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 627 毫秒
1.
Removal of biobutanol from acetone-butanolethanol (ABE) fermentation broth can be achieved by fixed-bed sorption by means of KA-I resin, and the relevant breakthrough curves would provide much valuable information to help design a continuous fixed-bed sorption process in field application. In the present study, the effects of several important design parameters, i.e., initial butanol concentration (C f: 3.0 ~ 30.0 g/L), inlet flow rate (Q f: 0.5 ~ 5.5 mL/min) and adsorbent bed height (Z: 4.2 ~ 18.0 cm), on the adsorption breakthrough curves of KA-I resin in a fixed-bed column were investigated. It was found that the amount of adsorbed butanol at breakthrough point was increased with an increase in the value of C f and Z; and with decrease in the value of Q f. However, the maximum sorption capacities of butanol at saturated point were basically unchanged. Three well-established fixed-bed adsorption models, namely Thomas, Yoon-Nelson and Adams-Bohart, were applied to predict the breakthrough curves and to determine the characteristic parameters of fixed-bed column, which are the basis for the process design at a real scale. Good agreement between the theoretical breakthrough curves and the experimental result were observed using Thomas and Yoon-Nelson models.  相似文献   

2.
The expanded bed characteristics of 75-103microm fluoride-modified zirconia (FmZr) particles synthesized by a fed batch oil emulsion process were investigated. These particles are distinguished from commercially available expanded-bed adsorbents by virtue of their high density (2.8 g/cc) and the mixed mode protein retention mechanism which allows for the retention of both cationic and anionic proteins. The linear velocity versus bed porosity data agree with the Richardson-Zaki relationship with the terminal velocity in infinite medium of 2858.4 cm/h and a bed expansion index of 5.1. Residence time distribution (RTD) studies and bovine serum albumin (BSA) adsorption studies were performed as a function of the height of the settled bed to the column diameter (H:D) ratio and degree of bed expansion with superficial velocities of 440 to 870 cm/h. The settled bed, a 2x expanded bed, and a 3x expanded bed were studied for the H:D ratios of 1:1, 2:1, and 3:1. The dynamic binding capacity (DBC) at 5% breakthrough was low (2-8 mg BSA/mL settled bed) and was independent of the H:D ratio or the degree of bed expansion. The saturation DBC was 32.3 +/- 7.0 mg BSA/mL settled bed. The adsorption-desorption kinetics and intraparticle diffusion for protein adsorption on FmZr (38-75 micrometer) were investigated by studying the packed bed RTD and BSA adsorption as a function of temperature and flow rate. The data show that the adsorption-desorption kinetics along with intraparticle diffusion significantly influence protein adsorption on FmZr. Low residence times ( approximately 0.8 min) of BSA result in a DBC at 5% breakthrough which is 3.5-fold lower compared to that at 6-fold higher protein residence time. At low linear velocity (45 cm/h) the breakthrough curve is nearly symmetrical and becomes asymmetrical and more dispersed at higher linear velocity (270 cm/h) due to the influence of slow adsorption-desorption kinetics and intraparticle diffusion.Copyright 1998 John Wiley & Sons, Inc. Bioeng 60: 333-340, 1998.  相似文献   

3.
This study reports the results of experiments on continuous adsorption and desorption of Cr(VI) ions by a chemically modified and polysulfone-immobilized biomass of the fungus Rhizopus nigricans. A fixed quantity of polymer-entrapped biomass beads corresponding to 2 g of dry biomass powder was employed in packed bed, fluidized bed, and stirred tank reactor for monitoring the continuous removal and recovery of Cr(VI) ions from aqueous solution and synthetic chrome plating effluent. Parameters such as flow rate (5, 10 and 15 mL/min), inlet concentration of Cr(VI) ions (50, 100, 150 and 250 mg/L) and the depth of biosorbent packing (22.8, 11.2 and 4.9 cm) were evaluated for the packed bed reactor. The breakthrough time and the adsorption rates in the packed bed column were found to decrease with increasing flow rate and higher Cr inlet concentrations and to increase with higher depths of sorbent packing. To have a comparative analysis of Cr adsorption efficiency in different types of reactors, the fluidized bed reactor and stirred tank reactor were operated using the same quantities of biosorbent material. For the fluidized bed reactor, Cr(VI) solution of 100 mg/L was pumped at 5 mL/min and fluidized by compressed air at a flow rate of 0.5 kg/cm.(2) The stirred tank reactor had a working volume of 200 mL capacity and the inlet/outlet flow rate was 5 mL/min. The maximum removal efficiency (mg Cr/g biomass) was obtained for the stirred tank reactor (159.26), followed by the fluidized reactor (153.04) and packed bed reactor (123.33). In comparison to the adsorption rate from pure chromate solution, approximately 16% reduction was monitored for synthetic chrome plating effluent in the packed bed. Continuous desorption of bound Cr ions from the reactors was effective with 0.01 N Na(2)CO(3) and nearly 80-94% recoveries have been obtained for all the reactors.  相似文献   

