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.
A new separation and purification process was developed for recovering 1,3‐propanediol (1,3‐PD) from crude glycerol‐based fermentation broth with high purity. The downstream process integrated chitosan flocculation, activated carbon decolorization, fixed bed cation exchange resin adsorption, and vacuum distillation. Breakthrough curves were measured considering the effect of sample concentration, flow rate, temperature, and resin stack height. Yoon–Nelson model was proposed to fit the fixed bed adsorption. The characteristic column parameters were calculated. Optimal condition for adsorption was 1,3‐PD, 30.0 g/L; flow rate, 1.00 mL/min; stacking height, 30.0 cm; and temperature, 298 K. Ethanol‐water (75%, 1 mL/min) was used as eluent to separate 1,3‐PD and glycerol with 95.3% 1,3‐PD elution rate. After vacuum distillation, the overall purity and yield of 1,3‐PD were 99.2% and 80.8% in the purification process, respectively. This is a simple and efficient downstream strategy for 1,3‐PD purification.  相似文献   

2.
Phosphoserine aminotransferase (SerC) from Escherichia coli (E. coli) MG1655 is engineered to catalyze the deamination of homoserine to 4‐hydroxy‐2‐ketobutyrate, a key reaction in producing 1,3‐propanediol (1,3‐PDO) from glucose in a novel glycerol‐independent metabolic pathway. To this end, a computation‐based rational approach is used to change the substrate specificity of SerC from l ‐phosphoserine to l ‐homoserine. In this approach, molecular dynamics simulations and virtual screening are combined to predict mutation sites. The enzyme activity of the best mutant, SerCR42W/R77W, is successfully improved by 4.2‐fold in comparison to the wild type when l ‐homoserine is used as the substrate, while its activity toward the natural substrate l ‐phosphoserine is completely deactivated. To validate the effects of the mutant on 1,3‐PDO production, the “homoserine to 1,3‐PDO” pathway is constructed in E. coli by coexpression of SerCR42W/R77W with pyruvate decarboxylase and alcohol dehydrogenase. The resulting mutant strain achieves the production of 3.03 g L?1 1,3‐PDO in fed‐batch fermentation, which is 13‐fold higher than the wild‐type strain and represents an important step forward to realize the promise of the glycerol‐independent synthetic pathway for 1,3‐PDO production from glucose.  相似文献   

3.
The ability of bacterial strains to assimilate glycerol derived from biodiesel facilities to produce metabolic compounds of importance for the food, textile and chemical industry, such as 1,3‐propanediol (PD), 2,3‐butanediol (BD) and ethanol (EtOH), was assessed. The screening of 84 bacterial strains was performed using glycerol as carbon source. After initial trials, 12 strains were identified capable of consuming raw glycerol under anaerobic conditions, whereas 5 strains consumed glycerol under aerobiosis. A plethora of metabolic compounds was synthesized; in anaerobic batch‐bioreactor cultures PD in quantities up to 11.3 g/L was produced by Clostridium butyricum NRRL B‐23495, while the respective value was 10.1 g/L for a newly isolated Citrobacter freundii. Adaptation of Cl. butyricum at higher initial glycerol concentration resulted in a PDmax concentration of ~32 g/L. BD was produced by a new Enterobacter aerogenes isolate in shake‐flask experiments, under fully aerobic conditions, with a maximum concentration of ~22 g/L which was achieved at an initial glycerol quantity of 55 g/L. A new Klebsiella oxytoca isolate converted waste glycerol into mixtures of PD, BD and EtOH at various ratios. Finally, another new C. freundii isolate converted waste glycerol into EtOH in anaerobic batch‐bioreactor cultures with constant pH, achieving a final EtOH concentration of 14.5 g/L, a conversion yield of 0.45 g/g and a volumetric productivity of ~0.7 g/L/h. As a conclusion, the current study confirmed the utilization of biodiesel‐derived raw glycerol as an appropriate substrate for the production of PD, BD and EtOH by several newly isolated bacterial strains under different experimental conditions.  相似文献   

