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.
By use of the microfluorimetric technique it is possible to study the diffusion of the fluorochrome di-dansylcystine (DDC) within isolated frog rod outer segments (ros) which are immobilysed in agarose gel. For this purpose, by a short hypotonic shock a leak is applied to one end of the ros. By this open end the DDC enters the rod and migrates through the whole outer segment. Following the propagation of the fluorescence boundary with time the cytoplasmatic diffusion constant can be determined if a Chromatographic model is used to allow for the considerable binding of DDC to the inner membrane surface. With a binding constant K=5·10–4 cm the cytoplasmatic diffusion constant was found to be D= 1.3·10–6cm2/s whereas D g =2·10–6cm2/s and D r =3.5·10–6cm2/s were found in agarose gel or ringer solution, respectively. Using the mobility reduction factor given by D/D r 0.4 to calculate the cytoplasmatic conductivity an inner resistance per length of 1.7 M / could be calculated for a frog rod which is in good agreement with corresponding data obtained from electrophysiological measurements.  相似文献   

2.
Heisenberg spin exchange rates and dipole-dipole spin lattice relaxation rates for deuterated 14N- and 15N-spin labels bound selectively to the histidine His15 and to the lysines Lys13, 96, 97 of the lysozyme molecule have been determined with the aid of electron spin resonance spectroscopy. The results can be interpreted in terms of a two dimensional translational diffusion of the nitroxide tips of the spin labels along the protein surface within restricted surface areas. The spin labels are regarded as models for long amino acid side chains and as probes for the dynamics of protein and water in the vicinity of the protein surface. The translational diffusion coefficient DPII is reduced by a factor of between six and thirty compared to the value of D found for the spin labels in bulk water, its value for T = 295 K is given by (1.3±0.6)·10–10m2s–1 D (2.4±0.3) 10–11 m2s–1. Offprint requests to: H.-J. Steinhoff  相似文献   

3.
We have used the fluorescence recovery after photobleaching technique to study the translational diffusion, in L phase multibilayers of 1-palmitoyl-2-oleoylphosphatidylcholine (POPC), of fluorescent derivatives of 1-palmitoyl-2-oleoylphosphatidylethanolamine (NBD-POPE) and a membrane-spanning phosphatidylethanolamine (NBD-MSPE). The latter derivative was prepared from a membrane-spanning glycerol-dialkyl-glycerol tetraether lipid isolated from the thermophilic and acidophilic archaebacterium Sulfolobus solfataricus. The translational diffusion was examined between about 15° and 45°C. It is shown that over this temperature range the translational diffusion coefficient for NBD-MSPE is 2/3 that for NBD-POPE which spans only one monolayer of the bilayer. The result is interpreted in terms of existing models for translational diffusion in lipid membranes.Abbreviations D t translational diffusion coefficient - FRAP fluorescence recovery after photobleaching - MSPE a membrane-spanning phosphatidylethanolamine derived from a glycerol-dialkyl-glycerol tetraether lipid isolated from Sulfolobus solfataricus - NBD 4-nitrobenz-2-oxa-1,3-diazolyl - PE phosphatidylethanolamine - POPC 1-palmitoyl-2-oleoylphosphatidylcholine - POPE 1-palmitoyl-2-oleoylphosphatidylethanolamine  相似文献   

