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 31 毫秒
1.
Chromatographic data pertaining to the enantioseparation of L- and D-phenylalanine anilide (PA) on a polymeric stationary phase imprinted with L-PA were studied from the viewpoints of phase equilibrium, mass transfer kinetics, and the thermodynamic properties of this enantiomeric separation system. The concentration dependence of the lumped mass transfer rate coefficient (k(m,L)) previously published was analyzed to obtain new information concerning the mass transfer characteristics in this chiral separation system. It was shown that intraparticle diffusion contributed much more to k(m,L) than adsorption/desorption. The positive concentration dependence of k(m,L) seemed to be interpreted by considering that of the surface diffusion coefficient, itself explained by the heterogeneous surface model. The characteristic features of the phase equilibrium, the mass transfer kinetics, and the thermodynamics of the enantiomeric separation system probably result from the adsorption energy distribution on the surface of the imprinted phase having an exponential decay.  相似文献   

2.
To study the effect of an unstirred layer (UL), we have investigated the exchange efflux kinetics of anandamide at 0 degrees C, pH 7.3, from albumin-free as well as from albumin-filled human red blood cell ghosts to media of various BSA concentrations ([BSA](o)). The rate constant (k(m)) of unidirectional flux from the outer membrane leaflet to BSA in the medium increased with the square root of [BSA](o) in accordance with the existence of a UL, which is a water layer adjacent to the membrane that is not subject to the same gross mixing that takes place in the rest of the medium. From k(m), it is possible to calculate the rate constant of anandamide dissociation from BSA (k(1)) if we know the membrane binding of anandamide, the equilibrium dissociation constant of BSA-anandamide complexes, and the diffusion constant of anandamide. We estimated k(1) to be 3.33 +/- 0.27 s(-1). The net flux of [(3)H]anandamide is balanced by an equal and opposite movement of nonradioactive anandamide in exchange efflux experiments. This means that our results are also valid for uptake. We show that for anandamide with rapid membrane translocation, UL causes a significant resistance to cellular uptake. Depicting the rate of anandamide uptake as a function of equilibrium water phase concentrations results in a parabolic uptake dependence. Such apparent "saturation kinetics" is often interpreted as indicating the involvement of transport proteins. The validity of such an interpretation is discussed.  相似文献   

3.
Permeable large-pore particles have many applications, in particular in perfusion chromatography for bioseparations. The objective of this paper is to elucidate the mass transport mechanisms in two commercial adsorbents-POROS Q/M and Q Hyper D- and to answer the question if intraparticle convection is present as a mass transfer mechanism. The paper contains three sections. In the first part, mass transfer inside porous particles is discussed. The mass transfer mechanism which allows improved performance of perfusion chromatography is intraparticle convection. The combined effect of intraparticle convection and diffusion is an “augmented” effective diffusivity. This is the key concept to explain the peak sharpening and modified Van Deemter plots found with large-pore particles. In the second part, column efficiencies in terms of HETP as a function of bed superficial velocity are experimentally measured for a non-retained protein (bovine serum albumine, BSA) in two adsorbents: POROS Q/M (PerSeptive Biosystems) and Q Hyper D (BioSepra). In the third section breakthrough curves for both materials are measured for a test protein (BSA) from which useful capacities and productivities as a function of flow-rate are calculated. Experimental results indicate that intraparticle convection plays indeed an important role in both adsorbents.  相似文献   

