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.
Studies of the degradation of the two isomeric forms of butyrate in different anaerobic environments showed isomerization betweenn- andi-butyrate. Degradation rates were similar for the different examined systems and degradation rates forn-butyrate degradation were generally higher than fori-butyrate. Degradation rates forn-butyrate ranged from 0.52 to 1.39 day–1, while the rates fori-butyrate were from 0.46 to 1.15 day–1. Production of isomers was not observed when the volatile fatty acid degradation was inhibited by addition of bromoethane sulfonic acid, indicating that isomerization was coupled to the methanogenic degradation of the acid. The degree of isomerization observed duringn-butyrate degradation was similar to the degree duringi-butyrate degradation. Experiments indicated that the isomerization degree was higher for the thermophilic than for the mesophilic inocula.  相似文献   

2.
《Biomass》1988,15(1):11-24
Calculations of the kinetics of volatile fatty acids removal in an anaerobic filter showed that methane production from n-butyrate was not consistent with β-oxidation when the reactor was operated under sequential feeding conditions. Under these conditions, n-butyrate might be anaerobically catabolised by decarboxylation. Iso-butyrate was formed at about 25% of the total amount of n-butyrate removed. After 13 days of continuous feeding, the material balance was consistent with the β-oxidation mechanism and the formation of iso-butyrate became negligible. N-butyrate decarboxylation may therefore be a transient catabolic pathway, although the reason why decarboxylation is preferred to β-oxidation under conditions such as sequential feeding is not clear. Results of similar experiments on acetate removal were consistent with theoretical expectations under conditions of both sequential and continous feeding.  相似文献   

3.
Two sulfur-dependent hyperthermophilic archaea, Desulfurococcus strain SY and Pyrococcus strain GB-D, which were isolated from deep-sea hydrothermal vents, utilized free amino acids and peptides obtained from various molecular size fractions of yeast extract. It was found that 11 amino acids were essential for growth. The metabolic products were acetate, i-butyrate, and i-valerate.  相似文献   

4.
Additional volatile compounds were isolated from muskmelon fruit by means of a water recycling apparatus, separated by GLC, and identified principally by MS and GLC retention data. Compounds reported for the first time as melon components are: n-hexanol, 1-octen-3-ol, cis-3-nonen-1-ol, n-butyl acetate, isobutyl acetate, 2-methylbutyl acetate, n-hexyl acetate, ethyl n-butyrate, ethyl 2-methylbutyrate, benzyl acetate, β-phenethyl acetate, and γ-phenylpropyl acetate. Muskmelon fruit stored frozen prior to steam distillation-extraction yielded an essence which, when compared with that obtained from freshly harvested fruit, contained considerably larger amounts of trans-2-nonenal, n-nonanol, cis-3-nonen-1-ol, cis-6-nonen-1-ol, and the methyl and ethyl esters of linoleic and linolenic acids. Marked decreases in the relative amounts of benzyl acetate, β-phenethyl acetate, and γ-phenylpropyl acetate resulted from freezing. All 21 compounds examined were present in the essences prepared from fresh, refrigerated, and frozen fruit.  相似文献   

5.
Samples taken from a fluidized-bed reactor revealed very low saturation constants for the degradation of acetate (2–12 mg/l) and propionate (<3 mg/l). The higher values for the acetate degradation appear to be caused by mass-transport limitation due to calcium carbonate precipitation within the biofilm. The intrinsic saturation constant is about 3 mg/l, which is significantly lower than previously published values for pure and mixed cultures. The influence of the pH on the saturation constant was investigated in fed-batch experiments. Contrary to the hypothesis that only the undissociated acid is the effective substrate, no significant influence of pH on the saturation constant (given as concentration of total acid) was observed. Batch experiments with n-butyrate revealed hyperbolic progress curves, which might be misinterpreted as a sign of a high saturation constant. However, fed-batch experiments showed that, for n-butyrate degradation, the saturation constant is very low. The isomerisation to isobutyrate and other side-reactions, for which indications were found, influence the progress curve such that an elevated saturation constant will result as an artifact. Thus saturation constants for n-butyrate degradation obtained from batch experiments have to be viewed critically. Received: 17 October 1997 / Received revision: 19 January 1998 / Accepted: 24 January 1998  相似文献   

