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 671 毫秒
1.
Gold adsorption from cyanide solution by bacterial (Bacillus subtilis), fungal (Penicillium chrysogenum) and seaweed (Sargassum fluitans) biomass was examined. At pH 2.0, these biomass types were capable of sequestering up to 8.0 µmol g−1, 7.2 µmol g−1 and 3.2 µmol g−1, respectively. An adverse effect of increasing solution ionic strength (NaNO3) on gold biosorption was observed. Gold‐loaded biomass could be eluted with 0.1 mol dm−3 NaOH with efficiencies higher than 90% at pH 5.0 at the Solid‐to‐Liquid ratio, S/L, = 4 (g dm−3). Cyanide mass balances for the adsorption, desorption as well as for the AVR process indicated the stability of the gold‐cyanide which did not dissociate either upon acidification or upon binding by biomass functional groups. Gold biosorption mainly involved anionic AuCN2 species bound by ionizable biomass functional groups carrying a positive charge when protonated. FTIR analyses indicated that the main biomass functional groups involved in gold biosorption are most probably nitrogen‐containing weak base groups. The present results confirmed that waste microbial biomaterials have some potential for removing and concentrating gold from solutions where it occurs as a gold‐cyanide complex. © 1999 Society of Chemical Industry  相似文献   

2.
The ability of Streptomyces rimosus biomass to bind zinc ions in batch mode was shown recently. The aim of this study was to determine the zinc uptake capacity by Streptomyces rimosus biomass in continuous mode. Bacterial biomass was able to bind more Zn(II) after pretreatment with sodium hydroxide (1 mol dm−3) than without treatment. The maximum adsorption capacity and the adsorption capacity at the saturation point calculated by means of both the exchange zone model and the Thomas model were practically identical of about 2.9 mgZn(II) g−1biomass. This result was lower than the batch adsorption capacity of Streptomyces rimosus, indicating that the packed‐bed is not the most appropriate process to exploit the bacterial biomass adsorption capacity. The effect of zinc concentration in the range of 10 to 200 mgZn(II) dm−3 on the biosorption capacity of the packed‐bed was not significant. Biomass regeneration with 0.1 mol dm−3 HCl gave a 90% recovery of the adsorbed Zn(II). © 1999 Society of Chemical Industry  相似文献   

3.
Zeolite NaY at 5 g dm−3 concentration, was selected to improve the production of ethanol fermentation by Saccharomyces bayanus from high glucose concentration media. The highest ethanol productivity (3·07 g dm−3 h−1) was obtained from a 220 g dm−3 initial glucose concentration, while the highest ethanol concentration (130 g dm−3) was obtained from a 350 g dm−3 glucose medium. The zeolite is believed to have acted as a pH regulator, maintaining the pH value around 3·7–3·8. Under these conditions cellular viability was preserved and metabolic activity was maintained. Thus all the glucose was consumed, and high ethanol productivity and concentration were obtained. Therefore, the addition of zeolite improved ethanol production from high concentrations of glucose by Saccharomyces bayanus. © 1998 Society of Chemical Industry  相似文献   

4.
Production of esterases (carboxyl esterase EC 3.1.1.1) and lipases (glycerol ester hydrolase EC 3.1.1.3) by Ophiostoma piliferum (CartapipTM), a fungus commercialized to decrease fatty acyl esters in wood, is described. The influence of various combinations of carbon and nitrogen sources, in the growth medium, was examined. Medium containing yeast extract as a nitrogen source and olive oil as a carbon source was found to be optimal for extracellular esterase (221 U dm−3) and lipase (152 U dm −3) activities. Further increases in those enzyme activities were achieved by decreasing medium pH from 6.5 to 5.5 (esterase 508 U dm−3; lipase 415 U dm −3) and increasing medium calcium content from 8 m mol dm−3 to 160 m mol dm−3 (esterase 4084 U dm−3; lipase 508 U dm −3) © 1999 Society of Chemical Industry  相似文献   

