morphological alterations from spindle to round shape and loss of flagella/cell integrity in promastigotes. Moreover, it induced DNA nicks, cell cycle arrest at sub G0/G1 phase and externalization of phosphatidylserine in dose and time dependent manner via increase in ROS and decrease in Ψm. Results of this study indicate that withanolides induce apoptotic like death through the production of ROS from mitochondria and disruption of Ψm in promastigotes of L donovani.  相似文献   

17.
Lanthanides such as La3+ and Gd3+ are well known to have large effects on the function of membrane proteins such as mechanosensitive ionic channels and voltage-gated sodium channels, and also on the structure of phospholipid membranes. In this report, we have investigated effects of La3+ and Gd3+ on the shape of giant unilamellar vesicle (GUV) of dioleoylphosphatidylcholine (DOPC-GUV) and GUV of DOPC/cholesterol by the phase-contrast microscopy. The addition of 10-100 μM La3+ (or Gd3+) through a 10-μm diameter micropipette near the DOPC-GUV (or DOPC/cholesterol-GUV) triggered several kinds of shape changes. We have found that a very low concentration (10 μM) of La3+ (or Gd3+) induced a shape change of GUV such as the discocyte via stomatocyte to inside budded shape transformation, the two-spheres connected by a neck to prolate transformation, and the pearl on a string to cylinder (or tube) transformation. To understand the effect of these lanthanides on the shape of the GUV, we have also investigated phase transitions of 30 μM dipalmitoylphosphatidylcholine-multilamellar vesicle (DPPC-MLV) by the ultra-sensitive differential scanning calorimetry (DSC). The chain-melting phase transition temperature and the Lβ′ to Pβ′ phase transition temperature of DPPC-MLV increased with an increase in La3+ concentration. This result indicates that the lateral compression pressure of the membrane increases with an increase in La3+ concentration. Thereby, the interaction of La3+ (or Gd3+) on the external monolayer membrane of the GUV induces a decrease in its area (Aex), whereas the area of the internal monolayer membrane (Ain) keeps constant. Therefore, the shape changes of the GUV induced by these lanthanides can be explained reasonably by the decrease in the area difference between two monolayers (ΔA=AexAin).  相似文献   

18.
《Inorganica chimica acta》1988,149(1):151-154
The extraction equilibrium of the hydronium-uranium(VI)-dicyclohexano-24-crown-8 complex was carried out in the crown ether1,2-dichloroethaneHCl aqueous solution system at different temperatures. The extraction complex has the overall composition (L)2·(H3O+·χH2O)2·UO2Cl42− (L = dicyclohexano-24-crown-8). The values of the extraction equilibrium constants (Kex) increase steadily with a decrease in temperature: 13.5 (298 K), 7.96 (301 K), 4.20 (303 K) and 2.07 (305 K). A plot of log Kex against 1/T shows a straight line. The value of the enthalpy change, ΔH°, was calculated from the slope and equals −212 kJ mol−1. The value of the entropy change, ΔS°, was calculated from ΔH° and Kex and equals −690 J K−1 mol−1, whereas ΔG° = −6.45 kJ mol−1. Comparing these thermodynamic parameters with those of the dicyclohexano-18-crown-6 isomer A [1] (ΔS° = −314 J K−1 mol−1, ΔH° = −101 kJ mol−1 and ΔG° = −8.37 kJ mol−1), it can be seen that ΔH° and ΔS° are more negative for the former than for the latter, and both are enthalpy-stabilized complexes. The molecular structure of the complex has the feature that there are two H5O2+ ions in it, in contrast to the H3O+ ions in the dicyclohexano-18-crown-6 isomer A complex [1]. Each of the H5O2+ ions is held in the crown ether cavity by four hydrogen bonds. The H5O2+ ion has a central bond. The uranium atom forms UO2Cl42− as a counterion away from the crown ether. The formation of this complex is in good agreement with more negative entropy change and less negative free energy change, as mentioned above.  相似文献   

19.
Ubiquinone (UQ) is the only natural compound which was reported both to generate and to scavenge oxygen-derived radicals. Redox-cycling of ubiquinone may yield six different species of the parent compound: UQH2, UQH, UQ2−, UQH, UQ•−, and UQ. Ubiquinol (UQH2) is unequivocally considered to be the ubiquinone species capable of scavenging oxygen-derived radicals. In contrast, the ubiquinone species responsible for one ereduction of dioxygen (O2) thereby initiating the cascade of oxidative stress is still a matter of controversial debate. In the present study this question was approached by following the effect of O2on the stability of the various reduced forms of UQ. For this purpose conditions were designed allowing the selective accumulation of the two protonated and of the two deprotonated forms of reduced ubiquinones. Our results exclude both protonated (ubiquinol, UQH2) and anionic (ubiquinol anion, UQH, and ubiquinol dianion, UQ2−) fully reduced ubiquinones as the source exerting one ereduction of O2. Ubisemiquinone (semiquinone radical, UQH), when protonated, underwent rapid disproportionation, while transition to the semireduced anionic form (semiquinone radical anion, UQ•−) was found to favor autoxidation. The results obtained in this study provide a chemical base for the assessment of one etransfer from redox-cycling UQ to O2in the respiratory chain and in biomembranes where ubihydroquinol is suggested to exert antioxidant activities.  相似文献   

20.
Dehalococcoides species have a highly restricted lifestyle and are only known to derive energy from reductive dehalogenation reactions. The lipid fraction of two Dehalococcoides isolates, strains BAV1 and FL2, and a tetrachloroethene-to-ethene-dechlorinating Dehalococcoides-containing consortium were analyzed for neutral lipids and phospholipid fatty acids. Unusual phospholipid modifications, including the replacement of unsaturated fatty acids with furan fatty acids, were detected in both Dehalococcoides isolates and the mixed culture. The following three furan fatty acids are reported as present in bacterial phospholipids for the first time: 9-(5-pentyl-2-furyl)-nonanoate (Fu18:2ω6), 9-(5-butyl-2-furyl)-nonanoate (Fu17:2ω5), and 8-(5-pentyl-2-furyl)-octanoate (Fu17:2ω6). The neutral lipids of the Dehalococcoides cultures contained unusually large amounts of benzoquinones (i.e., ubiquinones [UQ]), which is unusual for anaerobes. In particular, the UQ-8 content of Dehalococcoides was 5- to 20-fold greater than that generated in aerobically grown Escherichia coli cultures relative to the phospholipid fatty acid content. Naphthoquinone isoprenologues (MK), which are often found in anaerobically grown bacteria and archaea, were also detected. Dehalococcoides shows a difference in isoprenologue pattern between UQ-8 and MK-5 that is atypical of other bacteria capable of producing both quinone types. The difference in UQ-8 and MK-5 isoprenologue patterns strongly suggests a special function for UQ in Dehalococcoides, and Dehalococcoides may utilize structural modifications in its lipid armamentarium to protect against free radicals that are generated in the process of reductive dechlorination.  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