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.
2.
Butanol and butyric acid produced from acetone-butanol-ethanol (ABE) fermentation can be used to produce butyl butyrate, an important fragrance ester. However, low levels of butanol and butyric acid need to be purified from culture media first with energy-intensive distillation processes. In this study, a triphasic (organic/aqueous/fluorous) system is developed to esterify butanol and butyric acid in spent culture media into butyl butyrate directly without purification. The produced butyl butyrate forms a distinct organic phase floating on top and can then be separated easily. In a model system containing 37.1 g/L of butanol and 44.1 g/L of butyric acid, 57% of the butanol is converted to butyl butyrate after 8 h of esterification. With multiple cycles of esterification and product removal, butanol conversion can be further increased to 86%. When spent culture medium containing 7.12 g/L of butanol and 4.81 g/L of butyric acid is used for esterification, 38% of butanol (0.36 mmol) is consumed and 0.33 mmol of butyl butyrate is produced. However, when ABE fermentation and esterification are carried out simultaneously, only 0.042 mmol of butyl butyrate is produced, probably due to the incompatible pH requirements for cell growth (pH 5–7) and esterification (pH 2–3).  相似文献   

3.
The diffusion of the sodium salt of monocarboxylic fatty acids, from formate to stearate, has been studied as a function of water content and pH in lecithin--water lamellar phases. Evolution of the diffusion coefficients with increasing chain length reflects the different localizations of fatty acids in the system. From formate to butyrate, which are mainly restricted to the hydrophilic layer of the phase, diffusion rates decrease rapidly. From butyrate to stearate, fatty acids (anchored at the hydrophilic--lipophilic interface) undergo lateral diffusion and then the decrease of D with increasing chain length is much slower. The diffusion of stereate is already comparable to the diffusion of the lecithin molecule itself. The diffusion rates strongly depend upon phase hydration and pH: it is shown that both parameters control the fatty acid ionization. The variations in diffusion rates observed may be ascribed to the fact that, depending upon their state of ionization, fatty acids assume a different localization and therefore experience different interactions in the lamellar system.  相似文献   

4.
[gamma]-Aminobutyric acid (GABA) synthesis (L-glutamic acid + H+ -> GABA + CO2) is rapidly stimulated by a variety of stress conditions including hypoxia. Recent literature suggests that GABA production and concomitant H+ consumption ameliorates the cytosolic acidification associated with hypoxia or other stresses. This proposal was investigated using isolated asparagus (Asparagus sprengeri Regel) mesophyll cells. Cell acidification was promoted using hypoxia, H+/L-glutamic acid symport, and addition of butyrate or other permeant weak acids. Sixty minutes of all three treatments stimulated the levels of both intracellular and extracellular GABA by values ranging from 100 to 1800%. At an external pH of 5.0, addition of 5 mM butyrate stimulated an increase in overall GABA level from 3.86 (0.56 [plus or minus] SE) to 20.4 (2.16 [plus or minus] SE) nmol of GABA/106 cell. Butyrate stimulated GABA levels by 200 to 300% within 15 s, and extracellular GABA was observed after 10 min. The acid load due to butyrate addition was assayed by measuring [14C]butyrate uptake. After 45 s of butyrate treatment, H+-consuming GABA production accounted for 45% of the imposed acid load. The cytosolic location of a fluorescent pH probe was confirmed using fluorescent microscopy. Spectrofluorimetry indicated that butyrate addition reduced cytosolic pH by 0.60 units with a half-time of approximately 2 s. The proposal that GABA synthesis ameliorates cytosolic acidification is supported by the data. The possible roles of H+ and Ca2+ in stimulating GABA synthesis are discussed.  相似文献   

