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 78 毫秒
1.
Glucose uptake by Penicillium ochrochloron (formerly Penicillium simplicissimum) was studied from 0.01 to 400 mM glucose using chemostat culture and bioreactor batch culture. The characteristics of glucose uptake varied considerably with the conditions of growth, harvest and uptake assay. Glucose-limited grown mycelium showed one saturable transport system [K(S) below 0.01 mM; v(max) 1.1-1.2 mmol (g dry weight)(-1)h(-1)] plus a first order process (permeability P=1.2x10(-7)cm s(-1)). Ammonium-limited grown mycelium showed only one saturable transport system [K(S) 0.3-0.7 mM; v(max) 0.5-0.8 mmol (g dry weight)(-1)h(-1)]. During exponential growth at high glucose concentration (300-400 mM) a first order process was found with a P value of 5.6-9.3x10(-7)cm s(-1). After ammonium exhaustion a second first order phase showed a lower P value (6.1-9.3x10(-8)cm s(-1)). A similar change in permeability was also found after a re-evaluation of published data for Gibberella fujikuroi, Aspergillus niger, Aspergillus awamori and Saccharomycopsis lipolytica. For the first order processes simple diffusion was ruled out as a mechanism for glucose uptake. Glucose uptake by P. ochrochloron was controlled more strongly by metabolism than by transport and was not rate limiting for overflow metabolism.  相似文献   

2.
Intracellular diffusion of water   总被引:10,自引:0,他引:10  
Self-diffusion of cell water has been measured at diffusion times ranging from 0.3 ms to 1.0 s for human red cells, yeast, and brine shrimp using various pulsed gradient NMR methods. Intracellular diffusion coefficients and membrane permeabilities are calculated from these data with the aid of previous theoretical results for regularly spaced permeable planar barriers. The intracellular diffusion coefficients of water range from 1.2 X 10(-6) to 6 X 10(-6) cm2/s for the various samples. Outer-membrane permeabilities to water range from 0.0001 to 0.01 cm/s. The self-diffusion coefficient of lipid in a sample of human breast adipose tissue was found to be 1.5 X 10(-7) cm2/s.  相似文献   

3.
A sudden increase in the transmural pressure gradient across endothelial monolayers reduces hydraulic conductivity (L(p)), a phenomenon known as the sealing effect. To further characterize this endothelial adaptive response, we measured bovine aortic endothelial cell (BAEC) permeability to albumin and 70-kDa dextran, L(p), and the solvent-drag reflection coefficients (sigma) during the sealing process. The diffusional permeability coefficients for albumin (1.33 +/- 0.18 x 10(-6) cm/s) and dextran (0.60 +/- 0.16 x 10(-6) cm/s) were measured before pressure application. The effective permeabilities (measured when solvent drag contributes to solute transport) of albumin and dextran (P(ealb) and P(edex)) were measured after the application of a 10 cmH(2)O pressure gradient; during the first 2 h of pressure application, P(ealb), P(edex), and L(p) were significantly reduced by 2.0 +/- 0.3-, 2.1 +/- 0.3-, and 3.7 +/- 0.3-fold, respectively. Immunostaining of the tight junction (TJ) protein zonula occludens-1 (ZO-1) was significantly increased at cell-cell contacts after the application of transmural pressure. Cytochalasin D treatment significantly elevated transport but did not inhibit the adaptive response, whereas colchicine treatment had no effect on diffusive permeability but inhibited the adaptive response. Neither cytoskeletal inhibitor altered sigma despite significantly elevating both L(p) and effective permeability. Our data suggest that BAECs actively adapt to elevated transmural pressure by mobilizing ZO-1 to intercellular junctions via microtubules. A mechanical (passive) component of the sealing effect appears to reduce the size of a small pore system that allows the transport of water but not dextran or albumin. Furthermore, the structures of the TJ determine transport rates but do not define the selectivity of the monolayer to solutes (sigma).  相似文献   

