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 187 毫秒
1.
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   

2.
3.
Fe3+ doped TiO2 photocatalysts were prepared by hydrothermal treatment for the photocatalytic water splitting to produce stoichiometric hydrogen and oxygen under visible light irradiation. It was found that hydrothermal treatment at 110 °C for 10 h was essential for the synthesis of highly stabilized Fe3+ doped TiO2 photocatalysts. The synthesized photocatalysts were characterized by field emission scanning electron microscopy (FE-SEM), X-ray diffraction (XRD), ultraviolet–visible diffuse reflectance spectroscopy (UV–vis DRS) and BET surface area techniques. The doping of highly stabilized Fe3+ in the titania matrix leads to significant red shift of optical response towards visible light owing to the reduced band gap energy. Optimum amount of Fe3+ doped TiO2, 1.0 wt% Fe/TiO2, showed drastically improved hydrogen production performance of 12.5 μmol-H2/h in aqueous methanol and 1.8 μmol-H2/h in pure water, respectively. This Fe/TiO2 photocatalyst was stable for 36 h without significant deactivation in the water splitting reaction.  相似文献   

4.
Various metal ions play a key role in biohydrogen (H2) production by phototrophic bacteria through incorporation into or stimulating the responsible enzymes and/or related pathways. The Ni (II) and Mg (II) ions effects on growth and H2 production by Rhodobacter sphaeroides strain MDC6521 isolated from mineral springs in Armenia were established. The highest growth specific rate was obtained with 4–6 μM Ni2+ and 5 mM Mg2+. pH of the growth medium changed from 7.0 to 9.2–9.4 during the bacterial growth up to 72 h in spite of Ni2+ added but pH increased in different manner with Mg2+. In the presence of 2–4 μM Ni2+ external oxidation-reduction potential (ORP) decreased to more negative values (−800 ± 15 mV). This decrease of ORP indicated ∼2.7-fold enhanced H2 yield (9.80 mmol L−1) with Ni2+ compared with the control (without Ni2+). The H2 yield determined in the medium with Mg2+ was ∼2.2 fold higher than that with 1 mM Mg2+. These results reveal new regulatory ways to improve H2 production by R. sphaeroides those were depending on Ni2+ and Mg2+ of different concentrations.  相似文献   

5.
A new type of Li1−xFe0.8Ni0.2O2–LixMnO2 (Mn/(Fe + Ni + Mn) = 0.8) material was synthesized at 350 °C in air atmosphere using a solid-state reaction. The material had an XRD pattern that closely resembled that of the original Li1−xFeO2–LixMnO2 (Mn/(Fe + Mn) = 0.8) with much reduced impurity peaks. The Li/Li1−xFe0.8Ni0.2O2–LixMnO2 cell showed a high initial discharge capacity above 192 mAh g−1, which was higher than that of the parent Li/Li1−xFeO2–LixMnO2 (186 mAh g−1). We expected that the increase of initial discharge capacity and the change of shape of discharge curve for the Li/Li1−xFe0.8Ni0.2O2–LixMnO2 cell is the result from the redox reaction from Ni2+ to Ni3+ during charge/discharge process. This cell exhibited not only a typical voltage plateau in the 2.8 V region, but also an excellent cycle retention rate (96%) up to 45 cycles.  相似文献   

6.
In this paper, the effect of hydraulic retention time (HRT, 16 h–4 h) on fermentative hydrogen production by mixed cultures was firstly investigated in a sucrose-fed anaerobic continuous stirred tank reactor (CSTR) at 35 °C and initial pH 8.79. After stable operations at HRT of 16–6 h, the bioreactor became unstable when the HRT was lowered to 4 h. The maximum hydrogen yield reached 3.28 mol H2/mol-Sucrose at HRT 4 h. Supplementation of Cu2+ at HRT 4 h improved the operation stability through enhancement of substrate degradation efficiency. The effect of Cu2+ concentration ranging from 1.28 to 102.4 mg/L on fermentative hydrogen production was studied. The results showed that Cu2+ was able to enhance the hydrogen production yield with increasing Cu2+ concentration from 1.28 to 6.4 mg/L. The maximum hydrogen yield of 3.31 mol H2/mol-Sucrose and the maximum hydrogen production rate of 14.44 L H2/Day/L-Reactor were obtained at 6.4 mg/L Cu2+ and HRT 4 h Cu2+ at much higher concentration could inhibit the hydrogen production, but it could increase substrate degradation efficiency (12.8 and 25.6 mg/L Cu2+). The concentration of Cu2+ had effect on the distribution of soluble metabolite.  相似文献   