6.
This study focused on the VFA (volatile fatty acid) profile variation with organic loading rate (OLR) of a two stage thermophilic anaerobic membrane bioreactor (TAnMBR). The two stage TAnMBR treating high strength molasses-based synthetic wastewater was operated under a side-stream partial sedimentation mode at 55 °C. Reactor performances were studied at different OLR ranging from 5 to 12 kg COD m−3 d−1. Operational performance of TAnMBR was monitored by assessing biological activity, organic removal efficiency, and VFA. The major intermediate products of anaerobic digestion were identified as acetate, propionate, iso-butyrate, n-butyrate and valerate. Among them acetate and n-butyrate were identified as the most abundant components. Increase of OLR changes the predominant VFA type from acetic acid to n-butyric acid and the total VFA concentration was increased with increased OLR. Moreover, increased OLR increased organic removal efficiency up to second loading rate and dropped in third loading rate while biological activity was increased continuously.  相似文献   

7.
Acetate reduction is an alternative digestion process to convert organic waste into ethanol. Using acetate for fuel ethanol production offers the opportunity to use organic waste materials instead of sugar-containing feedstock. Methanogenesis, however, competes with acetate reduction for acetate and hydrogen and lowers the final efficiency. The aim of this research is to selectively inhibit methanogenesis and to enhance acetate reduction. Acetate reduction was stimulated in batch tests at pH between 4.5 and 8; and at pH 6 with and without thermal pre-treatment. It was found that methanogenesis was selectively inhibited while acetate reduction was enhanced after thermal pre-treatment incubated at pH 6. Initially the acetate reduction yielded 7.7 ± 3.2 mM ethanol with an efficiency of 60.2 ± 8.7%, but later on it was consumed to form 7.02 ± 0.85 mM n-butyrate with an efficiency of 76.2 ± 14.0%. It was the first time demonstrated that n-butyrate can be produced by mixed cultures from only acetate and hydrogen.  相似文献   

8.
In the presence of 1–5 mM n-butyrate, murine leukemic L1210 cells cease proliferation and become arrested in the G1A compartment of the G1 phase. Cells in this compartment, in comparison with the remaining cells of the G1 phase (G1B), are characterized by low RNA content and more condensed chromatin. During unperturbed growth the cell residence times in G1A are of indeterminate duration (exponentially distributed); the half-time of L1210 cell residence in G1A is about 1.4 h. The effect of n-butyrate in arresting cells in G1A was concentration-dependent. However, the sensitivity of L1210 cells to this drug was markedly enhanced when cells were treated for longer than one generation (12 h). Cells arrested in G1A remained viable and when n-butyrate was removed, after a lag period, they resumed progression through the cycle.The effect of n-butyrate on cell progression through various parts of the cycle was studied in a stathmokinetic experiment. The rate of cell entrance into mitosis was decreased by 30, 60 and 110%, in the presence of 1, 2.5 and 5 mM n-butyrate respectively, thus indicating a slowdown in cell progression through G2 and S. The duration of G2 was prolonged by 20, 70 and 140% at 1, 2.5 and 5 mM n-butyrate respectively. The half-time of cell residence in G1A was increased by as much as 1.5-, 6.3- and 15.6-fold by 1, 2.5 and 5 mM n-butyrate. Progression through late G1 (G1B) was not affected at 1 mM, and could not be estimated at higher drug concentrations. The effects on cell cycle progression were evident 1 h after addition of n-butyrate.DNA in situ in nuclei of n-butyrate-treated cells had lowered (by 2–8 °C) stability to thermal denaturation and increased (by 15%) accessibility to DNase I. The decrease in DNA stability to heat was more pronounced when permealized cells were heated in the presence of 1 mM MgCl2 rather than EDTA. DNA in situ in the nuclei of n-butyrate-treated cells also showed decreased sensitivity to acid-induced denaturation. Changes in chromatin were seen in all cells, regardless of cell cycle phase, within the first hours after addition of n-butyrate. Mitotic cells, however, reacted to n-butyrate more rapidly than interphase cells. The observed changes in L1210 cells are most likely a consequence of histone modifications (acetylation of inner histones, dephosphorylation of histone H1) induced by n-butyrate.  相似文献   