4.
Molecular diffusion of solutes, like sucrose in the xanthan gum fermentation, is important in order to understand the complex behavior of mass transfer mechanisms during the process. This work was focused to determine the diffusion coefficient of sucrose, a carbon source for xanthan production, using similar sucrose and xanthan concentrations to those occurring in a typical fermentation. The diaphragm cell method was used in experimental determinations. The data showed that diffusion coefficient of sucrose significantly decreases when xanthan gum concentration increases. Theoretical and semiempirical models were used to predict sucrose diffusivity in xanthan solutions. Molecular properties and rheological behavior of the system were considered in the modeling. The models tested fitted well the behavior of experimental data and that reported for oxygen in the same system.List of Symbols A constant in eq. (5) - C pg cm–3 polymer concentration - D cm2 s–1 diffusivity - D ABcm2 s–1 diffusivity of A through liquid solvent - D APcm2 s–1 diffusivity of A in polymer solution - D AWcm2 s–1 diffusivity of A in water - D Pcm2 s–1 diffusivity of polymer in liquid solvent - E D gradient of the activation energy for diffusion - H P hydratation factor of the polymer in water (g of bound water/g of polymer) - K dyn sn cm–2 consistency index - K 1 constant in eq. (5) - K P overall binding coefficient [g of bound solute/cm3 of solution]/[g of free solute/cm3 of polymer free solution] - n flow behavior index - M Bg g mol–1 molucular weight of liquid solvent - M Pg g mol–1 molecular weight of the polymer - M Sg g mol–1 Molecular weight of polymer solution (= M BXB+MPXP) - R cm3 atm g mol–1 K–1 ideal gas law constant - T K absolute temperature - V Bcm3 g mol–1 molar volume of liquid solvent - V Pcm3 g mol–1 molar volume of polymer - V Scm3 g mol–1 molar volume of polymer solution - X B solvent molar fraction - X P polymer molar fraction - polymer blockage shape factor - P volume fraction of polymer in polymer solution - g cm–1 s–1 viscosity - ag cm–1 s–1 apparent viscosity of the polymer solution - icm3 g–1 intrinsic viscosity - 0 g cm–1 s–1 solvent viscosity - Pg cm–1 s–1 polymer solution viscosity - R relative viscosity (= / 0) - =0 g cm–1 s–1 viscosity of polymer solution obtained at zero shear rate - 0 g cm–3 water density  相似文献   

5.
Summary Transport by an epithelium, possessing an accumulating, saturable transport system in the apical membrane as well as a finite Fick permeability to the transported solute, was considered in the steady state in the case of zerocis concentration, and in the presence of a peripheral diffusion resistance in a layer apposing thecis face of the tissue (unstirred solution or structural coating). Under suitable conditions, the combination of peripheral diffusion resistance and accumulating epithelial transport may lead to recycling of solute at thecis face of the epithelium. This causes a decrease of the effective permeability to diffusionaltrans-cis flow across the tissue. The phenomenon is discussed in terms of epidermald-glucose transport by the integument of aquatic animals with a collagenous cuticle, such as the seawater-acclimated polychaete wormNereis diversicolor. The recycling phenomenon may be of significance to other epithelia with the function of maintaining large concentration gradients of permeating substances.List of Symbols and Fixed Parameter Values C m Bulk medium solute concentration,cis face of epidermisC m=0 mol cm–3 - C i Concentration of solute at interface between cuticle and unstirred medium (mol cm–3) - C s Concentration of solute atcis face of apical epidermal membrane (mol cm–3) - C e Concentration of solute in extracellular fluid,trans-side of epidermisC e=1.0×10–6 mol cm–3 - D m Diffusion coefficient of solute in outside mediumD m=6.7×10–6 cm2 sec–1 - D c Diffusion coefficient of solute in cuticleD c=7.4×10–9 cm2 sec–1 - m Operative thickness of unstirred medium layer - c Thickness of cuticle - J Steady-state net flux of solute through cuticle or unstirred layer (flux is positive indirectioncis-trans) (mol cm–2 sec–1) - J i max Maximal influx through saturable transport system in apical membraneJ i max =2.0×10–12 mol cm–2 sec–1 - K t Transport constant, saturable systemK t=1.0×10–7 mol cm–3 - P Epithelial permeability (cm sec–1)  相似文献   

6.
Summary The leakage of Serratia marcescens cells immobilised in Ca-alginate beads was quantified and the experimental values fitted to the pore diffusion model in order to obtain the biomass diffusion coefficient (D x). This coefficient was then determined, resulting a value of 0.45 10–7 cm2 s–1 at the beginning of the fermentation process. The variation of this coefficient, together with the porosity of the particles (i), were also determined all along the process.  相似文献   

7.
The translational diffusion of pyrene, pyrene butyric acid and pyrene decanoic acid has been determined in phosphatidylcholine bilayers of different chain length and under pressure up to 200 bars. In the liquid crystalline phase and at a given temperature the diffusion decreases with increasing chain length. At a constant reduced temperature, T red (about 10 K above the transition temperature), long chain lipids exhibit the fastest diffusion which is in disagreement with hydrodynamic models but favours free volume models for diffusion in lipid bilayers. The volume of activation, V act, calculated from the decrease of the diffusion coefficient with pressure, ln D/P, depends on lipid chain length. V act decreases with decreasing lipid chain length at a given temperature, T=65°C, and increases at the reduced temperature. These results are again in agreement with the dependence of the diffusion on lipid chain length and therefore with the free volume model.Abbreviations DLPC Dilauroylphosphatidylcholine - DMPC Dimyristoylphosphatidylcholine - DPPC Dipalmitoylphosphatidylcholine - DSPC Distearoylphosphatidylcholine - LUV Large unilamellar vesicles - SUV Small unilamellar vesicles - Tris Tris(hydroxymethyl)aminomethan  相似文献   