5.
Liquid–liquid extraction of Ag(I) from nitrate solutions using N‐(N′,N′‐diethyl thiocarbamoyl)‐N″‐phenylbenzamidine (TCBA) and 1‐6,‐diethylcarbamoyl imino‐1,6‐diphenyl‐2,5 dithiahexane (TCTH) dissolved in cumene has been studied. The extraction of Ag(I) from 1 mol dm−3 NO3 solutions by TCTH and TCBA was investigated as a function of several variables: equilibration time, organic phase diluent, pH of aqueous phase, Ag(I) and NO3 concentration in aqueous phase as well as TCBA and TCTH concentrations. Experimental equilibrium data were analysed numerically using the programs LETAGROP‐DISTR and LETAPL and the results showed that Ag(I) extraction could be explained assuming the formation of AgL and AgNO3HL with TCBA (HL) and AgNO3S with TCTH (S). The metal extraction was not influenced significantly by the structures of the thiourea derivatives used as extractants. The back extraction of Ag(I) from loaded organic phase was performed using different strippants and 0.5 mol dm−3 NaSCN was found to be efficient for this purpose. © 2000 Society of Chemical Industry  相似文献   

6.
The conditions for immobilizing the new L -aminoacylase-producing bacterial strain, Pseudomonas sp. BA2, by entrapment in κ-carrageenan gel, were investigated. The optimal gel concentration and cell load were determined. The addition of CoCl2 and N-acetyl-L -alanine to the immobilizing matrix enhanced L -aminoacylase activity. The enzymatic properties of immobilized Pseudomonas sp. BA2 were investigated. Enzyme activity in immobilized cells was optimal at a pH of 6·5 using 0·15 mol dm−3 Tris–maleate buffer at 45°C. The presence of 0·7 mmol dm−3 CoCl2 in the enzymatic reaction mixture improved L -aminoacylase activity. The immobilized cell preparation was used for the production of L -alanine from N-acetyl-DL -alanine in a batch reactor. Conversions of 100% were obtained using substrate concentrations ranging from 20 to 200 mmol dm−3. The reactor production was 0·74 mol h−1 g cell−1 dm−3 which is noticeably higher than that previously reported in the literature. © 1998 Society of Chemical Industry  相似文献   

7.
Two novel amperometric biosensors for the determination of L -malic acid in food samples have been compared. Both sensors make use of a Clark-type O2-electrode but differ in the enzymes used. The first sensor is composed of malate dehydrogenase (decarboxylating), also known as ‘malic enzyme’ (MDH(dec.), EC 1.1.1.40) and pyruvate oxidase (POP, EC 1.2.3.3). It covers a linear detection range from 1 μmol dm−3 to 0·9 mmol dm−3 L -malate, with a response time of 1·5 min (t90) and a relative standard deviation of 3·5%. Measurements with real samples offered a good correlation with the standard enzymatic assay (difference ±7%). Stored at room temperature, the response of the sensor is constant for 8 days. The second biosensor is based on the three enzyme sequence malate dehydrogenase (MDH, EC 1.1.1.37), oxaloacetate decarboxylase (OAC, EC 4.1.1.3) and pyruvate oxidase (POP, EC 1.2.3.3). It has a non-linear calibration curve. Concentrations from 5 μmol dm−3 to 1 mmol dm−3 L -malate can be detected, within a response time of 1·5 min and with a relative standard deviation of 20%. The lower detection limit for L -malate is 2 μmol dm−3. The response is constant for 10 days when the sensor is stored at room temperature.  相似文献   

8.
Z. Cai  D. Zhang  K. Cheng  C. Song  Y. Li  K. Ye  P. Yan  D. Cao  G. Wang 《Fuel Cells》2016,16(5):646-651
In this paper, we employed the as‐prepared TiO2/C core/shell nanoarrays (TiO2/C) obtained by a facile thermal evaporation method as a three‐dimensional (3D) architecture to support Pt nanoparticles through an optimized electrodeposition process. The morphology and structure of the as‐prepared electrode are characterized by X‐ray diffraction (XRD), scanning electron microscopy (SEM) and transmission electron microscopy (TEM). Its catalytic performance towards H2O2 electroreduction in basic medium is evaluated by linear sweep voltammetry (LSV) and chronoamperometry (CV). Results revealed that the electrode exhibits significantly high catalytic activity. The current density reached –0.172 A cm−2 in 1 mol dm−3 NaOH and 0.5 mol dm−3 H2O2 at –1.1 V (vs. Ag/AgCl). This high performance might be due to the 3D electrode architecture inheriting the high electronic conductivity from carbon shell and providing a short pathway for the ion diffusion, and the using of Pt owning an excellent catalytic activity.  相似文献   