4.
The sorption of heavy metals ions by immobilized Trichoderma viride biomass in a packed-bed column was studied. Fungal biomass T. viride was immobilized to Ca-alginate used for removal of Cr(VI), Ni(II) and Zn(II) ions from synthetic solutions and electroplating effluent. The experiments were conducted to study the effect of important design parameters such as bed height, flow rate and initial concentration of metal ions. The maximum sorption capacity was observed at flow rate 5 ml/min, bed height 20 cm and metal ions concentration 50 mg/L with immobilized biomass. Whereas, breakthrough time and saturation time decreased with increase flow rate and metal ions concentration and an inverse condition was found in bed height. The bed depth service time (BDST) Adams-Bohart model was used to analyze the experimental data. The regeneration efficiency was observed 40.1%, 75% and 53% for Cr(VI), Ni(II) and Zn(II) without any significant alteration in sorption capacity after 5th sorption-desorption cycles.  相似文献   

5.
Removal of Pb(II) from an aqueous environment using biosorbents is a cost-effective and environmentally benign method. The biosorption process, however, is little understood for biosorbents prepared from plant materials. In this study, the biosorption process was investigated by evaluating four adsorption models. A fixed-bed column was prepared using a biosorbent prepared from the aquatic plant Hydrilla verticillata. The effect of bed height and flow rate on the biosorption process was investigated. The objective of the study was to determine the ability of H. verticillata to biosorb Pb(II) from an aqueous environment and to understand the process, through modeling, to provide a basis to develop a practical biosorbent column. Experimental breakthrough curves for biosorption of 50 mg L?1 aqueous Pb(II) using a fixed-bed column with 1.00 cm inner diameter were fitted to the Thomas, Adams-Bohart, Belter, and bed depth service time (BDST) models to investigate the behavior of each model according to the adsorption system and thus understand the adsorption mechanism. Model parameters were evaluated using linear and nonlinear regression methods. The biosorbent removed 65% (82.39 mg g?1 of biosorbent) of Pb(II) from an aqueous solution of Pb(NO3)2 at a flow rate of 5.0 ml min?1 in a 10 cm column. Na2CO3 was used to recover the adsorbed Pb(II) ions as PbCO3 from the biosorbent. The Pb(II) was completely desorbed at a bed height of 10.0 cm and a flow rate of 5.0 ml min?1. Fourier transform infrared (FT-IR) analysis of the native biosorbent and Pb(II)-loaded biosorbent indicated that the hydroxyl groups and carboxylic acid groups were involved in the metal bonding process. The FT-IR spectrum of Pb(II)-desorbed biosorbent showed an intermediate peak shift, indicating that Pb(II) ions were replaced by Na+ ions through an ion-exchange process. Of the four models tested, the Thomas and BDST models showed good agreement with experimental data. The calculated bed sorption capacity N0 and rate constant ka were 31.7 g L?1 and 13.6 × 10?4 L mg?1 min?1 for the Ct/C0 value of 0.02. The BDST model can be used to estimate the column parameters to design a large-scale column.  相似文献   