5.
6.
During the industrial stabilization process, lactic acid bacteria are subjected to several stressful conditions. Tolerance to dehydration differs among lactic acid bacteria and the determining factors remain largely unknown. Lactobacillus coryniformis Si3 prevents spoilage by mold due to production of acids and specific antifungal compounds. This strain could be added as a biopreservative in feed systems, e.g. silage. We studied the survival of Lb. coryniformis Si3 after freeze-drying in a 10% skim milk and 5% sucrose formulation following different fermentation pH values and temperatures. Initially, a response surface methodology was employed to optimize final cell density and growth rate. At optimal pH and temperature (pH 5.5 and 34 °C), the freeze-drying survival of Lb. coryniformis Si3 was 67% (±6%). The influence of temperature or pH stress in late logarithmic phase was dependent upon the nature of the stress applied. Heat stress (42 °C) did not influence freeze-drying survival, whereas mild cold- (26 °C), base- (pH 6.5), and acid- (pH 4.5) stress significantly reduced survival. Freeze-drying survival rates varied fourfold, with the lowest survival following mild cold stress (26 °C) prior to freeze-drying and the highest survival after optimal growth or after mild heat (42 °C) stress. Levels of different membrane fatty acids were analyzed to determine the adaptive response in this strain. Fatty acids changed with altered fermentation conditions and the degree of membrane lipid saturation decreased when the cells were subjected to stress. This study shows the importance of selecting appropriate fermentation conditions to maximize freeze-drying viability of Lb. coryniformis as well as the effects of various unfavorable conditions during growth on freeze-drying survival.  相似文献   

7.
We present a complete DNA sequence and metabolic analysis of the dominant oral bacterium Fusobacterium nucleatum. Although not considered a major dental pathogen on its own, this anaerobe facilitates the aggregation and establishment of several other species including the dental pathogens Porphyromonas gingivalis and Bacteroides forsythus. The F. nucleatum strain ATCC 25586 genome was assembled from shotgun sequences and analyzed using the ERGO bioinformatics suite (http://www.integratedgenomics.com). The genome contains 2.17 Mb encoding 2,067 open reading frames, organized on a single circular chromosome with 27% GC content. Despite its taxonomic position among the gram-negative bacteria, several features of its core metabolism are similar to that of gram-positive Clostridium spp., Enterococcus spp., and Lactococcus spp. The genome analysis has revealed several key aspects of the pathways of organic acid, amino acid, carbohydrate, and lipid metabolism. Nine very-high-molecular-weight outer membrane proteins are predicted from the sequence, none of which has been reported in the literature. More than 137 transporters for the uptake of a variety of substrates such as peptides, sugars, metal ions, and cofactors have been identified. Biosynthetic pathways exist for only three amino acids: glutamate, aspartate, and asparagine. The remaining amino acids are imported as such or as di- or oligopeptides that are subsequently degraded in the cytoplasm. A principal source of energy appears to be the fermentation of glutamate to butyrate. Additionally, desulfuration of cysteine and methionine yields ammonia, H(2)S, methyl mercaptan, and butyrate, which are capable of arresting fibroblast growth, thus preventing wound healing and aiding penetration of the gingival epithelium. The metabolic capabilities of F. nucleatum revealed by its genome are therefore consistent with its specialized niche in the mouth.  相似文献   

8.
The pH of continuous cultures of Clostridium acetobutylicum growing at pH 5.6 was allowed to decrease to 4.3 after acid production and thereby to shift the cultures from acetate and butyrate to acetone and butanol formation. Several parameters were determined during the shift. An increase in the intracellular acid concentration to 440 mM was recorded. An excess of undissociated butyric acid but not of acetic acid just before the shift to solventogenesis was followed by a decline in acid production and subsequently by the uptake of acids. The intracellular ATP concentration reached a minimum before the onset of solventogenesis; this presumably reflects the ATP-consuming proton extrusion connected with the increase in the ΔpH from 0.7 to 1.4 units. The pool of NADH plus NADPH exhibited a drastic increase until solventogenesis was induced. The changes in the ATP and ADP and NADH plus NADPH pools during these pH shift experiments were the beginning of a stable metabolic oscillation which could also be recorded as an oscillation of the culture redox potential under steady-state solventogenic conditions. Similar changes were observed when the shift was induced by the addition of butyrate and acetate (50 mM each) to the continuous culture. However, when methyl viologen was added, important differences were found: ATP levels did not reach a minimum, acetoacetate decarboxylase activity could not be measured, and butanol but not acetone was produced. A model for the shift is proposed; it assumes the generation of two signals, one by the changed ATP and ADP levels and the other by the increased NAD(P)H level.  相似文献   

