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.
Summary Power input measurements are carried out in a production bioreactor with a liquid volume up to 25 m3. The results show that the cavity formation principle is applicable to reactors at this scale. It can also be observed that empirical correlations are not useful to predict gassed power input accurately. It is found that at gas flow rates for normal production conditions (NQ =0.1), the gassed power input is about 30–40 % of the non gassed power input.Nomenclature Cp specific heat J/kgK - D impeller diameter m - Db1 impeller blade diameter m - d baffle diameter m - Fr Froude number - - g gravitation m/s2 - h impeller clearance m - H liquid height m - N stirrer speed s-1 - Np power number - - NQ gas flow (aeration) number - - NQ * critical gasflow number for 3 cavity formation - - Po ungassed power consumption W - Pg gassed power consumption W - Q gas flow rate (273 K, 105 N/m2) m3/s - Re Reynolds number - - T tankdiameter m temperature K - t time s - V liquid volume m3 - Vtip impeller tip speed m/s - Vs impeller correlated superficial gas flow rate m/s - W impeller blade width m - density kg/m3  相似文献   

2.
Gas hold-up (ɛg), sauter mean bubble diameter (d32) and oxygen transfer coefficient (kLa) were evaluated at four different alkane concentrations (0.05, 0.1, 0.3 and 0.5 vol.%) in water over the range of superficial gas velocity (ug) of (1.18–23.52) × 10−3 m/s at 25 °C in a laboratory-scale bubble column bioreactor. Immiscible hydrocarbons (n-decane, n-tridecane and n-hexadecane) were utilized in the experiments as impurity. A type of anionic surfactant was also employed in order to investigate the effect of addition of surfactant to organic-aqueous systems on sauter mean bubble diameter, gas hold-up and oxygen transfer coefficient. Influence of addition of alkanes on oxygen transfer coefficient and gas hold-up, was shown to be dependent on the superficial gas velocity. At superficial gas velocity below 0.5 × 10−3 m/s, addition of alkane in air–water medium has low influence on oxygen transfer coefficient and also gas hold-up, whereas; at higher gas velocities slight addition of alkane increases oxygen transfer coefficient and also gas hold-up. Increase in concentration of alkane resulted in increase in oxygen transfer coefficient and gas hold-up and roughly decrease in sauter mean bubble diameter, which was attributed to an increase in the coalescence-inhibiting tendency in the presence of surface contaminant molecules. Bubbles tend to become smaller with decreasing surface tension of hydrocarbon, thus, oxygen transfer coefficient increases due to increasing of specific gas–liquid interfacial area (a). Empirical correlations were proposed for evaluating gas hold-up as a function of sauter mean bubble diameter, superficial gas velocity and interfacial surface tension as well as evaluating Sherwood number as a function of Schmidt, Reynolds and Bond numbers.  相似文献   

3.
Oxygen transfer in a 0.35 m diameter stirred loop fermentor (a stirred tank with a concentric draft tube) has been studied with water containing a small amount of polymer(polyethylene oxide) as a drag-reducing additive.Power consumption was measured. It was found that the addition of polyethylene oxide causes an increase of power consumption. This is contrary to the results reported in the literature.Volumetric mass transfer coefficients (K La) were measured. In water the introduction of the draft tube increased the K La coefficient. The increase in K La became larger with impeller speed. On the other hand, mass transfer in dilute polymer solutions decreased due to the presence of the draft tube. An empirical correlation has been proposed for the volumetric mass transfer coefficient in stirred loop fermentors. It has a general applicability.List of Symbols a 1/m specific surface area - C constant in Eq. (6) - g m/s2 gravitational acceleration - K L m/s overall liquid-phase mass transfer coefficient - n 1/s impeller speed - P W aerated power input by mechanical agitation - P g W power input by sparged air - Q m3/min volumetric gas flow rate - U sg m/s superficial gas velocity - V m3 liquid volume Greek Symbols exponents in Eq. (3) - exponent in Eq. (6) - kg/m3 density  相似文献   