4.
Total internal reflection (TIR) microfluorimetry was established as a method to measure continuously the volume of adherent cells and applied to measure membrane permeabilities in cells transfected with water channel homologs. Cytosol was labeled with the membrane-impermeant fluorophore calcein. Fluorescence was excited by the TIR evanescent field in a thin section of cytosol (approximately 150 nm) adjacent to the cell-substrate interface. Because cytosolic fluorophore number per cell remains constant, the TIR fluorescence signal should be inversely related to cell volume. For small volume changes in Sf-9 and LLC-PK1 cells, relative TIR fluorescence was nearly equal to inverse relative cell volume; deviations from the ideal were modeled theoretically. To measure plasma membrane osmotic water permeability, Pf, the time course of osmotically induced cell volume change was inferred from the TIR fluorescence signal. LLC-PK1 cells expressing the CHIP28 water channel had an HgCl2-sensitive, threefold increase in Pf compared to nontransfected cells (Pf = 0.0043 cm/s at 10 degrees C). Solute permeability was measured from the TIR fluorescence time course in response to solute gradients. Glycerol permeability in Sf-9 cells expressing the water channel homolog GLIP was (1.3 +/- 0.2) x 10(-5) cm/s (22 degrees C), greater than that of (0.36 +/- 0.04) x 10(-5) cm/s (n = 4, p < 0.05) for control cells, indicating functional expression of GLIP. Water and urea permeabilities were similar in GLIP-expressing and control cells. The TIR method should be applicable to the study of water and solute permeabilities and cell volume regulation in cells of arbitrary shape and size.  相似文献   

5.
A procedure for high-yield spore production by Bacillus subtilis   总被引:1,自引:0,他引:1  
Bacillus subtilis spores have a number of potential applications, which include their use as probiotics and competitive exclusion agents to control zoonotic pathogens in animal production. The effect of cultivation conditions on Bacillus subtilis growth and sporulation was investigated in batch bioreactions performed at a 2-L scale. Studies of the cultivation conditions (pH, dissolved oxygen concentration, and media composition) led to an increase of the maximum concentration of vegetative cell from 2.6 x 10(9) to 2.2 x 10(10) cells mL(-)(1) and the spore concentration from 4.2 x 10(8) to 5.6 x 10(9) spores mL(-)(1). A fed-batch bioprocess was developed with the addition of a nutrient feeding solution using an exponential feeding profile obtained from the mass balance equations. Using the developed feeding profile, starting at the middle of the exponential growth phase and finishing in the late exponential phase, an increase of the maximum vegetative cell concentration and spore concentration up to 3.6 x 10(10) cells mL(-)(1) and 7.4 x 10(9) spores mL(-)(1), respectively, was obtained. Using the developed fed-batch bioreaction a 14-fold increase in the concentration of the vegetative cells was achieved. Moreover, the efficiency of sporulation under fed-batch bioreaction was 21%, which permitted a 19-fold increase in the final spore concentration, to a final value of 7.4 x 10(9) spores mL(-)(1). This represents a 3-fold increase relative to the highest reported value for Bacillus subtilis spore production.  相似文献   

6.
Pulsed high-field gradient in vivo NMR spectroscopy was used to measure diffusional water permeability in cell suspensions of the Gram-positive bacterium Corynebacterium glutamicum. Two different regions of H2O mobility were detected. One was characterized by the apparent coefficient of self-diffusion, D(1 app) = (4.6-12.7)x10(-8) cm(2) s(-1), depending on the observation time t. The other region was characterized by D(2) = 1.4x10(-5) cm(2) s(-1). The value of D(2) was similar to the diffusion coefficient of H2O in free water and in extracellular biological fluids. Restricted diffusion could be demonstrated for the slower process (D(1)). It was attributed to the cytoplasm of the cells. The membrane permeability, P(d H2O), for C. glutamicum was (4.8+/-0.4)x10(-3) cm s(-1). It compared favorably with values reported for human erythrocytes and was higher by a factor of about 100 compared to the diffusional permeability for ethanol, P(d ethanol), in Zymomonas mobilis. Addition of HgCl2, a water channel inhibitor in eukaryotes, decreased P(d H2O) in C. glutamicum by a factor of approximately 8. To our knowledge, these are the first functional studies of water transport in prokaryotes that yielded quantitative data, viz., transmembrane water permeability expressed through D(H2O) and P(d H2O).  相似文献   