9.
The aldehydes acetaldehyde, propanal, 2-methylpropanal, butanal, 3-methylbutanal, pentanal and hexanal, were reduced to their corresponding alcohols by incubation with strawberry fruit. The alcohols formed were then converted to their acetate, propionate, n-butyrate, isovalerate and n-caproate esters during the incubation with strawberry fruit. Simultaneous reaction of isobutyric acid, n-valeric acid and isocaproic acid with aldehyde and strawberry fruit resulted in the formation of esters of these acids. In all seven alcohols and 54 esters were produced by means of incubation of aldehydes and volatile fatty acids with strawberry fruit.  相似文献   

10.
The lipase of Pseudomonas cepacia was immobilized onto electrospun polyacrylonitrile (PAN) fibers and used for the conversion of (S)-glycidol with vinyl n-butyrate to glycidyl n-butyrate in isooctane. The rate of reaction with the adsorbed lipase was 23-fold higher than the initial material. After 10 recyclings, the initial reaction rate was 80% of the original rate. This system of enzyme immobilization is therefore suitable for carrying out transesterification reactions in nonaqueous solvents.  相似文献   

11.
Acetate kinase (ACK) catalyzes the reversible synthesis of acetyl phosphate by transfer of the γ-phosphate of ATP to acetate. Here we report the first biochemical and kinetic characterization of a eukaryotic ACK, that from the protist Entamoeba histolytica. Our characterization revealed that this protist ACK is the only known member of the ASKHA structural superfamily, which includes acetate kinase, hexokinase, and other sugar kinases, to utilize inorganic pyrophosphate (PPi)/inorganic phosphate (Pi) as the sole phosphoryl donor/acceptor. Detection of ACK activity in E. histolytica cell extracts in the direction of acetate/PPi formation but not in the direction of acetyl phosphate/Pi formation suggests that the physiological direction of the reaction is toward acetate/PPi production. Kinetic parameters determined for each direction of the reaction are consistent with this observation. The E. histolytica PPi-forming ACK follows a sequential mechanism, supporting a direct in-line phosphoryl transfer mechanism as previously reported for the well-characterized Methanosarcina thermophila ATP-dependent ACK. Characterizations of enzyme variants altered in the putative acetate/acetyl phosphate binding pocket suggested that acetyl phosphate binding is not mediated solely through a hydrophobic interaction but also through the phosphoryl group, as for the M. thermophila ACK. However, there are key differences in the roles of certain active site residues between the two enzymes. The absence of known ACK partner enzymes raises the possibility that ACK is part of a novel pathway in Entamoeba.  相似文献   

12.
Acetate Inhibition of Methanogenic, Syntrophic Benzoate Degradation   总被引:4,自引:4,他引:0       下载免费PDF全文
Acetate inhibited benzoate degradation by a syntrophic coculture of an anaerobic benzoate degrader (strain BZ-2) and Methanospirillum strain PM-1; the apparent Ki for acetate was approximately 40 mM. The addition of acetate resulted in a decrease in the hydrogen concentration in the coculture, indicating that phenomena related to interspecies hydrogen transfer affected this value and that the effect of acetate on the benzoate-degrading partner was probably greater than the apparent Ki for the coculture suggests.  相似文献   

13.
In the presence of a catalytic amount of RuHCl(CO)(PR3)n (R=iPr, n=2; R=Ph, n=3), 1,1,2,2-tetramethyl-1,2-divinyldisilane (1) undergoes unexpected and clean isomerization via the Si-Si bond cleavage to yield a mixture of 6- and 5-membered cyclic compounds, 1,1,4,4-tetramethyl-1,4-disilacyclohex-2-ene and 1,1,2,3,3,-pentamethyl-1,3-disilacyclopent-4-ene, the former being the major product.  相似文献   