4.
The scale up of the novel, pharmaceutically important pneumocandin (B(0)), from the filamentous fungus Glarea lozoyensis was successfully completed from pilot scale (0.07, 0.8, and 19 m(3)) to production scale (57 m(3)). This was accomplished, despite dissimilar reactor geometry, employing a combination of scale-up criteria, process sensitivity studies, and regime analysis using characteristic time constants for both oxygen mass transfer and bulk mixing. Dissolved oxygen tension, separated from the influence of agitation by gas blending at the 0.07 m(3)-scale, had a marked influence on the concentrations of pneumocandin analogs with different levels of hydroxylation, and these concentrations were used as an indicator of bulk mixing upon scale up. The profound impact of dissolved oxygen tension (DOT) (low and high levels) on analog formation dictated the use of constant DOT, at 80% air saturation, as a scale-up criterion. As a result k(L)a, Oxygen uptake rate (OUR) and hence the OTR were held constant, which were effectively conserved across the scales, while the use of other criterion such as P(g)/V(L), or mixing time were less effective. Production scale (57 m(3)) mixing times were found to be faster than those at 19 m(3) due to a difference in liquid height/tank diameter ratio (H(L)/D(T)). Regime analysis at 19 and 57 m(3) for bulk mixing (t(c)) and oxygen transfer (1/k(L)a) showed that oxygen transfer was the rate-limiting step for this highly shear thinning fermentation, providing additional support for the choice of scale-up criterion.  相似文献   

5.
The use of palm kernel fibre, a readily available agricultural waste product for the sorption of Methylene blue from aqueous solution and the possible mechanism of sorption has been investigated at various fibre doses. The extent of dye removal and the rate of sorption were analyzed using two kinetic rate models (pseudo-first and pseudo-second-order kinetic models) and two diffusion models (intraparticle and external mass transfer models).

Analysis of the kinetic data at different sorbent dose revealed that the pseudo-first order kinetics fitted to the kinetic data only in the first 5 min of sorption and then deviated from the experimental data. The pseudo-second-order kinetic model was found to better fit the experimental data with high correlation coefficients at the various fibre dose used. The dye sorption was confirmed to follow the pseudo-second-order model by investigating the relationship between the amount of dye sorbed and the change in hydrogen ion concentration of the dye solution and also the dependence of dye uptake with solution temperature. It was found that the change in hydrogen ion concentration and increase in sorption temperature were directly related to the amount of dye sorbed, and activation energy was calculated to be −39.57 kJ/mol, indicating that the dye uptake is chemisorption, involving valence forces through sharing or exchange of electrons between sorbent and sorbate as covalent forces.

The intraparticle diffusion plots showed three sections indicating that intraparticle diffusion is not solely rate controlling. The intraparticle diffusion and mass transfer rate constants where observed to be well correlated with sorbent dose in the first 5 min of sorption, indicating sorption process is complex. It was found that at low sorbent dose the mass transfer is the main rate controlling parameter. However at high sorbent dose, intraparticle diffusion becomes rate controlling.  相似文献   


6.
7.
J Tandori  P Sebban  H Michel  L Baciou 《Biochemistry》1999,38(40):13179-13187
The X-ray crystallographic structure of the photosynthetic reaction center from Rhodobacter sphaeroides obtained at high resolution has revealed a number of internal water molecules (Ermler, U., Fritzsch, G., Buchanan, S. K., and Michel, H. (1994) Structure 2, 925-936; Stowell, M. H. B., McPhillips, T. M., Rees, D. C., Soltis, S. M., Abresch, E., and Feher, G. (1997) Science 276, 812-816). Some of them are organized into distinct hydrogen-bonded water chains that connect Q(B) (the terminal quinone electron acceptor of the reaction center) to the aqueous phase. To investigate the role of the water chains in the proton conduction process, proline L209, located immediately adjacent to a water chain, was mutated to the following residues: F, Y, W, E, and T. We have first analyzed the effects of the mutations on the kinetic and thermodynamic properties of the rate constants of the second electron transfer (k(AB)(2)) and of the coupled proton uptake (k(H)+) at the second flash. In all aromatic mutants, k(AB)(2) and k(H)+ are notably and concomitantly decreased compared to the wild-type, while no effect is observed in the other mutants. The temperature dependence of these rates shows activation energy values (DeltaH) similar for the proton and electron-transfer processes in the wild-type and in most of the mutants, except for the L209PW and L209PF mutants. The analysis of the enthalpy factors related to the electron and proton-transfer processes in the L209PF and the L209PW mutants allows to distinguish the respective effects of the mutations for both transfer reactions. It is noteworthy that in the aromatic mutants a substantial increase of the free energies of activation is observed (DeltaG(L209PY) < DeltaG(L209PF) < DeltaG(L209PW)) for both proton and electron-transfer reactions, while in the other mutants, DeltaG is not affected. The salt concentration dependence of k(AB)(2) shows, in the L209PF and L209PW mutants, a higher screening of the protein surface potential experienced by Q(B). Our data suggest that residues F and W in position L209 increase the polarizability of the internal water molecules and polar residues by altering the organization of the hydrogen-bond network. We have also analyzed the rates of the first electron-transfer reaction (k(AB)(1)), in the 100 micros time domain. These kinetics have previously been shown to reflect protein relaxation events possibly including proton uptake events (Tiede, D. M., Vazquez, J., Cordova, J., and Marone, P. M. (1996) Biochemistry 35, 10763-10775). Interestingly, in the L209PF and L209PW mutants, k(AB)(1) is notably decreased in comparison to the wild type and the other mutants, in a similar way as k(AB)(2) and k(H)+. Our data imply that the dynamic organization of this web is tightly coupled to the electron transfer process that is kinetically limited by protonation events and/or conformational rearrangements within the protein.  相似文献   

