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 31 毫秒
1.
The kinetics of substitution reactions of [η-CpFe(CO)3]PF6 with PPh3 in the presence of R-PyOs have been studied. For all the R-PyOs (R = 4-OMe, 4-Me, 3,4-(CH)4, 4-Ph, 3-Me, 2,3-(CH)4, 2,6-Me2, 2-Me), the reactions yeild the same product [η5-CpFe(CO)2PPh3]PF6, according to a second-order rate law that is first order in concentrations of [η5-CpFe(CO)3]PF6 and of R-PyO but zero order in PPh3 concentration. These results, along with the dependence of the reaction rate on the nature of R-PyO, are consistent with an associative mechanism. Activation parameters further support the bimmolecular nature of the reactions: ΔH = 13.4 ± 0.4 kcal mol−1, ΔS = −19.1 ± 1.3 cal k−1 mol−1 for 4-PhPyO; ΔH = 12.3 ± 0.3 kcal mol−1, ΔS = 24.7 ±1.0 cal K−1 mol−1 for 2-MePyO. For the various substituted pyridine N-oxides studied in this paper, the rates of reaction increase with the increasing electron-donating abilities of the substituents on the pyridine ring or N-oxide basicities, but decrease with increasing 17O chemical shifts of the N-oxides. Electronic and steric factors contributing to the reactivity of pyridine N-oxides have been quantitatively assessed.  相似文献   

2.
The seasonal variability of specific growth rate and the carbon stable isotope ratio (δ13C) of leaf blades (δ13Cleaf) of a temperate seagrass, Zostera marina (within 10 days old) were measured simultaneously, together with the δ13C of dissolved inorganic carbon (δ13CDIC) at three sites in the semi-closed Akkeshi estuary system, northeastern Japan, in June, September, and November 2004. The δ13Cleaf ranged from −16.2 to −6.3‰ and decreased from summer to winter. The simultaneous measurement of the δ13Cleaf, growth rate, and morphological parameters (mean leaf length and width, mean number of leaves per shoot, and sheath length) of the seagrass and δ13CDIC in the surrounding water allowed us to compare directly the δ13Cleaf and specific growth rate of seagrass. The difference in the δ13C of seagrass leaves relative to the source DIC (Δδ13Cleaf − DIC) was the least negative (−11 to −7‰) in June at all three sites and became more negative (−17 to −8‰) as the specific growth rate decreased. This positive correlation between Δδ13Cleaf − DIC and specific growth rate can be used to diagnose the growth of seagrasses. Δδ13Cleaf − DIC changed by −1.7 ± 0.2‰ when the leaf specific growth rate decreased by 1% d−1.  相似文献   

3.
Statistical analysis reveals that the set of differences between the secondary shifts of the α- and β-carbons for residues i of a protein (Δδ13Cαi- Δδ13Cβi) provides the means to detect and correct referencing errors for 1H and 13C nuclei within a given dataset. In a correctly referenced protein dataset, linear regression plots of Δδ13Cαi,Δδ13Cβi, or Δδ1Hαi vs. (Δδ13Cαi- Δδ13Cβi) pass through the origin from two directions, the helix-to-coil and strand-to-coil directions. Thus, linear analysis of chemical shifts (LACS) can be used to detect referencing errors and to recalibrate the 1H and 13C chemical shift scales if needed. The analysis requires only that the signals be identified with distinct residue types (intra-residue spin systems). LACS allows errors in calibration to be detected and corrected in advance of sequence-specific assignments and secondary structure determinations. Signals that do not fit the linear model (outliers) deserve scrutiny since they could represent errors in identifying signals with a particular residue, or interesting features such as a cis-peptide bond. LACS provides the basis for the automated detection of such features and for testing reassignment hypotheses. Early detection and correction of errors in referencing and spin system identifications can improve the speed and accuracy of chemical shift assignments and secondary structure determinations. We have used LACS to create a database of offset-corrected chemical shifts corresponding to nearly 1800 BMRB entries: 300 with and 1500 without corresponding three-dimensional (3D) structures. This database can serve as a resource for future analysis of the effects of amino acid sequence and protein secondary and tertiary structure on NMR chemical shifts.Supplementary material to this paper is available in electronic form at http://dx.doi.org/10.1007/s10858-005-1717-0  相似文献   