7.
We reported increased water permeability and a low urea reflection coefficient in Xenopus oocytes expressing urea transporter UT-B (former name UT3), suggesting that water and urea share a common aqueous pathway (Yang, B., and Verkman, A. S. (1998) J. Biol. Chem. 273, 9369-9372). Although increased water permeability was confirmed in the Xenopus oocyte expression system, it has been argued (Sidoux-Walter, F., Lucien, N., Olives, B., Gobin, R., Rousselet, G., Kamsteeg, E. J., Ripoche, P., Deen, P. M., Cartron, J. P., and Bailly, P. (1999) J. Biol. Chem. 274, 30228-30235) that UT-B does not transport water when expressed at normal levels in mammalian cells such as erythrocytes. To quantify UT-B-mediated water transport, we generated double knockout mice lacking UT-B and the major erythrocyte water channel, aquaporin-1 (AQP1). The mice had reduced survival, retarded growth, and defective urinary concentrating ability. However, erythrocyte size and morphology were not affected. Stopped-flow light scattering measurements indicated erythrocyte osmotic water permeabilities (in cm/s x 0.01, 10 degrees C): 2.1 +/- 0.2 (wild-type mice), 2.1 +/- 0.05 (UT-B null), 0.19 +/- 0.02 (AQP1 null), and 0.045 +/- 0.009 (AQP1/UT-B null). The low water permeability found in AQP1/UT-B null erythrocytes was also seen after HgCl(2) treatment of UT-B null erythrocytes or phloretin treatment of AQP1 null erythrocytes. The apparent activation energy for UT-B-mediated water transport was low, <2 kcal/mol. Estimating 14,000 UT-B molecules per mouse erythrocyte, the UT-B-dependent P(f) of 0.15 x 10(-4) cm/s indicated a substantial single channel water permeability of UT-B of 7.5 x 10(-14) cm(3)/s, similar to that of AQP1. These results provide direct functional evidence for UT-B-facilitated water transport in erythrocytes and suggest that urea traverses an aqueous pore in the UT-B protein.  相似文献   

8.
Electron particle sizing (Coulter counter) was used to measure cell and protoplast volumes of Saccharomyces cerevisiae grown under different conditions designed to increase its cryotolerance. Membrane water permeabilities were estimated from those measurements. A relationship was obtained between the lower water permeability of yeast grown under microaerobic batch conditions and its weaker cryotolerance in water (cooling rate of 39·6°C/min), as compared to fed-batch cells. For the latter, cell water permeability was not related to the observed differences in survival for frozen-thawed cells grown under strong or partial (with temporary limitation of dissolved oxygen in growth media) aerobic conditions.  相似文献   

9.
An 11-L helical ribbon impeller (HRI) bioreactor was tested for the culture of Spodoptera frugiperda (Sf-9) cells. This impeller and surface baffling ensured homogeneous mixing and high oxygen transfer through surface aeration and surface-induced babble generation. Serum-supplemented and serum-free cultures, using TNMFH and IPL/41 media, respectively, grew a similar specific growth rates(0.031 and 0.028 h(-1)) to maximum cell densities of 5.5 x 10(6)-6.0 x 10(6) cells. mL(-1) with viability exceeding 98% during exponential growth phase. Growth limitation coincided with glucose and glutamine depletion and production of significant amounts of alanine. The bioreactor was further tested under more stringent conditions by infecting a serum-free medium culture with a recombinant baculovirus. Heterologous protein production of approximately 35 mug per 10(6) cells was comparable to yields obtained in serum-free cultures grown in spinner flasks and petri dishes. Average specific oxygen up-take and carbon dioxide production rates of the serum-free culture prior to infection as measured by on-line mass spectroscopy were 0.20 mumol O(2)mu.(10(6) cells)(-1) h(-1) and 0.22 mumol CO(2) . (10(6) cells)(-1)h(-1) and increased by 30-40% during infection. Therefore, the mixing and oxygenation conditions of this bioreactor were suitable for insect cell culture and recombinant protein production, with limitation being mainly attributed to nutrient depletion and toxic by-product generation.  相似文献   

