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 796 毫秒
1.
The suitability and limitation of yeast extract as nitrogen source to support cell growth and to enhance hydrogen photoproduction by Rhodobacter sphaeroides strains MDC6521 and MDC6522 isolated from mineral springs in Armenia was investigated during the anaerobic growth. Yeast extract (2 g L−1) was indicated to be an effective nitrogen source for bacterial cell growth stimulation and enhanced H2 production (compared to glutamate). Both strains followed similar growth patterns in medium with yeast extract as nitrogen source and succinate or malate as carbon source. The highest growth rate was obtained for bacterial cells with yeast extract: the latter added gave a stimulated (2–3.5 fold) growth rate than using glutamate. R. sphaeroides suspension oxidation–reduction potential (ORP), which was measured with a platinum electrode, decreased down to low negative values with nitrogen source for both strains. ORP decreased down to more negative values (−610 ± 25 mV) in the presence of yeast extract than when adding glutamate (−405 ± 15 mV) compared to the control (without nitrogen source addition): the significant decrease of ORP indicated enhanced (∼6 fold) H2 yield. The noticeable ORP decrease measured with the titanium-silicate electrode and simultaneously the increase of extracellular pH ([pH]out) were observed; ORP was more negative at alkaline [pH]out. Thus, the optimal culture conditions with nitrogen and carbon sources for bacterial growth stimulation and enhanced H2 production were established. The ORP decrease together with the increase of [pH]out point out a significant role of reduction processes in cell growth and ability of bacteria to live.  相似文献   

2.
Cheese whey wastewater diluted to 10 g lactose/L was initially subjected to dark-fermentation by Enterobacter aerogenes MTCC 2822, and the VFAs-rich spent medium (acetic acid 1900 mg/L, butyric acid 537 mg/L, and traces of propionic acid) was subjected to photo-fermentation through enrichment by Ni2+ (0–8 μmol/L), Fe2+ (0–100 μmol/L) or Mg2+ (0–15 mmol/L) in batch mode by Rhodopseudomonas BHU 01 strain. The maximum cumulative H2 production (144 ml) and yield (58 mmol) was obtained at 4 μmol Ni2+/L. Likewise, Fe2+ (60 μmol/L) resulted in maximum cumulative H2 production (139 ml) and yield (56 mmol). Nevertheless, 6 mmol of Mg2+ did not significantly affect H2 production (110 ml) or yield (44 mmol); the latter value in close proximity with the control (37 mmol). The concomitant reduction in COD was maximum (15.61%) for 4 μmol Ni2+/L, followed by 15.33% for 60 μmol Fe2+/L, and the least for 6 mmol Mg2+/L (14.5%). The observations suggest the role of Fe2+ and Ni2+ in regulation of nitrogenase and hydrogenase, while that of Mg2+ mainly in the biosynthesis of photopigment bacteriochlorophyll (Bchl).  相似文献   

3.
Rhodobacter sphaeroides MDC 6521 isolated from Arzni mineral springs in Armenia is able to produce bio-hydrogen (H2) in anaerobic conditions upon illumination in the presence of various metal ions. The significant aspect in regulation of H2 production by these bacteria and its energetics is the requirement for F0F1-ATPase, the main membrane enzyme responsible for generation of proton motive force under anaerobic conditions. In order to determine the mediatory role of F0F1 in H2 production, the effects of various metal ions (Mn2+, Mg2+, Fe2+, Ni2+, and Mo6+) on N,N′-dicyclohexylcarbodiimide inhibited ATPase activity of R. sphaeroides membrane vesicles were investigated. These ions in appropriate concentrations considerably enhanced H2 production, which was not observed in the absence of Fe2+, indicate the requirement for Fe2+. The R. sphaeroides membrane vesicles demonstrated significant ATPase activity. In the absence of Fe2+ inhibition (∼80%) of ATPase activity was observed, which was increased by addition of metal ions. A higher ATPase activity was detected in the presence of Fe2+ (80 μM) and Mo6+ (16 μM). These results indicate a relationship between the F0F1-ATPase activity and H2 production that might be a significant pathway to provide novel evidence of a requirement for F0F1-ATPase in H2 production by R. sphaeroides.  相似文献   

