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.
Chlorella cells were examined in a modulated oxygen polarograph under aerobic and anaerobic conditions. At light intensities below about 600 ergs · cm?2 · s?1 of 650 nm light, the oxygen yield and phase lag are lower under anaerobic conditions. Addition of 25 mM sodium nitrite increases both these parameters to values close to those found in the presence of oxygen. It is proposed that nitrite is reduced by Photosystem I thus diverting electrons from the cyclic electron transport pathway. The intersystem electron transport chain becomes more oxidized and this suppresses a backflow of electrons to the oxidizing side of Photosystem II, hence increasing the oxygen yield and the phase lag. The removal of oxygen from the bathing medium also alters the response of dark adapted Chlorella to a series of saturating light flashes. In terms of the Kok model of Photosystem II (Kok, B., Forbush, B. and McGloin, M. (1970) Photochem. Photobiol. 11, 457–475) there is a large increase in the parameter α. Addition of nitrite reverses this change and virtually restores the response seen in the presence of oxygen. The deactivation of the S2 state is greatly speeded up in the absence of oxygen but the addition of nitrite again reverses this.  相似文献   

2.
H Kühne  V A Szalai  G W Brudvig 《Biochemistry》1999,38(20):6604-6613
The binding of chloride and acetate to photosystem II (PSII) was examined to elucidate the mechanism of acetate inhibition. The mode of inhibition was studied, and individual binding sites were assigned by steady-state O2 evolution measurements in correlation with electron paramagnetic resonance (EPR) results. Two binding sites were found for acetate, one chloride-sensitive on the electron donor side and one chloride-insensitive on the electron acceptor side. The respective binding constants were as follows: KCl = 0.5 +/- 0.2 mM (chloride binding to the donor side), KI = 16 +/- 5 mM (acetate binding to the donor side), and KI' = 130 +/- 40 mM (acetate binding to the acceptor side). When acetate was bound to the acceptor side of PSII, 200 K illumination induced a narrowed form of the QA-FeII EPR signal, the yield of which was independent of the chloride concentration. When acetate was bound to the donor side, room-temperature illumination produced the S2YZ* state. EPR measurements showed that both the yield and formation rate of this state increased with acetate concentration. Increasing chloride concentrations slowed the rate of formation of the S2YZ* state, but did not affect the steady-state yield of the S2YZ* state. These findings indicate that the light-induced reactions in acetate-inhibited PSII are modulated by both donor side and acceptor side binding of acetate, while the steady-state yield of the S2YZ* state at the high PSII concentrations used for EPR measurements depends primarily on acceptor side turnover. Our data further support a close proximity of chloride to YZ*, indicating a possible role for chloride in the electron-transfer mechanism at the O2-evolving complex.  相似文献   

3.
Acetaldehyde inhibited the oxidation of fatty acids by rat liver mitochondria as assayed by oxygen consumption and CO2 production. ADP-stimulated oxygen uptake was more sensitive to inhibition by acetaldehyde than was uncoupler-stimulated oxygen uptake, suggesting an effect of acetaldehyde on the electron transport-phosphorylation system. This conclusion is supported by the decrease in the respiratory control ratio, associated with fatty acid oxidation. Acetaldehyde depressed ketone body production as well as the content of acetyl CoA during palmitoyl-1-carnitine oxidation. Acetaldehyde was considerably more inhibitory toward fatty acid oxidation than was acetate. Therefore, the inhibition by acetaldehyde is not mediated by acetate, the direct product of acetaldehyde oxidation by the mitochondria. Oxygen uptake was depressed by acetaldehyde to a slightly, but consistently, greater extent in the absence of fluorocitrate, than in its presence. This suggests inhibition of oxygen consumption from β-oxidation to acetyl CoA and that which arises from citric acid cycle activity. The inhibition of fatty acid oxidation is not due to any effect on the activation or translocation of fatty acids into the mitochondria.The depression of the end products of fatty acid oxidation (CO2, ketones, acetyl CoA) as well as the greater sensitivity of palmitate oxidation compared to acetate oxidation, suggests inhibition by acetaldehyde of β-oxidation, citric acid cycle activity, and the respiratory-phosphorylation chain. Neither the activities of palmitoyl CoA synthetase nor carnitine palmitoyltransferase appear to be rate limiting for fatty acid oxidation.  相似文献   