7.
The effect of ferrous ion (0–3.2 mg/l) on photo heterotrophic hydrogen production was studied in batch culture using sodium lactate as substrate. The results showed that hydrogen production by Rhodobacter sphaeroides   was significantly suppressed when Fe2+Fe2+ was limited. Hydrogen production increased linearly with an increase in Fe2+Fe2+ concentration in the range of 0–1.6 mg/l; reaching a maximum at 2.4 mg/l. When hydrogen production was suppressed in the above medium, a pH increase to 8.9 was observed, and the ratio of lactate utilized to total organic carbon removal was found to be increased, indicating that more soluble organic products were produced. Under the Fe2+Fe2+ limited conditions, ferrous iron was shown to have a greater effect on hydrogen production by Rb. sphaeroides than that by the anaerobic heterotrophic bacterium Clostridium butyricum.  相似文献   

8.
This study evaluated hydrogen production by co-culture of Ethanoligenens harbinense B49 and immobilized Rhodopseudomonas faecalis RLD-53 with different control strategies. To enhance cooperation of dark and photo-fermentation bacteria during hydrogen production process, the glucose concentration, phosphate buffer concentration and initial pH were controlled at 6 g/l, 50 mmol/l and 7.5, respectively. The maximum yield and rate of hydrogen production were 3.10 mol H2/mol glucose and 17.2 mmol H2/l/h, respectively. Ethanol from E. harbinense B49 in acetate medium can enhance hydrogen production by R. faecalis RLD-53 except the ratio of ethanol to acetate (RE/A) among 0.8 to 1.0. Control of the proper phosphate buffer concentration (50 mmol/l) not only increased acetic acid production by E. harbinense B49, but also maintained stable pH of co-culture system. Therefore, the results showed that co-culture of E. harbinense B49 and immobilized R. faecalis RLD-53 was a promising way of converting glucose into hydrogen.  相似文献   

9.
In this study, hydrogen gas was produced from starch feedstock via combination of enzymatic hydrolysis of starch and dark hydrogen fermentation. Starch hydrolysis was conducted using batch culture of Caldimonas taiwanensis On1 able to hydrolyze starch completely under the optimal condition of 55 °C and pH 7.5, giving a yield of 0.46–0.53 g reducing sugar/g starch. Five H2-producing pure strains and a mixed culture were used for hydrogen production from raw and hydrolyzed starch. All the cultures could produce H2 from hydrolyzed starch, whereas only two pure strains (i.e., Clostridium butyricum CGS2 and CGS5) and the mixed culture were able to ferment raw starch. Nevertheless, all the cultures displayed higher hydrogen production efficiencies while using the starch hydrolysate, leading to a maximum specific H2 production rate of 116 and 118 ml/g VSS/h, for Cl. butyricumCGS2 and Cl. pasteurianum CH5, respectively. Meanwhile, the H2 yield obtained from strain CGS2 and strain CH5 was 1.23 and 1.28 mol H2/mol glucose, respectively. The best starch-fermenting strain Cl. butyricum CGS2 was further used for continuous H2 production using hydrolyzed starch as the carbon source under different hydraulic retention time (HRT). When the HRT was gradually shortened from 12 to 2 h, the specific H2 production rate increased from 250 to 534 ml/g  VSS/h, whereas the H2 yield decreased from 2.03 to 1.50  mol H2/mol glucose. While operating at 2 h HRT, the volumetric H2 production rate reached a high level of 1.5 l/h/l.  相似文献   