4.
The main parameters which influence the behaviour of phase separation in a single-stage Kühni-type aqueous two-phase extraction column containing polyethylene (PEG) and di-potassium hydrogen phosphate were characterised. Two aqueous two-phase system (ATPS) composed of 12% (w/w) PEG 1450 and 12% (w/w) di-potassium hydrogen phosphate (designated as 12/12) and 12% (w/w) PEG 1450 and 11% (w/w) di-potassium hydrogen phosphate (designated as 12/11) were chosen in this study. The hold-up ɛD increased with increasing impeller speeds and mobile phase flow rates. Phase separation for the 12/11 system was slower than that for the 12/12 system, which resulted in higher dispersed phase hold-up values for the 12/11 system. For 12/12 system, mass transfer of plasmid DNA (pDNA) from the dispersed mobile phase to the stationary phase increased rapidly with increasing impeller speeds of 130, 160 and 200 rpm which was reflected in the decreased values for CT/CTo. The degree of back-mixing quantified by the axial dispersion coefficient Dax was estimated to be 2.7 × 10−6 m2 s−1.  相似文献   

5.
Power requirements in the agitation of non-Newtonian fermentation broths with and without aeration were measured by a strain gage-type dynamometer. Broth from the production of gluc-amylase by Endomyces species and carboxymethyl cellulose solutions were used as non-Newtonian fluids. In gas–liquid agitation systems, the correlation between Pg and P02 ND3/Q0.56 observed by Michel and Miller was found to be applicable to non-Newtonian fluids in laminar and transition regions. This was particularly true for fluids with apparent viscosities larger than 300 cp. The impeller diameter and impeller blade width had considerable effects on power consumption in a nongassed system. It was suggested, therefore, that Pg/P0 should be correlated by a dimensionless term involving some impeller-size factors.  相似文献   

6.
For three types of self-sucking impellers (fourand six-pipe and disk impellers) mixing power, initial point, amount of gas leaving the impeller and mass transfer coefficient were determined experimentally. Investigations were performed for two systems: water and biomass solution.From the point of view of a minimum mixing power and maximum mass transfer coefficient the best impeller has been chosen. Fuzzy multiobjective optimization for determination of optimum operating conditions is proposed.List of Symbols c concentration of oxygen - D tank diameter - d impeller diameter - g acceleration of gravity - H height of liquid in the tank - H height of liquid above impeller, H=H-y - k consistency coefficient - k L a volumetric mass transfer coefficient - N rotational speed of impeller - n flow behaviour index - P mixing power for pure liquid - P G mixing power for aerated liquid - V G volumetric air flow rate - y distance of impeller from the tank bottom - v a apparent kinematic viscosity of liquid - density of liquid - time - gas hold-up - Eu=P/N 3 d 5 or EuG=P G /N 3 d 5 Euler Number for non-gassed or aerated liquid - Fr=N 2 d/g Froude Number - Fr*=N 2 d 2 /g(H -y) modified Froude Number - KG=V G /N d 3 gas flow number - Re=N d 2 /v a Reynolds Number - Sh=k K a/(g 2 /v a )1/3 Sherwood Number  相似文献   

7.
In photobioreactors, which are usually operated under light limitation,sufficient dissolved inorganic carbon must be provided to avoid carbonlimitation. Efficient mass transfer of CO2 into the culture mediumisdesirable since undissolved CO2 is lost by outgassing. Mass transferof O2 out of the system is also an important consideration, due tothe need to remove photosynthetically-derived O2 before it reachesinhibitory concentrations. Hydrodynamics (mixing characteristics) are afunctionof reactor geometry and operating conditions (e.g. gas and liquid flow rates),and are a principal determinant of the light regime experienced by the culture.This in turn affects photosynthetic efficiency, productivity, and cellcomposition. This paper describes the mass transfer and hydrodynamics within anear-horizontal tubular photobioreactor. The volume, shape and velocity ofbubbles, gas hold-up, liquid velocity, slip velocity, axial dispersion,Reynoldsnumber, mixing time, and mass transfer coefficients were determined intapwater,seawater, and algal culture medium. Gas hold-up values resembled those ofvertical bubble columns, and the hydraulic regime could be characterized asplug-flow with medium dispersion. The maximum oxygen mass transfer coefficientis approximately 7 h–1. A regime analysisindicated that there are mass transfer limitations in this type ofphotobioreactor. A methodology is described to determine the mass transfercoefficients for O2 stripping and CO2 dissolution whichwould be required to achieve a desired biomass productivity. This procedure canassist in determining design modifications to achieve the desired mass transfercoefficient.  相似文献   