4.
Equilibrium constants for the binding of a series of alkyl isocyanides to ferrous cytochrome c' from Rhodospirillum molischianum have been measured spectrophotometrically. The equilibrium constants range from 3.3 M-1 to 2.6 x 10(2) M-1 and follow the order methyl greater than ethyl less than n-propyl less than tert-butyl less than n-butyl less than amyl less than cyclohexyl less than n-hexyl. The decrease in equilibrium constant from methyl to ethyl isocyanide provides evidence for a steric interaction between the ligand and the protein. The increase in equilibrium constant from ethyl to n-hexyl isocyanide is accounted for by a favorable partitioning of the ligand into a hydrophobic heme coordination site. The effect of steric interactions on the differences in the binding constants has been further evaluated by comparing the alkyl isocyanide and CO binding constants for the ferrous cytochrome c' to those of a sterically unconstrained model heme complex in a detergent micelle. The results indicate that the heme coordination site of the ferrous cytochrome c' is severely sterically hindered, similar to that of the reported crystal structure of Rs. molischianum ferric cytochrome c'.  相似文献   

5.
The trophic ecology of marine vertebrates has been increasingly studied via stable isotope analysis of body tissues. However, the theoretical basis for using stable isotopes to elucidate consumer–prey relationships remains poorly validated for most taxa despite numerous studies using this technique in natural systems. In this study, we measured stable carbon and stable nitrogen diet-tissue discrimination (Δdt) in whole blood, red blood cells, blood plasma solutes, and skin of leatherback sea turtles (Dermochelys coriacea; N = 7) maintained in captivity for up to 424 days and fed an isotopically consistent control diet with a mean C:N ratio of 2.94:1.00 and an energetic content of 20.16 ± 0.39 kJ g− 1 Dry Mass. We used a random-effect repeated measure model to evaluate isotopic consistency among tissue samples collected on days 276, 348, and 424. Both δ13C and δ15N remained consistent among sampling events in all tissues (all 95% posterior intervals for the slopes of a linear model included zero), indicating that all tissues had fully integrated diet-derived stable isotope compositions. Mean tissue-specific δ13C ranged from − 18.30 ± 0.16‰ (plasma solutes) to − 15.54 ± 0.14‰ (skin), whereas mean δ15N was from 10.06 ± 0.22‰ (whole blood) to 11.46 ± 0.10‰ (plasma solutes). The computed Δdt factors for carbon ranged from − 0.58‰ (plasma solutes) to + 2.25‰ (skin), whereas Δdt for nitrogen was from + 1.49 (red blood cells) to + 2.85 (plasma solutes). As the only discrimination factors available for leatherback turtles, our data will be useful for future interpretations of field-derived stable isotope data for this species. The inherent variability in Δdt values among individuals was low, which supports the value of these data for dietary reconstructions. However, it is important to note that tissue-specific discrimination factors for leatherbacks contrast with the widely accepted values for endothermic species (0–1‰ for C, 3–5‰ for N), and are also different from values established for hard-shelled turtles. This underscores the need for species- and tissue-specific discrimination factors before interpreting trophic studies of wild animals, including marine turtles.  相似文献   

6.
Three series of new cannabinoids were prepared and their affinities for the CB1 and CB2 cannabinoid recptors were determined. These are the 1-methoxy-3-(1′,1′-dimethylalkyl)-, 1-deoxy-11-hydroxy-3-(1′,1′-dimethylalkyl)- and 11-hydroxy-1-methoxy-3-(1′,1′-dimethylalkyl)-Δ8-tetrahydrocannabinols, which contain alkyl chains from dimethylethyl to dimethylheptyl appended to C-3 of the cannabinoid. All of these compounds have greater affinity for the CB2 receptor than for the CB1 receptor, however only 1-methoxy-3-(1′,1′-dimethylhexyl)-Δ8-THC (JWH-229, 6e) has effectively no affinity for the CB1 receptor (Ki=3134±110 nM) and high affinity for CB2 (Ki=18±2 nM).  相似文献   

7.
Association constants for N---H+O hydrogen bond formation between substituted ammonium dications and phenolate ion were measured in water and deuterium oxide at 25°C and 2.0 ionic strength. In combination with isotopic fractionation factors for phenol and the conjugate diacid of 1,2-ethanediamine determined by 13C NMR spectroscopy, these yield isotopic fractionation factors for amine dication-phenolate ion hydrogen bonds in water: φAB = 0.69 for 1,2-propanediamine dication with a pK difference between donor and acceptor, ΔpKa = −0.45, φAB = 0.88 for 1,2-ethanediamine dication (ΔpKa = −2.1), and φAB = 1.1 for piperizine dication (ΔpKa = −3.5). The hydrogen bond association constants follow Brønsted correlations α = 0.19 in water and α = 0.27 in deuterium oxide. The results are consistent with a double-minimum potential with a significant barrier for motion across the hydrogen bond.  相似文献   