14.
A phenol-degrading methanogenic enrichment was successfully immobilized in agar as shown by the stoichiometric conversion of phenol to CH4 and CO2. The enrichment contained members of three physiological groups necessary for the syntrophic mineralization of phenol: a phenol-oxidizing bacterium, a Methanothrix-like bacterium, and an H2-utilizing methanogen. The immobilization technique resulted in the cells being embedded in a long, thin agar strand (1 mm in diameter by 2 to 50 cm in length) that resembled spaghetti. Immobilization had three effects as shown by a comparative kinetic analysis of phenol degradation by free versus immobilized cells. (i) The maximum rate of degradation was reduced from 14.8 to 10.0 μg of phenol per h; (ii) the apparent Km for the overall reaction was reduced from 90 to 46 μg of phenol per ml, probably because of the retention of acetate, H2 and CO2 in the proximity of immobilized methanogens; and (iii) the cells were protected from substrate inhibition caused by high concentrations of phenol, which increased the apparent Ki value from 900 to 1,725 μg of phenol per ml. Estimates for the kinetic parameters Km, Ki, and Vmax were used in a modified substrate inhibition model that simulated rates of phenol degradation for given phenol concentrations. The simulated rates were in close agreement with experimentally derived rates for both stimulatory and inhibitory concentrations of phenol.  相似文献   

15.
A mineral salts medium containing 1% (w/v) glucose was subjected to anaerobic digestion in an upflow reactor. Performance with respect to utilization of glucose was monitored by collection of fermentation gases and calculation of carbon mass balances. Sub-samples of bacterial supennsions from the upflow reactor were incubated with (U-14C)-glucose, (U-14C)-acetate, (2-14C)-propionate, (1-14C)-butyrate or 14C-carbonate. Individual radioactive products in samples from incubation mixtures were analysed by radio gas chromatography.Quantitatively, acetate and propionate were the only important intermediates in glucose degradation by glucose-adapted sludge, with acetate accounting for the largest part of intermediary fatty acid flux.Abbreviations used ATP Adenosine Triphosphate - TIC Total Inorganic Carbon - TOC Total Organic Carbon - VFA Volatile Fatty Acids The research described in this publication was supported by the Dutch Ministry of Public Health and Environmental Protection.  相似文献   

16.
17.
Biological availability of 106 esters of alcohols and aliphatic mono-, di- and tri-carboxylic acids and diethylene glycol succinate was compared by the mini-test with chicks. Chicks can utilize methyl esters of saturated fatty acids of carbon chain from 10 to 14, ethyl esters of those from 9 to 12, propyl caprate, n-butyl esters of those from 8 to 12, n-amyl esters of those from 6 to 12, n-hexyl n-butyrate and i-vaterate, and n-octyl and n-decyl acetates. Only 3 dicarboxylates, i.e. di-octyl and di-lauryl succinates and di-methyl cis-cyclopropane-l,2-dicarboxylate, were available among the dicarboxylates tested. Availability of ethyl esters of succinic, fumaric and citric, acid was unexpectedly low.  相似文献   

18.
Human glutathione transferase A1-1 (GST A1-1) has a flexible C-terminal segment that forms a helix (α9) closing the active site upon binding of glutathione and a small electrophilic substrate such as 1-chloro-2,4-dinitrobenzene (CDNB). In the absence of active-site ligands, the C-terminal segment is not fixed in one position and is not detectable in the crystal structure. A key residue in the α9-helix is Phe 220, which can interact with both the enzyme-bound glutathione and the second substrate, and possibly guide the reactants into the transition state. Mutation of Phe 220 into Ala and Thr was shown to reduce the catalytic efficiency of GST A1-1. The mutation of an additional residue, Phe 222, caused further decrease in activity. The presence of a viscosogen in the reaction medium decreased the kinetic parameters kcat and kcat/Km for the conjugation of CDNB catalyzed by wild-type GST A1-1, in agreement with the view that product release is rate limiting for the substrate-saturated enzyme. The mutations cause a decrease of the viscosity dependence of both kinetic parameters, indicating that the motion of the α9-helix is linked to catalysis in wild-type GST A1-1. The isomerization reaction with the alternative substrate Δ5-androstene-3,17-dione (AD) is affected in a similar manner by the viscosogens. The transition state energy of the isomerization reaction, like that of the CDNB conjugation, is lowered by Phe 220 as indicated by the effects of the mutations on kcat/Km. The results demonstrate that Phe 220 and Phe 222, in the dynamic C-terminal segment, influence rate-determining steps in the catalytic mechanism of both the substitution and the isomerization reactions.  相似文献   