8.
The absorption of oxygen in aqueous–organic solvent emulsions was studied in a laboratory-scale bubble reactor at a constant gas flow rate. The organic and the gas phases were dispersed in the continuous aqueous phase. Volumetric mass transfer coefficients (kLa) of oxygen between air and water were measured experimentally using a dynamic method. It was assumed that the gas phase contacts preferentially the water phase. It was found that addition of silicone oils hinders oxygen mass transfer compared to air–water systems whereas the addition of decane, hexadecane and perfluorocarbon PFC40 has no significant influence. By and large, the results show that, for experimental conditions (organic liquid hold-up ≤10% and solubility ratio ≤10), the kLa values of oxygen determined in binary air–water systems can be used for multiphase (gas–liquid–liquid) reactor design with applications in environmental protection (water and air treatment processes).  相似文献   

9.
Radial flow Rushton impellers were compared qualitatively with axial flow hydrofoil impellers (Maxflo T and A315) at the pilot scale. Six types of impellers were compared for qualitative differences in mass transfer. Measurements were conducted using three model systems: water, glycerol and Melojel (soluble starch). Power measurements were obtained using watt transducers, which although limited in accuracy and prone to interferences, were able to provide useful qualitative monitoring results. While there was little effect of impeller type on mass transfer as measured by the rapid pressure increase technique, significant qualitative differences were observed using the rapid temperature increase technique specifically for the Melojel and glycerol model systems. The Miller correlation, relating gassed-to-ungassed power, was used effectively to qualitatively evaluate the power drop upon gassing for both the model systems and a Streptomyces fermentation for the various impeller types. A high oxygen demand Streptomcyes fermentation then was conducted in fermenters possessing each type of impeller. Performance was not adequate with the A315 impellers pumping upwards and the small diameter Maxflo T impellers. Peak titers and profiles of the estimated apparent broth viscosity varied depending upon the impeller type. Mass transfer rates generally declined with higher viscosities when other fermentation operating conditions where held constant. Overall, values for OUR, k L a, P g /V L and other calculated mass transfer and power input quantities for the A315 pumping upwards and undersized Maxflo T (D T /D I ?=?2.3) impellers were at the lower end of the range obtained for the larger Maxflo T (D T /D I ?=?1.8–2.0) and A315 impellers pumping downwards. Rushton impellers generally behaved qualitatively similar to hydrofoil impellers based on these calculated quantities.  相似文献   

10.
 Human thermal physiological and comfort models will soon be able to simulate both transient and spatial inhomogeneities in the thermal environment. With this increasing detail comes the need for anatomically specific convective and radiative heat transfer coefficients for the human body. The present study used an articulated thermal manikin with 16 body segments (head, chest, back, upper arms, forearms, hands, pelvis, upper legs, lower legs, feet) to generate radiative heat transfer coefficients as well as natural- and forced-mode convective coefficients. The tests were conducted across a range of wind speeds from still air to 5.0 m/s, representing atmospheric conditions typical of both indoors and outdoors. Both standing and seated postures were investigated, as were eight different wind azimuth angles. The radiative heat transfer coefficient measured for the whole-body was 4.5 W/m2 per K for both the seated and standing cases, closely matching the generally accepted whole-body value of 4.7 W/m2 per K. Similarly, the whole-body natural convection coefficient for the manikin fell within the mid-range of previously published values at 3.4 and 3.3 W/m2 per K when standing and seated respectively. In the forced convective regime, heat transfer coefficients were higher for hands, feet and peripheral limbs compared to the central torso region. Wind direction had little effect on convective heat transfers from individual body segments. A general-purpose forced convection equation suitable for application to both seated and standing postures indoors was h c=10.3v 0.6 for the whole-body. Similar equations were generated for individual body segments in both seated and standing postures. Received: 21 May 1996/Accepted: 27 November 1996  相似文献   