4.
Apple pomace (AP), before and after chemical modification (CM), was assessed for the removal of cadmium (Cd2+) ions from aqueous solution by equilibrium, kinetics, and thermodynamics studies. The chemical modification of AP was done with succinic anhydride by a simple ring opening mechanism for providing a large surface area for cadmium adsorption. The surface area of chemically modified apple pomace (CMAP) increased about 18% after the treatment. The amount of CMAP required for cadmium removal was 50 times less than the unmodified AP. The Langmuir adsorption isotherm equation was found to be more suitable for the AP and CMAP adsorption experimental data with a correlation coefficient of r2 = 0.99 than was the Freundlich isotherm. The FTIR spectra of CMAP, with or without cadmium loading, indicated that ester (–COO), carboxyl (–CO), and amine (–NH2) groups were involved in the cadmium adsorption mechanism. The adsorption of cadmium ions onto AP and CMAP followed pseudo-second-order kinetics. The ΔG° value, at different temperatures, was calculated by applying the Van't Hoff equation and found to be negative, indicating that the reaction is spontaneous in nature. The value of ΔH° indicated that the adsorption was exothermic (−6.93 KJ mol−1) and endothermic (3.171 KJ mol−1) for AP and CMAP, respectively. CMAP could be reutilized for up to three cycles with a removal efficiency of 76.6%, while AP efficiency lessened significantly after a single cycle.  相似文献   

5.
In this paper we present a model for the ion exchange effects in protein adsorption. The model is applied to chromatography of lysozyme on strong cation exchanger ‘mono S’. The experimental and general thermodynamic aspects have been discussed in Part 1, the preceding paper. The main modelling assumptions are (i) the charge regulation is confined to the small layer of contact between adsorbed protein and exchanger surface, (ii) the contact layer as a whole is electroneutral and (iii) the number of protein acid/base groups and exchanger surface acid groups which participate in the ion exchange is proportional to the area of the contact layer. The model is fitted to the experimental data by adjustment of only two or three parameters. The experimental co-adsorption numbers are very well reproduced. A few conspicuous features emerge: (i) the number of protein acid/base groups and exchanger surface acid groups in the contact layer varies with the medium conditions, such that the number is higher when the interaction between protein and exchanger surface is stronger. (ii) There is indirect evidence for structural alterations in the upper layers of the exchanger surface: the adsorbed protein is probably partly ‘buried’ in the surface.  相似文献   

6.
1,3‐Propanediol (1,3‐PD) is a versatile bulk chemical and widely used as a monomer to synthesis polymers, such as polyesters, polyethers and polyurethanes. 1,3‐PD can be produced by microbial fermentation with the advantages of the environmental protection and sustainable development. Low substrate tolerance and wide by‐product profile limit microbial production of 1,3‐PD by Klebsiella pneumonia on industrial scale. In this study, microbial consortia were investigated to overcome some disadvantages of pure fermentation by single strain. Microbial consortium named DL38 from marine sludge gave the best performance. Its bacterial community composition was analyzed by 16S rRNA gene amplicon high‐throughput sequencing and showed that Enterobacteriaceae was the most abundant family. Compared with three K. pneumonia strains isolated from DL38, the microbial consortium could grow well at an initial glycerol concentration of 200 g/L to produce 81.40 g/L of 1,3‐PD with a yield of 0.63 mol/mol. This initial glycerol concentration is twice the highest concentration by single isolated strain and more than the critical value (188 g/L) extrapolated from the fermentation kinetics for K. pneumonia. On the other hand, a small amount of by‐products were produced in batch fermentation of microbial consortium DL38,  especially no 2,3‐butanediol detected. The mixed culture of strain W3, Y5 and Y1 improved the tolerance to glycerol and changed the metabolite profile of single strain W3. The batch fermentation with the natural proportion (W3: Y5: Y1 = 208: 82: 17) was superior to that with other proportions and single strain. This study showed that microbial consortium DL38 possessed excellent substrate tolerance, narrow by‐product profile and attractive potential for industrial production of 1,3‐PD.  相似文献   

7.
The kinetic, thermodynamic and isotherm modeling studies were carried out on adsorptive removal of Victoria blue (VB) dye using activated carbon, Ba/alginate and modified carbon/Ba/alginate polymer beads. The feasibility of sorption process was determined by varying the experimental parameters viz., dye concentration (4–20 mg L−1), contact time (10–90 min), pH (3–10), adsorbent dose (0.5–2.5 g) and temperature (303–343 K). Freundlich and Langmuir isotherms were determined and ascertained with the dimensionless separation factor (RL). Lagergren's pseudo-first order, pseudo-second order and intraparticle diffusion model equations were used to analyze the kinetics of the adsorption process. The thermodynamic consistency of adsorption was found with Gibbs free energy (ΔG°), changes in enthalpy (ΔH°) and entropy (ΔS°) were calculated using the Van’t Hoff plot. The polymer beads were characterized using Fourier Transform Infrared Spectroscopy (FTIR) and their morphology was determined by scanning electron microscopy (SEM).  相似文献   