8.
In 2003, 50 game carcasses (ungulates) originating from one Austrian hunting ground were subject to visual examination for (fecal) contamination of the body cavities and microbiological testing of the body cavities in order to assess variations in microbial surface contamination in the season June–August compared to October–December. No carcass tested positive for the bacterial pathogens Salmonella or Listeria. Bacterial surface counts in October–December (median values: total aerobic count: 4.12 log10 colony-forming-units (cfu)/cm2; Enterobacteriaceae: 2.48 log10 cfu/cm2) were significantly lower than those in June–August (median values: total aerobic count: 5.65 log10 cfu/cm2; Enterobacteriaceae: 3.45 log10 cfu/cm2). The cooling regime (0.4 °C, 62% relative humidity) allowed no microbial growth for 96 h but was associated with weight loss of the carcasses. All carcasses had undergone a precooling phase of 8–12 h, with temperatures of 17.8±1.2 °C in the season June–August and 9.8±1.2 °C in October–December. This temperature difference was identified as the most probable effector for the observed seasonal variation. The results demonstrate the need for a continuous cool chain after evisceration of game carcasses.  相似文献   

9.
Inactivation of urease (25 nM) in aqueous solutions (pH 5.0–6.0) treated with low-frequency ultrasound (LFUS; 27 kHz, 60 W/cm2, 36–56°C) or high-frequency ultrasound (HFUS; 2.64 MHz, 1 W/cm2, 36 or 56°C) has been characterized quantitatively, using first-order rate constants: k in, total inactivation; k in *, thermal inactivation; and k in(us), ultrasonic inactivation. Within the range from 1 nM to 10 M, propyl gallate (PG) decreases by approximately threefold the rate of LFUS-induced inactivation of urease (56°C), whereas resorcinol poly-2-disulfide stops this process at 1 nM or higher concentrations. PG completely inhibits HFUS-induced inactivation of urease at 1 nM (36°C) or 10 nM (56°C). At 0.2–1.0 M, human serum albumin (HSA) increases the resistance of urease treated with HFUS to temperature- and cavitation-induced inactivation. Complexes of gallic acid polydisulfide (GAPDS) with HSA (GAPDS–HSA), formed by conjugation of 1.0 nM GAPDS with 0.33 nM HSA, prevent HFUS-induced urease inactivation (56°C).  相似文献   

10.
Summary The decreasing effect of -adrenergic blockade on skin resistance to vapor diffusion and the onset of cutaneous water evaporation in the pigeon (Columba livia) was investigated. Oral administration of 1, 2.3 and 5 mg propranolol to pigeons (268±53 g) initiated intensive trans-cutaneous water evaporation (CWE) up to 29.1 mg H2O·cm–2·h–1 in resting birds at 30°C air temperature (Ta), but had only a slight effect on CWE of birds exposed to 50 °C Ta.After 7 h of effective -adrenergic blockade (oral administration of 5 mg propranolol), skin and body temperature stabilized at 39.0±0.5 °C and 41.0±0.7 °C, compared to 40.2±0.8 °C and 41.9±0.6 °C in the control group, respectively. A slight hypothermia was accompanied by feather fluffing.Intradermal injection of 0.001, 0.01 and 0.12 mg propranolol also caused intensive CWE. Local -adrenergic blockade in relatively low blocker doses (0.001 and 0.01 mg propranolol) decreased skin resistance from a high value of 44.5 s·cm–1 to about 6.0 s·cm–1, and caused a sharp increase in CWE from a control value of about 4 to a high of 26.4 mg H2O·cm–2·h–1 during the first two hours of exposure to 30°C Ta.The possible role of -adrenergic blockade in regulation of trans-cutaneous water evaporation of latent heat dissipation is discussed.  相似文献   

