首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
E. Schäfer  B. Marchal  D. Marmé 《Planta》1971,101(3):265-276
Summary The in vivo phototransformation kinetics of mustard hook and cotyledon phytochrome exhibit a deviation from a single first order curve, quite similar to that for pumpkin hooks as reported in a previous paper (Boisard, Marmé and Schäfer, 1971). The P frPrkinetics can be characterized by the ratios fr, I · P fr I / fr, II · P fr, II and where P fr I and P fr II are two populations of phytochrome molecules which convert to P rwith a first order half-life of and . These ratios depend on the length of time of etiolation. The ratio fr, I · P fr I / fr, II · P fr, II is independent of the amount of total P frpresent at the beginning of the P frPrphototransformation after a non-saturating dose of red light. The half-lives of the two populations, however, depend on the concentration of total P frinitially present. P frPrphototransformation kinetics with different light intensities show that reciprocity holds.  相似文献   

2.
Behavioral and physiological responses to hypoxia were examined in three sympatric species of sharks: bonnethead shark Sphyrna tiburo, blacknose shark, Carcharhinus acronotus, and Florida smoothhound shark, Mustelus norrisi, using closed system respirometry. Sharks were exposed to normoxic and three levels of hypoxic conditions. Under normoxic conditions (5.5–6.4mg l–1), shark routine swimming speed averaged 25.5 and 31.0cm s–1 for obligate ram-ventilating S. tiburo and C. acronotus respectively, and 25.0cm s–1 for buccal-ventilating M. norrisi. Routine oxygen consumption averaged about 234.6 mg O2kg–1h–1 for S. tiburo, 437.2mg O2kg–1h–1 for C. acronotus, and 161.4mg O2 kg–1 h–1 for M. norrisi. For ram-ventilating sharks, mouth gape averaged 1.0cm whereas M. norrisi gillbeats averaged 56.0 beats min–1. Swimming speeds, mouth gape, and oxygen consumption rate of S. tiburo and C. acronotus increased to a maximum of 37–39cm s–1, 2.5–3.0cm and 496 and 599mg O2 kg–1 h–1 under hypoxic conditions (2.5–3.4mg l–1), respectively. M. norrisi decreased swimming speeds to 16cm s–1 and oxygen consumption rate remained similar. Results support the hypothesis that obligate ram-ventilating sharks respond to hypoxia by increasing swimming speed and mouth gape while buccal-ventilating smoothhound sharks reduce activity.  相似文献   

3.
The effect of varying light regimes on in vitro rooting of microcuttings of two pear (Pyrus communis L.) cultivars was investigated. Cultures of the easy to-root Conference and the difficult-to-root Doyenne d'Hiver were incubated for 21 days with or without indole-3-butyric acid (IBA) in the medium in darkness or under continuous far-red (8 µmol m–2 s–1), blue, white or red (15 or 36 µmol m–2 s–1) light. Conference rooted without IBA when exposed to red, blue or white light while no rooting was observed under far-red light and in darkness. The high rooting efficiency under red and, by contrast, the inhibition under far-red light and darkness suggest the involvement of the phytochrome system in rhizogenesis. The addition of IBA to the culture medium enhanced root production under all light regimes in both cultivars. Red light, especially at the lower photon fluence rate, had a positive effect by increasing root extension (number × length of roots) and stimulating secondary root formation.Abbreviations IBA Indole-3-butyric acid - R red light - B blue light - FR far-red light - W white light - D darkness - Pfr active (far-red light absorbing) form of phytochrome - Ptot total phytochrome - BA benzyl-adenine  相似文献   