8.
The estimation of volumetric mass transfer coefficient, k(L)a, in stirred tank reactors using artificial neural networks has been studied. Several operational conditions (N and V(s)), properties of fluid (μ(a)) and geometrical parameters (D and T) have been taken into account. Learning sets of input-output patterns were obtained by k(L)a experimental data in stirred tank reactors of different volumes. The inclusion of prior knowledge as an approach which improves the neural network prediction has been considered. The hybrid model combining a neural network together with an empirical equation provides a better representation of the estimated parameter values. The outputs predicted by the hybrid neural network are compared with experimental data and some correlations previously proposed in the literature for tanks of different sizes.  相似文献   

9.
The intermolecular electron transfer from Achromobacter cycloclastes pseudoazurin (AcPAZ) to wild-type and mutant Alcaligenes xylosoxidans nitrite reductases (AxNIRs) was investigated using steady-state kinetics and electrochemical methods. The affinity and the electron transfer reaction constant (k(ET)) are considerably lower between AcPAZ and AxNIR (K(m) = 1.34 mM and k(ET) = 0.87 x 10(5) M(-1) s(-1)) than between AcPAZ and its cognate nitrite reductase (AcNIR) (K(m) = 20 microM and k(ET) = 7.3 x 10(5) M(-1) s(-1)). A negatively charged hydrophobic patch, comprising seven acidic residues around the type 1 copper site in AcNIR, is the site of protein-protein interaction with a positively charged hydrophobic patch on AcPAZ. In AxNIR, four of the negatively charged residues (Glu-112, Glu-133, Glu-195, and Asp-199) are conserved at the corresponding positions of AcNIR, whereas the other three residues are not acidic amino acids but neutral amino acids (Ala-83, Ala-191, and Gly-198). Seven mutant AxNIRs with additional negatively charged residues surrounding the hydrophobic patch of AxNIR (A83D, A191E, G198E, A83D/A191E, A93D/G198E, A191E/G198E, and A83D/A191E/G198E) were prepared to enhance the specificity of the electron transport reaction between AcPAZ and AxNIR. The k(ET) values of these mutants become progressively larger as the number of mutated residues increases. The K(m) and k(ET) values of A83D/A191E/G198E (K(m) = 88 microM and k(ET) = 4.1 x 10(5) M(-1) s(-1)) are 15-fold smaller and 4.7-fold larger than those of wild-type AxNIR, respectively. These results suggest that the introduction of negatively charged residues into the docking surface of AxNIR facilitates both the formation of electron transport complex and the electron transfer reaction.  相似文献   

