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.
In this study a suitable alcohol/salt aqueous two‐phase (ATP) system was selected for the recovery of 1,3‐propandiol (1,3‐PD) from fermentation broth. From the different alcohol/salt systems studied the ethanol and dipotassium hydrogen phosphate ATP system appeared to be favorable. To examine the potential of this ATP system the partition coefficient of 1,3‐PD in synthetic solutions was first optimized with the response surface methodology. The parameters studied were concentrations of ethanol (21.99–38.81% w/w), dipotassium hydrogen phosphate (14.99–31.81% w/w) and 1,3‐PD (6.36–73.64 g/L). The optimum conditions were found to be 35.39% w/w for ethanol, 28.40% w/w for dipotassium hydrogen phosphate and 73.6 g/L for 1,3‐PD. Under these conditions the maximum partition coefficient of 1,3‐PD and the extraction yield were determined as 23.14 and 97.82%, respectively. The optimum extraction conditions were then used to guide the recovery of 1,3‐PD from a real fermentation broth. The partition coefficient and extraction yield of 1,3‐PD reached 20.28–97.20% in this case, respectively. A favorable partition of the organic acids lactate, acetate and butyrate in the bottom phase was also achieved. We have also studied the removal of cells and macromolecules from the broth. Removal ratio of cells and proteins were 96.47 and 93.05%, respectively. Thus, the ethanol/dipotassium hydrogen phosphate ATP system appears to be an interesting alternative or can be used as one useful step in the downstream processing of 1,3‐PD from fermentation broth.  相似文献   

2.
Flurbiprofen is a kind of nonsteroidal anti‐inflammatory drug, which has been widely used in clinic for treatment of rheumatoid arthritis and osteoarthritis. It has been reported that S‐flurbiprofen shows good performance on clinic anti‐inflammatory treatment, while R‐enantiomer almost has no pharmacological activities. It has important practical values to obtain optically pure S‐flurbiprofen. In this work, chiral ionic liquids, which have good structural designability and chiral recognize ability, were selected as the extraction selector by the assistance of quantum chemistry calculations. The distribution behaviors of flurbiprofen enantiomers were investigated in the extraction system, which was composed of organic solvent and aqueous phase containing chiral ionic liquid. The results show that maximum enantioselectivity up to 1.20 was attained at pH 2.0, 25°C using 1,2‐dichloroethane as organic solvent, 1‐butyl‐3‐methylimidazole L‐tryptophan ([Bmim][L‐trp]) as chiral selector. The racemic flurbiprofen initial concentration was 0.2 mmol L?1, and [Bmim][L‐trp] concentration was 0.02 mol L?1. Furthermore, the recycle of chiral ionic liquids has been achieved by reverse extraction process of the aqueous phase with chiral selector, which is significant for industrial application of chiral ionic liquids and scale‐up of the extraction process.  相似文献   

3.
《Chirality》2017,29(11):708-715
A liquid–liquid extraction resolution of 4‐chloro‐mandelic acid (4‐ClMA) was studied by using 2‐chloro‐N‐carbobenzyloxy‐L‐amino acid (2‐Cl‐Z‐AA) as a chiral extractant. Important factors affecting the extraction efficiency were investigated, including the type of chiral extractant, pH value of aqueous phase, initial concentration of chiral extractant in organic phase, initial concentration of 4‐ClMA in aqueous phase, and resolution temperature. It was observed that the concentration of (R)‐4‐ClMA was much higher than that of (S)‐4‐ClMA in organic phase due to a higher stability of the complex formed between (R)‐4‐ClMA and 2‐Cl‐Z‐AA. A separation factor (α) of 3.05 was obtained at 0.02 mol/L 2‐Cl‐Z‐Valine dissolved in dichloromethane, pH of 2.0, concentration of 4‐ClMA of 0.11 mmol/Land T of 296.7K.  相似文献   

4.
This article presents results of continuous multistage aqueous two‐phase extraction of an immunoglobulin G1 from cell supernatant in a mixer‐settler unit. An aqueous two‐phase system consisting of polyethylene glycol 2000, phosphate salt, and water was applied without and with sodium chloride (NaCl). Influences of different parameters such as throughput, phase ratio, and stage number on the extraction performance were analyzed. For systems without NaCl, the extraction was carried out as a washing step. An increase of stage number from one to five stages enabled to increase the immunoglobulin G1 purity from 11.8 to 32.6% at a yield of nearly 90%. Furthermore, a reduction of product phase volume due to a higher phase ratio led to an increase of purity from 20.8 to 29.6% in a three‐stage countercurrent extraction. For experiments with NaCl moderate partitioning conditions were adjusted by adding 8 wt% NaCl. In that case, the extraction was carried out as a stripping step. © 2015 American Institute of Chemical Engineers Biotechnol. Prog., 31:925–936, 2015  相似文献   