4.
The obligate shade plant, Tradescantia albiflora Kunth grown at 50 mol photons · m–2 s–1 and Pisum sativum L. acclimated to two photon fluence rates, 50 and 300 mol · m–2 · s–1, were exposed to photoinhibitory light conditions of 1700 mol · m–2 · s–1 for 4 h at 22° C. Photosynthesis was assayed by measurement of CO2-saturated O2 evolution, and photosystem II (PSII) was assayed using modulated chlorophyll fluorescence and flash-yield determinations of functional reaction centres. Tradescantia was most sensitive to photoinhibition, while pea grown at 300 mol · m–2 · s–1 was most resistant, with pea grown at 50 mol · m–2 · s–1 showing an intermediate sensitivity. A very good correlation was found between the decrease of functional PSII reaction centres and both the inhibition of photosynthesis and PSII photochemistry. Photoinhibition caused a decline in the maximum quantum yield for PSII electron transport as determined by the product of photochemical quenching (qp) and the yield of open PSII reaction centres as given by the steady-state fluorescence ratio, FvFm, according to Genty et al. (1989, Biochim. Biophys. Acta 990, 81–92). The decrease in the quantum yield for PSII electron transport was fully accounted for by a decrease in FvFm, since qp at a given photon fluence rate was similar for photoinhibited and noninhibited plants. Under lightsaturating conditions, the quantum yield of PSII electron transport was similar in photoinhibited and noninhibited plants. The data give support for the view that photoinhibition of the reaction centres of PSII represents a stable, long-term, down-regulation of photochemistry, which occurs in plants under sustained high-light conditions, and replaces part of the regulation usually exerted by the transthylakoid pH gradient. Furthermore, by investigating the susceptibility of differently lightacclimated sun and shade species to photoinhibition in relation to qp, i.e. the fraction of open-to-closed PSII reaction centres, we also show that irrespective of light acclimation, plants become susceptible to photoinhibition when the majority of their PSII reaction centres are still open (i.e. primary quinone acceptor oxidized). Photoinhibition appears to be an unavoidable consequence of PSII function when light causes sustained closure of more than 40% of PSII reaction centres.Abbreviations Fo and Fo minimal fluorescence when all PSII reaction centres are open in darkness and steady-state light, respectively - Fm and Fm maximal fluorescence when all PSII reaction centres are closed in darkand light-acclimated leaves, respectively - Fv variable fluorescence - (Fm-Fo) under steady-state light con-ditions - Fs steady-state fluorescence in light - QA the primary,stable quinone acceptor of PSII - qNe non-photochemical quench-ing of fluorescence due to high energy state - (pH); qNi non-photochemical quenching of fluorescence due to photoinhibition - qp photochemical quenching of fluorescence To whom correspondence should be addressedThis work was supported by the Swedish Natural Science Research Council (G.Ö.) and the award of a National Research Fellowship to J.M.A and W.S.C. We thank Dr. Paul Kriedemann, Division of Forestry and Forest Products, CSIRO, Canberra, Australia, for helpful discussions.  相似文献   

5.
New and known homo- and heterodinuclear RuII and OsII complexes with 4,4-bipyridine (4,4-bpy), pyrazine, and 4-pyCH=CHpy-4 as bridging ligands (LL) of the type [Cl(bpy)2M(LL)MCl(bpy)2]X2 (bpy=2,2-bipyridine; X=PF6 or BF4) have been studied in their capacity to exchange electrons with a reduced active site of glucose oxidase (GO) from Aspergillus niger. Cyclic voltammograms (CVs) of the dimers in the aqueous buffered solution, when compared with CVs of the parent monomeric species [MCl(LL)(bpy)2]BF4 and [MCl2(bpy)2] which could be generated at pH7, if the dimers undergo monomerization, indicate that the dimers are the dominating species under such conditions. All electrochemically oxidized dinuclear complexes studied show high rates of oxidation of GO reduced by d-glucose and the corresponding observed second-order rate constants are in the range (5–64)×105 M–1 s–1 at 25 °C. However, these values are lower than that for the mononuclear complex [OsCl(4,4-bpy)(bpy)2]BF4 (1.1×107 M–1 s–1), suggesting that potentially two-electron dimeric mediators have no advantage compared with corresponding monomeric complexes of RuII and OsII. The structure of [OsCl(4,4-bpy)(bpy)2]BF4 was confirmed by X-ray crystallography. The monodentate 4,4-bpy ligand is coordinated cis to the chloride. Its higher reactivity toward reduced GO is accounted for in terms of the antenna effect of the monodentate 4,4-bpy ligand. The antenna length equals 9.2 Å and matches the depth of the enzyme active site pocket of ca. 10 Å. The mechanism of the antenna effect is discussed  相似文献   