11.
Summary Diffusion coefficients for FITC-molecular probes in intercellular pores (D) and rate of molecular probe loss into the vacuole (k1) have been obtained for FITC molecular probes in staminal hairs ofSetcreasea purpurea. The kinetic curves of FITC-Gly, -Ala, -Leu,-Ser, -Thr, -Cys, -Met, -Tyr, -Asp, -Glu, -Asn, -Gln, -Lys, -His,-Arg, -(Asp)2, -(Glu)2, -(Lys)2, -(Asp)3, -(Glu)3, -(Gln)2, -(Gln)3, -(Gln)4, and carboxyfluorescein (group I probes) matched the curves calculated for simple diffusion through a chain of cells, while the majority of kinetic curves of FITC-Phe, and -Try (group II probes) did not. None of the kinetic curves for FITC-(Met)2 and -(His)2 (group III probes) matched. Average Ds for group I probes ranged from 0.77× 10–8cm2/s to 3.75× 10–8cm2/s and for group II probes were 0.50× 10–8cm2/s. A meaningful average D for group III probes could not be calculated. Average k1 for group I probes ranged from 1.62× 10–7/m2/s to 13.21× 10–7/m2/s, and for group II probes were 5.42 and 11.54× 10–7/m2/s. Average k1s for group III probes could not be calculated. Symplastic transport occurred by cell-to-cell diffusion for most of the probes (e.g., group I probes) but not always for some (e.g., group II probes) and never for others (group III probes). The rate of cell-to-cell diffusion and loss within the vacuole depended upon the molecule's specific structure, molecular weight and charge. We concluded that plasmodesmata select for molecules that are hydrophilic, small and have a charge of from — 2 to — 4, and against molecules that contain either Phe, Try, Met or His groups.Abbreviations CF carboxyfluorescein - D diffusion coefficients for FITC-molecular probes in intercellular pores - k1 rate of FITC-molecular probe loss  相似文献   

12.
The aim of the study was to determine properties of selected ion-exchange resins for citric and lactic acids recovery, to define sorption isotherms for these acids at different temperatures (in the range of 20–60°C) and to determine diffusion coefficients inside sorbent particles. A mathematical model of the ion-exchange process in the chromatographic column and its experimental verification is also presented. During investigations 18 types of ion-exchange resins were tested. It was found that weakly basic resins were more suitable for the recovery process than strongly basic ones. The best resin for the separation of citric acid was Amberlite IRA-67 and for lactic acid Amberlite IRA-92. As a result of transient-state sorption experiments diffusion coefficients of the citric acid inside the sorbent particle at different temperatures were obtained. It was found that Dp increased with the temperature by two times in the range of 20–60°C, and its value at 60°C was 7.2×10–10 m2/s. The proposed mathematical model was applied to identify bed operation parameters in the column for the needs of the simulated moving bed chromatography method.List of symbols b Equilibrium constant in Langmuir equation, [dm3/g] - c Acid concentration in the liquid phase inside the particle pores, [g/dm3] - C Acid concentration in the liquid, [g/dm3] - DL Axial dispersion coefficient, [m2/s] - Dp Intraparticle diffusion coefficient, [m2/s] - kf Liquid film mass transfer coefficient, [m/s] - L Ion-exchanger bed height, [m] - q Acid concentration in the adsorbent phase, [g/dm3] - Rp Particle radius, [m] - U Volumetric flow rate of the feeding solution, [dm3/s] - V Volume of the solution, [dm3] - W Weight of the wet resin particles, [g] - The ion-exchanger bed porosity, [-] - p Particle porosity, [-] - Linear liquid velocity, [m/s] - Apparent density of the wet resin, [g/dm3]  相似文献   

13.
Summary Particle supported biofilms of uniform thickness were generated in an aerobic fluidized-bed reactor with phenol as the carbon source. A method was developed for determining the effective diffusivities of oxygen and phenol using trypan blue, a vital stain as the tracer. The effective diffusivities of oxygen and phenol were found to be 2.72×10–6 cm2/s and 1.12×10–6 cm2/s respectively.Nomenclature Ci initial solute concentration in bulk, g/cm3 - Ct solute concentration in bulk at time t, g/cm3 - C bulk solute concentration at equilibrium, g/cm3 - D molecular diffusivity, cm2/s - D effective diffusivity, cm2/s - Do Dp Dtb molecular diffusivity of oxygen, phenol and trypan blue, cm2/s - Do, Dp, Dtb effective diffusivity of oxygen, phenol and trypan blue, cm2/s - Ds molecular diffusivity of substrate, cm2/s - Ds effective diffusivity of substrate, cm2/s - K partition coefficient - Mt amount of solute in the particle at time t, g - M amount of solute in the particle at equilibrium, g - r particle radius, cm - r bp radius of the particle with biofilm, cm - S substrate concentration, g/cm3 - Sb substrate concentration in bulk, g/cm3 - Si initial substrate concentration, g/cm3 - V1 solute molar volume, cm3/g mol Greek Symbols bf porosity of the biofilm - tortuosity factor  相似文献   