4.
1. The oxidation of NN-dimethyl-p-phenylenediamine (DPD) by inorganic oxidants and by caeruloplasmin was studied. Some experiments were also made with NNN'N'-tetramethyl-p-phenylenediamine (TPD). 2. E(mM) (550) of the first free radical oxidation product of DPD (DPD(+)) was 9.8 and E(mM) (563) of the corresponding product of TPD (TPD(+)) was 12.5. 3. The non-enzymic decomposition of DPD(+) was studied with respect to temperature, pH, concentration and DPD/DPD(+) ratio, thus defining conditions for enzyme experiments under which DPD(+) extinction at 550mmu was proportional to enzyme activity. 4. Rates of oxidation of DPD to DPD(+) by caeruloplasmin were constant over a range of DPD concentrations. At low DPD concentrations a lag period occurred, which was eliminated by addition of DPD(+). 5. A lag period was not observed with TPD, but at low TPD concentrations the rate of TPD(+) formation was greater when TPD(+) was added. This suggests that TPD(+) may compete weakly as a substrate with TPD and may be oxidized further by the enzyme before a non-enzymic reaction with TPD to form more TPD(+). 6. With DPD sulphate or acetate or TPD sulphate as substrate, Lineweaver-Burk plots were curved. With DPD hydrochloride the chloride ion caused inhibition at higher concentrations, opposing the curvature. 7. Curved Lineweaver-Burk plots were interpreted in terms of two types of substrate binding site with different K(m) values but similar V(max.) values. 8. The apparent thermodynamic changes associated with enzyme-substrate-complex formation at the sites with higher K(m) suggest that considerable conformational change may occur on binding at these sites. 9. With substrate concentrations at which only the low-K(m) sites are involved 2mol. of DPD(+)/mol. of caeruloplasmin are formed before a steady state is established. At higher substrate concentrations up to 3.2mol. of DPD(+)/mol. of caeruloplasmin are formed at this initial stage. 10. Results are discussed in relation to caeruloplasmin structures in which (a) two valence-changing and two permanently cuprous copper atoms are more accessible than the remaining four copper atoms or (b) binding of substrate at one site hinders access of substrate to another site.  相似文献   

5.
A mathematical model was developed to describe the biodegradation kinetics of perchlorate in the presence of nitrate and oxygen as competing electron acceptors. The rate of perchlorate degradation is described as a function of the electron donor (acetate) degradation rate, the concentration of the alternate electron acceptors, and rates of biomass growth and decay. The kinetics of biomass growth are described using a modified Monod model, and inhibition factors are incorporated to describe the influence of oxygen and nitrate on perchlorate degradation. In order to develop input parameters for the model, a series of batch biodegradation studies were performed using Azospira suillum JPLRND, a perchlorate-degrading strain isolated from groundwater. This strain is capable of utilizing oxygen, nitrate, or perchlorate as terminal electron acceptors. The maximum specific growth rate (μmax) and half-saturation constant (K S don) for the bacterium when utilizing either perchlorate or nitrate were similar; 0.16 per h and 158 mg acetate/L, respectively. However, these parameters were different when the strain was growing on oxygen. In this case, μmax and K S don were 0.22 per h and 119 mg acetate/L, respectively. The batch experiments also revealed that nitrate inhibits perchlorate biodegradation by this strain. This finding was incorporated into the model by applying an inhibition coefficient (K i nit) value of 25 mg nitrate/L. Combined with appropriate groundwater transport models, this model can be used to predict perchlorate biodegradation during in situ remediation efforts.  相似文献   

6.
Rabbit muscle pyruvate kinase is shown not to be activated by several anions (acetate, chloride, lactate, nitrate and sulphate) when tested as their tetramethylammonium salts. This contradicts the conclusions of Focant & Watts (1973).  相似文献   

7.
We have investigated hydroxyl free radical mediated damage to pBR322 DNA produced by ascorbate/iron and oxygen in a phosphate-buffered in vitro system. An observed lag phase in DNA nicking suggests a multi-target model of hydroxyl free radical attack on DNA. In the present report we further examine the model system and show that there is a "heat labile" component of the ascorbate/iron system which can be completely restored by the readdition of ascorbate. These observations have allowed us to rule out the possibility that intermediates build up in the reaction and act independently of ascorbate to increase the reaction rate. We have investigated the initial rate of OH production with two OH trapping agents, salicylate and deoxyguanosine, and find that the lag in DNA nicking is not due to a corresponding lag in the production of OH as assessed by formation of the products, dihydroxybenzoic acids and 8-hydroxydeoxyguanosine, respectively. We have found that the energy of activation for DNA supercoiled nicking is 13.9 kcal/mole and for OH trapping by salicylate is 21.1 kcal/nmole. These two activation energies are sufficiently different to suggest that the rate-limiting steps of these two reactions are different. Investigation of the rate of oxygen consumption during the ascorbate/iron-mediated DNA damage showed that oxygen was not a limiting component at any point in the reaction. The addition of catalase slowed down oxygen consumption by 31% and this data taken together with our previous observations on the model implicate hydrogen peroxide as a key intermediate in DNA damage caused by hydroxyl free radical.  相似文献   