9.
The permeability characteristics of the strongly basic polymer membrane Neosepta® AFN‐7, (Tokuyama Soda) have been studied for acetic, propionic, lactic, tartaric, oxalic, and citric acid. The results were interpreted by using the model of transport in reactive membranes. The specific constants, that is, the maximum flux Jmax, the reactivity constant K, and the permeability coefficient (P), were calculated using the experimental quasi‐stationary fluxes and the equation derived as a sum of reaction–diffusion (Michaelis–Menten‐type), and the solution–diffusion transport equation. The constants K and Jmax were found to range from 0.1 to 5 dm3 mol−1 and from 0.4 × 10−7 to 2.5 × 10−7 mol cm−2 s−1 depending, on the acid properties. The values of K and Jmax were correlated with the dissociation constants Kdis.acid, and the diffusion coefficients Daq.acid in aqueous media, respectively. It was found that the reaction–diffusion flux is predominating for all acids, except for the lactic one, when the feed concentration is lower than 0.5 mol dm−3. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 71: 2179–2190, 1999  相似文献   

10.
The sorption of linuron on bentonite desiccated at 110°C untreated, and acid‐treated with H2SO4 solutions over a concentration range between 0.25 M and 1.00 M from aqueous solution at 25°C has been studied by using batch experiments. In addition, column experiments were carried out with the bentonite sample treated with the 1.00 M H 2SO4 solution [B‐A(1.00)] by using two aqueous solutions of linuron of different concentrations (C=4.97 mg dm−3 and C=7.63 mg dm−3 ). The experimental data points have been fitted to the Langmuir equation in order to calculate the sorption capacities (Xm) of the samples; Xm values range from 0.02 g kg−1 for the untreated bentonite [B‐N] up to 0.20 g kg−1 for the sample acid‐treated with the 1.00 M H2 SO4 solution. The removal efficiency (R ) has also been calculated; R values ranging from 15.86% for the [B‐N] sample up to 41.54% for [B‐A(1.00)]. The batch experiments show that the acid‐treated bentonite is more effective than the natural bentonite in relation to sorption of linuron. The column experiments show that the B‐A(1.00) sample might be reasonably used in removing linuron, the column efficiency increasing from 61.8% for the C=7.63 mg dm−3 aqueous solution of linuron up to 77.6% for the C=4.97 mg dm−3 one. © 1999 Society of Chemical Industry  相似文献   

11.
The partially purified chlorophyllase, obtained from the alga Phaeodactylum tricornutum, was assayed for its hydrolytic activity towards the pheophytin in ternary micellar systems of hexane/Tris–HCl/surfactant. A wide range of surfactants, sorbitans (Span 20, 40, 60, 80 and 85) and polysorbates (Tween 20, 40, 60, 80 and 85), was used. The use of either 50 μmol dm−3 of Span 85 or 1 μmol dm−3 of Tween 80 increased the hydrolytic activity of chlorophyllase by 110 and 23%, respectively. The optimum values of pH, enzyme content, incubation time and temperature for the hydrolytic activity of chlorophyllase were determined as 8·25, 8·00 μg protein cm−3, 60 min and 27·5°C, respectively. The Vmax and Km values were 6·91 nmole hydrolyzed pheophytin mg−1 protein min−1 and 47·2 nmole pheophytin dm−3, respectively, in the Span 85 medium and 10·04 and 121·00, respectively, in the Tween 80 medium. The addition of optimized amounts of individual membrane lipids, L -α-phosphatidylcholine, L -α-phosphatidyl-DL -glycerol and β-carotene increased the hydrolytic activity of chlorophyllase by 50, 36 and 10%, respectively, for Span 85 and 30, 48 and 15%, respectively, for Tween 80. Phytol showed a competitive inhibitory effect on chlorophyllase activity in both Span and Tween systems with a Ki value of 15·5 and 14·3 μmol dm−3, respectively. High-performance liquid chromatography and spectrophotometry analyses were used to characterize the end-products of chlorophyllase hydrolytic reaction. © 1997 SCI.  相似文献   