8.
The kinetics of formation of the complex ion, μ-carbonato-di-μ-hydroxo-bis((1,5-diamino-3-aza-pentane) cobalt(III), from the tri-μ-hydroxo-bis((1,5-diamino-3-aza-pentane(III)cobalt(III)) ion in aqueous buffered carbonate solution have been studied spectrophotometrically at 295 nm over the ranges 20.0θ°C34.8, 8.03pH9.44, 5 mM [CO32−35 mM and at an ionic strength of 0.1 M (LiClO4). On the basis of the kinetic results a mechanism, involving rapid cleavage of an hydroxo bridge followed by carbon dioxide uptake with subsequent bridge formation, has been proposed. At 25 °C, the rate of the carbon dioxide uptake is 0.58 M−1 s−1 with ΔH≠ = (13.2±0.7) kcal mol−1 and ΔS≠ = (−15.1 ± 0.7) cal deg−1 mol−1. The results are composed with those obtained for several mononuclear cobalt(III) and one dinuclear cobalt(III) complexes.  相似文献   

9.
We studied the effect of intracellular glutathione (GSH), which was known to conjugate readily with an α, β-unsaturated carbonyl of 9-deoxy-Δ9,12-13,14-dihydro PGD212-PGJ2), on the cytotoxicity of Δ12-PGJ2. Δ12-PGJ2 caused DNA fragmentation in human hepatocellular carcinoma Hep 3B cells, which was blocked by cycloheximide (CHX). The Δ12-PGJ2-induced apoptosis was augmented by GSH depletion resulted from pretreatment with buthioninine sulfoximine (BSO), an inhibitor of γ-glutamylcysteine synthetase. On the contrary, N-acetyl-cysteine (NAC), a precursor of cysteine, elevated the GSH level and protected cells from initiating apoptosis by Δ12-PGJ2. Sodium arsenite, a thiol-reactive agent, also induced apoptosis, which was potentiated or attenuated by BSO or NAC treatment respectively. These results suggest that the apoptosis-inducing activity of Δ12-PGJ2 is due to thiol-reactivity and intracellular GSH modulates the Δ12-PGJ2-induced apoptosis by regulating the accessibility of Δ12-PGJ2 to target proteins containing thiol groups.  相似文献   

10.
The synthesis and pharmacology of 15 1-deoxy-Δ8-THC analogues, several of which have high affinity for the CB2 receptor, are described. The deoxy cannabinoids include 1-deoxy-11-hydroxy-Δ8-THC (5), 1-deoxy-Δ8-THC (6), 1-deoxy-3-butyl-Δ8-THC (7), 1-deoxy-3-hexyl-Δ8-THC (8) and a series of 3-(1′,1′-dimethylalkyl)-1-deoxy-Δ8-THC analogues (2, n=0–4, 6, 7, where n=the number of carbon atoms in the side chain−2). Three derivatives (1719) of deoxynabilone (16) were also prepared. The affinities of each compound for the CB1 and CB2 receptors were determined employing previously described procedures. Five of the 3-(1′,1′-dimethylalkyl)-1-deoxy-Δ8-THC analogues (2, n=1–5) have high affinity (Ki=<20 nM) for the CB2 receptor. Four of them (2, n=1–4) also have little affinity for the CB1 receptor (Ki=>295 nM). 3-(1′,1′-Dimethylbutyl)-1-deoxy-Δ8-THC (2, n=2) has very high affinity for the CB2 receptor (Ki=3.4±1.0 nM) and little affinity for the CB1 receptor (Ki=677±132 nM).
Scheme 3. (a) (C6H5)3PCH3+ Br, n-BuLi/THF, 65°C; (b) LiAlH4/THF, 25°C; (c) KBH(sec-Bu)3/THF, −78 to 25°C then H2O2/NaOH.  相似文献   