14.
D. Meyer  A. Jungk 《Plant and Soil》1993,149(2):235-243
To predict the contribution of soil K fractions of different mobility to K supply of plants, the kinetics of K release from soil was related to the kinetics of K uptake of young sugar beet and wheat plants. For this purpose K release rates from soil were measured by continuously percolating samples of a luvisol with 0.01 M CaCl2 solution and effective diffusion coefficients, De, were determined. Two soil K fractions of different mobility were obtained. De values of the more mobile exchangeable K and the less mobile non-exchangeable K fraction were found to be 58.9 × 10–9 and 8.2 × 10–9 cm2 s–1, respectively. In a pot experiment, sugar beet and wheat plants were grown, for 15 days and both root growth and K uptake were measured. K uptake kinetics of both crops was determined in a separate experiment using flowing solution culture. To integrate these data quantitatively, the simulation model of Claassen et al. (1986) was applied. Results show that calculated total K uptake agreed closely with real K uptake of the plants. On this basis, 64 and 79% of the K taken up by wheat and sugar beet plants was derived from the rapidly released exchangeable and 21–36% from the less mobile non-exchangeable soil K fraction.  相似文献   

15.
Summary Diffusion of inorganic mercury (Hg2+) through planar lipid bilayer membranes was studied as a function of chloride concentration and pH. Membranes were made from egg lecithin plus cholesterol in tetradecane. Tracer (203Hg) flux and conductance measurements were used to estimate the permeabilities to ionic and nonionic forms of Hg. At pH 7.0 and [Cl] ranging from 10–1000mm, only the dichloride complex of mercury (HgCl2) crosses the membrane at a significant rate. However, several other Hg complexes (HgOHCl, HgCl 3 and HgCl 4 2– ) contribute to diffusion through the aqueous unstirred layer adjacent to the membrane. The relation between the total mercury flux (J Hg), Hg concentrations, and permeabilities is: 1/J Hg=1/P ul[Hg t ]+1/P m [HgCl2], where [Hg t ] is the total concentration of all forms of Hg,P ul is the unstirred layer permeability, andP m is the membrane permeability to HgCl2. By fitting this equation to the data we find thatP m =1.3×10–2 cm sec–1. At Cl concentrations ranging from 1–100mm, diffusion of Hg t through the unstirred layer is rate limiting. At Cl concentrations ranging from 500–1000mm, the membrane permeability to HgCl2 becomes rate limiting because HgCl2 comprises only about 1% of the total Hg. Under all conditions, chemical reactions among Hg2+, Cl and/or OH near the membrane surface play an important role in the transport process. Other important metals, e.g., Zn2+, Cd2+, Ag+ and CH3Hg+, form neutral chloride complexes under physiological conditions. Thus, it is likely that chloride can facilitate the diffusion of a variety of metals through lipid bilayer and biological membranes.  相似文献   

16.
The diffusion translational coefficient DT of core particles in monodisperse solutions has been measured by the quasielastic light scattering method in a large scale of salinities over the range 6.10−4 to 2M Na+ or K+. The observed values of DT are independent of particle concentration in the range 0.1–2 mg/ml and do not vary with the scattering vector q corresponding to scattering angles between 40°–120°. When the salinity is progressively raised an increase of DT from 1.9.10−7 cm2s−1 to 3.2.10−7 cm2s−1 was observed at about 2.10−3 M NaCl followed by a decrease of DT beyond 0.6 M NaCl.The various possible causes of the changes of DT such as interactions between particles or between particles and salt ions are discussed. We show that the single low ionic strength change is due to a conformational transition of the core particles, while the second variation of DT accompanies the disorganization of the core particles.  相似文献   