12.
The effect of reaction conditions on the grafting parameter during grafting of acrylamide onto guar gum has been studied using peroxydiphosphate–metabisulphite redox pair at 35 °C. Grafting ratio, efficiency and add‐on all increase as the concentrations of peroxydiphosphate and acrylamide increase up to 40.0 × 10−3 mol dm−3 and 40.0 × 10−2 mol dm−3, respectively. It has been observed that the optimum concentrations of metabisulphite and guar gum for obtaining high grafting ratio, efficiency, add‐on and conversion are 6.0 × 10−3mol dm−3 and 91.7 × 10−2 g dm−3, respectively. © 2000 Society of Chemical Industry  相似文献   

13.
A new phase separation phenomenon was observed in which the perfluorooctanesulfonate ion (PFOS) and calcium ion form an ion‐pair associator and the sedimented liquid phase occurs from the homogeneous aqueous solution. This phenomenon was observed in the neutral pH region at room temperature (25 °C). The optimum concentration conditions for the reagents were [PFOS]T = 7 × 10−3 mol dm‐3 and [Ca2+]T = 1.1 mol dm‐3. When these findings were applied to the homogeneous liquid–liquid extraction of vitamin B12, the extraction percentage (E) was 83% and the concentration ratio (ie Va/Vs, where Va is the volume of the aqueous phase and Vs is the volume of the sedimented liquid phase) was a maximum of 149. The recovery of vitamin B12 was achieved by adding the propanol–acetone (20 : 80 v/v%) mixed solvent to the sedimented liquid phase; the vitamin B12 precipitated and was filtered. Both the PFOS and Ca2+ were removed by dissolution in the mixed solvent. The recovery percentage of vitamin B12 was 78%. © 1999 Society of Chemical Industry  相似文献   

14.
A simulated textile effluent (STE) was generated for use in laboratory biotreatment studies; this effluent contained one reactive azo dye, PROCION Red H‐E7B (1.5 g dm−3); sizing agent, Tissalys 150 (1.9 g dm−3); sodium chloride (1.5 g dm−3) and acetic acid (0.53 g dm−3) together with nutrients and trace elements, giving a mean COD of 3480 mg dm−3. An inclined tubular anaerobic digester (ITD) was operated for 9 months on the STE and a UASB reactor for 3 months. For a 57 day period anaerobic effluent from two reactors, a UASB and an ITD, was mixed and treated in an aerobic stage. In days 77–247 68% of the true colour of PROCION Red H‐E7B was removed by anaerobic treatment with no colour removal aerobically and up to 37% COD was removed anaerobically, with a corresponding BOD removal of 71%. For combined anaerobic and aerobic treatment a mean COD removal of 57% and BOD removal of 86% was achieved. Operation of the ITD at a 2.8 day HRT (volumetric loading rate (B v) 1.24 g COD dm−3day−1) and the UASB at a 2 day HRT (B v 1.74 g COD dm−3day−1) gave comparable COD removals but the UASB gave better true colour removal. Effluent from the combined process operating on this simulated waste still contained an average 1500 mg COD dm−3, and further treatment would be required to meet consent standards. © 1999 Society of Chemical Industry  相似文献   