11.
Complexation of M+=Li+, Na+, Ag+ and TI+ by the cryptands 4, 7, 13, 18-tetraoxa-l, 10-diazabicyclo[8.5.5]eicosane (C211) and 4,7,13-trioxa-1,10-diazabicyclo[8.5.5]eicosane (C21C5) to form the cryptates [M.C211]+ and [M.C21C5]+ has been studied in trimethyl phosphate by potentiometric titration and 7Li and 23Na NMR spectroscopy. For [M.C211]+ the logarithm of the apparent stability constants, log K (dm3 mol-1)=6.98±0.05, 5.38±0.05, 9.82±0.02 and 3.95±0.02 for M+ =Li+, Na+, Ag+ and TI+, respectively; and for [M.C21C5]+ log K (dm3 mol-1)=2.40±0.10, 1.90±0.05, 6.04±0.02 and 2.42±0.10 for M+=Li+, Na+, Ag+ and Tl+, respectively. The decomplexation kinetic parameters for [Na.C211]+ are: kd (298.2 K)=6.924±0.50 s-l, ΔHd≠=62.2±0.9 kJ mol-1, and ΔSd≠= -20.3±2.7 J K-1 mol-1; and those for [Li.C21C5]+ are: kd (298.2 K)=23.3±0.4 s-1, ΔHd≠ =61.2±1.1 kJ mol-1, and ΔSd≠= -13.6±3.6 J K-1 mol-1. Metal ion exchange on [Li.C211]+ is in the very slow extreme of the NMR timescale up to 390 K and kd « 4 s-1 at 298.2 K, while in contrast exchange on [Na.C21C5]+ is in the fast extreme of the NMR timescale at 298.2 K (kd≈ 104 s-1). These data are compared with those obtained in other solvents.  相似文献   

12.
Ingestion of red kidney bean phytohemagglutinin causes impaired growth and intestinal malabsorption, and facilitates bacterial colonization in the small intestine of weanling rats. We have studied interactions of the highly purified phytohemagglutinin erythroagglutinating (E4) and mitogenic (L4) isolectins with microvillous membrane vesicles prepared from rat small intestines. E4 and L4 were radioiodinated with 125I by the chloramine-T technique. E4 and L4 isolectins both bound to microvillous membrane vesicles. Binding was saturable and reversible. Each mg of membrane protein bound 744±86 μg E4 and 213±21 μg L4. The apparent Ka for E4 and L4 binding was 2.5·10−6 and 13.0·10−6 M−1, respectively. Binding of each 125I-labelled isolectin was abolished by 100-fold excess of unlabelled isolectin. In each case binding also was inhibited by appropriate oligosaccharide inhibitors, indicating that isolectin-microvillous membrane interactions were mediated by carbohydrate recognition. Patterns of saccharide inhibition of isolectin binding were different for E4 and L4. Competitive binding experiments demonstrated mutual noncompetitive inhibition of E4 and L4 binding consistent with steric hindrance. Therefore, E4 and L4 each bound to its own set of receptors. Based on the known saccharide specificities of E4 and L4, these data indicate that there are differences in expression of complex asparagine-linked biantennary and tri- or tetraantennary oligosaccharides at the microvillous surface. The data also provide the possibility that direct interactions of one or more phytohemagglutinin isolectins with intestinal mucosa in vivo may contribute to the antinutritional effects associated with ingestion of crude red kidney beans.  相似文献   

13.
γ-Aminobutyraldehyde dehydrogenase from Escherichia coli K-12 has been purified and characterized from cell mutants able to grow in putrescine as the sole carbon and nitrogen source. The enzyme has an Mr of 195 000±10 000 in its dimeric form with an Mr of 95 000±1000 for each subunit, a pH optimum at 5.4 in sodium citrate buffer, and does not require bivalent cations for its activity. Km values are 31.3±6.8 μM and 53.8±7.4 μM for Δ-1-pyrroline and NAD+, respectively. An inhibitory capacity for NADH is also shown using the purified enzyme.  相似文献   

14.
Amniotic fluid from 15–17-day rat fetuses bound estrone and 17β-estradiol specifically. Related steroids such as estriol, 6-ketoestradiol, 17α-estradiol and testosterone were not bound to any significant extent. The apparent Ka for 17β-estradiol was 2.6·108 M at 4°C; 6 nmoles of 17β-estradiol were bound per ml of amniotic fluid. The binding component appears to be α-fetoprotein in that it migrates as an α1-globulin upon polyacrylamide gel electrophoresis and has an isoelectric pH of 4.7 as determined by isoelectric focusing. Furthermore, binding activity was precipitated by antiserum which was shown by immuno-electrophoresis to be specific for α-fetoprotein. Binding activity, partially purified by isoelectric focusing of amniotic fluid, was associated with one of two bands seen by polyacrylamide gel electrophoresis. This band migrated as an α1-globulin.  相似文献   