10.
To define aspects of lipid composition and bilayer asymmetry critical to barrier function, we examined the permeabilities of liposomes that model individual leaflets of the apical membrane of a barrier epithelium, Madin-Darby canine kidney type 1 cells. Using published lipid compositions we prepared exofacial liposomes containing phosphatidylcholine, sphingomyelin, glycosphingolipids, and cholesterol; and cytoplasmic liposomes containing phosphatidylethanolamine, phosphatidylserine, and cholesterol. The osmotic permeability of cytoplasmic liposomes to water (P(f)), solutes, and NH(3) was 18-90-fold higher than for the exofacial liposomes (P(f(ex)) = 2.4 +/- 0.4 x 10(-4) cm/s, P(f(cy)) = 4.4 +/- 0.3 x 10(-3) cm/s; P(glycerol(ex)) = 2.5 +/- 0.3 x 10(-8) cm/s, P(glycerol(cy)) = 2.2 +/- 0.02 x 10(-6) cm/s; P(NH3(ex)) = 0. 13 +/- 0.4 x 10(-4) cm/s, P(NH3(cy)) = 7.9 +/- 1.0 x 10(-3) cm/s). By contrast, the apparent proton permeability of exofacial liposomes was 4-fold higher than cytoplasmic liposomes (P(H+(ex)) = 1.1 +/- 0. 1 x 10(-2) cm/s, P(H+(cy)) = 2.7 +/- 0.6 x 10(-3) cm/s). By adding single leaflet permeabilities, we calculated a theoretical P(f) for a Madin-Darby canine kidney apical membrane of 4.6 x 10(-4) cm/s, which compares favorably with experimentally determined values. In exofacial liposomes lacking glycosphingolipids or sphingomyelin, permeabilities were 2-7-fold higher, indicating that both species play a role in barrier function. Removal of cholesterol resulted in 40-280-fold increases in permeability. We conclude: 1) that we have reconstituted the biophysical properties of a barrier membrane, 2) that the barrier resides in the exofacial leaflet, 3) that both sphingomyelin and glycosphingolipids play a role in reducing membrane permeability but that there is an absolute requirement for cholesterol to mediate this effect, 4) that these results further validate the hypothesis that each leaflet offers an independent resistance to permeation, and 5) that proton permeation was enhanced by sphingolipid/cholesterol interactions.  相似文献   

11.
A new stop-flow technique was employed to quantify the impact of internal unstirred layers on the measurement of the solute permeability coefficient (P(s)) across the plasma membrane of internodes of the giant-celled alga Chara corallina using a cell pressure probe. During permeation experiments with rapidly permeating solutes (acetone, 2-propanol, and dimethylformamide), the solute concentration inside the cell was estimated and the external medium was adjusted to stop solute transport across the membrane, after which responses in turgor were measured. This allowed estimation of the solute concentration right at the membrane. Stop-flow experiments were also simulated with a computer. Both the stop-flow experiments and simulations provided quantitative data about internal concentration gradients and the contribution of unstirred layers to overall measured values of P(meas)(s) for the three solutes. The stop-flow experimental results agreed with stop-flow simulations assuming that solutes diffused into a completely stagnant cell interior. The effects of internal unstirred layers on the underestimation of membrane P(s) declined with decreasing P(s). They were no bigger than 37% in the presence of the most rapidly permeating solute, acetone (P(meas)(s) =4.2 x 10(-6) m s(-1)), and 14% for the less rapidly permeating dimethylformamide (P(meas)(s) =1.6x10(-6) m s(-1)). It is concluded that, even in the case of rapidly permeating solutes such as isotopic water and, even when making pessimistic assumptions about the internal mixing of solutes, an upper limit for the underestimation of P(s) due to internal unstirred layers was 37%. The data are discussed in terms of recent theoretical estimates of the effect of internal unstirred layers and in terms of some recent criticism of cell pressure probe measurements of water and solute transport coefficients. The current stop-flow data are in line with earlier estimations of the role of unstirred layers in the literature on cell water relations.  相似文献   