8.
The inhibition of the oxidase and respiratory nitrate reductase activity in membrane preparations from Klebsiella aerogenes by 2-n-heptyl-4-hydroxyquinoline-N-oxide (HQNO) has been investigated. Addition of HQNO only slightly affected the aerobic steady-state reduction of cytochrome b559 with NADH, but caused a significantly lower nitrate reducing steady-state of this cytochrome. The changes in the redox states of the cytochromes during a slow transition from anaerobic to aerobic conditions in the presence and absence of HQNO showed that the inhibition site of HQNO is located before cytochrome d. Inhibition patterns obtained upon titration of the NADH oxidase and NADH nitrate reductase activity with HQNO indicated one site of inhibitor interaction in the NADH nitrate reductase pathway and suggested a multilocated inhibition of the NADH oxidase pathway. Difference spectra with ascorbate-dichlorophenolindophenol as electron donor indicated the presence of a cytochrome b563 component which was not oxidized by nitrate, but was rapidly oxidized by oxygen. The latter oxidation was prevented by HQNO. A scheme for the electron transport to oxygen and nitrate is presented. In the pathway to oxygen, HQNO inhibits both at the electron-accepting side of cytochrome b559 and at the electron-donating side of cytochrome b563, whereas in the pathway to nitrate, inhibition occurs only at the electron-accepting side of cytochrome b559.  相似文献   

9.
《BBA》1985,807(1):81-95
(1) The apparent Km for nitrate of the electron-transport system in intact cells of Paracoccus denitrificans was less than 5 μM. In contrast the apparent Km for nitrate of inverted membrane vesicles oxidising NADH was greater than 50 μM. When azide, a competitive inhibitor, was present, the apparent Km observed with the vesicles was raised to 0.64 mM, consistent with values previously reported for purified preparations of the reductase. In membrane vesicles the nitrate reductase is probably not rate-limiting for NADH-nitrate oxido-reductase activity, and thus a lower limit for Km (NO3) is obtained. It is suggested that the very low Km (NO3) in intact cells must arise from either a transport process or a nitrate-specific pore that allows access of nitrate directly to the active site of its reductase from the periplasm. (2) The swelling of spheroplasts has been studied under both aerobic and anaerobic conditions to probe possible mechanisms of nitrate and nitrite transport across the plasma membrane of P. denitrificans. Nitrate reductase was inhibited by azide to prevent reduction of internal nitrate. No evidence for operation of either nitrate-nitrite antiport or proton-nitrate symport was obtained. (3) Measurements from the fluorescence intensity of 8-anilino-naphthalene-1-sulphonate of the rates of decay of diffusion potentials generated by addition of potassium salts to valinomycin-treated plasma membrane vesicles from P. denitrificans showed that the permeability of the membrane to anions is SCN > NO3, NO2, pyruvate, acetate > CI > SO42−. In the presence of a protonophore the rate of decay of the diffusion potential was considerably enhanced with potassium acetate or potassium nitrite, but not with potassium salts of nitrate, chloride or pyruvate. This result indicates that HNO2 and CH3COOH can rapidly and passively diffuse across the cell membrane. This finding suggests that transport systems for nitrite are in general probably not required in bacteria. The failure of a protonophore to enhance the dissipation of the diffusion potential generated by potassium nitrate is evidence against the operation of a proton-nitrate symporter. (4) Low concentrations of added nitrite very strongly inhibit electron flow to oxygen in anaerobically grown cells, provided that they have been treated with Triton X-100 or an uncoupler. This inhibition is not observed with aerobically grown cells. It is concluded that the inhibitory species is a reaction product or an intermediate of the nitrite reductase reaction. The requirement for collapse of protonomotive force by uncoupler or permeabilising the plasma membrane suggests that any such species could be negatively charged. Nitroxyl anion (NO) can be considered, as its conjugate acid is a postulated intermediate between nitrite and nitrous oxide; nitroxyl anion can bind to heme centres to give nitrosyl derivatives. (5) The basis for the ability of permeabilised, but not intact, cells of P. denitrificans to reduce oxygen and nitrate simultaneously is discussed.  相似文献   