11.
The batch productivity (Q TM) of the production of the nucleoside antibiotic toyocamycin (TM) by Streptomyces chrestomyceticus was increased ten-fold by selection of a UV generated mutant, optimization of pH, increasing incubation temperature from 28 °C to 36 °C, and addition of soy oil. Initial high oxygen transfer rates stimulated Q TM maxima two-fold. Antibiotic production by the mutant strain, U190, however, appeared more shear sensitive than the parent culture FCRF 341 with maximum antibiotic titer being inversely related to impellor tip velocity, T v . For this reason, scale-up could not be done at constant P/V or constant volumetric oxygen transfer. Instead, programming of impeller speed was evaluated in order to maintain optimal impeller tip velocity during scale-up. It was found that a low constant T v maintained in scale-up in geometrically similar vessels was most beneficial for duplication of optimal antibiotic productivity, Q TM. Pilot fermentations (120 dm3 scale) were used to determine coefficients of Q TM variation from oxygen uptake rate (OUR) and total CO2 evolution data for monitoring of Q TM variation during scale-up to the 12,000 dm3 scale. This technique allowed for on-line prediction of antibiotic titer and Q TM from fermentor exhaust gas data.List of Symbols A scale constant - B shape constant - C location of maximum constant - D m impeller diameter (m) - H m liquid height (m) - OTR MmolO2·(dm3)–1min–1 oxygen transfer rate - OUR MmolO2·(dm3)–1min–1 oxygen uptake rate - PCV cm3 packed cell volume - P/V watts/dm3 volumetric power consumption - Q 1 · min–1 corrected to standard conditions of temperature, pressure aeration rate - Q TM g/(cm3 · h) or kg/(m3 · h) antibiotic productivity - T m tank diameter - T mix s mixing time - T v cm · s–1 impeller tip velocity - TM g/cm3 Toyocamycin concentration - TNP Tricyclic nucleoside phosphate  相似文献   

12.
Power input is an important engineering and scale‐up/down criterion in stirred bioreactors. However, reliably measuring power input in laboratory‐scale systems is still challenging. Even though torque measurements have proven to be suitable in pilot scale systems, sensor accuracy, resolution, and errors from relatively high levels of friction inside bearings can become limiting factors at smaller scales. An experimental setup for power input measurements was developed in this study by focusing on stainless steel and single‐use bioreactors in the single‐digit volume range. The friction losses inside the air bearings were effectively reduced to less than 0.5% of the measurement range of the torque meter. A comparison of dimensionless power numbers determined for a reference Rushton turbine stirrer (NP = 4.17 ± 0.14 for fully turbulent conditions) revealed good agreement with literature data. Hence, the power numbers of several reusable and single‐use bioreactors could be determined over a wide range of Reynolds numbers between 100 and >104. Power numbers of between 0.3 and 4.5 (for Re = 104) were determined for the different systems. The rigid plastic vessels showed similar power characteristics to their reusable counterparts. Thus, it was demonstrated that the torque‐based technique can be used to reliably measure power input in stirred reusable and single‐use bioreactors at the laboratory scale.  相似文献   

13.
An Amycolatopsis fastidiosa culture, which produces the nocathiacin class of antibacterial compounds, was scaled up to the 15,000 L working volume. Lower volume pilot fermentations (600, 900, and 1,500 L scale) were conducted to determine process feasibility at the 15,000 L scale. The effects of inoculum volume, impeller tip speed, volumetric gas flow rate, superficial gas velocity, backpressure, and sterilization heat stress were examined to determine optimal scale‐up operating conditions. Inoculum volume (6 vs. 2 vol %) and medium sterilization (Ro of 68 vs. 92 min?1) had no effect on productivity or titer, and higher impeller tip speeds (2.1 vs. 2.9 m/s) had a slight effect (20% decrease). In contrast, higher backpressure, incorporating increased head pressure at the 15,000 L scale (1.2 vs. 0.7 kg/cm2) and low gas flow rates (0.25 vs. 0.8 vvm), appeared to be problematic (40–50% decrease). High off‐gas CO2 levels were likely reasons for observed lower productivity. Consequently, air flow rate for this 25‐fold scale‐up (600–15,000 L) was controlled to match off‐gas CO2 profiles of acceptable smaller scale batches to maintain levels below 0.5%. The 15,000 L‐scale fermentation achieved an expected nocathiacin I titer of 310 mg/L after 7 days. Other on‐line data (i.e., pH, oxygen uptake rate, and CO2 evolution rate) and off‐line data (i.e., analog production, glucose utilization, ammonium production, and dry cell weight) at the 15,000 L scale also tracked similarly to the smaller scale, demonstrating successful fermentation scale‐up. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2009  相似文献   