12.
The purpose of this study was to develop a cell culture process in a bioreactor for the production of a viral insecticide for the spruce budworm, Choristoneura fumiferana . Several cell lines were tested for their growth in serum-free medium suspension cultures. One cell line, CF-124T-2C1 (CF-2C1), was successfully adapted to grow in suspension cultures in SFM. Serum-free Ex-Cell 405 medium produced a much higher cell density (6.3 x 10 6 cells ml -1 ) than the Grace's medium supplemented with 10% fetal bovine serum (2.5 x 10 6 cells ml -1 ). Also, a higher yield of virus was obtained in the former medium. Ex-Cell 405, was used to study the growth of CF-2C1 cells and the production of C. fumiferana nucleopolyhedrovirus (CfMNPV) in a 3 l bioreactor. Under these conditions, a specific growth rate ( μ) of 0.027 h -1 was obtained during the exponential growth phase, and the specific carbon dioxide evolution rate, as determined by on-line measurement, was 0.9 x 10 -16 mol cell -1 s -1 and 1.78 x 10 -16 mol cell -1 s -1 during growth and infection phases, respectively. Virus production in bioreactor cultures infected at 1.3 x 10 6 cells ml -1 was consistently lower than that obtained in Erlenmeyer shake flasks. Only 26% of the cells were infected in the bioreactor compared to 44% in the shake flasks. However, a higher yield of occluded virus was obtained in the bioreactor cultures than in shake flasks. The production of occlusion bodies (OB) achieved in bioreactor cultures was 2 x 10 6 OB ml -1 .  相似文献   

13.
Members of the major intrinsic protein (MIP) family, described in plants as water-selective channels (aquaporins), can also transport small neutral solutes in other organisms. In the present work, we characterize the permeability of plant vacuolar membrane (tonoplast; TP) and plasma membrane (PM) to non-electrolytes and evaluate the contribution of MIP homologues to such transport. PM and TP vesicles were purified from tobacco suspension cells by free-flow electrophoresis, and membrane permeabilities for a wide range of neutral solutes including urea, polyols of different molecular size, and amino acids were investigated by stopped-flow spectrofluorimetry. For all solutes tested, TP vesicles were found to be more permeable than their PM counterparts, with for instance urea permeabilities from influx experiments of 74.9 +/- 9.6 x 10(-6) and 1.0 +/- 0.3 x 10(-6) cm sec-1, respectively. Glycerol and urea transport in TP vesicles exhibited features of a facilitated diffusion process. This and the high channel-mediated permeability of the same TP vesicles to water suggested a common role for MIP proteins in water and solute transport. A cDNA encoding a novel tonoplast intrinsic protein (TIP) homologue named Nicotiana tabacum TIPa (Nt-TIPa) was isolated from tobacco cells. Immunodetection of Nt-TIPa in purified membrane fractions confirmed that the protein is localized in the TP. Functional expression of Nt-TIPa in Xenopus oocytes showed this protein to be permeable to water and solutes such as urea and glycerol. These features could account for the transport selectivity profile determined in purified TP vesicles. These results support the idea that plant aquaporins have a dual function in water and solute transport. Because Nt-TIPa diverges in sequence from solute permeable aquaporins characterized in other organisms, its identification also provides a novel tool for investigating the molecular determinants of aquaporin transport selectivity.  相似文献   