17.
Zusammenfassung Es wurde der Transport von45Ca,85Sr und32P in polykristallinen Sinterkörpern von synthetischem Hydroxylapatit im Temperaturbereich 1000 bis 1400 °C untersucht. Nach sorgfältiger Berücksichtigung von Korngrenzen-Diffusionseffekten ergaben sich für die Diffusion von45Ca und85Sr gleiche Werte für die Aktivierungsenthalpien und Frequenzfaktoren, und zwar beipH2O<30 Torr:Q=3,50 ± 0,02 eV;D 0=41 ± 5 cm2s–1 und beipH2O=230 Torr:Q=3,55 ± 0,02 eV;D 0=20 ± 3 cm2s–1 Die Abhängigkeit des Kationen-Diffusionskoeffizienten vom Wasserdampfpartialdruck ist vermutlich dadurch bedingt, daß im untersuchten Temperaturbereich feste Lösungen von Hydroxylapatit und Oxyapatit entstehen und Leerstellen im OH-Teilgitter den Kationentransport beschleunigen. Der32P-Transport wurde nur bei 1360 °C undpH2O < 30 Torr gemessen. Der Diffusionskoeffizient ist um einen Faktor 400 ± 50 kleiner als der entsprechende Diffusionskoeffizient der Kationen.Die Ergebnisse der Diffusionsuntersuchungen werden in Verbindung mit einer einfachen Modellvorstellung zum Retentionsmechanismus der Erdalkalien im Skelett diskutiert.
Diffusion of45Ca,85Sr, and32P in hydroxyapatite
Summary The transport of45Ca,85Sr, and32P in polycrystalline sinter pellets of synthetic hydroxyapatite has been investigated in the temperature range 1000 to 1400 °C. After subtraction of activity transports by grain boundary diffusion processes, equal values of activation enthalpy and frequency factor were found for the lattice diffusion of45Ca and85Sr: atpH2O<30 Torr:Q=3,50 ± 0,02 eV,D 0=41 ± 5 cm2s–1 and atpH2O=230 Torr:Q=3,55 ± 0,02 eV,D 0=20 ± 3 cm2s.The dependence of the cation diffusion coefficient on the partial vapour pressure is probably caused by formation of solid solutions of hydroxyapatite and oxyapatite where vacancies of the OH sublattice accelerate the cation transport. The diffusion of32P was investigated only atT=1360 °C andpH2O<30 Torr. The value obtained is smaller by a factor of 400 ± 50 then the cation diffusion coefficient.The results of the diffusion experiments are discussed in the light of a simple model for the retention mechanism of the alkaline earth metals in the skeleton.
  相似文献   

18.
Summary Single fast fibres and small bundles of slow fibres were isolated from the trunk muscles of an Antarctic (Notothenia neglecta) and various warm water marine fishes (Blue Crevally,Carangus melampygus; Grey Mullet,Mugil cephalus; Dolphin Fish,Coryphaena hippurus; Skipjack-tuna,Katsuwonus pelamis and Kawakawa,Euthynuus affinis). Fibres were chemically skinned with the nonionic detergent Brij 58.For warm water species, maximum Ca2+-activated tension (P 0) almost doubled between 5–20°C with little further increase up to 30°C. However, when measured at their normal body temperatures,P 0 values for fast fibres were similar for all species examined, 15.7–22.5 N · cm–2. Ca2+-regulation of contraction was disrupted at temperatures above 15°C in the Antarctic species, but was maintained at up to 30°C for warm water fish.Unloaded (maximum) contraction speeds (V max) of fibres were determined by the slacktest method. In general,V max was approximately two times higher in white than red muscles for all species studied, except Skipjack tuna. For Skipjack tuna,V max of superficial red and white fibres was similar (15.7 muscle lengths · s–1 (L 0 · s–1)) but were 6.5 times faster than theV max of internal red muscle fibres (2.4±0.2L 0 · s–1) (25°C). V max forN. neglecta fast fibres at 0–5°C (2–3L 0 · s–1) were similar to that of warm water species measured at 10–20°C. However, when measured at their normal muscle temperatures, theV max for the fast muscle fibres of the warm water species were 2–3 times higher than that forN. neglecta.In general,Q 10(15–30°C) values forV max were in the range 1.8–2.0 for all warm water species studied except Skipjack tuna.V max for the internal red muscle fibres of Skipjack tuna were much more temperature dependent (Q 10(15–30°C)=3.1) (P<0.01) than for superficial red or white muscle fibres. The proportion of slower red muscle fibres in tuna (28% for 1 kg Skipjack) is 3–10 times higher than for most teleosts and is related to the tuna's need to sustain high cruising speeds. We suggest that the 8–10°C temperature gradient that can exist in Skipjack tuna between internal red and white muscles allows both fibre types to contract at the same speed. Therefore, in tuna, both red and white muscle may contribute to power generation during high speed swimming.  相似文献   