9.
Kwon YM  Ricke SC 《Anaerobe》1998,4(6):251-256
Propionic acid is commonly found as a fermentation product in the gastrointestinal tracts of food animals and has also been used to limit the microbial contaminants in animal feeds. Because propionic acid is known to have antibacterial activity, the propionic acid encountered by foodborne pathogens during their life cycles may play an important role in inhibiting the survival of the pathogens. The survival patterns of Salmonella typhimurium poultry isolate were determined both in aerobic and anaerobic tryptic soy broth (TSB; pH 5.0 or 7.0) containing various concentrations of propionic acid (0-200 mM). The levels of recovered cells were consistently greater at pH 7.0 compared to those at pH 5.0. For the first 4 days, the levels were significantly decreased by incubation under anaerobic conditions as compared to aerobic condition at pH 7.0 (P<0.05). However, there were fluctuations of cell populations with different patterns depending on both concentrations and growth conditions. To characterize the nature of the capability which allowed the cell multiplication following decreases in cell population during incubation at pH 7.0, the cells isolated from the outgrowth cultures were tested for survival in aerobic or anaerobic TSB (pH 5.0 or pH 7.0) containing propionic acid (50 mM). The outgrowth isolates did not show significant differences in the level of recovered cells in the presence of propionic acid when compared to the wild type strain (P>0.05), suggesting that the cells in the outgrowth cultures did not harbour mutation(s) conferring increased resistance to propionic acid. In addition, the level of recovered cells of isogenic rpoS mutant strain of S. typhimurium was not significantly different from that of the wild type strain in the same assay conditions (P<0.05). The results of this study show that the bactericidal activity of propionic acid on S. typhimurium can be affected by environmental conditions such as acidic pH levels and anaerobiosis in food materials and gastrointestinal tracts. However, S. typhimurium is also able to multiply in the presence of sublethal concentrations of propionic acid at neutral pH during prolonged incubation under both aerobic and anaerobic conditions.  相似文献   

10.
利用厌氧菌群生物合成己酸被认为是一种非常有潜力的新型废弃物资源化技术,但是其合成效能的提高是目前亟待解决的关键问题。本研究以实际果蔬废弃物为原料,对两相厌氧发酵产己酸的效能进行了研究。首先优化接种比以提高酸化相的水解转化效率;在此基础上通过调控醇酸比和pH以强化产己酸相的发酵效能。结果显示,果蔬废弃物厌氧产酸的最佳接种比为2∶1,此时水解率和酸化率分别可达到98.1%和83.2%,乙酸和丁酸产量分别达到5.4 g/L和3.3 g/L。合理控制醇酸比和pH对提高产己酸相的发酵效能非常关键。当醇酸比和pH控制为4∶1和7.5时,己酸生成量可达14.9 g/L,约占液相总COD的80.84%;而低醇酸比和低pH易造成丁酸的累积,从而降低了己酸产量。己酸发酵过程属于非生长偶联型,己酸菌(Clostridium kluyveri)指数增长期伴随着丁酸的生成,而己酸合成主要发生在生长中后期。此外,己酸菌对于pH变化较为敏感,适当提高pH有助于减轻有机酸毒性,提高生物量;但是碱性环境会严重抑制己酸菌的生长繁殖。研究表明,通过分别对酸化相和产己酸相进行优化和调控,两相发酵策略更有利于提高己酸合成效能。  相似文献   