10.
The aerated stirred reactor (ASR) has been widely used in biochemical and wastewater treatment processes. The information describing how the activated sludge properties and operation conditions affect the hydrodynamics and mass transfer coefficient is missing in the literature. The aim of this study was to investigate the influence of flow regime, superficial gas velocity (U(G)), power consumption unit (P/V(L)), sludge loading, and apparent viscosity (mu(ap)) of activated sludge fluid on the mixing time (t(m)), gas hold-up (epsilon), and volumetric mass transfer coefficient (k(L)a) in an activated sludge aerated stirred column reactor (ASCR). The activated sludge fluid performed a non-Newtonian rheological behavior. The sludge loading significantly affected the fluid hydrodynamics and mass transfer. With an increase in the U(G) and P/V(L), the epsilon and k(L)a increased, and the t(m), decreased. The epsilon, k(L)a, and t(m), were influenced dramatically as the flow regime changed from homogeneous to heterogeneous patterns. The proposed mathematical models predicted the experimental results well under experimental conditions, indicating that the U(G), P/V(L), and mu(ap) had significant impact on the t(m), epsilon, and k(L)a. These models were able to give the t(m), epsilon, and k(L)a values with an error around +/-8%, and always less than +/-10%.  相似文献   

11.
Cathepsin C, or dipeptidyl peptidase I, is a lysosomal cysteine protease of the papain family that catalyzes the sequential removal of dipeptides from the free N-termini of proteins and peptides. Using the dipeptide substrate Ser-Tyr-AMC, cathepsin C was characterized in both steady-state and pre-steady-state kinetic modes. The pH(D) rate profiles for both log k cat/ K m and log k cat conformed to bell-shaped curves for which an inverse solvent kinetic isotope effect (sKIE) of 0.71 +/- 0.14 for (D)( k cat/ K a) and a normal sKIE of 2.76 +/- 0.03 for (D) k cat were obtained. Pre-steady-state kinetics exhibited a single-exponential burst of AMC formation in which the maximal acylation rate ( k ac = 397 +/- 5 s (-1)) was found to be nearly 30-fold greater than the rate-limiting deacylation rate ( k dac = 13.95 +/- 0.013 s (-1)) and turnover number ( k cat = 13.92 +/- 0.001 s (-1)). Analysis of pre-steady-state burst kinetics in D 2O allowed abstraction of a normal sKIE for the acylation half-reaction that was not observed in steady-state kinetics. Since normal sKIEs were obtained for all measurable acylation steps in the presteady state [ (D) k ac = 1.31 +/- 0.04, and the transient kinetic isotope effect at time zero (tKIE (0)) = 2.3 +/- 0.2], the kinetic step(s) contributing to the inverse sKIE of (D)( k cat/ K a) must occur more rapidly than the experimental time frame of the transient kinetics. Results are consistent with a chemical mechanism in which acylation occurs via a two-step process: the thiolate form of Cys-234, which is enriched in D 2O and gives rise to the inverse value of (D)( k cat/ K a), attacks the substrate to form a tetrahedral intermediate that proceeds to form an acyl-enzyme intermediate during a proton transfer step expressing a normal sKIE. The subsequent deacylation half-reaction is rate-limiting, with proton transfers exhibiting normal sKIEs. Through derivation of 12 equations describing all kinetic parameters and sKIEs for the proposed cathepsin C mechanism, integration of both steady-state and pre-steady-state kinetics with sKIEs allowed the provision of at least one self-consistent set of values for all 13 rate constants in this cysteine protease's chemical mechanism. Simulation of the resulting kinetic profile showed that at steady state approximately 80% of the enzyme exists in an active-site cysteine-acylated form in the mechanistic pathway. The chemical and kinetic details deduced from this work provide a potential roadmap to help steer drug discovery efforts for this and other disease-relevant cysteine proteases.  相似文献   

12.
A new type of bubble aeration column called a hollow fiber membrane (HFM) aeration column was proposed, which was featured in the use of hollow fiber membranes and gave a high bubble density in the column. The value of k(L)a was increased by modifying the membrane surface for making the pore size smaller. The Sauter mean diameter of bubbles (D(vs)) was 2.0 +/- 0.1 mm in the range of the superficial gas velocity from 0.02 m s(-1) to 0.065 m s(-1), while that obtained for the bubbles near the membrane was 811 mum at the superficial gas velocity of 4.0 x 10(-4) m s(-1). The difference was ascribed to the effect of coalescence of bubbles. The value of K(L)a increased in proportion to the superficial gas velocity up to 0.02 m s(-1), and was almost constant above 0.03 m s(-1). The maximum value of k(L)a, 2.5 s(-1), was higher than those of the other aeration columns reported previously. The pneumatic power consumption per unit liquid volume (P(v)) for obtaining the same k(L)a was the smallest in the HFM aeration columns. P(v), for obtaining the same interfacial area of bubbles per liquid volume, was also lower than those for other types of aeration columns. It was suggested from the measurement of bubble diameter that the larger interfacial area generated in the HFM aeration column ascribes to the larger gas holdup than the smaller D(vs). (c) 1992 John Wiley & Sons, Inc.  相似文献   