8.
Euglena gracilis is shown to be able to grow on potato liquor as the main medium component leading to an interesting biotechnological product represented by paramylon – a β‐1,3‐glucan – and, at the same time, revaluating an otherwise annoying waste stream of the potato‐starch industry. Paramylon mass fractions of about 75% are obtained for biomass concentrations of 15 g/L during simple batch cultivation under heterotrophic conditions. Supplementation of the growth medium with glucose and the vitamins B1 and B12 are shown to improve growth rate as well as paramylon content. E. gracilis grows best at about 27.5°C without requiring pH control.  相似文献   

9.
Recently, it had been shown that Euglena gracilis was able to grow heterotrophically not only on synthetic media, but also on media based on potato liquor. Supplementation with glucose in both cases led to the accumulation of paramylon, a β‐1,3‐glucan. Thus, such a process may yield a valuable product accompanied by the revaluation of an otherwise annoying waste stream of the potato‐starch industry. Actually, process strategies have been evaluated in order to optimise the concentration of paramylon obtained at the end of the cultivation process. Therefore, cultivation processes based on fed‐batch and in particular repeated‐batch strategies have been studied. It is shown that repeated‐batch operation maybe particularly suited for such a process since E. gracilis seems to adapt gradually to the cultivation medium so that the concentration of media components may be increased step by step. Repeated‐batch cultivation of E. gracilis leads to biomass concentrations in access of 20 g/L with a consistent paramylon mass fraction of about 75%. Cultivations have been carried out at an operating temperature of 27.5°C. As had been found earlier already, pH control is not required during cultivation. On the basis of these results it is clear that repeated‐batch cultivation represent a simple and economic way for the production of paramylon by heterotrophic cultivation of E. gracilis.  相似文献   

10.
A freshwater filamentous green alga Spirogyra sp. was used as an inexpensive and efficient biosorbent for the removal of C.I. Acid Orange 7 (AO7), C.I. Basic Red 46 (BR46) and C.I. Basic Blue 3 (BB3) dyes from contaminated water. The effects of various physico–chemical parameters on dye removal efficiency were investigated, e.g. contact time, pH, initial dyes concentration, the amount of alga, temperature and biosorbent particle size. Dyes biosorption was a quick process and reactions reached to equilibrium conditions within 60 min. The biosorption capacity of three dyes onto alga was found in the following order: BR46 > BB3> AO7. The values of thermodynamic parameters, including ΔG, ΔH and ΔS, indicated that the biosorption of the dyes on the dried Spirogyra sp. biomass was feasible, spontaneous and endothermic. The pseudo-first order, pseudo-second order and the intraparticle diffusion models were applied to the experimental data in order to kinetically describe the removal mechanism of dyes, with the second one showing the best fit with the experimental kinetic biosorption data (R2 = 0.99). It was also found that the adsorption process followed the Freundlich isotherm model with the highest value of correlation coefficients (0.99) and the biosorption capacity being estimated to be 13.2, 12.2 and 6.2 mg g−1 for BR46, BB3 and AO7, respectively.  相似文献   

11.
Guozhen Wu  Peijie Wang 《Chirality》2015,27(11):820-825
A bond polarizability algorithm was developed and applied to interpret the Raman optical activity (ROA) intensity. It is demonstrated that for the chiral molecule such as S(+)2,2‐dimethyl‐1,3‐dioxolane‐4‐methanol there exists approximate (or symmetry breaking) mirror reflection that reverses the signs of the differential bond polarizabilities of the pair bond coordinates that are related to each other by the mirror reflection, just like that between the right and left enantiomers. The magnitude difference of the differential bond polarizabilities of the pair bond coordinates becomes smaller as they are farther away from the asymmetric atom. Hence, that the asymmetric atom (center) plays a central role in ROA is confirmed from a spectroscopic viewpoint. Meanwhile, the concept of intramolecular enantiomerism is proposed. Chirality 27:820–825, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