6.
The possibility of solving the mass balances to a multiplicity of substrates within a CSTR in the presence of a chemical reaction following Michaelis-Menten kinetics using the assumption that the discrete distribution of said substrates is well approximated by an equivalent continuous distribution on the molecular weight is explored. The applicability of such reasoning is tested with a convenient numerical example. In addition to providing the limiting behavior of the discrete formulation as the number of homologous substrates increases, the continuous formulation yields in general simpler functional forms for the final distribution of substrates than the discrete counterpart due to the recursive nature of the solution in the latter case.List of Symbols C{N. M} mol/m3 concentration of substrate containing N monomer residues each with molecular weight M - {N, M} normalized value of C{N. M} - C {M} mol/m3 da concentration of substrate of molecular weight M - in normalized value of C {M} at the i-th iteration of a finite difference method - {M} normalized value of C {M} - C 0{N.M} mol/m3 inlet concentration of substrate containing N monomer residues each with molecular weight M - {N ·M} normalized value of C0{N. M} - 0 i normalized value of C 0 {M} at the i-th iteration of a finite difference method - C 0 {M} mol/m3 da initial concentration of substrate of molecular weight M - C tot mol/m3 (constant) overall concentration of substrates (discrete model) - C tot mol/m3 (constant) overall concentration of substrates (continuous model) - D deviation of the continuous approach relative to the discrete approach - i dummy integer variable - I arbitrary integration constant - j dummy integer variable - k dummy integer variable - K m mol/m3 Michaëlis-Menten constant for the substrates - l dummy integer variable - M da molecular weight of substrate - M normalized value of M - M da maximum molecular weight of a reacting substrate - N number of monomer residues of a reacting substrate - N maximum number of monomer residues of a reacting substrate - N total number of increments for the finite difference method - Q m3/s volumetric flow rate of liquid through the reactor - S inert product molecule - S i substrate containing i monomer residues - V m3 volume of the reactor - v max mol/m3 s reaction rate under saturating conditions of the enzyme active site with substrate - v max{N. M} mol/m3 s reaction rate under saturating conditions of the enzyme active site with substrate containing N monomer residues with molecular weight M - max{N · M} dimensionless value of vmax{N. M} (discrete model) - max{M} dimensionless value of v max {M} (continuous model) - mol/m3 s molecular weight-averaged value of vmax (discrete model) - mol.da/m3s molecular weight-averaged value of vmax (continuous model) - v max {M} mol.da/m3s reaction rate under saturating conditions of the enzyme active site with substrate with molecular weight M - max {M} dimensionless value of vmax{M} - max, (i) dimensionless value of vmax{M} at the i-th iteration of a finite difference method - v max mol/m3 s reference constant value of v max Greek Symbols dimensionless operating parameter (discrete distribution) - dimensionless operating parameter (continuous distribution) - M da (average) molecular weight of a monomeric subunit - M selected increment for the finite difference method - auxiliary corrective factor (discrete model)  相似文献   

7.
Measurement of the light response of photosynthetic CO2 uptake is often used as an implement in ecophysiological studies. A method is described to calculate photosynthetic parameters, such as the maximum rate of whole electron transport and dissimilative respiration in the light, from the light response of CO2 uptake. Examples of the light-response curves of flag leaves and ears of wheat (Triticum aestivum cv. ARKAS) are shown.Abbreviations and symbols A net photosynthesis rate - D 1 rate of dissimilative respiration occurring in the light - f loss factor - I incident PPFD - I effective absorbed PPFD - J rate of whole electron transport - J m maximum rate of whole electron transport - p c intercellular CO2 partial pressure - PPFD photosynthetic photon flux density - q effectivity factor for the use of light (electrons/quanta) - absorption coefficient - I * CO2 compensation point in the absence of dissimilative respiration (bar) - II conversion factor for calculation of CO2 uptake from the rate of whole electron transport - convexity factor Gas-exchange rates relate to the projective area and are given in mol·m-2·s-1. Electron-transport rates are given in mol electrons·m-2·s-1; PPFD is given in mol quanta·m-2·s-1.  相似文献   