15.
Reaction of sodium or potassium molybdate and excess malic acid in a wide range of pH values (pH 4.0–7.0) resulted in the isolation of two cis-dioxo-bis(malato)-Mo(VI) complexes, viz. Na3[MoO2H(S-mal)2] and K3[MoO2H(S-mal)2]·H2O (H3mal=malic acid). The sodium complex is also characterized by an X-ray structure analysis, showing that the mononuclear Mo units are linked together via very strong symmetric CO2···H··· O2C-hydrogen bond [2.432(5) Å], forming a polymeric chain. The molybdenum atoms are quasi-octahedrally coordinated by two cis-oxo groups and two bidentate malate ligands via its alkoxy and α-carboxyl groups, while the β-carboxylic and carboxylate groups remain uncomplexed, as the coordination of vicinal carboxylate and alkoxide of homocitrate in FeMo cofactor of nitrogenase. The absolute configuration of the metal center in this S-malato complex is assigned as Λ and the homochirality within the chain is established as a homochiral form ···ΛS–ΛS–ΛS–ΛS···. It is proposed that the chiral configuration of the metal center in wild-type FeMo-co biosynthesis might be induced by the early coordination of the chiral R-homocitric acid, while a mixture of raceme might be obtained in the biosynthesis of NifV FeMo-cofactor. The absolute configuration of wild-type FeMo-cofactor is assigned as ΔR.  相似文献   

16.
The cross-reactivity of the PGI3 metabolite, Δ17-6-keto-PGF, with antibodies against 6-keto-PGF for radioimmunoassays (RIA) has been investigated. Δ17-6-keto-PGF was obtained either from commercial sources or after its purification from endothelial cells. In the latter case, primary cultured bovine aortic endothelial cells were incubated for 20 min at 37°C with 10 μM eicosapentaenoic acid (EPA) in the presence of 2 μM 13-hydroperoxy-octadecadienoic acid, an activator of the EPA cyclooxygenation, and the 6-keto-PGF and Δ17-6keto-PGF produced were separated by RP-HPLC. Then, cross-reactivities of the commercial and purified Δ17-6-keto-PGF with 6-keto-PGF antibodies were determined and found not to exceed 10%. In addition, the amounts of prostacyclin-related compounds detected by direct measurements in media of cells loaded with EPA were compared with those obtained after purification of 6-keto-PGF. In accordance with the cross-reactivity data, we found that RIA in media mainly measured 6-keto-PGF, the Δ17-6-keto-PGF formed being undetected at 90%. It is concluded that 6-keto-PGF antibodies generally used for RIA of 6-keto-PGF are highly specific since they can discriminate a metabolite bearing an additional double bond such as the PGI3 metabolite Δ17-6-keto-PGF.  相似文献   

17.
, originally introduced as an inadvertent contaminant in solutions used for evaluating the stability of prostaglandins, proved to lead to the rapid disappearance of the cyclopentenone unit of PGA2 (as monitored by circular dichroic spectroscopy). The cyclopentenone unit is converted, in various metabolites, to a 9-keto, 9α or 9β-hydroxy group lacking the ring unsaturation. The major EtoAc-soluble 9-hydroxy metabolite (Compound-I) was shown to be 9α, 15α-dihydroxy-2,3,4,5-tetranor-13- -prostenoic acid. Similar tetranor 9-hydroxy metabolites with one additional degree of unsaturation, and with a 9β-hydroxy group, also occur but these have not been fully characterized. Only two of the wide range of 9-keto metabolites are fully characterized by mass spectral (MS) data: 9,15-oxo-2,3,4,5-tetranorprostanoic acid and 9,15-oxo-2,3,4,5-tetranor-13- -prostenoic acid. The water soluble metabolites have not been characterized further.The fully characterized metabolites together with MS data from mixtures of minor metabolites indicate that can perform the following transformations: β-oxidation, dehydrogenation at C-15, reduction of the enone carbon-carbon double bonds (both Δ10,11 and Δ13,14), reduction of the 9-ketone, and possibly migration of the cyclopentyl double bond (Δ10,11 → Δ11,12). metabolizes 15-epimeric PGA2 equally readily with the production of similar products. PGA1 affords less 9-keto metabolites with compound I constituting 33% of the product by HPLC analysis. displays some enantioselectivity, PGA2 and 15-epi-PGA2 are each metabolized more rapidly than their enantiomers. Other prostaglandins appear to be less readily metabolized.  相似文献   