10.
In this study, anaerobic mixed microbial consortium isolated from a local sewage treatment plant in Guwahati, India, was used to convert carbon monoxide (CO) to hydrogen. The consortium was initially grown in acetate containing medium and later acclimatized to utilize CO as the sole carbon source for hydrogen production. By 16S rDNA analysis, the consortium was identified to be predominantly Petrobacter sp. Statistically designed experiments were then applied to optimize the CO conversion and hydrogen production by the anaerobic mixed consortium. To evaluate the significant factors that influenced the biohydrogen production, Plackett–Burman screening design of experiments was applied, which revealed that temperature and Fe2+ influenced the most on hydrogen production with P values less than 0.05 each. The effect due to pH and Ni2+ was less with P values 0.120 and 0.132, respectively. Concentration of Fe2+ and Ni2+ in the medium was then subsequently optimized by using Central Composite Design (CCD) of experiments followed by response surface methodology (RSM) which yielded the optimum value of 213 mg/L for Fe2+ and 2.2 mg/L for Ni2+. At these optimum conditions, 60.8 mol hydrogen production was achieved which was 8% higher than that observed from the screening experiment.  相似文献   

11.
A new hydrogen-producing bacterial strain Ethanoligenens harbinense B49 was examined for its capability of H2 production with glucose as sole carbon source. The H2 production was significantly affected by the concentration of the yeast powder and phosphate in the synthetic medium. The optimized concentration of yeast powder was 0.3–0.5 g/L and the maximum hydrogen yield was obtained at the concentration of phosphate about 100–150 mmol/L. The dynamics of hydrogen production showed that rapid evolution of hydrogen appeared to start after the middle-phase of exponential growth (about 8 h). The maximum H2 yield and specific hydrogen production rate were estimated to be 2.26 mol H2/mol glucose and 27.74 mmol H2/g cell, respectively, when 10 g/L of glucose was present in the medium. The possible pathway of hydrogen production by Ethanoligenens sp. B49 during glucose fermentation was oxidative decarboxylation of pyruvate and the NADH pathway.  相似文献   

12.
The present work focused on the investigation of the hydrogen generation through the ethanol steam reforming over the core–shell structured NixOy–, FexOy–, and CoxOy–Pd loaded Zeolite Y catalysts. The transmission electron microscopy (TEM) image of NixOy–Pd represented a very clear core–shell structure, but the other two catalysts, CoxOy– and FexOy–Pd, were irregular and non-uniform. The catalytic performances differed according to the added core metal and the support. The core–shell structured CoxOy–Pd/Zeolite Y provided a significantly higher reforming reactivity compared to the other catalysts. The H2 production was maximized to 98% over CoxOy–Pd(50.0 wt%)/Zeolite Y at the conditions of reaction temperature 600 °C, CH3CH2OH:H2O = 1:3, and GHSV (gas hourly space velocity) 8400 h−1. In the mechanism that was suggested in this work, the cobalt component played an important role in the partial oxidation and the CO activation for acetaldehyde and CO2 respectively, and eventually, cobalt increased the hydrogen yield and suppressed the CO generation.  相似文献   

13.
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   

14.
The relationship between hydrogen generation and the age of culture was investigated under fed-batch growth conditions. The specific growth rate (μe) was determined during the log phase of the growth curve and the μeMax was 0.02643 h−1. Boltzmann's sigmoidal regression model was used to determine the specific rate of hydrogen evolution (μH): the maximum was 0.04440 h−1. At low irradiance (36–75 W m−2), an inverse relationship was found between μH and I; after increasing the irradiance further, μH reached a plateau (0.00916 h−1). The maximum reactor yield of cumulative hydrogen (4.5 l) was obtained at an irradiance of 320 W m−2, but the highest hydrogen evolution rate (17.217 ml h−1) was achieved at 500 W m−2. The light conversion efficiency reached its maximum (6.91%) at the lowest irradiance investigated (36 W m−2); when the irradiance increased further, it decreased progressively down to 0.36%.  相似文献   