19.
A method for isolating extracellular glucose oxidase from the fungus Penicillium funiculosum 46.1 using ultrafiltration membranes was developed. Two samples of the enzyme with a specific activity of 914–956 IU were obtained. The enzyme exhibited a high catalytic activity at pH above 6.0. The effective rate constant of glucose oxidase inactivation at pH 2.6 and 16°C was 2.74 × 10–6 s–1. This constant decreased significantly as the pH of the medium increased (4.0–10.0). The temperature optimum for glucose oxidase–catalyzed -D-glucose oxidation was in the range 30–65°C. At temperatures below 30°C, the activation energy for -D-glucose oxidation was 6.42 kcal/mol; at higher temperatures, this parameter was equal to 0.61 kcal/mol. Kinetic parameters of glucose oxidase–catalyzed -D-glucose oxidation depended on the initial concentration of the enzyme in the solution. Glucose oxidase also catalyzed the oxidation of 2-deoxy-D-glucose, maltose, and galactose.  相似文献   

20.
E. Steudle  J. S. Boyer 《Planta》1985,164(2):189-200
Hydraulic resistances to water flow have been determined in the cortex of hypocotyls of growing seedlings of soybean (Glycine max L. Merr. cv. Wayne). Data at the cell level (hydraulic conductivity, Lp; half-time of water exchange, T 1/2; elastic modulus, ; diffusivity for the cell-to-cell pathway, D c) were obtained by the pressure probe, diffusivities for the tissue (D t) by sorption experiments and the hydraulic conductivity of the entire cortex (Lpr) by a new pressure-perfusion technique. For cortical cells in the elongating and mature regions of the hypocotyls T 1/2=0.4–15.1 s, Lp=0.2·10-5–10.0·10-5 cm s-1 bar-1 and D c=0.1·10-6–5.5·10-6 cm2 s-1. Sorption kinetics yielded a tissue diffusivity D t=0.2·10-6–0.8·10-6 cm2 s-1. The sorption kinetics include both cell-wall and cell-to-cell pathways for water transport. By comparing D c and D t, it was concluded that during swelling or shrinking of the tissue and during growth a substantial amount of water moves from cell to cell. The pressure-perfusion technique imposed hydrostatic gradients across the cortex either by manipulating the hydrostatic pressure in the xylem of hypocotyl segments or by forcing water from outside into the xylem. In segments with intact cuticle, the hydraulic conductance of the radial path (Lpr) was a function of the rate of water flow and also of flow direction. In segments without cuticle, Lpr was large (Lpr=2·10-5–20·10-5 cm s-1 bar-1) and exceeded the corticla cell Lp. The results of the pressure-perfusion experiments are not compatible with a cell-to-cell transport and can only the explained by a preferred apoplasmic water movement. A tentative explanation for the differences found in the different types of experiments is that during hydrostatic perfusion the apoplasmic path dominates because of the high hydraulic conductivity of the cell wall or a preferred water movement by film flow in the intercellular space system. For shrinking and swelling experiments and during growth, the films are small and the cell-to-cell path dominates. This could lead to larger gradients in water potential in the tissue than expected from Lpr. It is suggested that the reason for the preference of the cell-to-cell path during swelling and growth is that the solute contribution to the driving force in the apoplast is small, and tensions normally present in the wall prevent sufficiently thick water films from forming. The solute contribution is not very effective because the reflection coefficient of the cell-wall material should be very small for small solutes. The results demonstrate that in plant tissues the relative magnitude of cell-wall versus cell-to-cell transport could dependent on the physical nature of the driving forces (hydrostatic, osmotic) involved.Abbreviations and symbols D c diffusivity of the cell-to-cell pathway - D t diffusivity of the tissue - radial flow rate per cm2 of segment surface - Lp hydraulic conductivity of plasma-membrane - Lpr radial hydraulic conductance of the cortex - T 1/2 half-time of water exchange between cell and surroundings - volumetric elastic modulus  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