19.
Inhibition of the fermentation of propionate to methane and carbon dioxide by hydrogen, acetate, and propionate was analyzed with a mesophilic propionate-acclimatized sludge that consisted of numerous flocs (size, 150 to 300 μm). The acclimatized sludge could convert propionate to methane and carbon dioxide stoichiometrically without accumulating hydrogen and acetate in a propionate-minimal medium. Inhibition of propionate utilization by propionate could be analyzed by a second-order substrate inhibition model (shown below) given that the substrate saturation constant, Ks, was 15.9 μM; the substrate inhibition constant, Ki, was 0.79 mM; and the maximum specific rate of propionate utilization, qm, was 2.15 mmol/g of mixed-liquor volatile suspended solids (MLVSS) per day: qs = qmS/[Ks + S + (S2/Ki)], where qs is the specific rate of propionate utilization and S is the initial concentration of undissociated propionic acid. For inhibition by hydrogen and acetate to propionate utilization, a noncompetitive product inhibition model was used: qs = qm/[1 + (P/Kp)n], where P is the initial concentration of hydrogen or undissociated acetic acid and Kp is the inhibition constant. Kinetic analysis gave, for hydrogen inhibition, Kp(H2) = 0.11 atm (= 11.1 kPa, 71.5 μM), qm = 2.40 mmol/g of MLVSS per day, and n = 1.51 and, for acetate inhibition, Kp(HAc) = 48.6 μM, qm = 1.85 mmol/g of MLVSS per day, and n = 0.96. It could be concluded that the increase in undissociated propionic acid concentration was a key factor in inhibition of propionate utilization and that hydrogen and acetate cooperatively inhibited propionate degradation, suggesting that hydrogenotrophic and acetoclastic methanogens might play an important role in enhancing propionate degradation to methane and carbon dioxide.  相似文献   

20.
Amino Acid Metabolism of Lemna minor L. : II. Responses to Chlorsulfuron   总被引:7,自引:6,他引:1  
Chlorsulfuron, an inhibitor of acetolactate synthase (EC 4.1.3.18) (TB Ray 1984 Plant Physiol 75: 827-831), markedly inhibited the growth of Lemna minor at concentrations of 10−8 molar and above, but had no inhibitory effects on growth at 10−9 molar. At growth inhibitory concentrations, chlorsulfuron caused a pronounced increase in total free amino acid levels within 24 hours. Valine, leucine, and isoleucine, however, became smaller percentages of the total free amino acid pool as the concentration of chlorsulfuron was increased. At concentrations of chlorsulfuron of 10−8 molar and above, a new amino acid was accumulated in the free pool. This amino acid was identified as α-amino-n-butyrate by chemical ionization and electron impact gas chromatography-mass spectrometry. The amount of α-amino-n-butyrate increased from undetectable levels in untreated plants, to as high as 840 nanomoles per gram fresh weight (2.44% of the total free pool) in plants treated with 10−4 molar chlorsulfuron for 24 hours. The accumulation of this amino acid was completely inhibited by methionine sulfoximine. Chlorsulfuron did not inhibit the methionine sulfoximine induced accumulations of valine, leucine, and isoleucine, supporting the idea that the accumulation of the branched-chain amino acids in methionine sulfoximine treated plants is the result of protein turnover rather than enhanced synthesis. Protein turnover may be primarily responsible for the failure to achieve complete depletion of valine, leucine, and isoleucine even at concentrations of chlorsulfuron some 104 times greater than that required to inhibit growth. Tracer studies with 15N demonstrate that chlorsulfuron inhibits the incorporation of 15N into valine, leucine, and isoleucine. The α-amino-n-butyrate accumulated in the presence of chlorsulfuron and [15N]H4+ was heavily labeled with 15N at early time points and appeared to be derived by transamination from a rapidly labeled amino acid such as glutamate or alanine. We propose that chlorsulfuron inhibition of acetolactate synthase may lead to accumulation of 2-oxobutyrate in the isoleucine branch of the pathway, and transamination of 2-oxobutyrate to α-amino-n-butyrate by a constitutive transaminase utilizing either glutamate or alanine as α-amino-N donors.  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