5.
Preparative enantioseparation of four β‐substituted‐2‐phenylpropionic acids was performed by countercurrent chromatography with substituted β‐cyclodextrin as chiral selectors. The two‐phase solvent system was composed of n‐hexane‐ethyl acetate‐0.10 mol L‐1 of phosphate buffer solution at pH 2.67 containing 0.10 mol L‐1 of hydroxypropyl‐β‐cyclodextrin (HP‐β‐CD) or sulfobutylether‐β‐cyclodextrin (SBE‐β‐CD). The influence factors, including the type of substituted β‐cyclodextrin, composition of organic phase, concentration of chiral selector, pH value of the aqueous phase, and equilibrium temperature were optimized by enantioselective liquid–liquid extraction. Under the optimum separation conditions, 100 mg of 2‐phenylbutyric acid, 100 mg of tropic acid, and 50 mg of 2,3‐diphenylpropionic acid were successfully enantioseparated by high‐speed countercurrent chromatography, and the recovery of the (±)‐enantiomers was in the range of 90–91% for (±)‐2‐phenylbutyric acid, 91–92% for (±)‐tropic acid, 85–87% for (±)‐2,3‐diphenylpropionic acid with purity of over 97%, 96%, and 98%, respectively. The formation of 1:1 stoichiometric inclusion complex of β‐substituted‐2‐phenylpropionic acids with HP‐β‐CD was determined by UV spectrophotometry and the inclusion constants were calculated by a modified Benesi‐Hildebrand equation. The results showed that different enantioselectivities among different racemates were mainly caused by different enantiorecognition between each enantiomer and HP‐β‐CD, while it might be partially caused by different inclusion capacity between racemic solutes and HP‐β‐CD. Chirality 27:795–801, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

6.
Proteins (bovine serum albumin (BSA), α‐chymotrypsin, cytochrome c, and lysozyme) were extracted from 0.5 to 2.0 g L?1 aqueous solution by adding an equal volume of isooctane solution that contained a surfactant mixture (Aerosol‐OT, or AOT, and a 1,3‐dioxolane (or cyclic ketal) alkyl ethoxylate, CK‐2,13‐E5.6), producing a three‐phase (Winsor‐III) microemulsion with a middle, bicontinuous microemulsion, phase highly concentrated in protein (5–13 g L?1) and small in volume (12–20% of entire volume). Greater than 90% forward extraction was achieved within a few minutes. Robust W‐III microemulsion systems were formulated at 40°C, or at 25°C by including a surfactant with shorter ethoxylate length, CK‐2,13‐E3, or 1.5% NaCl (aq). Successful forward extraction correlated with high partitioning of AOT in the middle phase (>95%). The driving force for forward extraction was mainly electrostatic attractions imposed by the anionic surfactant AOT, with the exception of BSA at high ionic strength, which interacted via hydrophobic interactions. Through use of aqueous stripping solutions of high ionic strength (5.0 wt %) and/or pH 12.0 (to negate the electrostatic attractive driving force), cytochrome c and α‐chymotrypsin were back extracted from the middle phase at >75% by mass, with the specific activity of recovered α‐chymotrypsin being >90% of its original value. © 2011 American Institute of Chemical Engineers Biotechnol. Prog., 2011  相似文献   

7.
In the present study, hydroxypropyl‐β‐cyclodextrin and an ionic liquid (1‐ethyl‐3‐methylimidazolium‐l ‐lactate) were used as additives in capillary electrophoresis for the enantioseparation of 10 analytes, including ofloxacin, propranolol hydrochloride, dioxopromethazine hydrochloride, isoprenaline hydrochloride, chlorpheniramine maleate, liarozole, tropicamide, amlodipine benzenesulfonate, brompheniramine maleate, and homatropine methylbromide. The effects of ionic liquid concentrations, salt effect, cations, and anions of ionic liquids on enantioseparation were investigated and the results proved that there was a synergistic effect between hydroxypropyl‐β‐cyclodextrin and the ionic liquid, and the cationic part of the ionic liquid played an important role in the increased resolution. With the developed dual system, all the enantiomers of 10 analytes were well separated in resolutions of 5.35, 1.76, 1.85, 2.48, 2.88, 1.43, 5.45, 4.35, 2.76, and 2.98, respectively. In addition, the proposed method was applied to the determination of the enantiomeric purity of S‐ofloxacin after validation of the method in terms of selectivity, repeatability, linearity range, accuracy, precision, limit of detection (LOD), and limit of quality (LOQ). Chirality 25:409–414, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