11.
Short chain fatty acids (SCFA) also named volatile fatty acids, mainly acetate, propionate and butyrate, are the major end-products of the microbial digestion of carbohydrates in the alimentary canal. The highest concentrations are observed in the forestomach of the ruminants and in the large intestine (caecum and colon) of all the mammals. Butyrate and caproate released by action of gastric lipase on bovine milk triacylglycerols ingested by preruminants or infants are of nutritional importance too. Both squamous stratified mucosa of rumen and columnar simple epithelium of intestine absorb readily SCFA. The mechanisms of SCFA absorption are incompletely known. Passive diffusion of the unionized form across the cell membrane is currently admitted. In the lumen, the necessary protonation of SCFA anions could come first from the hydration of CO2. The ubiquitous cell membrane process of Na+-H+ exchange can also supply luminal protons. Evidence for an acid microclimate (pH = 5.8-6.8) suitable for SCFA-protonation on the surface of the intestinal lining has been provided recently. This microclimate would be generated by an epithelial secretion of H+ ions and would be protected by the mucus coating from the variable pH of luminal contents. Part of the absorbed SCFA does not reach plasma because it is metabolized in the gastrointestinal wall. Acetate incorporation in mucosal higher lipids is well-known. However, the preponderant metabolic pathway for all the SCFA is catabolism to CO2 except in the rumen wall where about 80% of butyrate is converted to ketone bodies which afterwards flow into bloodstream. Thus, SCFA are an important energy source for the gut mucosa itself.  相似文献   

12.
Coenzyme A (CoA) transferase from Clostridium acetobutylicum ATCC 824 was purified 81-fold to homogeneity. This enzyme was stable in the presence of 0.5 M ammonium sulfate and 20% (vol/vol) glycerol, whereas activity was rapidly lost in the absence of these stabilizers. The kinetic binding mechanism was Ping Pong Bi Bi, and the Km values at pH 7.5 and 30 degrees C for acetate, propionate, and butyrate were, respectively, 1,200, 1,000, and 660 mM, while the Km value for acetoacetyl-CoA ranged from about 7 to 56 microM, depending on the acid substrate. The Km values for butyrate and acetate were high relative to the intracellular concentrations of these species; consequently, in vivo enzyme activity is expected to be sensitive to changes in those concentrations. In addition to the carboxylic acids listed above, this CoA transferase was able to convert valerate, isobutyrate, and crotonate; however, the conversion of formate, n-caproate, and isovalerate was not detected. The acetate and butyrate conversion reactions in vitro were inhibited by physiological levels of acetone and butanol, and this may be another factor in the in vivo regulation of enzyme activity. The optimum pH of acetate conversion was broad, with at least 80% of maximal activity from pH 5.9 to greater than 7.8. The purified enzyme was a heterotetramer with subunit molecular weights of about 23,000 and 25,000.  相似文献   

13.
Coenzyme A (CoA) transferase from Clostridium acetobutylicum ATCC 824 was purified 81-fold to homogeneity. This enzyme was stable in the presence of 0.5 M ammonium sulfate and 20% (vol/vol) glycerol, whereas activity was rapidly lost in the absence of these stabilizers. The kinetic binding mechanism was Ping Pong Bi Bi, and the Km values at pH 7.5 and 30 degrees C for acetate, propionate, and butyrate were, respectively, 1,200, 1,000, and 660 mM, while the Km value for acetoacetyl-CoA ranged from about 7 to 56 microM, depending on the acid substrate. The Km values for butyrate and acetate were high relative to the intracellular concentrations of these species; consequently, in vivo enzyme activity is expected to be sensitive to changes in those concentrations. In addition to the carboxylic acids listed above, this CoA transferase was able to convert valerate, isobutyrate, and crotonate; however, the conversion of formate, n-caproate, and isovalerate was not detected. The acetate and butyrate conversion reactions in vitro were inhibited by physiological levels of acetone and butanol, and this may be another factor in the in vivo regulation of enzyme activity. The optimum pH of acetate conversion was broad, with at least 80% of maximal activity from pH 5.9 to greater than 7.8. The purified enzyme was a heterotetramer with subunit molecular weights of about 23,000 and 25,000.  相似文献   