12.
A series of 1,3‐bis‐chalcone derivatives ( 3a‐i, 6a‐i and 8 ) were synthesized and evaluated antimicrobial, antibiofilm and carbonic anhydrase inhibition activities. In this evaluation, 6f was found to be the most active compound showing the same effect as the positive control against Bacillus subtilis and Streptococcus pyogenes in terms of antimicrobial activity. Biofilm structures formed by microorganisms were damaged by compounds at the minimum inhibitory concentration value between 0.5% and 97%.1,3‐bis‐chalcones ( 3a‐i, 6a‐i and 8 ) showed good inhibitory action against human (h) carbonic anhydrase (CA) isoforms I and II. hCA I and II were effectively inhibited by these compounds, with K i values in the range of 94.33 ± 13.26 to 787.38 ± 82.64 nM for hCA I, and of 100.37 ± 11.41 to 801.76 ± 91.11 nM for hCA II, respectively. In contrast, acetazolamide clinically used as CA inhibitor showed K i value of 1054.38 ± 207.33 nM against hCA I, and 983.78 ± 251.08 nM against hCA II, respectively.  相似文献   

13.
考察了H^ 和Na^ 变化对紫外吸收单波长法、双波长法和考马斯亮蓝法的影响。结果表明双波长法比考马斯亮蓝法和单波长法稳定性要好。选择双波长法用于磺酸型离子交换树脂吸附胰蛋白酶(trypsin)和牛血清白蛋白(BSA)的检测,结果表明:交换容量为4-88mmol/g的树脂对BSA和trypsin的吸附速率和饱和吸附量均高于交换容量为3.17mmol/g的树脂。  相似文献   

14.
Inducible plant defences against pathogens are stimulated by infections and comprise several classes of pathogenesis‐related (PR) proteins. Endo‐β‐1,3‐glucanases (EGases) belong to the PR‐2 class and their expression is induced by many pathogenic fungi and oomycetes, suggesting that EGases play a role in the hydrolysis of pathogen cell walls. However, reports of a direct effect of EGases on cell walls of plant pathogens are scarce. Here, we characterized three EGases from Vitis vinifera whose expression is induced during infection by Plasmopara viticola, the causal agent of downy mildew. Recombinant proteins were expressed in Escherichia coli. The enzymatic characteristics of these three enzymes were measured in vitro and in planta. A functional assay performed in vitro on germinated P. viticola spores revealed a strong anti‐P. viticola activity for EGase3, which strikingly was that with the lowest in vitro catalytic efficiency. To our knowledge, this work shows, for the first time, the direct effect against downy mildew of EGases of the PR‐2 family from Vitis.  相似文献   

15.
16.
A growing body of evidence supports that pyrimidine derivatives, in which the sugar residues have been replaced by acyclic side chains, might be developed as promising anticancer agents that interfere with tumor cell proliferation, survival, and metastatic formation. In this work, we prepared novel pyrimidines bearing i‐Bu (i.e., 3, 4 , and 7 – 9 ) and isobutenyl (i.e., 5 and 10 ) side chains at C(6) and examined their in vitro effects on tumor cell lines. The dihydropyrrolo[1,2‐c]pyrimidine‐1,3‐diones 6 and 11 were obtained as products of intramolecular cyclization, which occurred during the removal of Bn in 5 or MeO protecting groups in 10 . Fluorination of 3 with diethylaminosulfur trifluoride (DAST) and then dehydrohalogenation of the resulting fluorinated derivative 4 afforded 6‐isobut‐2′‐enyl pyrimidine derivative 5 with a C(2′)C(3′) bond. For the preparation of 6‐isobut‐1′‐en‐1‐yl pyrimidine 10 , a synthetic strategy involving acetylation of the 1,3‐diols was applied. Antitumor evaluation of compounds 3 – 11 showed that 2,4‐dimethoxypyrimidine containing 6‐[(1,3‐dibenzyloxy)‐2‐hydroxy]methyl side chain, 3 , exerted a strong antiproliferative effect on the studied tumor cell lines. Additionally, it was shown that the mechanism of antiproliferative effect of 3 in HeLa cells include early G2/M arrest and apoptosis, as well as a p53‐independent S‐phase arrest upon prolonged treatment.  相似文献   