8.
Age and growth of the whiskery shark, Furgaleus macki, from southwestern Australia were examined using vertebral ageing and tag-recapture data. The readability of bands on the vertebral centra varied markedly between individuals. Four readers were used to make band counts, with the most experienced reader having the lowest index of average percent error and the highest level of agreement with final counts. Marginal increment analysis indicated that opaque bands form in January. With parturition occurring from August to October, size data suggests that the first band is probably formed 15–17 months after birth. The age at maturity was estimated to be 4.5 years for males, and 6.5 years for females. The oldest male was 10.5 years, and oldest female was 11.5 years. Von Bertalanffy growth parameters for males were L =121.5cm fork length, K=0.423 year–1, t 0=–0.472 years, were L =120.7cm fork length, K=0.369 year–1, t 0=–0.544 years for females, and were L =118.1cm fork length, K=0.420 year–1, t 0=–0.491 years for combined sexes. Data from a tag recapture study were analysed using a maximum likelihood method to verify the estimates of growth parameters from vertebral ageing. Von Bertalanffy growth parameters from the tag recapture study were L =128.2cm fork length, K=0.288 year–1, t 0=–0.654 years. The two methods of estimating growth parameters produced similar results, with rapid growth until approximately 5 years of age, after which there was little increase in length.  相似文献   

9.
A functional model for the S4/IV -helix of the action potential sodium channel is described by means of a thermodynamic approach. The model is based on a phase transition between the -helix and an ion conducting channel-helix which is similar to the well established helix-coil transition in solution. The right hand channel-helix is a peptide chain with an alternating sequence of torsional angles (11)=(87°, 315°) and (22)=(22°, 107°) which yields a helix of 13.5 Å per turn. The axial dipole moments of the peptide bonds of this chain of l-amino acids nearly cancel each other out in similar way to those in the gramicidin A channel, which is formed by alternating d-and l-amino acids. The helix, which does not contain any H-bonds, is stabilized by a helical file of water molecules which includes the permeating ion(s). This file turns around the channel-helix to form a relatively stable double helix structure which corresponds to the open channel. Since every third side chain in the S4/IV helix carries a positive charge their environments must be polarized. These polarized regions form a left hand screening-helix around the -helix are broken and the internal -carbon atom is considered as fixed, the outer ten residues leave the membrane while the internal ten residues form the channel-helix. In this configuration every positively charged side chain matches nearly exactly every second polarized region of the screening-helix leaving the three regions in-between exposed to the water file containing the ion(s). This further stabilizes the channel and agrees nicely with the idea of cationic selectivity. An analysis of the energetics of the -helix-channel-helix transition showed that the voltage-independent part of the free energy per helix residue could well be close to 0 kcal/mol and thus be in the range where a transition could occur. Two voltage-dependent contributions were included: the break down of the considerable dipole of the -helix and the outward shift of the positive charges of the side chains upon channel-helix formation. Taking into account the fact that the formation of an -helix is a highly cooperative process the degree of voltage dependence of the probability of formation of a channel-helix proved to be in the same range as experimental values for the open probability of modified Na channels whose inactivation had been removed. With regard to gating currents, the model predicts that 2.74 positive charges are moved in an outward direction. Consequences of the model for other experimental findings are discussed.  相似文献   

10.
The orientation behavior of walking flies, Drosophila melanogaster, towards a single 6° wide black vertical stripe (elementary stripe) can be explained by use of the turning tendency function H(). This function is characterized by maximal values at an angular distance of =25° from the stable zero position (=orienting direction), a sharp decline from this maximum to =60°, and a very slow approach to the unstable zero position (Horn and Wehner, 1975). The shape of this function is influenced by both translatory and rotatory components of movement. If the translatory component is minimized by measuring the turning function W() (see 2.3) at a distance of 10 mm (C1) from the center of the arena, a change in the strength of this decline is caused. But with increasing translatory component, i.e. at a greater distance from the center of the arena, W() approximates the heuristical function H() (Fig. 12). The turning functions W() are pattern-specific; the angular positions of the maximum responses shift to greater angles with increasing width of the patterns (Fig. 2). In the twopattern configuration with double or single stripes, there is always a coincidence between the stable zero positions of W (), the mean of the frequency distributions P() of the flies' positions and n g() of the straight courses, and the stable zero positions of H () obtained from an additive superposition of two or more angular shifted turning tendency functions H() (Fig. 5, 7). Therefore, the mean positions of the flies in a multi-stripe experiment composed of elementary stripes can be predicted from the addition of many angular shifted turning tendency functions H(). Between H() and the frequency distribution P() of the flies' positions , the following formula holds: P() =C·H()d (Fig. 13). With this equation, the spontaneous preference of the broader of two double stripes can be explained presuming lateral interactions between the components of the patterns (Fig. 8, 10). The strength x i * of this lateral interaction depends on the width of the double stripes. The greater , the smaller is x i * . x i * is a pattern-specific value (Table 1, 2).Supported by the Deutsche Forschungsgemeinschaft, Ho 664/2  相似文献   