14.
In Vitro Lactate Metabolism by Ruminal Ingesta   总被引:1,自引:1,他引:0  
Ruminal ingesta (300 ml) obtained from a fistulated cow fed alfalfa hay (H), 3.6 kg of grain mixture with corn silage fed ad libitum (S), 2.5:1 grain-alfalfa hay mixture (G), or a 2.5:1 grain-alfalfa hay mixture providing 545 g of sodium and calcium lactate daily (L) were incubated for 8 hr with nonpolymerized sodium lactate or 17% polymerized lactic acid neutralized to pH 6.7. Polymerization had no effect on the rate of lactate utilization. The initial rates of lactate metabolism for the H, G, S, and L ingesta were 0.72, 0.95, 1.8, and 3.4 meq per 100 ml of rumen fluid per hr, respectively. Lactate-2-(14)C was incubated for 4 hr with each type of ruminal ingesta. Of the label recovered in the volatile fatty acids (VFA), 74.1, 61.2, 49.3, and 38.9% was recovered in acetate, and 9.4, 19.8, 23.3, and 51.9% was recovered in propionate with H, G, S, and L ingesta, respectively. The balance of label was distributed between butyrate and valerate. The titratable VFA did not follow this pattern of production. With the hay ingesta, lactate metabolism resulted in a net loss of acetate and a large increase in butyrate. Little propionate was produced. The G, S, and L ingesta metabolized lactate to yield progressively more propionate and less butyrate. Evidence was gathered to suggest that acetate was the primary end product of lactate metabolism but that oxidation of lactate to pyruvate dictated the synthesis of butyrate from acetate to maintain an oxidation-reduction balance. It was noted that acetate and butyrate production from lactate was pH-dependent, with acetate production maximal at pH 7.4 and butyrate at 6.2. Propionate production was largely unaffected within this pH range.  相似文献   

15.
Clostridium acetobutylicum ATCC 824 cells harvested from a phosphate-limited chemostat culture maintained at pH 4.5 had intracellular concentrations of acetate, butyrate, and butanol which were 13-, 7-, and 1.3-fold higher, respectively, than the corresponding extracellular concentrations. Cells from a culture grown at pH 6.5 had intracellular concentrations of acetate and butyrate which were only 2.2-fold higher than the respective external concentrations. The highest intracellular concentrations of these acids were attained at ca. pH 5.5. When cells were suspended in anaerobic citrate-phosphate buffer at pH 4.5, exogenous acetate and butyrate caused a concentration-dependent decrease in the intracellular pH, while butanol had relatively little effect until the external concentration reached 150 mM. Acetone had no effect at concentrations up to 200 mM. These data demonstrate that acetate and butyrate are concentrated within the cell under acidic conditions and thus tend to lower the intracellular pH. The high intracellular butyrate concentration presumably leads to induction of solvent production, thereby circumventing a decrease in the intracellular pH great enough to be deleterious to the cell.  相似文献   

16.
Microbiological safety has been a critical issue for acid and acidified foods since it became clear that acid-tolerant pathogens such as Escherichia coli O157:H7 can survive (even though they are unable to grow) in a pH range of 3 to 4, which is typical for these classes of food products. The primary antimicrobial compounds in these products are acetic acid and NaCl, which can alter the intracellular physiology of E. coli O157:H7, leading to cell death. For combinations of acetic acid and NaCl at pH 3.2 (a pH value typical for non-heat-processed acidified vegetables), survival curves were described by using a Weibull model. The data revealed a protective effect of NaCl concentration on cell survival for selected acetic acid concentrations. The intracellular pH of an E. coli O157:H7 strain exposed to acetic acid concentrations of up to 40 mM and NaCl concentrations between 2 and 4% was determined. A reduction in the intracellular pH was observed for increasing acetic acid concentrations with an external pH of 3.2. Comparing intracellular pH with Weibull model predictions showed that decreases in intracellular pH were significantly correlated with the corresponding times required to achieve a 5-log reduction in the number of bacteria.  相似文献   