15.
Mercuric reductase which originated from a recombinant Escherichia coli PWS1 was purified and immobilized on a chemically modified diatomaceous earth support. The mercury reduction kinetics, pH dependence, storage stability, and reusability of the immobilized enzyme were investigated. Four dyes were examined for their electron transfer efficiency with the soluble and bound mercuric reductase. Continuous mercury detoxification by the immobilized mercuric reductase was also performed in fixed‐bed processes. The effects of bed‐length, mercury loading rate, and electron donor on the performance of the fixed beds were assessed. Immobilized mercuric reductase exhibited substrate‐inhibition‐type kinetics with a maximal activity (1.2 nmol Hg mg−1 protein s−1) occurring at an initial Hg2+ concentration of 50 µmol dm−3. The optimal pH was 7.0 for the soluble and immobilized mercuric reductase, but the immobilized enzyme maintained higher relative activity for less favorable pH values. Immobilization of the enzyme appeared to significantly enhance its storage stability and reusability. Of four artificial electron donors tested, azure A (5 mmol dm−3) demonstrated the highest relative activity (78%) for soluble mercuric reductase. For the immobilized enzyme, neutral red (5 mmol dm−3) gave a relative activity of nearly 82%. With a fixed‐bed, the mercury‐reducing efficiency of using neutral red was only 30–40% of that obtained using NADPH. Fixed‐bed operations also showed that increased bed length facilitated mercury reduction rate, and the optimal performance of the beds was achieved at a flow rate of approximately 100–200 cm3 h−1. © 1999 Society of Chemical Industry  相似文献   

16.
The influence of adding ortho‐phenylenediamine (OPDA) during the polymerization of aniline on the characteristics of the resulting polymer film was examined. When using a platinum electrode, the deposits were obtained from solutions containing 0.1 mol dm?3 aniline and 1, 5, or 10 mmol dm?3 OPDA. The deposits were also prepared from solutions containing 0.5 mol dm?3 aniline and 5, 10, or 50 mmol dm?3 OPDA. In both cases, 3 mol dm?3 phosphoric acid solution was used as a supporting electrolyte. The characteristics of the obtained layers were investigated through the catalytic effect of different polymer layers on hydroquinone/quinone (H2Q/Q) test redox system. The results obtained confirm the earlier established catalytic effect on the potential of the redox reaction by shifting it to more reversible values. However, as the concentration of OPDA was increased, the resulting limiting current decreased, thus indicating in the presence of OPDA a lower population of the available active centers necessary for the catalytic reaction to proceed. The influence of OPDA on polymer characteristics was also studied by using scanning electron microscopy as well as electrochemical impedance spectroscopy. The polymer was synthesized on a stainless steel electrode (13% Cr) from a solution containing 0.5 mol dm?3 aniline and 5, 10, or 50 mmol dm?3 OPDA. The layers were tested in chloride‐containing solutions by monitoring the open circuit potential. The results obtained suggest that, by increasing the concentration of OPDA, the time of OCP in the passive region of stainless steel is prolonged. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

17.
A recombinant plasmid was constructed by inserting a DNA fragment with the coding region of Cu/Zn–superoxide dismutase (Cu/Zn–SOD) cDNA from sweet potato, Ipomoea batatas (l) Lam cv Tainong 57, into the 3′ end of the open reading frame of the glutathione S‐transferase (GST) gene in an expression vector, pGEX‐2T. The constructed plasmid was transformed into E coli XL1 Blue. Fusion proteins of Cu/Zn–SOD and GST (GST–SOD) were produced from the recombinant E coli. About 6 mg of GST–SOD fusion proteins could be obtained from 1 dm3 of cultural broth after induction with 0.075 mmol dm−3 Isopropyl‐β‐D ‐thiogalactoside (IPTG). Lactose was not an efficient inducer. High cell density culture was performed by fed‐batch fermentation using a glucose analyzer to control glucose concentration at 1 g dm−3. The cell density of the fed‐batch culture reached an OD600 of 30, the total amount of GST–SOD fusion protein was 100 mg dm−3 which is about 14 times more than that of the batch culture. Most of the fusion proteins were shown to be in an active monomeric form, and the molecular weight was estimated to be 45 kDa by SDS–PAGE and 47 kDa by gel filtration. The specific activity of the purified fusion proteins was about 1200 mg−1 and equal to 3200 unit per mg of SOD domain only. © 2000 Society of Chemical Industry  相似文献   