14.
Summary Conditions for the production of microbial uricase byCandida utilis were studied. For the selected strain, hypoxanthine proved to be the most effective inducer of uricase formation. The highest values of biomass as well as uricase activity in the mechanically agitated fermentor were obtained under the following conditions: 50 h, rotation impeller speed 7 s–1, air flow rate 1.25×10–5 m3s–1, concentration of inducer 0.1%.List of symbols b width of baffle, m - c length of baffle, m - D diameter of cylindrical fermentor, m - d diameter of impeller, m - d 1 diameter of impeller disc, m - Fr m impeller Froud number - g gravitional acceleration, ms–2 - H height of batch surface above bottom, m - H 2 height of impeller disc above bottom, m - h height of impeller blade, m - Kp g flow rate number - L length of impeller blade, m - N rotational speed of impeller, s–1 - Re m impeller Reynolds number - T time, h - V volume of batch, m3 - V g air (gas) flow rate, m3s–1 - x mass fraction of the dry matter of cells - x 0 initial value of the mass fraction of the dry matter of cells - r volume fraction of the dry matter of cells - <eta<1 viscosity of pure liquid, Pa s - viscosity of batch (suspension of microbial suspension), Pa s - density of batch, kg m–3  相似文献   

15.
 The rates of convection and evaporation at the interface between the human body and the surrounding air are expressed by the parameters convective heat transfer coefficient h c, in W m–2°C–1 and evaporative heat transfer coefficient h e, W m–2 hPa–1. These parameters are determined by heat transfer equations, which also depend on the velocity of the airstream around the body, that is still air (free convection) and moving air (forced convection). The altitude dependence of the parameters is represented as an exponential function of the atmospheric pressure p: h cp n and h ep 1–n, where n is the exponent in the heat transfer equation. The numerical values of n are related to airspeed: n=0.5 for free convection, n=0.618 when airspeed is below 2.0 ms–1 and n=0.805 when airspeed is above 2.0 ms–1. This study considers the coefficients h c and h e with respect to the similarity of the two processes, convection and evaporation. A framework to explain the basis of established relationships is proposed. It is shown that the thickness of the boundary layer over the body surface increases with altitude. As a medium of the transfer processes, the boundary layer is assumed to be a layer of still air with fixed insulation which causes a reduction in the intensity of heat and mass flux propagating from the human body surface to its surroundings. The degree of reduction is more significant at a higher altitude because of the greater thickness of the boundary layer there. The rate of convective and evaporative heat losses from the human body surface at various altitudes in otherwise identical conditions depends on the following factors: (1) during convection – the thickness of the boundary layer, plus the decrease in air density, (2) during evaporation (mass transfer) – the thickness of the boundary layer, plus the increase with altitude in the diffusion coefficient of water vapour in the air. The warming rate of the air volume due to convection and evaporation is also considered. Expressions for the calculation of altitude dependences h c (p) and h e (p) are suggested. Received: 23 June 1998 / Accepted: 10 February 1999  相似文献   

16.
Sugarcane bagasse is a low-cost and abundant by-product generated by the bioethanol industry, and is a potential substrate for cellulolytic enzyme production. The aim of this work was to evaluate the effects of air flow rate (Q AIR), solids loading (%S), sugarcane bagasse type, and particle size on the gas hold-up (ε G) and volumetric oxygen transfer coefficient (k L a) in three different pneumatic bioreactors, using response surface methodology. Concentric tube airlift (CTA), split-cylinder airlift (SCA), and bubble column (BC) bioreactor types were tested. Q AIR and  %S affected oxygen mass transfer positively and negatively, respectively, while sugarcane bagasse type and particle size (within the range studied) did not influence k L a. Using large particles of untreated sugarcane bagasse, the loop-type bioreactors (CTA and SCA) exhibited higher mass transfer, compared to the BC reactor. At higher  %S, SCA presented a higher k L a value (0.0448 s?1) than CTA, and the best operational conditions in terms of oxygen mass transfer were achieved for  %S < 10.0 g L?1 and Q AIR > 27.0 L min?1. These results demonstrated that pneumatic bioreactors can provide elevated oxygen transfer in the presence of vegetal biomass, making them an excellent option for use in three-phase systems for cellulolytic enzyme production by filamentous fungi.  相似文献   