10.
The effects of red and far-red light on the enhancement of in vitro nitrate reductase activity and on nitrate accumulation in etiolated excised maize leaves were examined. Illumination for 5 min with red light followed by a 4-h dark period caused a marked increase in nitrate reductase activity, whereas a 5-min illumination with far-red light had no effect on the enzyme activity. The effect of red light was completely reversed by a subsequent illumination with the same period of far-red light. Continuous far-red light also enhanced nitrate reductase activity. Both photoreversibility by red and far-red light and the operation of high intensity reaction under continuous far-red light indicated that the induction of nitrate reductase was mediated by phytochrome. Though nitrate accumulation was slightly enhanced by red and continuous far-red light treatments by 17% and 26% respectively, this is unlikely to account for the entire increase of nitrate reductase activity. The far-red light treatments given in water, to leaves preincubated in nitrate, enhanced nitrate reductase activity considerably over the dark control. The presence of a lag phase and inhibition of increase in enzyme activity under continuous far-red light-by tungstate and inhibitors of RNA synthesis and protein synthesis-rules out the possibility of activation of nitrate reductase and suggests de novo synthesis of the enzyme affected by phytochrome.  相似文献   

11.
The ability of the yeast Kluyveromyces marxianus to convert lactose into ethyl acetate offers good opportunities for the economical reuse of whey. The formation of ethyl acetate as a bulk product depends on aerobic conditions. Aeration of the bioreactor results in discharge of the volatile ester with the exhaust gas that allows its process‐integrated recovery. The influence of aeration (varied from 10 to 50 L/h) was investigated during batch cultivation of K. marxianus DSM 5422 in 0.6 L whey‐borne medium using a stirred reactor. With lower aeration rates, the ester accumulated in the bioreactor and reached higher concentrations in the culture medium and the off gas. A high ester concentration in the gas phase is considered beneficial for ester recovery from the gas, while a high ester concentration in the medium inhibited yeast growth and slowed down the process. To further investigate this effect, the inhibition of growth by ethyl acetate was studied in a sealed cultivation system. Here, increasing ester concentrations caused a nearly linear decrease of the growth rate with complete inhibition at concentrations greater than 17 g/L ethyl acetate. Both the cultivation process and the growth rate depending on ethyl acetate were described by mathematical models. The simulated processes agreed well with the measured data.  相似文献   

12.
Cytochromes b of anaerobically nitrate-grown Escherichia coli cells are analysed. Ascorbate phenazine methosulfate distinguishes low and high potential cytochromes b. Reduction kinetics performed at 559 nm presents a very complex pattern which can be analysed assuming that at least four b-type cytochromes are present. The electron transport chain from formate to oxygen would contain a low potential cytochrome b-556, a cytochrome b-558 associated to the oxidase, and a cytochrome d as the principle oxidase. Cytochrome o is also present, but seems to be functional only at low oxygen concentrations. A cytochrome b-556 associated to nitrate reductase is shown to belong to a branch of the formate-oxidase chain. 2-N-Heptyl-4-hydroxyquinoline-N-oxide affects the reduction kinetics in a very complex way. One inhibition site is in evidence between cytochrome b-558 and cytochrome d; another between the cytochrome associated to nitrate reductase and the nitrate reductase. A third inhibition site is located in the common part of the formate-nitrate and the formate-oxidase systems. Ascorbate phenazine methosulfate is shown to donate electrons near cytochrome b-558.  相似文献   