6.
Dead cells of Saccharomyces cerevisiae 54 were immobilized by entrappment in polyacrylonitrile. The beads obtained were used to adsorb copper in an up-flow fixed-bed column. The effect of polymer content and cell loading were studied to optimize the porosity and the efficiency in copper removal of the biosorbent beads in a batch system. The optimal concentration of the polyacrylonitrile was assumed to be 12%(w/v) and a concentration of 0.5 g cell dry weight in 1 g polymer was most effective in adsorption of Cu2+. The adsorption capacity of this biosorbent was 27 mg Cu2+/g dry biomass at 200 mg/l initial concentration of copper ions. Adsorption of Cu2+ in a batch system was studied using different initial concentrations of the solute. The optimal conditions in the up-flow column of the following parameters were determined: flow rate, bed height, and initial concentration of Cu2+ of the solutions. Results of fixed-bed biosorption showed that breakthrough and saturation time appeared to increase with the bed height, but decrease with the flow rate and the initial concentration. The linearized form of the Thomas equation was used to describe dynamic adsorption of metal ions. As a result, the adsorption capacity of the batch system and the column system was compared. Desorption of copper ions was achieved by washing the column biomass with 0.1 M HCl at an eluent flow rate of 1 ml/min. The reusability of the immobilized biomass was tested in five consecutive adsorption-desorption cycles. The regenerated beads retained over 45% of their original adsorption capacity after five A/D cycles. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

7.
Arsenic is a toxic element and may be found in natural as well as in industrial water; therefore, before using water for drinking purpose, its proper treatment is required. Thus, the aim of this work was to evaluate the potential of chitosan nanoparticles, in a continuous-flow method, for the removal of arsenic (III) and (V) from aqueous solutions. All experiments were conducted in fixed-bed columns. Experiments were carried out as a function of varying liquid flow rate (0.3–1.0 ml/min), initial metal concentration (0.5–1.5 mg/L), and bed height (3–9 cm) of adsorbent. The total adsorbed quantity, equilibrium uptake, and total percentage removal of arsenic ions were determined by evaluating the breakthrough curves obtained at different flow rates, initial concentrations, and bed heights. The results showed that the column performed well at the lowest flow rate. Also, column bed capacity and exhaustion time were found to increase with increasing bed height. When initial metal ion concentration was increased from 0.5 to 1.5 mg/L, the corresponding adsorption bed capacity decreased from 0.076 to 0.028 mg/g. The bed depth service time model (BDST) model was used to analyze the experimental data and the model parameters were evaluated. The calculated values of N o and K a were found to be 19.28 × 10?2 mg/L and 0.662 L/mg·min, respectively. Good agreement was found between the experimental breakthrough curves and the model predictions.  相似文献   

8.
Marine algae Gelidium and algal composite material were investigated for the continuous removal of Cu(II) from aqueous solution in a packed bed column. The biosorption behaviour was studied during one sorption–desorption cycle of Cu(II) in the flow through column fed with 50 and 25 mg l−1 of Cu(II) in aqueous solution, at pH 5.3, leading to a maximum uptake capacity of ≈13 and 3 mg g−1, respectively, for algae Gelidium and composite material. The breakthrough time decreases as the inlet copper concentration increases, for the same flow rate. The pH of the effluent decreases over the breakthrough time of copper ions, which indicates that ion exchange is one of the mechanisms involved in the biosorption process. Temperature has little influence on the metal uptake capacity and the increase of the ionic strength reduces the sorption capacity, decreasing the breakthrough time. Desorption using 0.1 M HNO3 solution was 100% effective. After two consecutive sorption–desorption cycles no changes in the uptake capacity of the composite material were observed. A mass transfer model including film and intraparticle resistances, and the equilibrium relationship, for adsorption and desorption, was successfully applied for the simulation of the biosorption column performance.  相似文献   