14.
To understand how plasma membranes may limit water flux, we have modeled the apical membrane of MDCK type 1 cells. Previous experiments demonstrated that liposomes designed to mimic the inner and outer leaflet of this membrane exhibited 18-fold lower water permeation for outer leaflet lipids than inner leaflet lipids (Hill, W.G., and M.L. Zeidel. 2000. J. Biol. Chem. 275:30176-30185), confirming that the outer leaflet is the primary barrier to permeation. If leaflets in a bilayer resist permeation independently, the following equation estimates single leaflet permeabilities: 1/P(AB) = 1/P(A) + 1/P(B) (Eq. l), where P(AB) is the permeability of a bilayer composed of leaflets A and B, P(A) is the permeability of leaflet A, and P(B) is the permeability of leaflet B. Using for the MDCK leaflet-specific liposomes gives an estimated value for the osmotic water permeability (P(f)) of 4.6 x 10(-4) cm/s (at 25 degrees C) that correlated well with experimentally measured values in intact cells. We have now constructed both symmetric and asymmetric planar lipid bilayers that model the MDCK apical membrane. Water permeability across these bilayers was monitored in the immediate membrane vicinity using a Na+-sensitive scanning microelectrode and an osmotic gradient induced by addition of urea. The near-membrane concentration distribution of solute was used to calculate the velocity of water flow (Pohl, P., S.M. Saparov, and Y.N. Antonenko. 1997. Biophys. J. 72:1711-1718). At 36 degrees C, P(f) was 3.44 +/- 0.35 x 10(-3) cm/s for symmetrical inner leaflet membranes and 3.40 +/- 0.34 x 10(-4) cm/s for symmetrical exofacial membranes. From, the estimated permeability of an asymmetric membrane is 6.2 x 10(-4) cm/s. Water permeability measured for the asymmetric planar bilayer was 6.7 +/- 0.7 x 10(-4) cm/s, which is within 10% of the calculated value. Direct experimental measurement of P(f) for an asymmetric planar membrane confirms that leaflets in a bilayer offer independent and additive resistances to water permeation and validates the use of.  相似文献   

15.
Permeation of glycerol and propane-1,2-diol into human platelets   总被引:3,自引:2,他引:1  
F G Arnaud  D E Pegg 《Cryobiology》1990,27(2):107-118
The permeability of human platelets to glycerol (GLY) and propane-1,2-diol (propylene glycol, PG) has been determined by measuring the time course of their change in volume following abrupt immersion in solutions of these solutes. A simple light-scattering method, and its calibration to measure mean platelet volume is described. The data are analyzed by means of the Kedem-Katchalsky (K-K) equations, modified to take into account the nonideal behavior of both intracellular and extracellular solutes. The values of the K-K parameters at 2, 21, and 37 degrees C, respectively, were as follows: the hydraulic conductivities (Lp) were 1 x 10(-7), 7 x 10(-7) and 3 x 10(-6) cm.sec-1.atm-1; the solute permeabilities for PG (omega RTPG) were 1.9 x 10(-6), 2.8 x 10(-5), and 1.3 x 10(-4) cm.sec-1; the solute permeabilities for GLY (omega RTGLY), at 21 and 37 degrees C only, were 2.6 x 10(-7) and 1.4 x 10(-6) cm.sec-1. The reflection coefficient (sigma) was 1 throughout. The relevant activation energies were -Lp, 16.5 kcal.mol-1; omega RTPG, 20.5 kcal.mol-1; and omega RTGLY, 17.9 kcal.mol-1. The use of these data is illustrated by computing schedules for the addition and removal of GLY and PG so that the amplitudes of changes in platelet volume are held within predetermined limits.  相似文献   

16.
The effects of pH strategy on endo- and exo-metabolome profiling of beta-lactamase-producing Bacillus licheniformis were investigated at controlled-pH (pH(C) = 6.5, 6.75, 7.0, 7.25, 7.5) and uncontrolled-pH (pH(UC) = 7.5) values using a glucose-based defined medium. The cell concentration profiles were not affected by the pH considerably within the investigated range. The highest enzyme activities were obtained as A = 54 U cm(-)(3) at pH(C) = 6.75 among the controlled-pH operations and as A = 57 U cm(-3) at the uncontrolled-pH pH(UC) = 7.5. At all conditions, oxygen transfer resistances were more effective, whereas the limitation increased in the beta-lactamase production phase. Total intracellular amino acid concentrations ranged between 0.142 and 6.766 kg m(-3) (0.0058-0.277 g g(cell)(-1)), and their concentrations in terms of kg m(-3) were, at most, 580-fold higher than the extracellular concentrations. Methionine/cysteine concentrations were generally higher than the other intracellular amino acids, whereas asparagine concentration was the highest in the fermentation broth. From Na(+), K(+), and H(+) ion profiles, Na(+)-K(+) antiport and Na(+)-H(+) symport were found to be present within the system, and a correlation was found between organic acid transport and Na(+)-H(+) symport. Intracellular organic acid concentrations in terms of kg m(-3) were, at most, 20-fold higher than that of the extracellular, and with the increase in pH, extracellular acetic acid concentration increased and lactic acid concentration decreased. Average permeability coefficient values of organic acids were found to be in the range from 4.10 x 10(-7) to 4.32 x 10(-6) cm s(-1) for the growth phase (0 < t < 6 h) and decreased at least 3-fold in the beta-lactamase production phase (8 < t < 15 h), indicating the considerable structural change of the lipid membrane during the fermentation.  相似文献   