13.
Microalgae are extensively used in the remediation of heavy metals like iron. However, factors like toxicity, bioavailability and iron speciation play a major role in its removal by microalgae. Thus, in this study, toxicity of three different iron salts (FeSO4, FeCl3 and Fe(NO3)3) was evaluated towards three soil microalgal isolates, Chlorella sp. MM3, Chlamydomonas sp. MM7 and Chlorococcum sp. MM11. Interestingly, all the three iron salts gave different EC50 concentrations; however, ferric nitrate was found to be significantly more toxic followed by ferrous sulphate and ferric chloride. The EC50 analysis revealed that Chlorella sp. was significantly resistant to iron compared to other microalgae. However, almost 900 μg g?1 iron was accumulated by Chlamydomonas sp. grown with 12 mg L?1 ferric nitrate as an iron source when compared to other algae and iron salts. The time-course bioaccumulation confirmed that all the three microalgae adsorb the ferric salts such as ferric nitrate and ferric chloride more rapidly than ferrous salt, whereas intracellular accumulation was found to be rapid for ferrous salts. However, the amount of iron accumulated or adsorbed by algae, irrespective of species, from ferrous sulphate medium is comparatively lower than ferric chloride and ferric nitrate medium. The Fourier transform infrared spectroscopy (FTIR) analysis shows that the oxygen atom and P?=?O group of polysaccharides present in the cell wall of algae played a major role in the bioaccumulation of iron ions by algae.  相似文献   

14.
Ferric horseradish peroxidase reacts with nitrate and acetate in acidic solution to form weakly bound complexes. Competitive binding experiments with cyanide show that the nitrate binding site is not at the sixth coordination position of the heme iron. The nitrate inhibits compound I formation apparently by binding inside the heme pocket. One physical manifestation of this binding is to increase the apparent pKa value of the conjugate acid of a catalytic distal group.  相似文献   

15.
—Mercuric chloride, silver acetate and cupric sulphate (0·1 mm ) completely inhibited purified choline acetyltransferase from bovine caudate nuclei. At the same concentration cadmium chloride and zinc acetate gave a 50 per cent inhibition. Potassium and sodium salts more than doubled the enzymatic activity while creatinine hydrochloride more than tripled it. Guanidine hydrochloride was less effective than creatinine hydrochloride but more effective than KCl and NaCl. Sodium chloride and creatinine hydrochloride had a synergistic effect on the enzyme. When ammonium sulphate was used to fractionate the choline acetyltransferase that had been extracted from bovine caudate nuclei, the enzyme aggregated into different molecular sizes as determined by exclusion chromatography on Bio-gel A-1·5 m. The molecular weight of the largest aggregate was at least 106 daltons. The initial tissue extract contained only one molecular species of ChAc as did a partially purified preparation in which ammonium sulphate was not used in the purification.  相似文献   

16.
This study, based on field and laboratory work, investigates the biogeochemical activity below the organic top soil horizons, particularly the potential for nitrate removal processes in the deep vadose zone (1–2.5 m depth) of a weathered granite. An experimental site located in the Kerbernez agricultural catchment (Brittany) has been equipped with ceramic cups from 0.5 to 2.5 m depth since November 2001. This arrangement allowed collection of water samples from the soil profile and the upper part of the unsaturated weathered granite. Samples were analysed twice a month for chloride, nitrate and sulphate concentrations over a period of 2.5 years. Laboratory measurements were carried out on three soil horizons and four weathered granite facies sampled in October 2003 for hydrolasic activity, potential nitrification, potential denitrification and batch experiments to study nutrient dynamics. Anion analyses in the field show that the nitrate and chloride concentrations were linearly correlated at each depth. The nitrate/chloride ratio decreased with depth in the upper part of the weathered granite from 4.93 to 2.82. This suggests that nitrate was removed during its vertical transport in the unsaturated zone. The laboratory experiments show that the bacterial activity decreased with depth. However, a significant potential for biogeochemical reactions exists below the organic soil layers. The denitrification rates obtained in the laboratory were significant, up to 800 ng  N h−1 g−1 after about 100 h of incubation for the most reactive weathered granite facies. These rates agree with effective rates usually measured in riparian zones, but they were 50 times higher than those observed in the field. This difference suggests that the denitrification processes occurring in the field were spatially limited to localised anaerobic microsites, where the bacterial activities are controlled by the availability of N and C substrate. Finally, the laboratory measurements lead us to assume that heterotrophic denitrification was clearly the predominant process occurring in the field because of the good correlation between nitrate concentration variation and carbon content (r = −0.94). Moreover, the slight increase in sulphate concentrations observed in the field and in the laboratory was insufficient to explain the complete removal of nitrate.  相似文献   