13.
Miniature parallel bioreactors are becoming increasingly important as tools to facilitate rapid bioprocess design. Once the most promising strain and culture conditions have been identified a suitable scale-up basis needs to be established in order that the cell growth rates and product yields achieved in small scale optimization studies are maintained at larger scales. Recently we have reported on the design of a miniature stirred bioreactor system capable of parallel operation [Gill et al. (2008); Biochem Eng J 39:164-176]. In order to enable the predictive scale-up of miniature bioreactor results the current study describes a more detailed investigation of the bioreactor mixing and oxygen mass transfer characteristics and the creation of predictive engineering correlations useful for scale-up studies. A Power number of 3.5 for the miniature turbine impeller was first established based on experimental ungassed power consumption measurements. The variation of the measured gassed to ungassed power ratio, P(g)/P(ug), was then shown to be adequately predicted by existing correlations proposed by Cui et al. [Cui et al. (1996); Chem Eng Sci 51:2631-2636] and Mockel et al. [Mockel et al. (1990); Acta Biotechnol 10:215-224]. A correlation relating the measured oxygen mass transfer coefficient, k(L)a, to the gassed power per unit volume and superficial gas velocity was also established for the miniature bioreactor. Based on these correlations a series of scale-up studies at matched k(L)a (0.06-0.11 s(-1)) and P(g)/V (657-2,960 W m(-3)) were performed for the batch growth of Escherichia coli TOP10 pQR239 using glycerol as a carbon source. Constant k(L)a was shown to be the most reliable basis for predictive scale-up of miniature bioreactor results to conventional laboratory scale. This gave good agreement in both cell growth and oxygen utilization kinetics over the range of k(L)a values investigated. The work described here thus gives further insight into the performance of the miniature bioreactor design and will aid its use as a tool for rapid fermentation process development.  相似文献   

14.
The apparent catalytic constant (k(cat)) of artichoke (Cynara scolymus L.) peroxidase (AKPC) with 2,2'-azino-bis(3-ethylbenzthiazoline-6-sulphonic acid) (ABTS) increased 130-fold in the presence of calcium ions (Ca2+) but the affinity (K(m)) of the enzyme for ABTS was 500 times lower than for Ca2+-free AKPC. AKPC is known to exhibit an equilibrium between 6-aquo hexa-coordinate and penta-coordinate forms of the haem iron that is modulated by Ca2+ and affects compound I formation. Measurements of the Ca2+ dissociation constant (K(D)) were complicated by the water-association/dissociation equilibrium yielding a global value more than 1000 times too high. The value for the Ca2+ binding step alone has now been determined to be K(D) approximately 10 nM. AKPC-Ca2+ was more resistant to inactivation by hydrogen peroxide (H(2)O(2)) and exhibited increased catalase activity. An analysis of the complex H(2)O(2) concentration dependent kinetics of Ca2+-free AKPC is presented.  相似文献   