4.
Rhodobacter sphaeroides MDC6522 is able to produce hydrogen (H2) during photofermentation. Metronidazole (2-methyl-5-nitroimidazole-1-ethanol) has been shown to affect bacterial growth under anaerobic nitrogen-limited conditions in a concentration-dependent manner (within the range of 0.1–2 mM) by increasing lag phase duration and decreasing growth rate. The addition of metronidazole into the growth medium resulted in a delayed decrease of redox potential (Eh): by the addition of 0.1 mM metronidazole Eh decreased to −590 ± 25 mV, whereas in the presence of 2 mM Eh drop down was to −175 ± 15 mV. H2 production during R. sphaeroides growth disappeared in the presence of metronidazole. By addition of 0.5 mM metronidazole H2 yield was ∼8 fold lower in comparison with control; whereas the bacterium was unable to produce H2 in the presence of 1–2 mM. The effects of metronidazole on nitrogenase-dependent H2 production by R. sphaeroides might be explained by change of general photosynthetic electron transport with metronidazole as an alternative electron acceptor instead of nitrogenase. ATPase activity of membrane vesicles was determined with and without N,N'-dicyclohexylcarbodiimide (DCCD), an inhibitor of the F0F1-ATPase. It was revealed that DCCD-inhibited ATPase activity increased in the presence of metronidazole. It is possible that this effect may be resulted in either by direct affect of metronidazole on the F0F1-ATPase, or by its effect on Eh regulating ATPase activity.  相似文献   

5.
In the present study, the effect of Ni2+ (0–10 μmol/l), Fe2+ (0–200 μmol/l) and Mg2+ (0–15 mmol/l) concentration on photo-hydrogen production from acetate was investigated by batch culture. Results showed that under a proper concentration range, Ni2+ was able to enhance the hydrogen production rate and the hydrogen yield; Fe2+ was able to increase the hydrogen yield, and hydrogen production rate was enhanced only when the culturing time was 24–72 h. Ni2+ and Fe2+ at a higher concentration inhibited cell growth. When Ni2+ and Fe2+ concentrations were 4 μmol/l and 80 μmol/l, respectively, maximal hydrogen yield of 2.87 and 2.78 mol H2/mol acetate was obtained when batch culturing at 35 °C with initial pH 7.0. Mg2+ did not significantly affect hydrogen production and hydrogen yield which maintained at about 2.45 mol H2/mol acetate, but it was favorable to cell growth.  相似文献   

6.
7.
In this study we described the isolation of eight new strains of purple non-sulfur bacteria resistant to salinity ≥30 g L−1 and high concentration of VFAs (200 mM). These strains were characterized by their general physiological properties and the occurrence of hupSL genes. Some correlation was observed between the rate of H2 photoproduction, the absence of hupSL genes and hydrogenase activity. Two fast-growing strains without hupSL genes showed high nitrogenase activity and hydrogen accumulation during growth on Ormerod medium. These strains were capable of H2 photoproduction using non-treated dark culture (75% in water) after dark fermentation of starch at 30 g L−1, unlike control strains, Rhodobacter capsulatus B10 and Rb. sphaeroides GL. New N7 and 13 strains identified as Rb. sphaeroides can be recommended for application in a two-stage H2 production system.  相似文献   

8.
Rhodobacter sphaeroides O.U.001 is one of the candidates for photobiological hydrogen production among purple non-sulfur bacteria. Hydrogen is produced by Mo-nitrogenase from organic acids such as malate or lactate. A hupSL in frame deletion mutant strain was constructed without using any antibiotic resistance gene. The hydrogen production potential of the R. sphaeroides O.U.001 and its newly constructed hupSL deleted mutant strain in acetate media was evaluated and compared with malate containing media. The hupSLR. sphaeroides produced 2.42 l H2/l culture and 0.25 l H2/l culture in 15 mM malate and 30 mM acetate containing media, respectively, as compared to the wild type cells which evolved 1.97 l H2/l culture and 0.21 l H2/l culture in malate and acetate containing media, correspondingly. According to the results, hupSLR. sphaeroides is a better hydrogen producer but acetate alone does not seem to be an efficient carbon source for photoheterotrophic H2 production by R. sphaeroides.  相似文献   