17.
Chemical permeabilization has been widely studied for the release useful metabolites from many types of plant cells and tissues. In this study, the effect of 0-30% (v/v) of aqueous methanol solutions were used to permeabilize soybeans for the release of two isoflavonoids: daidzein and genistein. The release of these metabolites increases with increasing methanol concentrations. The amounts of daidzein and genistein released can increase up to 40- and 86-fold, respectively, when incubated in a 30% (v/v) methanol solution for 24 h compared with those incubated with water only. The effect of methanol on the release rates is primarily due to an increase in solubility of the stored daidzein and genistein (14- to 18-fold) inside the seeds, thus maximizing the concentration gradients for metabolite release. However, the viability of the seeds dropped with increase in methanol concentrations and the incubation time. The viability of soybeans (indicated by their ability to germinate) after permeabilization treatment with 0-20% (v/v) methanol solutions was maintained above 80% throughout the 24 h, whereas no seeds were found to be viable when 30% (v/v) methanol solution was used. The permeability coefficients (P) of daidzein and genistein were found to increase as the methanol concentration used was increased. These P values were estimated to range from 1.1 x 10(-)(9) to 1.9 x 10(-)(8) m/s and 1.0 x 10(-)(9) to 1.7 x 10(-)(8) m/s, respectively. The increase in P can be attributed primarily to an increase in the partition coefficient of the metabolites in the soybean seedcoats. An empirical correlation is proposed in which the log P values are described as a function of the metabolite molecular weights and the partition coefficients of the metabolites between octanol and water, K(oct/water), which was modified to include the effect of methanol present. Knowledge obtained from this study will help provide useful selection criteria for chemical permeabilization of plant tissues, such as seeds, with minimal loss in their viability.  相似文献   

18.
Y. Konishi  S. Asai    N. Yoshida 《Applied microbiology》1995,61(10):3617-3622
The growth kinetics of Thiobacillus thiooxidans on elemental sulfur in batch cultures at 30(deg)C and pH 1.5 was studied by measuring the time courses of the concentration of adsorbed cells on sulfur, the concentration of free cells suspended in liquid medium, and the amount of sulfur oxidized. As the elemental sulfur was oxidized to sulfate ions, the surface concentration of adsorbed cells per unit mass of sulfur approached a maximum value (maximum adsorption capacity of sulfur particles) whereas the concentration of free cells continued to increase with time. There was a close relationship between the concentrations of free and adsorbed cells during the microbial sulfur oxidation, and the two cell concentrations were well correlated by the Langmuir isotherm with adsorption equilibrium constant K(infA) and maximum adsorption capacity X(infAm) of 2.10 x 10(sup-9) ml per cell and 4.57 x 10(sup10) cells per g, respectively. The total concentration of free and adsorbed cells increased in parallel with the amount of sulfate formed. The total growth on elemental sulfur gave a characteristic growth curve in which a linear-growth phase followed the period of an initial exponential phase. The batch rate data collected under a wide variety of inoculum levels (about 10(sup5) to 10(sup8) cells per ml) were consistent with a kinetic model assuming that the growth rate of adsorbed bacteria is proportional to the product of the concentration, X(infA), of adsorbed cells and the fraction, (theta)(infV), of adsorption sites unoccupied by cells. The kinetic and stoichiometric parameters appearing in the model were estimated from the experimental data, and the specific growth rate, (mu)(infA), and growth yield, Y(infA), were 2.58 day(sup-1) and 2.05 x 10(sup11) cells per g, respectively. The proposed model and the parameter values allowed us to predict quantitatively the surface attachment of T. thiooxidans cells on elemental sulfur and the bacterial growth in both initial exponential and subsequent linear phases. The transition from exponential to linear growth was a result of two competing factors: an increase in the adsorbed-cell concentration, X(infA), permitted a decrease in the unoccupied-site fraction, (theta)(infV).  相似文献   