9.
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   

10.
The sorption of Cd(II) and Pb(II) ions was conducted in a continuous fixed-bed column by using dead calcareous skeletons (CS). The column performances were evaluated by varying the adsorbent bed height, influent flow rate and metals initial concentration. The breakthrough curve for the bed height indicated that a longer bed column prolonged the life span of the column with a maximum capacity of 26.447 and 38.460 mg/g for the Cd(II) and Pb(II) column, respectively. The increased flow rate and initial concentration caused the column exhaustion time to occur earlier. The experimental column data were also expressed in column adsorption models, namely, the Thomas, Yoon–Nelson and Adam–Bohart models. The Thomas model fitted well with the Cd(II) data with the correlated curve (r2 > 0.9). The Yoon–Nelson model was selected to predict the 50% breakthrough time achieved by the column system and provided the estimated breakthrough time for the columns that were not exhausted during the operation. The Adam–Bohart model was applicable for the initial part of adsorption with the saturation concentration data at the equilibrium. The saturation index of aragonite and calcite depicted that dissolution of calcium occurred in the aqueous solution. The experimental and theoretical data were correlated with a significant relationship trend (p < 0.01), which showed that the trend of experimental data fit well with the modeling trend. The trends of both the experimental and theoretical data were strongly and significantly correlated due to involving the column parameters and the components of CS.  相似文献   

11.
12.
The influence of matrix properties and operating conditions on the performance in fluidized-bed adsorption has been studied using Streamline diethyl-aminoethyl (DEAE), an ion exchange matrix based on quartz-weighted agarose, and bovine serum albumin (BSA) as a model protein. Three different particle size fractions (120-160 mum, 120-300 mum, and 250-300 mum) were investigated. Dispersion in the liquid phase was reduced when particles with a wide size distribution were fluidized compared to narrow particle size distributions. When the mean particle diameter was reduced, the breakthrough capacities during frontal adsorption were enlarged due to a shorter diffusion path length within the matrix. At small particle diameters the effect of film mass transfer became more relevant to the adsorption performance in comparison to larger particles. Therefore matrices designed for fluidized-bed adsorption should have small particle diameter and increased mean particle density to ensure small diffusion path length in the particle and a high interstitial velocity to improve film mass transfer. Studies on the influence of sedimented matrix height on axial mixing showed an increased Bodenstein number with increasing bed length. Higher breakthrough capacities were also found for longer adsorbent beds due to reduced dispersion and improved fluid and particle side mass transfer. With increasing bed height the influence of flow rate on breakthrough capacity was reduced. For a settled bed height of 50 cm breakthrough capacities of 80% of the equilibrium capacity for flow rates varying from 3 to 9 cm/min could be achieved. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 55: 54-64, 1997.  相似文献   

13.
The application of fixed bed adsorption is an important separation technique used for heavy metals in environmental pollution control. To design a fixed bed column, it is necessary to find dynamics data in the breakthrough curve form. The objective of this study was to model the biosorption process of nickel by using biomass of Sargassum filipendula in a fixed bed column. Experimental data were generated at 30°C, pH 3, flow rate of 6 ml/min and feed concentrations of 1, 1.5, 2, 3, 5 and 6 mequiv/l. Langmuir isotherm was used to represent the equilibrium data in the column (q max = 2.496 mequiv/g, b = 0.456 l/mequiv) and in a batch operation of the system (q max = 1.577 mequiv/g, b = 0.269 l/mequiv). These Langmuir parameters were used to simulate the continuous adsorption process of nickel. The partial differential equations model has taken into consideration the mass transfer resistance in the biosorbent as the key controlling phenomenon, which adequately represented the dynamic biosorption process of nickel.  相似文献   