9.
This study was devoted to investigate production of hydrogen gas from acid hydrolyzed molasses by Escherichia coli HD701 and to explore the possible use of the waste bacterial biomass in biosorption technology. In variable substrate concentration experiments (1, 2.5, 5, 10 and 15 g L−1), the highest cumulative hydrogen gas (570 ml H2 L−1) and formation rate (19 ml H2 h−1 L−1) were obtained from 10 g L−1 reducing sugars. However, the highest yield (132 ml H2 g−1 reducing sugars) was obtained at a moderate hydrogen formation rate (11 ml H2 h−1 L−1) from 2.5 g L−1 reducing sugars. Subsequent to H2 production, the waste E. coli biomass was collected and its biosorption efficiency for Cd2+ and Zn2+ was investigated. The biosorption kinetics of both heavy metals fitted well with the pseudo second-order kinetic model. Based on the Langmuir biosorption isotherm, the maximum biosorption capacities (qmax) of E. coli waste biomass for Cd2+ and Zn2+ were 162.1 and 137.9 (mg/g), respectively. These qmax values are higher than those of many other previously studied biosorbents and were around three times more than that of aerobically grown E. coli. The FTIR spectra showed an appearance of strong peaks for the amine groups and an increase in the intensity of many other functional groups in the waste biomass of E. coli after hydrogen production in comparison to that of aerobically grown E. coli which explain the higher biosorption capacity for Cd2+ or Zn2+ by the waste biomass of E. coli after hydrogen production. These results indicate that E. coli waste biomass after hydrogen production can be efficiently used in biosorption technology. Interlinking such biotechnologies is potentially possible in future applications to reduce the cost of the biosorption technology and duplicate the benefits of biological H2 production technology.  相似文献   

10.
The present study deals with the optimization of pretreatment conditions followed by thermophilic dark fermentative hydrogen production using Anabaena PCC 7120 as substrate by mixed microflora. Different airlift photobioreactors with ratio of area of downcomer and riser (Ad/Ar) in range of 0.4–3.2 were considered. Maximum biomass concentration of 1.63 g L−1 in 9 d under light intensity of 120 μE m−2 s−1 was observed at Ad/Ar of 1.6. The mixing time of the reactors was inversely proportional to Ad/Ar. Maximal H2 production was found to be 1600 mL L−1 upon pretreatment with amylase followed by thermophilic fermentation for 24 h compared to other methods like sonication (200 mL L−1), autoclave (600 mL L−1) and HCl treatment (1230 mL L−1). The decrease of pH from 6.5 to 5.0 during fermentation was due to the accumulation of volatile fatty acids. Amylase pretreatment gave higher reducible sugar content of 7.6 g L−1 as compare to other pretreatments. Thermophilic fermentation of pretreated Anabaena biomass by mixed bacterial culture was found suitable for H2 production.  相似文献   

11.
12.
In this study, a new outer-cycle flat-panel photobioreactor was designed for an anaerobic, photo-fermentation process by Rhodobacter sphaeroides ZX-5. In order to obtain the high hydrogen yield, photo-hydrogen production by fed-batch culture with on-line oxidation-reduction potential (ORP) feedback control was investigated. Meanwhile, the effects of feeding malic acid concentration and pH adjustment on the growth and hydrogen production of R. sphaeroides ZX-5 were studied. In the entire fed-batch culture, biomass (i.e., OD660) rapidly increased up to 1.79 within 18 h, and then OD660 value stayed constant within a range of 1.85-2.18 until the end of the photo-fermentation. The cumulative hydrogen volumes in each phase of fed-batch process were 2339, 1439, 1328, and 510 ml H2/l-culture, respectively. Throughout the entire repeated fed-batch photo-fermentation, the maximum substrate conversion efficiency of 73.03% was observed in the first fed-batch process, obviously higher than that obtained from batch culture process (59.81%). In addition, compared to the batch culture, a much higher maximum hydrogen production rate (102.33 ml H2/l h) was achieved during fed-batch culture. The results demonstrated that photo-hydrogen production using fed-batch operation based on ORP feedback control is a favorable choice of sustainable and feasible strategy to improve phototrophic hydrogen production efficiency.  相似文献   

13.
The effect of culture parameters on hydrogen production using strain GHL15 in batch culture was investigated. The strain belongs to the genus Thermoanaerobacter with 98.9% similarity to Thermoanaerobacter yonseiensis and 98.5% to Thermoanaerobacter keratinophilus with a temperature optimum of 65–70 °C and a pH optimum of 6–7. The strain metabolizes various pentoses, hexoses, and disaccharides to acetate, ethanol, hydrogen, and carbon dioxide. However substrate inhibition was observed above 10 mM glucose concentration. Maximum hydrogen yields on glucose were 3.1 mol H2 mol−1 glucose at very low partial pressure of hydrogen. Hydrogen production from various lignocellulosic biomass hydrolysates was investigated in batch culture. Various pretreatment methods were examined including acid, base, and enzymatic (Celluclast® and Novozyme 188) hydrolysis. Maximum hydrogen production (5.8–6.0 mmol H2 g−1 dw) was observed from Whatman paper (cellulose) hydrolysates although less hydrogen was produced by hydrolysates from other examined lignocellulosic materials (maximally 4.83 mmol H2 g−1 dw of grass hydrolysate). The hydrogen yields from all lignocellulosic hydrolysates were improved by acid and alkaline pretreatments, with maximum yields on grass, 7.6 mmol H2 g−1 dw.  相似文献   