19.
We hypothesized that a high circulating free fatty acid (FFA) concentration is involved in the pathogenesis of hyposomatotropism associated with obesity. To evaluate this hypothesis, 10 healthy premenopausal women (body mass index 33.8 +/- 1.0 kg/m(2)) were studied in the follicular phase of their menstrual cycle at two occasions with a time interval of at least 8 wk, where body weight remained stable. Subjects were randomly assigned to treatment with either acipimox (an inhibitor of lipolysis, 250 mg orally 4 times daily) or placebo in a double-blind crossover design, starting 1 day before admission until the end of the blood sampling period. Blood samples were taken during 24 h with a sampling interval of 10 min for assessment of growth hormone (GH) concentrations, and GH secretion was estimated by deconvolution analysis. Identical methodology was used to study GH secretion in a historical control group of age-matched normal weight women. GH secretion was clearly blunted in obese women (total daily release 66 +/- 10 vs. lean controls: 201 +/- 23 mU x l(Vd)(-1) x 24 h(-1), P = 0.005, where l(Vd) is lite of distribution volume). Acipimox considerably enhanced total (113 +/- 50 vs. 66 +/- 10 mU x l(Vd)(-1) x 24 h(-1), P = 0.02) and pulsatile GH secretion (109 +/- 49 vs. 62 +/- 30 mU x l(Vd)(-1) x 24 h(-1), P = 0.02), but GH output remained lower compared with lean controls. Further analysis did not show any relationship between the effects of acipimox on GH secretion and regional body fat distribution. In conclusion, acipimox unleashes spontaneous GH secretion in obese women. It specifically enhances GH secretory burst mass. This might mean that lowering of systemic FFA concentrations by acipimox modulates neuroendocrine mechanisms that orchestrate the activity of the somatotropic ensemble.  相似文献   

20.
Recent reports suggest the expression of aquaporin (AQP)-type water channels in mitochondria from liver (AQP8) (Calamita, G., Ferri, D., Gena, P., Liquori, G. E., Cavalier, A., Thomas, D., and Svelto, M. (2005) J. Biol. Chem. 280, 17149-17153) and brain (AQP9) (Amiry-Moghaddam, M., Lindland, H., Zelenin, S., Roberg, B. A., Gundersen, B. B., Petersen, P., Rinvik, E., Torgner, I. A., and Ottersen, O. P. (2005) FASEB J. 19, 1459-1467), where they were speculated to be involved in metabolism, apoptosis, and Parkinson disease. Here, we systematically examined the functional consequence of AQP expression in mitochondria by measurement of water and glycerol permeabilities in mitochondrial membrane preparations from rat brain, liver, and kidney and from wild-type versus knock-out mice deficient in AQPs -1, -4, or -8. Osmotic water permeability, measured by stopped-flow light scattering, was similar in all mitochondrial preparations, with a permeability coefficient P(f) approximately 0.009 cm/s. Glycerol permeability was also similar ( approximately 5 x 10(-6) cm/s) in the various preparations. HgCl(2) slowed osmotic equilibration comparably in mitochondria from wild-type and AQP-deficient mice, although the slowing was explained by altered mitochondrial size rather than reduced P(f). Immunoblot analysis of mouse liver mitochondria failed to detect AQP8 expression, with liver homogenates from wild-type/AQP8 null mice as positive/negative controls. Our results provide evidence against functionally significant AQP expression in mitochondria, which is consistent with the high mitochondrial surface-to-volume ratio producing millisecond osmotic equilibration, even when intrinsic membrane water permeability is not high.  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