8.
Superoxide dismutase (SOD; EC 1.15.1.1) is an antioxidant enzyme that represents the primary cellular defense against superoxide radicals and has interesting applications in the medical and cosmetic industries. In the present work, the partition behavior of SOD in aqueous two‐phase systems (ATPS) (using a standard solution and a complex extract from Kluyveromyces marxianus as sample) was characterized on different types of ATPS (polymer–polymer, polymer–salt, alcohol–salt, and ionic liquid (IL)–salt). The systems composed of PEG 3350‐potassium phosphate, 45% TLL, 0.5 M NaCl (315 U/mg, 87% recovery, and 15.1‐fold purification) and t‐butanol‐20% ammonium sulfate (205.8 U/mg, 80% recovery and 9.8‐fold purification), coupled with a subsequent 100 kDa ultrafiltration stage, allowed the design of a prototype process for the recovery and partial purification of the product of interest. The findings reported herein demonstrate the potential of PEG‐salt ATPS for the potential recovery of SOD. © 2014 American Institute of Chemical Engineers Biotechnol. Prog., 30:1326–1334, 2014  相似文献   

9.
The possibility of creating a biorefinery using inexpensive biomass has attracted a great deal of attention, which is mainly focused on the improvement of strains and fermentation, whereas few resources have been spent on downstream processing. Bio‐based chemical downstream processing can become a bottleneck in industrial production because so many impurities are introduced into the fermentation broth. This review introduces a technique referred to as salting‐out extraction, which is based on the partition difference between chemicals in two phases consisting of salts and polymers or hydrophilic solvents, hydrophobic solvents, and amphipathic chemicals. The effects of solvents and salts on the formation of two phases were discussed, as was the use of this method to recover bio‐based chemicals. This review focused on the separation of hydrophilic chemicals (1,3‐propanediol, 2,3‐butanediol, acetoin, and lactic acid) from fermentation broths. Diols could be recovered at a high yield from fermentation broths without pretreatment especially with a hydrophilic solvent‐based system, whereas the recovery of organic acids was slightly lower. Most of the impurities (cells and proteins) were removed during the same step. Extractive fermentations were also used for polymer‐based aqueous two‐phase systems.  相似文献   

10.
To improve the effects of protease finishing on wool, 1‐butyl‐3‐methylimidazolium chloride ionic liquid was employed as a pretreatment reagent. It was found that ionic liquid pretreatment significantly changed the wool surface characteristics. The Allwördern reaction showed that the epicuticle layer was damaged by the ionic liquid, and X‐ray photoelectron spectroscopy analysis further demonstrated that the surface elemental composition was significantly changed. Ionic liquid pretreatment remarkably improved the accessibility of protease to the wool and thus accelerated the hydrolysis rate of keratin. The properties of wool fabric after combined processing were also changed. Dyeability results showed that the color depth was increased but the wet rubbing and washing fastness of wool fabrics showed a decreased half grade. The wettability results demonstrate that the contact angle was further reduced after the comprehensive treatment because of the exposure of more proteins under the fatty‐acid layer. In addition, the shrink proofing of wool fabric was also enhanced after combined processing. In summary, ionic liquid modification presents a promising pretreatment method for protease processing of wool.  相似文献   

11.
Xiong Liu  Yu Ma  Longqi Xu  Qi Liu 《Chirality》2019,31(9):750-758
(S,S)‐DIOP, a common catalyst used in asymmetric reaction, was adopted as chiral extractant to separate 3‐chloro‐phenylglycine enantiomers in liquid‐liquid extraction. The factors affecting extraction efficiency were studied, including metal precursors, organic solvents, extraction temperature, chiral extractant concentration, and pH of aqueous phase. (S,S)‐DIOP‐Pd exhibited good ability to recognize 3‐chloro‐phenylglycine enantiomers, and the operational enantioselectivity (α) is 1.836. The highest performance factor (pf) was obtained under the condition of extraction temperature of 9.1°C, (S,S)‐DIOP‐Pd concentration of 1.7 mmol/L, and pH of aqueous phase of 7.0. In addition, the possible recognition mechanism of (S,S)‐DIOP‐Pd towards 3‐chloro‐phenylglycine enantiomers was discussed.  相似文献   