15.
Dark fermentation, photo fermentation, and autotrophic microalgae cultivation were integrated to establish a high-yield and CO2-free biohydrogen production system by using different feedstock. Among the four carbon sources examined, sucrose was the most effective for the sequential dark (with Clostridium butyricum CGS5) and photo (with Rhodopseudomonas palutris WP3-5) fermentation process. The sequential dark–photo fermentation was stably operated for nearly 80 days, giving a maximum H2 yield of 11.61 mol H2/mol sucrose and a H2 production rate of 673.93 ml/h/l. The biogas produced from the sequential dark–photo fermentation (containing ca. 40.0% CO2) was directly fed into a microalga culture (Chlorella vulgaris C–C) cultivated at 30 °C under 60 μmol/m2/s illumination. The CO2 produced from the fermentation processes was completely consumed during the autotrophic growth of C. vulgaris C–C, resulting in a microalgal biomass concentration of 1999 mg/l composed mainly of 48.0% protein, 23.0% carbohydrate and 12.3% lipid.  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.
We evaluated the feasibility of improving the scale of hydrogen (H2) production from sugar cane distillery effluent using co-cultures of Citrobacter freundii 01, Enterobacter aerogenes E10 and Rhodopseudomonas palustris P2 at 100 m3 scale. The culture conditions at 100 ml and 2 L scales were optimized in minimal medium and we observed that the co-culture of the above three strains enhanced H2 productivity significantly. Results at the 100 m3 scale revealed a maximum of 21.38 kg of H2, corresponding to 10692.6 mol, which was obtained through batch method at 40 h from reducing sugar (3862.3 mol) as glucose. The average yield of H2 was 2.76 mol mol−1 glucose, and the rate of H2 production was estimated as 0.53 kg/100 m3/h. Our results demonstrate the utility of distillery effluent as a source of clean alternative energy and provide insights into treatment for industrial exploitation.  相似文献   

18.
Pure hydrogen can be stored and supplied directly to polymer electrolyte fuel cell by the redox of iron oxide: Fe3O4 + 4H2 → 3Fe + 4H2O and 4H2O + 3Fe → Fe3O4 + 4H2. Four bimetal-modified samples were prepared by impregnation. The hydrogen storage properties of the samples were investigated. The result shows that the Fe2O3–Mo–Al sample presented the most excellent catalytic activity and cyclic stability. H2 forming temperature and H2 forming rate could be surprisingly decreased and enhanced, respectively. The average H2 forming temperature at the rate of 250 μmol min−1·Fe-g−1 for Fe2O3–Mo–Al in the first 4 cycles could be decreased from 469 °C before the addition of Mo–Al to 273 °C after the addition of Mo–Al. The reason for it may be that the Mo–Al additive in the sample can prevent from the sintering of the particles and accelerate the H2O decomposition due to Mo taking part in the redox reaction. The average storage capacity of Fe2O3–Mo–Al was up to 4.68 wt%.  相似文献   

19.
A hydrogen producing facultative anaerobic alkaline tolerant novel bacterial strain was isolated from crude oil contaminated soil and identified as Enterobacter cloacae DT-1 based on 16S rRNA gene sequence analysis. DT-1 strain could utilize various carbon sources; glycerol, CMCellulose, glucose and xylose, which demonstrates that DT-1 has potential for hydrogen generation from renewable wastes. Batch fermentative studies were carried out for optimization of pH and Fe2+ concentration. DT-1 could generate hydrogen at wide range of pH (5–10) at 37 °C. Optimum pH was; 8, at which maximum hydrogen was obtained from glucose (32 mmol/L), when used as substrate in BSH medium containing 5 mg/L Fe2+ ion. Decrease in hydrogen partial pressure by lowering the total pressure in the fermenter head space, enhanced the hydrogen production performance of DT-1 from 32 mmol H2/L to 42 mmol H2/L from glucose and from 19 mmol H2/L to 33 mmol H2/L from xylose. Hydrogen yield efficiency (HY) of DT-1 from glucose and xylose was 1.4 mol H2/mol glucose and 2.2 mol H2/mol xylose, respectively. Scale up of batch fermentative hydrogen production in proto scale (20 L working volume) at regulated pH, enhanced the HY efficiency of DT-1 from 2.2 to 2.8 mol H2/mol xylose (1.27 fold increase in HY from laboratory scale). 84% of maximum theoretical possible HY efficiency from xylose was achieved by DT-1. Acetate and ethanol were the major metabolites generated during hydrogen production.  相似文献   

20.
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   

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