17.
Klebsiella pneumoniae HR526, a new isolated 1,3‐propanediol (1,3‐PD) producer, exhibited great productivity. However, the accumulation of lactate in the late‐exponential phase remained an obstacle of 1,3‐PD industrial scale production. Hereby, mutants lacking D ‐lactate pathway were constructed by knocking out the ldhA gene encoding fermentative D ‐lactate dehydrogenase (LDH) of HR526. The mutant K. pneumoniae LDH526 with the lowest LDH activity was studied in aerobic fed‐batch fermentation. In experiments using pure glycerol as feedstock, the 1,3‐PD concentrations, conversion, and productivity increased from 95.39 g L?1, 0.48 and 1.98 g L?1 h?1 to 102. 06 g L?1, 0.52 mol mol?1 and 2.13 g L?1 h?1, respectively. The diol (1,3‐PD and 2,3‐butanediol) conversion increased from 0.55 mol mol?1 to a maximum of 0.65 mol mol?1. Lactate would not accumulate until 1,3‐PD exceeded 84 g L?1, and the final lactate concentration decreased dramatically from more than 40 g L?1 to <3 g L?1. Enzymic measurements showed LDH activity decreased by 89–98% during fed‐batch fermentation, and other related enzyme activities were not affected. NADH/NAD+ enhanced more than 50% in the late‐exponential phase as the D ‐lactate pathway was cut off, which might be the main reason for the change of final metabolites concentrations. The ability to utilize crude glycerol from biodiesel process and great genetic stability demonstrated that K. pnemoniae LDH526 was valuable for 1,3‐PD industrial production. Biotechnol. Bioeng. 2009; 104: 965–972. © 2009 Wiley Periodicals, Inc.  相似文献   

18.
A multi‐tiered approach to determine the binding mechanism of viral clearance utilizing a multi‐modal anion exchange resin was applied to a panel of four viral species that are typically used in validating viral clearance studies (i.e., X‐MuLV, MVM, REO3, and PrV). First, virus spiked buffer‐only experiments were conducted to evaluate the virus's affinity for single mode and multi‐modal chromatography resins under different buffer conditions in a chromatography column setting. From these results we hypothesize that the mechanisms of binding of the viruses involve binding to both the hydrophobic and anionic functional groups. This mechanistic view agreed with the general surface characteristics of the different virus species in terms of isoelectric point and relative hydrophobicity values. This hypothesized mechanistic binding was then tested with commercially relevant, in‐process materials, in which competitive binding occurred between the load components (e.g., viruses, target product, and impurities) and the resin. © 2018 American Institute of Chemical Engineers Biotechnol. Prog., 34:1019–1026, 2018  相似文献   

19.
Low-cost activated carbon was prepared from Spartina alterniflora by phosphoric acid activation for the removal of Pb(II) from dilute aqueous solution. The effect of experimental parameters such as pH, initial concentration, contact time and temperature on the adsorption was studied. The obtained data were fitted with the Langmuir and Freundlich equations to describe the equilibrium isotherms. The kinetic data were fitted with the Lagergren-first-order, pseudo-second-order and Elovich models. It was found that pH played a major role in the adsorption process. The maximum adsorption capacity for Pb(II) on S. alterniflora activated carbon (SAAC) calculated from Langmuir isotherm was more than 99 mg g−1. The optimum pH range for the removal of Pb(II) was 4.8–5.6. The Freundlich isotherm model was found to best describe the experimental data. The kinetic rates were best fitted to the pseudo-second-order model. Thermodynamic study showed the adsorption was a spontaneous exothermic process.  相似文献   

20.
The first geometric enantiomers in the cyclic compounds series are reported. The investigated compounds are 2,2‐disubstituted‐5‐methyl‐1,3‐dioxane derivatives in which the substituents at position 2 bear chiral centers with identical substituents but with opposite configurations. The structure of the unlike isomers was determined from the solid state molecular structure of the compounds obtained by single crystal X‐ray diffractometry and the enantiomers of these diastereoisomers were observed by chiral column HPLC base‐line separation. The investigated compounds were obtained by a diastereoselective bromination reaction of the corresponding 2,2‐dialkyl and 2,2‐dibenzyl‐5‐methyl‐1,3‐dioxanes. Chirality, 2011. © 2010 Wiley‐Liss, Inc.  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