12.
In biotechnology, extraction by means of aqueous biphasic systems (ABS) is known as a promising tool for the recovery and purification of bio‐molecules. Over the past decade, the increasing emphasis on cleaner and environmentally benign extraction procedures has led to enhanced interest in the ABS containing ionic liquids (ILs)—a new class of non‐volatile alternative solvents. ABS composed of the hydrophilic IL {1‐butyl‐3‐methylimidazolium bromide ([C4mim]Br)} and potassium citrate—which is easily degraded—represents a clean media to green separation of bio‐molecules. In this regard, here, the extraction capability of this ABS was evaluated through its application to the extraction of some amino acids. To gain an insight into the driving forces of amino acid partitioning in the studied IL ‐based ABS, the distribution of five model amino acids (L ‐tryptophan, L ‐phenylalanine, L ‐tyrosine, L ‐leucine, and L ‐valine) at different aqueous medium pH values and different phase compositions was investigated. The studies indicated that hydrophobic interactions were the main driving force, although electrostatic interactions and salting‐out effects were also important for the transfer of the amino acids. Moreover, based on the statistical analysis of the driving forces of amino acid partitioning in the studied IL ‐based ABS, a model was established to describe the partition coefficient of three model amino acids, L ‐tryptophan, L ‐phenylalanine, and L ‐valine, and employed to predict the partition coefficient of two other model amino acids, L ‐tyrosine and L ‐leucine. © 2011 American Institute of Chemical Engineers Biotechnol. Prog., 2011  相似文献   

13.
Aims: Saccharomyces cerevisiae is a safe micro‐organism used in fermentation industry. 1,3‐Propanediol is an important chemical widely used in polymer production, but its availability is being restricted owing to its expensively chemical synthesis. The aim of this study is to engineer a S. cerevisiae strain that can produce 1,3‐propanediol at low cost. Methods and Results: By using d ‐glucose as a feedstock, S. cerevisiae could produce glycerol, but not 1,3‐propanediol. In this study, we have cloned two genes yqhD and dhaB required for the production of 1,3‐propanediol from glycerol, and integrated them into the chromosome of S. cerevisiae W303‐1A by Agrobacterium tumefaciens‐mediated transformation. Both genes yqhD and dhaB functioned in the engineered S. cerevisiae and led to the production of 1,3‐propanediol from d ‐glucose. Conclusion: Saccharomyces cerevisiae can be engineered to produce 1,3‐propanediol from low‐cost feedstock d ‐glucose. Significance and Impact of the Study: To our knowledge, this is the first report on developing S. cerevisiae to produce 1,3‐propanediol by using A. tumefaciens‐mediated transformation. This study might lead to a safe and cost‐efficient method for industrial production of 1,3‐propanediol.  相似文献   

14.
1,3‐propanediol is an important chemical widely used in polymer production. In this study, two strains, Zygosacharomyces rouxii JL2011 and Klebsiella pneumoniae S6, were used as a mixed culture for 1,3‐propanediol production directly from glucose. Two important parameters including inoculation time of K. pneumoniae S6 at stage of mixed culture and initial cell ratio of Z. rouxii JL2011 to K. pneumoniae S6 in mixed fermentation were optimized in culture flasks. In those experiments, the best results were obtained with a yield of 6.8 g/L 1,3‐propanediol from glucose when K. pneumoniae S6 was inoculated after 48 h in the culture of Z. rouxii JL2011 by mixed culture of Z. rouxii JL2011 and K. pneumoniae S6 with initial cell ratio of 1:200. In a 7‐L bioreactor, the maximum 1,3‐propanediol production could reach up to 15.2 g/L. Thus, this study presents an effective process for 1,3‐propanediol microbial production from glucose by using mixed culture of Z. rouxii JL2011 and K. pneumoniae S6. This work does not only demonstrate a new way to produce 1,3‐propanediol from a low‐cost feedstock, but may also make a valuable contribution to the development of a cost‐effective fermentation based on renewable resources.  相似文献   

15.
Ionic liquids have been used as exceptional nonaqueous reaction media for enzymatic transformation. The ring‐opening polymerization of ε‐caprolactone catalyzed by Novozyme‐435 lipase was successfully conducted in 1‐butyl‐3‐methylimidazolium hexafluorophosphate ([Bmim]PF6) ionic liquid. 1H‐NMR and MALDI‐TOF analyses of poly(ε‐caprolactone) (PCL) formed by Novozyme‐435 lipase‐catalyzed reaction revealed an asymmetric telechelic α‐hydroxy‐ω‐carboxylic acid end group. The effects of enzyme concentration, temperature, reaction time, and water activities on monomer conversion and Mn were systematically evaluated. Through the optimization of reaction conditions, PCL was produced in 85% monomer conversion, with an Mn of 5942, in [Bmim]PF6 at 60°C for 48 h. DSC results demonstrated that high‐molecular‐weight PCL exhibited an excellent thermal property. SEM results showed that PCL had a clear spherulites structure, which could provide a large surface area for cell adhesion. These results showed that [Bmim]PF6 ionic liquid was suitable for the biocatalytic synthesis of PCL using Novozyme‐435 lipase, and could be used as alternative environmentally friendly media to replace the traditional organic solvents.  相似文献   