15.
Oxygen transfer in broths of plant cells at high density   总被引:2,自引:0,他引:2  
The rheological properties of the culture broths of some plant cells (Cudrania tricuspidata, Vinca rosea, and Agrostemma githago) at high density (10-18 g dry wt/L) were measured, and oxygen transfer in the broths in various bioreactors was investigated. The rheological properties of the broths were dependent on the size, specific gravity, and concentration of the cell aggregates contained in the broths. The broths were non-Newtonian and pseudoplastic fluids. The flow behavior index n was fairly constant (0.53) and the consistency index K varied in proportion to the sixth-to-seventh power of the cell mass concentration M. The apparent viscosity mu(a) of the broths was in proportion to the 6.5th power of M. The oxygen transfer in the broths was discussed on the basis of the results obtained for suspensions of granulated agars (agar concentration, 5.8%) in water, which were similar to the broths in rheological properties. The volumetric oxygen transfer coefficient k(L)a in the broths was dependent on mu(a)(k(L)a proportional, variant mu(a) (-m)) and decreased greatly at a certain apparent viscosity, mu(ac). The values of m and mu(ac) were closely related to the aeration-agitation mechanisms of the bioreactors. The values of mu(ac) in aeration-agitation type bioreactors was larger than that in aeration-type bioreactors, whereas for m, the reverse was true.  相似文献   

16.
Mao Y  Wei W  He D  Nie L  Yao S 《Analytical biochemistry》2002,306(1):23-30
A new method for monitoring, in real time, the drug-binding process to protein with piezoelectric quartz crystal impedance (PQCI) is proposed. The method was used to monitor the binding process of berberine hydrochloride to bovine serum albumin (BSA). BSA was immobilized on the silver electrode surface of a piezoelectric quartz crystal and the optimized experimental conditions were established. The BSA-coated piezoelectric sensor was in contact with berberine solution. The time courses of the resonant frequency and equivalent circuit parameters of the sensor during the protein-drug binding were simultaneously obtained. On the basis of the analysis of the multidimensional information provided by PQCI, it was concluded that the observed frequency decrease was mainly ascribed to the mass increase of the sensor surface resulting from the binding. According to the frequency decrease with time, the kinetics of the binding process were quantitatively studied. A piezoelectric response model for the binding was theoretically derived. Fitting the experimental data to the model, the kinetic parameters, such as the binding and dissociation rate constants (k(1) and k(-1)) and the binding equilibrium constant (K(a)), were determined. The k(1), k(-1), and K(a) values obtained at 25 degrees C were 67.5 (+/-0.1) (mol liter(-1))(-1) s(-1), 1.7 (+/- 0.1) x 10(-3) s(-1), and 3.97 (+/- 0.06) x10(4) (mol liter(-1))(-1), respectively.  相似文献   

17.
The static light scattering and sedimentation equilibrium of solutions of Dextran 70 were measured as functions of concentration up to 100 g/L in pH 7.4 phosphate-buffered saline at temperatures between 5 and 37 °C. The concentration dependence of scattering intensity and the apparent molar mass obtained from sedimentation equilibrium were found to be nearly independent of temperature over this range to within the uncertainty of measurement. Global analysis of the concentration dependence of both properties yielded a reliable estimate of the concentration-dependent thermodynamic activity coefficient, a quantitative measure of the free energy of self-interaction. The self-interaction between Dextran molecules is compared with that of a globular protein (BSA) and a highly crosslinked polymer of similar molar mass (Ficoll 70). The observed concentration dependence of the free energy of Dextran self-interaction may be quantitatively accounted for by a semi-empirical model in which the polymer molecule is represented by a compressible sphere.  相似文献   