18.
Process variables were optimized for the production of lactic acid from pretreated beet molasses by Lactobacillus delbrueckii IFO 3202 for batch and continuous fermentations. In the batch fermentation, maximum yields (95·4% conversion, 77·1% effective) and maximum lactic acid volumetric productivity (4·83 g dm−3 h−1) was achieved at 45°C, pH 6·0, 78·2 g dm−3 sugar concentration with 10 g dm−3 yeast extract. Various cheaper nitrogen sources were replaced with yeast extract on equal nitrogen bases in batch fermentation. Of all the nitrogen sources tested, yeast extract yielded the highest and malt sprouts yielded the second highest level of lactic acid. In the continuous fermentation, maximum lactic acid (4·15%) was obtained at a dilution rate of 0·1 h−1. Maximum volumetric lactic acid productivity (11·20 g dm−3 h−1) occurred at D = 0·5 h−1 dilution rate. © 1997 SCI.  相似文献   

19.
The alkane monooxygenase AlkBGT from Pseudomonas putida GPo1 constitutes a versatile enzyme system for the ω‐oxyfunctionalization of medium chain‐length alkanes. In this study, recombinant Escherichia coli W3110 expressing alkBGT was investigated as whole‐cell catalyst for the regioselective biooxidation of fatty acid methyl esters to terminal alcohols. The ω‐functionalized products are of general economic interest, serving as building blocks for polymer synthesis. The whole‐cell catalysts proved to functionalize fatty acid methyl esters with a medium length alkyl chain specifically at the ω‐position. The highest specific hydroxylation activity of 104 U gCDW−1 was obtained with nonanoic acid methyl ester as substrate using resting cells of E. coli W3110 (pBT10). In an optimized set‐up, maximal 9‐hydroxynonanoic acid methyl ester yields of 95% were achieved. For this specific substrate, apparent whole‐cell kinetic parameters were determined with a Vmax of 204±9 U gCDW−1, a substrate uptake constant (KS) of 142±17 μM, and a specificity constant Vmax/KS of 1.4 U gCDW−1 μM −1 for the formation of the terminal alcohol. The same E. coli strain carrying additional alk genes showed a different substrate selectivity. A comparison of biocatalysis with whole cells and enriched enzyme preparations showed that both substrate availability and enzyme specificity control the efficiency of the whole‐cell bioconversion of the longer and more hydrophobic substrate dodecanoic acid methyl ester. The efficient coupling of redox cofactor oxidation and product formation, as determined in vitro, combined with the high in vivo activities make E. coli W3110 (pBT10) a promising biocatalyst for the preparative synthesis of terminally functionalized fatty acid methyl esters.  相似文献   

20.
The production of ligninolytic enzymes by Phanerochaete chrysosporium BKM‐F‐1767 (ATCC 24725) in laboratory‐scale bioreactors was studied. The cultivations were carried out in semi‐solid‐state conditions, employing corncob as carrier, which functioned both as a place of attachment and as a source of nutrients. Several bioreactor configurations were investigated in order to determine the most suitable one for ligninolytic enzyme production: a 1‐dm3‐static‐bed bioreactor, a 1‐dm3‐static‐bed bioreactor with air diffusers into the bed, a 0.5‐dm3‐static‐bed bioreactor with air diffusers into the bed and a tray bioreactor. Although the static‐bed configurations produced maximum individual lignin peroxidase (LiP) activities about 400 U dm−3 (1.0‐dm3 bioreactor) and about 700 U dm−3 (0.5‐dm3 bioreactor), manganese‐dependent peroxidase (MnP) was not detected throughout the cultures. Nevertheless, the tray configuration led to maximum individual MnP and LiP activities of about 200 U dm−3 and 300 U dm−3, respectively. Therefore, this configuration is the most adequate of the different bioreactor configurations tested in the present work, since the ligninolytic complex formed by MnP and LiP is more efficient for its application to bio‐processing systems. In addition, the results indicated the influence of the oxygen in ligninolytic enzyme production. © 2001 Society of Chemical Industry  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