16.
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   

17.
Aiming at understanding the molecular mechanism of the lignin dissolution in imidazolium‐based ionic liquids (ILs), this work presents a combined quantum chemistry (QC) calculation and molecular dynamics (MD) simulation study on the interaction of the lignin model compound, veratrylglycerol‐β‐guaiacyl ether (VG) with 1‐allyl‐3‐methylimidazolium chloride ([Amim]Cl). The monomer of VG is shown to feature a strong intramolecular hydrogen bond, and its dimer is indicated to present important π‐π stacking and intermolecular hydrogen bonding interactions. The interactions of both the cation and anion of [Amim]Cl with VG are shown to be stronger than that between the two monomers, indicating that [Amim]Cl is capable of dissolving lignin. While Cl anion forms a hydrogen‐bonded complex with VG, the imidazolium cation interacts with VG via both the π‐π stacking and intermolecular hydrogen bonding. The calculated interaction energies between VG and the IL or its components (the cation, anion, and ion pair) indicate the anion plays a more important role than the cation for the dissolution of lignin in the IL. Theoretical results provide help for understanding the molecular mechanism of lignin dissolution in imidazolium‐based IL. The theoretical calculations on the interaction between the lignin model compound and [Amim]Cl ionic liquid indicate that the anion of [Amim]Cl plays a more important role for lignin dissolution although the cation also makes a substantial contribution.  相似文献   

18.
The rose‐like aroma compound 2‐phenylethanol (2‐PE) is an important fragrance and flavor ingredient. Several yeast strains are able to convert l ‐phenylalanine (l ‐phe) to 2‐PE among which Kluyveromyces marxianus has shown promising results. The limitation of this process is the low product concentration and productivity primarily due to end product inhibition. This study explored the possibility and benefits of using a solid–liquid Two‐Phase Partition Bioreactor (TPPB) system as an in situ product removal technique. The system applies polymer beads as the sequestering immiscible phase to partition 2‐PE and reduce the aqueous 2‐PE concentration to non‐inhibitory levels. Among six polymers screened for extracting 2‐PE, Hytrel® 8206 performed best with a partition coefficient of 79. The desired product stored in the polymer was ultimately extracted using methanol. A 3 L working volume solid–liquid batch mode TPPB using 500 g Hytrel® as the sequestering phase generated a final overall 2‐PE concentration of 13.7 g/L, the highest reported in the current literature. This was based on a polymer phase concentration of 88.74 g/L and aqueous phase concentration of 1.2 g/L. Even better results were achieved via contact with more polymers (approximately 900 g) with the aqueous phase applying a semi‐continuous reactor configuration. In this system, a final 2‐PE concentration (overall) of 20.4 g/L was achieved with 1.4 g/L in the aqueous and 97 g/L in the polymer phase. The overall productivities of these two reactor systems were 0.38 and 0.43 g/L h, respectively. This is the first report in the literature of the use of a polymer sequestering phase to enhance the bioproduction of 2‐PE, and exceeds the performance of two‐liquid phase systems in terms of productivity as well as ease of operation (no emulsions) and ultimate product recovery. Biotechnol. Bioeng. 2009; 104: 332–339 © 2009 Wiley Periodicals, Inc.  相似文献   

19.
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   

20.
In this study, a novel pH‐sensitive terpolymer PADB was synthesized by random terpolymerization of 2‐(dimethylamino) ethyl methacrylate, acrylic acid, and butyl methacrylate. The terpolymer PADB could form aqueous two‐phase systems (ATPS) with a light‐sensitive terpolymer PNBC, which was synthesized in our laboratory, using n‐isopropylacrylamide, n‐butyl acrylate, chlorophyllin sodium copper salt as monomers. More than 97% of the PADB terpolymer could be recovered by adjusting the pH to isoelectric point (PI) 4.1. The terpolymer PNBC could be recovered by using light radiation at 488 nm, with recovery ratio of 98%. BSA and lysozyme were partitioned in the PNBC–PADB ATPS to examine this new system. It was found that the partition coefficient of BSA and lysozyme could reach 4.46 and 0.49 in the systems, respectively. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2009  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