18.
Stewart RC  Jahreis K  Parkinson JS 《Biochemistry》2000,39(43):13157-13165
The histidine protein kinase CheA plays a central role in the bacterial chemotaxis signal transduction pathway. Autophosphorylated CheA passes its phosphoryl group to CheY very rapidly (k(cat) approximately 750 s(-)(1)). Phospho-CheY in turn influences the direction of flagellar rotation. The autophosphorylation site of CheA (His(48)) resides in its N-terminal P1 domain. The adjacent P2 domain provides a high-affinity binding site for CheY, which might facilitate the phosphotransfer reaction by tethering CheY in close proximity to the phosphodonor located in P1. To explore the contribution of P2 to the CheA --> CheY phosphotransfer reaction in the Escherichia coli chemotaxis system, we examined the transfer kinetics of a mutant CheA protein (CheADeltaP2) in which the 98 amino acid P2 domain had been replaced with an 11 amino acid linker. We used rapid-quench and stopped-flow fluorescence experiments to monitor phosphotransfer to CheY from phosphorylated wild-type CheA and from phosphorylated CheADeltaP2. The CheADeltaP2 reaction rates were significantly slower and the K(m) value was markedly higher than the corresponding values for wild-type CheA. These results indicate that binding of CheY to the P2 domain of CheA indeed contributes to the rapid kinetics of phosphotransfer. Although phosphotransfer was slower with CheADeltaP2 (k(cat)/K(m) approximately 1.5 x 10(6) M(-)(1) s(-)(1)) than with wild-type CheA (k(cat)/K(m) approximately 10(8) M(-)(1) s(-)(1)), it was still orders of magnitude faster than the kinetics of CheY phosphorylation by phosphoimidazole and other small molecule phosphodonors (k(cat)/K(m) approximately 5-50 M(-)(1) s(-)(1)). We conclude that the P1 domain of CheA also makes significant contributions to phosphotransfer rates in chemotactic signaling.  相似文献   

19.
The use of confocal scanning laser microscopy (CSLM) has recently been described for the visualization of intraparticle protein profiles during single-protein finite bath uptake experiments. By coupling of fluorescent molecules to proteins the penetration of porous media by labeled macromolecules could be detected by scanning single adsorbent particles for fluorescence emission after laser excitation. Thus the internal protein distribution profile, which is a central element in modeling of protein transport in porous adsorbents, became experimentally accessible. Results from the simultaneous visualization of two proteins by this technology are shown here. The use of two different fluorescent dyes for protein labeling and two independent detectors in the CSLM allowed for the first time ever the direct observation of a two-component diffusion process within a porous stationary phase. The finite bath uptake of human immunoglobulin G (hIgG) and bovine serum albumin (BSA) to two different ion exchange adsorbents (SP Sepharose Fast Flow and Source 30S) and to an affinity adsorbent (Protein A Sepharose) was measured using Cy5 and Oregon Green as labels. Single adsorbent particles were scanned for intensity distribution of fluorescence emission from the two fluorophors. The intraparticle profiles obtained from the confocal images were translated into a relative protein concentration thus allowing the calculation of protein uptake kinetics from direct measurement in the stationary phase. The confocal technique may prove to be a very powerful means of data generation for modeling of multi-component mass transfer phenomena in protein adsorption.  相似文献   

20.
Mass transfers of O(2), CO(2), and water vapor are among the key processes in the aerobic, autotrophic cultivation of moderate and extreme thermophiles. The dynamics and kinetics of these processes are, in addition to the obvious microbial kinetics, of crucial importance for the industrial desulfurization of high-pyritic coal by such thermophiles. To evaluate the role of the temperature on the gas mass transfer, k(L)a measurements have been used to supplement the existing published data. Oxygen mass transfer from gas (air) to liquid (5 mM H(2)SO(4) in water) phase as a function of the temperature has been studied in a laboratory-scale fermentor. At 15, 30, 45, and 70 degrees C, (k(L)a)(o) values (for oxygen) were determined under three different energy input conditions by the dynamic gassing in/out method. The (k(L)a)(o) was shown to increase under these conditions with increasing temperature, and straight lines were obtained when the logarithm of (k(L)a)(o) was plotted against the temperature. By multiplying the equilibrium concentration of O(2) in water with (k(L)a)(o) maximal, O(2) transfer capacities were calculated. It appeared that in finite of a decreased solubility of O(2) at elevated temperature in mechanically mixed fermentors the calculated transfer capacities showed only minor changes for the range between 15 and 70 degrees C. However, in an air-mixed fermentor the transfer capacity of O(2) decreased slowly but steadily.Carbon dioxide mass transfer was predicted by calculations on the basis of the data for oxygen transfer. The maximal CO(2) transfer capacity, calculated as the product of the equilibrium CO(2) concentration times (k(L)a)(c), decreased slowly as the temperature increased over the range 15-70 degrees C under all three energy input conditions. Subsequent process design calculations showed that for aerobic, autotrophic cultures, CO(2) limitation is more likely to occur than O(2) limitation.  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