14.
Various adsorbents were screened for in situ recovery of betalain pigments effluxed from hairy root cultures of red beet, Beta vulgaris. Alumina/silica (1:1) appeared ideal, showing in situ adsorption of 97% in a unit time of 30 min accounting for in situ recovery of 71.39% of the total betalaine effluxed. Other adsorbents such as Amberlite series (XAD-2 and -4), cyclodextrin, maltodextrin, dextrin white, and starches such as wheat starch and corn starch exhibited very poor in situ adsorption properties. Pretreatment of adsorbents with methanol significantly improved the adsorption capacities of some of the adsorbents, with a highest adsorption of 97.2% for alumina followed by alumina/silica (1:1) and higher adsorption by XAD-2 and -4. Complete in situ adsorption equilibrium was reached in 20 min for a solution containing 2.5 mg mL(-)(1) of betalain in adsorbents alumina, silica, and a mixture of alumina and silica. In situ betalain adsorption parameters for alumina/silica were determined using the Langmuir isotherm model where the adsorption capacity was found to be 0.174 mg g(-)(1) and the adsorption energy was 0.9 at pH 5.5 and 25 degrees C. Desorption of pigments from the adsorbents was invariably highest in poor adsorbents, indicating their poor adsorption energy for betalaines. Similarly, recovery by desorption was low in those adsorbents having high adsorption capacity, indicating that adsorbents such as activated ones with highest adsorption capacity with zero desorption property were unsuitable for the recovery of effluxed pigments. Ex situ recovery of betalain done using various combinations of alumina/silica and processed sand and different column geometries indicated that alumina with processed sand at a 2:1 ratio (w/w) and a minimum column material of 2 cm height and 2 cm diameter was good enough to cause 97% pigment adsorption from a solution containing 1.6 mg mL(-)(1). Desorption and recovery of pigments ex situ from columns were affected by various elution mixtures, where a gradient elution with ascending levels of HCl/ethanol in water resulted in 100% recovery of adsorbed pigments in a significantly lesser volume of eluent in a short period of 1 h. Different pigment flow rates of 0.2, 0.3, and 3.1 mL s(-)(1) through a column of alumina/processed sand indicated that a pigment equilibrium concentration of 0.18 mg mL(-)(1) at flow rates of 0.02 and 0.3 mL s(-)(1) resulted in a breakthrough at 110 and 14 min adsorbing 16.9 and 16.91 mg g(-)(1) betalain, respectively. From the breakthrough curves, the column capacities for respective flow rates were calculated as 8.86 and 9.6 mg g(-)(1), and the higher flow rates resulted in earlier breakthrough with lower capacity. Observations made in the present study are useful to develop a process for the on-line recovery of betalains effluxed from hairy roots.  相似文献   

15.
Abstract

The ability of dried bacterial strain Bacillus sp. S14 to adsorb Malathion in a packed bed column reactor was studied. The effects of important design parameters such as bed height, flow rate and influent Malathion concentration on Malathion removal from an aqueous solution was studied using a packed bed column reactor. The optimised conditions for maximum Malathion removal were found to be: flow rate: 5 mL min-1, bed height: 6.0 cm and influent Malathion concentration: 25 mg L-1. The Adams-Bohart model, Wolborska model, Thomas model, Yoon and Nelson Model were employed to determine characteristic parameters such as N0 (saturation concentration, mg L-1), βo (external mass transfer coefficient, min-1), k Th(Thomas rate constant, mL min-1mg-1), q0 (maximum solid phase concentration of the solute, mg L-1), kYN (rate constant, min-1) and τ (time required for 50 % adsorbate breakthrough time, min) which are useful for process design. Data were fitted with Adams-Bohart model at lower region of (C/C0) values but more accurately fitted with Wolborska and Thomas model.  相似文献   

16.
A combination of olive pomace after solvent extraction and charcoal produced from the solid waste of olive oil press industry was used as an adsorbent for the removal of methylene blue (MB) dye from aqueous solutions. Batch tests showed that up to 80% of dye was removed when the dye concentration was 10 mg/ml and the sorbent concentration was 45 mg/ml. An increase in the olive pomace concentration resulted in greater dye removal from aqueous solution, and an increase in MB dye concentration at constant adsorbent concentration increased the dye loading per unit weigh of adsorbent. In the kinetic of the adsorbent process, the adsorption data followed the second-order kinetic model better than first order kinetic model. Charcoal showed higher sorption capacity (uptake) than that of olive pomace. In the fixed bed adsorption experiment, the breakthrough curves showed constant pattern behavior, typical of favorable isotherms. The breakthrough time increased with increasing bed height, decreasing flow rate and decreasing influent concentration and methylene blue dye uptake. The uptake of MB dye was significantly increased when a mixture of olive pomace and charcoal was packed in the column in a multi-layer fashion. Different models were used to describe the behavior of this packed-sorption process.  相似文献   