17.
Short chain fatty acids (SCFAs) are fermentation products of anaerobic bacteria. More than just being an important energy source for intestinal epithelial cells, these compounds are modulators of leukocyte function and potential targets for the development of new drugs. The aim of this study was to evaluate the effects of SCFAs (acetate, propionate and butyrate) on production of nitric oxide (NO) and proinflammatory cytokines [tumor necrosis factor α (TNF-α) and cytokine-induced neutrophil chemoattractant-2 (CINC-2αβ)] by rat neutrophils. The involvement of nuclear factor κB (NF-κB) and histone deacetylase (HDAC) was examined. The effect of butyrate was also investigated in vivo after oral administration of tributyrin (a pro-drug of butyrate). Propionate and butyrate diminished TNF-α, CINC-2αβ and NO production by LPS-stimulated neutrophils. We also observed that these fatty acids inhibit HDAC activity and NF-κB activation, which might be involved in the attenuation of the LPS response. Products of cyclooxygenase and 5-lipoxygenase are not involved in the effects of SCFAs as indicated by the results obtained with the inhibitors of these enzymes. The recruitment of neutrophils to the peritonium after intraperitoneal administration of a glycogen solution (1%) and the ex vivo production of cytokines and NO by neutrophils were attenuated in rats that previously received tributyrin. These results argue that this triglyceride may be effective in the treatment of inflammatory conditions.  相似文献   

18.
Mutants of Escherichia coli K-12 which grow on butyrate and valerate were studied with respect to uptake of these substrates. To utilize short-chain and medium-chain fatty acids, E. coli must synthesize the beta-oxidation enzymes constitutively. In addition, growth on the C(4) and C(5) acids requires a second mutation which permits entry of these substrates. At pH 5, both in the parent and mutant strains, butyrate and valerate penetrate as the undissociated acids but appear not to be activated and thus inhibit growth. At pH 7, the parent strain is not permeable to the anions, whereas the mutant concentrates these substrates. There appear to be two components of the uptake system, a nonspecific diffusion component and an energy-linked activating enzyme. Two mutant types which take up short-chain fatty acids are described. One synthesizes the uptake system constitutively and is inhibited by 4-pentenoate when cultured on acetate. In the other, the uptake system is inducible, and the strain is pentenoate-resistant when grown on acetate but pentenoate-sensitive when cultured on butyrate or valerate.  相似文献   

19.
在已有的酶法拆分丁酸缩水甘油酯单因素试验最适条件的基础上, 采用Plackett-Burrman设计在影响酶法拆分的6个因素中, 有效筛选出主效因素: 底物浓度、酶用量和温度。在此基础上, 再利用响应面分析法(RSM)对以上三个显著因子的最佳水平范围进行研究, 通过对二次多项回归方程求解得知, 酶法拆分最适条件为: 底物浓度0.499 mol/L、酶用量30.23 mg/(g底物)和温度29.68oC; 并结合单因素最适条件: pH 7.6; 时间4 h; 转速150 r/min进行酶促拆分实验, 得到R-酯的对映体过量值为93.28%。对比优化之前的单因素试验最适条件的结果, 最大对映体过量值84.65%, 有了显著的提高, 证明RSM法优化酶法拆分丁酸缩水甘油酯工艺是可行的。  相似文献   

20.
在已有的酶法拆分丁酸缩水甘油酯单因素试验最适条件的基础上, 采用Plackett-Burrman设计在影响酶法拆分的6个因素中, 有效筛选出主效因素: 底物浓度、酶用量和温度。在此基础上, 再利用响应面分析法(RSM)对以上三个显著因子的最佳水平范围进行研究, 通过对二次多项回归方程求解得知, 酶法拆分最适条件为: 底物浓度0.499 mol/L、酶用量30.23 mg/(g底物)和温度29.68oC; 并结合单因素最适条件: pH 7.6; 时间4 h; 转速150 r/min进行酶促拆分实验, 得到R-酯的对映体过量值为93.28%。对比优化之前的单因素试验最适条件的结果, 最大对映体过量值84.65%, 有了显著的提高, 证明RSM法优化酶法拆分丁酸缩水甘油酯工艺是可行的。  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