14.
Three different Rhodobacter sphaeroides (RS) strains (RS–NRRL, RS–DSMZ and RS–RV) and their combinations were used for light fermentation of dark fermentation effluent of ground wheat containing volatile fatty acids (VFA). In terms of cumulative hydrogen formation, RS–NRRL performed better than the other two strains producing 48 ml H2 in 180 h. However, RS–RV resulted in the highest hydrogen yield of 250 ml H2 g−1 TVFA. Specific hydrogen production rate (SHPR) with the RS–NRRL was also better in comparison to the others (13.8 ml H2 g−1 biomass h−1). When combinations of those three strains were used, RS–RV + RS–DSMZ resulted in the highest cumulative hydrogen formation (90 ml H2 in 330 h). However, hydrogen yield (693 ml H2 g−1 TVFA) and SHPR (12.1 ml H2 g−1 biomass h−1) were higher with the combination of the three different strains. On the basis of Gompertz equation coefficients mixed culture of the three different strains gave the highest cumulative hydrogen and formation rate probably due to synergistic interaction among the strains. The effects of initial TVFA and NH4–N concentrations on hydrogen formation were investigated for the mixed culture of the three strains. The optimum TVFA and NH4–N concentrations maximizing the hydrogen formation were determined as 2350 and 47 mg L−1, respectively.  相似文献   

15.
Purple non-sulfur (PNS) bacteria can convert volatile fatty acids into hydrogen with a high substrate conversion efficiency. However, when PNS bacteria utilize sugars as a carbon source, such as glucose and sucrose, the substrate conversion efficiency is relatively low. In order to investigate the contributions of the glucose catabolic pathways in Rhodobacter sphaeroides 6016 to its hydrogen production, the cfxA gene from the Embden–Meyerhof–Parnas (EMP) pathway, edd from the Entner–Doudoroff (ED) pathway, and kdg from the semi-phosphorylative ED bypass were knocked out to construct the mutant strains edd, cfxA, and kdg, respectively. Additionally, two of these three genes were knocked out to construct the mutant strains kdgedd, kdgcfxA, and cfxAedd. Hydrogen productions by these mutant strains were compared to that of the wild type strain 6016 using 25 mM glucose as a carbon source. Compared to 6016, variations in hydrogen production and growth were detected in the edd mutant strains (kdgedd, cfxAedd, and edd), while no obvious changes were detected in the others. Notably, the kdgedd mutant did not produce hydrogen, and its maximum growth was 70% less than that of R. sphaeroides 6016. These results indicate that the ED pathway and semi-phosphorylative ED bypass have a governing impact on cell growth and hydrogen production from glucose in R. sphaeroides 6016. The potential synergistic function of the ED pathway and semi-phosphorylative ED bypass and the reasons for the low hydrogen yield from sugar carbon sources in R. sphaeroides 6016 are discussed.  相似文献   

16.
In this study, a pilot solar tubular photobioreactor was successfully implemented for fed batch operation in outdoor conditions for photofermentative hydrogen production with Rhodobacter capsulatus (Hup) mutant. The bacteria had a rapid growth with a specific growth rate of 0.052 h−1 in the batch exponential phase and cell dry weight remained in the range of 1–1.5 g/L throughout the fed batch operation. The feeding strategy was to keep acetic acid concentration in the photobioreactor at the range of 20 mM by adjusting feed acetate concentration. The maximum molar productivity obtained was 0.40 mol H2/(m3 h) and the yield obtained was 0.35 mol H2 per mole of acetic acid fed. Evolved gas contained 95–99% hydrogen and the rest was carbon dioxide by volume.  相似文献   