17.
Summary The hydrodynamics and mass transfer behaviour of an airlift fermentor with an external loop (height 10m) has been investigated by measuring gas and liquid velocities, gas hold-up, liquid mixing and oxygen transfer coefficients. Liquid phase properties, i.e., ionic strength, viscosity and surface tension have been varied by altering the fermentation media. Results are compared with those from bubble column experiments performed in the same unit. It is shown, that more uniform two-phase flow in the airlift leads to advantages in scale-up and operation.Nomenclature a Specific interfacial area per volume of dispersion (m2/m3) - c Local concentration of tracer (kmol/m3) - c Concentration of tracer at infinite time (kmol/m3) - CL Concentration of oxygen in the liquid bulk (kmol/m3) - CL * Concentration of oxygen in the interface (kmol/m3) - Dax Axial dispersion coefficient (cm2/s) - I Ionic strength (kmol/m3) - i Inhomogeneity [defined in Eq. (2)] - Rate of oxygen transfer (kmol/s) - tc Circulation time (s) - tM Mixing time (s) - VR Volume of gas-liquid dispersion (m3) - VSG Superficial gas velocity in up-flow column (m/s) Greek letter symbols L Oxygen transfer coefficient (m/s) - Dynamic viscosity (m Pa s) - Surface tension (m N/m) Presented at the First European Congress on Biotechnology, Interlaken, September 25–29, 1978  相似文献   

18.
An atmospheric-pressure dc discharge in air (i = 10–50 mA) with metal and liquid electrolyte electrodes was studied experimentally. An aqueous solution of sodium chloride (0.5 mol/L) was used as the cathode or anode. The electric field strength in the plasma and the cathode (anode) voltage drops were obtained from the measured dependences of the discharge voltage on the electrode gap length. The gas temperature was deduced from the spectral distribution of nitrogen emission in the band N2(C3Π u B3Π g , 0–2). The time dependences of the temperatures of the liquid electrolyte electrodes during the discharge and in its afterglow, as well as the evaporation rate of the solution, were determined experimentally. The contributions of ion bombardment and heat flux from the plasma to the heating of the liquid electrode and transfer of solvent (water) into the gas phase are discussed using the experimental data obtained.  相似文献   

19.
Harper JE 《Plant physiology》1981,68(6):1488-1493
Studies were conducted to quantitate the evolution of nitrogen oxides (NO(x)) from soybean [Glycine max (L.) Merr.] leaves during in vivo nitrate reductase (NR) assays with aerobic and anaerobic gas purging. Anaerobic gas purging (N2 and argon) consistently resulted in greater NO(x) evolution than did aerobic gas purging (air and O2). The evolution of NO(x) was dependent on gas flow rate and on NO2 formation in the assay medium; although a threshold level of NO2 appeared to exist beyond which the rate of NO(x) evolution did not increase further.  相似文献   

20.
谭鑫  李超  郭美锦 《生物工程学报》2022,38(12):4692-4704
红霉素(erythromycin)是由绛红色糖多胞菌(Saccharopolyspora erythraea)发酵生产的次级代谢产物,其生产水平不仅受发酵工艺的影响,也受反应器结构影响。为解决红霉素发酵过程放大问题,本研究采用时间常数法和计算流体力学(computational fluid dynamics,CFD)数值模拟验证相结合的方法设计了500m3超大规模红霉素耗氧发酵生物反应器。首先,通过对50L反应器红霉素发酵过程研究,发现溶氧是关键性限制因素,通过氧消耗速率(oxygen uptake rate,OUR)等参数分析计算得到设备的氧供应时间常数tmt需小于6.25s。然后,基于时间常数法和经验关联式理性设计500m3反应器搅拌桨叶组合方式,即底层BDT8桨叶+两层MSX4桨叶的搅拌桨组合,并通过经验公式及CFD方法对设计结果进行了模拟验证。两种验证方法结果均表明500m³反应器采取底层BDT8桨叶+两层MSX4桨叶的组合方式时设备的氧供应时间常数小于6.25s,且反应器内流场特性(如持气率、剪切率和速度矢量等)均能满足红霉素大规模发酵的需要。经实际发酵验证,设计的生物反应器能够满足红霉素的工业规模发酵应用。  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