17.
A new moderately halophilic Micrococcus sp. 4, isolated from salt-pan water from India, produced extracellular amylase when cultivated aerobically in medium containing wheat bran, peptone, beef extract and sodium chloride. Other salts, such as sodium nitrate, potassium nitrate and sodium sulphate, were also found to be suitable for growth and enzyme production. Maximum amylase activity (1.2 IU ml-1) was secreted in the presence of 1 mol 1-1 sodium chloride. The enzyme requires the presence of either sodium chloride, potassium chloride, sodium nitrate, sodium citrate or sodium acetate for its activity. Maximum activity was found in the presence of 1 mol 1-1 sodium chloride. The pH and temperature optima for enzyme activity were 7.5 and 50°C, respectively.  相似文献   

18.
A study was made with a modulated oxygen electrode of the effect of variations of oxygen concentration on photosynthetic oxygen evolution from algal cells. When Chlorella vulgaris is examined with a modulated 650 nm light at 22 degrees C, both the oxygen yield and the phase lag between the modulated oxygen signal and the light modulations have virtually constant values between 800 and 120 ergs . cm-1 . s-1 if the bathing medium is in equilibrium with the air. Similar results are obtained at 32 degrees C between 1600 and 120 ergs . cm-2 . s-1. Under anaerobic conditions both the oxygen yield and the phase lag decrease if the light intensity is lowered below about 500 ergs . cm-2 . s-1 at 22 degrees C or about 1000 ergs . cm-2 . s-1 at 32 degrees C. A modulated 706 nm beam also gives rise to these phenomena but only at significantly lower rates of oxygen evolution. The cells of Anacystis nidulans and Porphyridium cruentum appear to react in the same way to anaerobic conditions as C. vulgaris. An examination of possible mechanisms to explain these results was performed using a computer simulation of photosynthetic electron transport. The simulation suggests that a backflow of electrons from a redox pool between the Photosystems to the rate-limiting reaction between Photosystem II and the water-splitting act can cause a decrease in oxygen yield and phase lag. If the pool between the Photosystems is in a very reduced state a significant cyclic flow is expected, whereas if the pool is largely oxidized little or no cyclic flow should occur. It is shown that the effects of 706 nm illumination and removal of oxygen can be interpreted in accordance with these proposals. Since a partial inhibition of oxygen evolution by 3-(3.4-dichlorophenyl)-1,1-dimethylurea (10(-8) M) magnifies the decreases in oxygen yield and phase lag, it is proposed that the pool which cycles back electrons is in front of the site of 3-(3,4-dichlorophenyl)-1,1-dimethylurea inhibition and is probably the initial electron acceptor pool after Photosystem II.  相似文献   

19.
Crystal structure, determinations on the four complexes MO2 (4,4-bipyridyl)X2 with M = U, Np; X = nitrate, acetate, show the normal contraction in M-ligand distances on replacing U by Np, when X = nitrate. When X = acetate, the Np complex has a larger unit cell volume and longer MpN distances (UN, 2.636(7)); NpN, 2.838(10) Å). This is explained by overcrowding caused by the large bite bipyridyl ligand.  相似文献   

20.
Characterization of Clostridium thermocellum JW20   总被引:9,自引:3,他引:6       下载免费PDF全文
Clostridium thermocellum JW20 (ATCC 31549), which was isolated from a Louisiana cotton bale, grew on cellulose, cellobiose, and xylooligomers and, after adaptation, on glucose, fructose, and xylose in the pH range of 7.5 to 6.1 with Topt of 60°C, Tmax of 69°C, and Tmin of above 28°C. Doubling times during growth on cellulose and cellobiose were 6.5 and 2.5 h, respectively. The G+C content of the DNA was 40 mol% (chemical analysis). Growth on cellulose as substrate was totally inhibited in the presence of more than 125 mM sodium sulfate, 300 mM sodium chloride, 250 mM potassium chloride, 200 mM calcium chloride, 125 mM magnesium chloride, 40 mM lactate, or 250 mM acetate. The ratio of the fermentation products ethanol to acetate plus H2 decreased when the culture was agitated. Agitation otherwise increased the rate of cellulose degradation in a growing culture but not under nongrowth conditions or with cell-free culture supernatant containing the extracellular cellulase. Shaking lowered the concentration of H2 in the culture broth and thus minimized inhibition by the H2 formed. Externally added H2 caused an increased formation of ethanol during growth on cellulose or cellobiose. However, at an atmospheric pressure as high as 355 kPa (50 lb/in2), H2 did not cause significant growth inhibition beyond an increasing lag phase (up to 24 h). Several criteria to specifically prove the purity of C. thermocellum cultures were suggested.  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