17.
Microalgal biofuel alternatives have been hindered by their cost and energy intensive production. In the microalgal harvesting process, the intermediate step of flocculation shows potential in drastically reducing the need for costly centrifugation processes. Moringa oleifera seeds, which have been used for water treatment due to their high flocculation potential, low cost and low toxicity, are presented in this paper as strong candidate for flocculating Chlorella vulgaris, a microalgae with high biodiesel production potential. Early results of our group showed a very high flocculation (around 85% of biomass recovery). The aim of this work was to investigate the influence of Moringa oleifera seed flour concentration, sedimentation time and pH on the flocculation efficiency. Cell suspensions treated with Moringa seed flour (1 g L-1) had their flocculation significantly increased with the rise of pH, reaching 89% of flocculation in 120 min at pH 9.2. Sedimentation time of 120 min and a concentration of 0.6 g L-1 proved to be ample for substantial flocculation efficiency. In spite of the need for more research to ensure the economic viability and sustainability of this process, these results corroborate Moringa oleifera seeds as a strong candidate as a bioflocculant for Chlorella vulgaris cells and indicate optimal pH range of its action.  相似文献   

18.
Performance of affinity chromatography columns was studied by measuring the rates of adsorption and elution of trypsin in a Sepharose 4B-soybean trypsin inhibitor column and a Sepharose 4B-arginine peptides column. The volumetric coefficient for trypsin transfer was evaluated from the break-through curves of trypsin, and elution profiles bed height of Sepharose 4B-STI column was estimated based on these results.  相似文献   

19.
Ten different seaweed species were compared on the basis of lead uptake at different pH conditions. The brown seaweed, Turbinaria conoides, exhibited maximum lead uptake (at pH 4.5) and hence was selected for further studies. Sorption isotherms, obtained at different pH (4–5) and temperature (25–35 °C) conditions were fitted using Langmuir and Sips models. According to the Langmuir model, the maximum lead uptake of 439.4 mg/g was obtained at optimum pH (4.5) and temperature (30 °C). The Sips model better described the sorption isotherms with high correlation coefficients at all conditions examined. Various thermodynamic parameters such as ΔG°, ΔH° and ΔS° were calculated indicating that the present system was a spontaneous and endothermic process. Through potentiometric titrations, number of binding sites (carboxyl groups) and pK1 were determined as 4.1 mmol/g and 4.4, respectively. The influence of co-ions (Na+, K+, Mg2+ and Ca2+) on lead uptake was well pronounced in the case of divalent ions compared to monovalent ions. The solution of 0.1 M HCl successfully eluted all lead ions from lead-loaded T. conoides biomass. The regeneration experiments revealed that the alga could be successfully reused for five cycles without any loss in lead biosorption capacity. A glass column (2 cm i.d. and 35 cm height) was used to study the continuous lead biosorption performance of T. conoides. At 25 cm (bed height), 5 ml/min (flow rate) and 100 mg/l (initial lead concentration), T. conoides exhibited lead uptake of 220.1 mg/g. The column was successfully eluted using 0.1 M HCl, with elution efficiency of 99.7%.  相似文献   

20.
Human epidermal growth factor is a polypeptide hormone having many diverse biological functions. This paper first presents the recovery results of human epidermal growth factor (hEGF) immediately from the fermentation broth of recombinant Escherichia coli by using an expanded bed system (a couple of STREAMLINE25 and ÄKTA explorer 100). The influences of operational conditions such as linear flow rate, gradient length of NaCl concentration, pH and sample concentration on the purification performances of hEGF in expanded and packed bed modes with STREAMLINE DEAE resin were systematically evaluated. After optimization, the practical recovery procedure in the expanded bed mode was carried out on a scaled-up system under the conditions of linear flow rates of 183 cm/h (upward) and 37 cm/h (downward), sample volume of 300 ml and column bed height of 13.8 cm which yielded a primary product of hEGF from the cell-free supernatant containing hEGF after centrifugation at 4000 rev/min for 15 min. As a result, the hEGF concentration in the product was higher than 20% (w/v), the concentration factor was greater than 4.3 and the total yield was higher than 80%, respectively. At the same time, the results of hEGF recovery by using expanded bed adsorption (EBA), packed bed chromatography (PBC) and salting out were compared. The results show that the procedure of hEGF recovery in expanded bed adsorption has some advantages over the other two procedures, because of its higher concentration factor, recovery yield, productivity, hEGF concentration in the primary product and shorter duration of purification run.  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