17.
Carbon monoxide (CO) is highly toxic but is an abundant carbon source that can be utilized for the production of hydrogen (H2). CO-dependent H2 production is catalyzed by a unique enzyme complex composed of carbon monoxide dehydrogenase (CODH) and CO-dependent hydrogenase (CO–H2ase), both of which contain metal cluster(s). In this study, CODH and the required maturation proteins from the novel facultative anaerobic bacterium Citrobacter amalonaticus Y19 were cloned and heterologously expressed in Escherichia coli. For functional expression of CODH in E. coli, only CooF (ferredoxin-like protein) and CooS (CODH), not the maturation proteins, were needed. The recombinant E. coli BL21(DE3)-cooFS showed a 3.5-fold higher specific CODH activity (4.9 U mg protein−1) compared to C. amalonaticus Y19 (Y19) (1.4 U mg protein−1). Purified heterologous CODH from the soluble cell-free extract of the recombinant E. coli showed a specific activity of 170.6 U mg protein−1. Recombinant E. coli harboring Y19 CODH and maturation proteins did not produce H2 from CO, suggesting that the native hydrogenases present in E. coli could not substitute the Y19 CO–H2ase for CO-dependent H2 production.  相似文献   

18.
A novel marine hyperthermophile, Thermococcus onnurineus NA1, was found to grow on C1 carbon compounds, such as formate and carbon monoxide (CO), and produce hydrogen (H2). In the present study, the growth and H2 production of NA1 were examined to determine its potential as H2 producer. NA1 showed relatively high specific growth rates, 0.48 h−1 and 0.40 h−1 with CO (20%, v/v) and formate (100 mM), respectively, when cultivated in batch mode in a minimal salt medium fortified with 1.0 g L−1 yeast extract. On the other hand, cell growth in both cases stopped at approximately 6 h and the final cell densities were extremely low at 18.2 and 12.1 mg protein L−1 with CO and formate, respectively. The maximum final cell density could be improved greatly to 36.0 mg protein L−1 by optimizing CO content (50%, v/v) and yeast extract concentration (4.0 g L−1), but it was still very low. During the cell growth, formate and CO were used as energy source rather than carbon source. In the resting cell experiments, NA1 exhibited remarkably high H2 production activities as 385.0 and 207.5 μmol mg protein−1 h−1 for CO and formate, respectively. When formate (100 mM) or CO (100%, v/v) was added repeatedly at 30–35 h intervals, NA1 showed consistent H2 production for 3 cycles with a yield of approximately 1.0 mol H2 mol−1 for both CO and formate. This study suggests that T. onnurineus NA1 has a high H2 production potential from formate or CO but a method for achieving a high cell density culture is needed.  相似文献   

19.
Photofermentative H2 production at higher rate is desired to make H2 viable as cheap energy carrier. The process is influenced by C/N composition, pH levels, temperature, light intensity etc. In this study, Rhodobacter sphaeroides strain O.U 001 was used in the annular photobioreactor with working volume 1 L, initial pH of 6.7 ± 0.2, inoculum age 36 h, inoculum volume 10% (v/v), 250 rpm stirring and light intensity of 15 ± 1.1 W m−2. The effect of parameters, i.e. variation in concentration of DL malic acid, L glutamic acid and temperature on the H2 production was noted using three factor three level full factorial designs. Surface and contour plots of the regression models revealed optimum H2 production rate of 7.97 mL H2 L−1 h−1 at 32 °C with 2.012 g L−1 DL malic acid and 0.297 g L−1 L glutamic acid, which showed an excellent correlation (99.36%) with experimental H2 production rate of 7.92 mL H2 L−1 h−1.  相似文献   

20.
A type of Yb2O3 doped Ni–ZrO2 catalyst for ethanol steam reforming was developed, and displayed excellent catalyzing performance for the selective formation of H2 and CO2. Over a Ni1.25Zr1Yb0.8 catalyst, STY(H2) can maintain stable at the level of 0.396 mol h−1 g−1 (data taken 120 h after the reaction started) under the reaction conditions of 0.5 MPa and 723 K, which was 1.6 times that (0.247 mol h−1 g−1) of the Yb-free counterpart Ni1.25Zr1. Characterization of the catalyst revealed that dissolution of an appropriate amount of Yb3+ ions in the zirconia host resulted in the formation of the Zr–Yb composite oxide with cubic-ZrO2 structure, c-(Zr–Yb)Oz, which inhibited effectively the transformation of c-ZrO2 to thermodynamically more stable m-ZrO2, thus avoiding sintering of the (Zr–Yb)Oz composite. It was demonstrated that the doping of Yb2O3 to Ni–ZrO2 changed also the valence states or the micro-environments of the Ni-species at the quasi-active surface of the tested catalyst, which was conducive to inhibiting agglomeration of the Nix0–Nin+ species active catalytically, with resulting in maintaining the high metallic nickel dispersion and inhibiting coking. The aforementioned two factors both contributed to improving the activity and operating stability as well as heat-resistant quality of the catalyst.  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