18.
α-Synuclein is an intrinsically disordered protein of 140 residues that switches to an α-helical conformation upon binding phospholipid membranes. We characterize its residue-specific backbone structure in free solution with a novel maximum entropy procedure that integrates an extensive set of NMR data. These data include intraresidue and sequential HN–Hα and HN–HN NOEs, values for 3JHNHα, 1JHαCα, 2JCαN, and 1JCαN, as well as chemical shifts of 15N, 13Cα, and 13C′ nuclei, which are sensitive to backbone torsion angles. Distributions of these torsion angles were identified that yield best agreement to the experimental data, while using an entropy term to minimize the deviation from statistical distributions seen in a large protein coil library. Results indicate that although at the individual residue level considerable deviations from the coil library distribution are seen, on average the fitted distributions agree fairly well with this library, yielding a moderate population (20–30%) of the PPII region and a somewhat higher population of the potentially aggregation-prone β region (20–40%) than seen in the database. A generally lower population of the αR region (10–20%) is found. Analysis of 1H–1H NOE data required consideration of the considerable backbone diffusion anisotropy of a disordered protein.  相似文献   

19.
The direct involvement of manganese peroxidase (MnP) in the mineralization of natural and xenobiotic compounds was evaluated. A broad spectrum of aromatic substances were partially mineralized by the MnP system of the white rot fungus Nematoloma frowardii. The cell-free MnP system partially converted several aromatic compounds, including [U-14C]pentachlorophenol ([U-14C]PCP), [U-14C]catechol, [U-14C]tyrosine, [U-14C]tryptophan, [4,5,9,10-14C]pyrene, and [ring U-14C]2-amino-4,6-dinitrotoluene ([14C]2-AmDNT), to 14CO2. Mineralization was dependent on the ratio of MnP activity to concentration of reduced glutathione (thiol-mediated oxidation), a finding which was demonstrated by using [14C]2-AmDNT as an example. At [14C]2-AmDNT concentrations ranging from 2 to 120 μM, the amount of released 14CO2 was directly proportional to the concentration of [14C]2-AmDNT. The formation of highly polar products was also observed with [14C]2-AmDNT and [U-14C]PCP; these products were probably low-molecular-weight carboxylic acids. Among the aliphatic compounds tested, glyoxalate was mineralized to the greatest extent. Eighty-six percent of the 14COOH-glyoxalate and 9% of the 14CHO-glyoxalate were converted to 14CO2, indicating that decarboxylation reactions may be the final step in MnP-catalyzed mineralization. The extracellular enzymatic combustion catalyzed by MnP could represent an important pathway for the formation of carbon dioxide from recalcitrant xenobiotic compounds and may also have general significance in the overall biodegradation of resistant natural macromolecules, such as lignins and humic substances.  相似文献   

20.
The pathway of propionate conversion in a syntrophic coculture of Smithella propionica and Methanospirillum hungatei JF1 was investigated by 13C-NMR spectroscopy. Cocultures produced acetate and butyrate from propionate. [3-13C]propionate was converted to [2-13C]acetate, with no [1-13C]acetate formed. Butyrate from [3-13C]propionate was labeled at the C2 and C4 positions in a ratio of about 1:1.5. Double-labeled propionate (2,3-13C) yielded not only double-labeled acetate but also single-labeled acetate at the C1 or C2 position. Most butyrate formed from [2,3-13C]propionate was also double labeled in either the C1 and C2 atoms or the C3 and C4 atoms in a ratio of about 1:1.5. Smaller amounts of single-labeled butyrate and other combinations were also produced. 1-13C-labeled propionate yielded both [1-13C]acetate and [2-13C]acetate. When 13C-labeled bicarbonate was present, label was not incorporated into acetate, propionate, or butyrate. In each of the incubations described above, 13C was never recovered in bicarbonate or methane. These results indicate that S. propionica does not degrade propionate via the methyl-malonyl-coenzyme A (CoA) pathway or any other of the known pathways, such as the acryloyl-CoA pathway or the reductive carboxylation pathway. Our results strongly suggest that propionate is dismutated to acetate and butyrate via a six-carbon intermediate.  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