11.
Two intracellular -glucosidases (E.C. 3.2.1.21) were purified from the filamentous fungus Neurospora crassa, mutant cell-1 (FGSC no. 4335) and characterized. The extent of purification were 2.55- and 28.89-fold for -glucosidase A and -glucosidase B, respectively. -Glucosidase A was a dimeric protein, and B a monomeric protein, with molecular masses of 178 and 106 kDa, respectively. Both isoenzymes were glycoproteins with relatively high carbohydrate contents (-glucosidase A, 29.2%; -glucosidase B, 34.2%). The isoelectric points determined by IEF were 6.27 and 4.72, respectively. pH optima for activity were determined to be 5.0 and 5.5, and temperature optima to be 55 and 60 °C, for -glucosidases A and B, respectively. Both purified -glucosidases. especially -glucosidase B, showed relatively high stability against pH and temperature. Both enzymes were stable in the pH range of 5.0–9.0. The activities were completely retained up to 48 h at temperatures below 40 °C. At higher temperatures, enzymes were relatively unstable and lost their activities at 60 °C after 24 h. Both -glucosidases were highly activated by CuCl2, and inhibited by SnCl2 and KMnO4. Hg2+ and Ag+ also inhibited severely -glucosidase B. The K m and V max values of the isoenzymes against cellobiose as substrate were 1.50 mM and 12.2mol min–1 mg–1 for -glucosidase A and 2.76 mM and 143.5 mol min–1 mg–1 for -glucosidase B.  相似文献   

12.
Summary The fluorescent intercalation complex of ethidium bromide (ETB) with DNA was used as a probe to compare the effects of various radicals with respect to impairment of the DNA base-pair region.OH radicals inhibit up to 0.7 dye intercalations perOH at low salt concentration, and for various oxidizing species the effect decreases in the orderOH > Br 2 > N 3 > $$ " align="middle" border="0"> (SCN) 2 . DNA impairment by theOH product of Met-Gly is comparable to that of N 3 , but no effect was found due to the interaction between DNA and Lys-Tyr-Lys phenoxyl radicals. The reducing speciese aq , H, O 2 , and CO 2 hardly affect the DNA-ETB intercalation.  相似文献   

13.
The sugar conformation of a DNA decamer was studied with proton-proton 3J coupling constants. Two samples, one comprising stereospecifically labeled 2-R-2H for all residues and the other 2-S-2H, were prepared by the method of Kawashima et al. [J. Org. Chem. (1995) 60, 6980–6986; Nucleosides Nucleotides (1995) 14, 333–336], the deuterium labeling being highly stereospecific 99% for all 2-2H, 98% for 2-2H of A, C, and T, and 93% for 2-2H of G). The 3J values of all H1-H2 and H1-H2 pairs, and several H2-H3 and H2-H3 pairs were determined by line fitting of 1D spectra with 0.1–0.2 Hz precision. The observed J coupling constants were explained by the rigid sugar conformation model, and the sugar conformations were found to be between C3-exo and C2-endo with m values of 26° to 44°, except for the second and 3 terminal residues C2 and C10. For the C2 and C10 residues, the lower fraction of S-type conformation was estimated from JH1H2 and JH1H2 values. For C10, the N–S two-site jump model or Gaussian distribution of the torsion angle model could explain the observed J values, and 68% S-type conformation or C1-exo conformation with 27° distribution was obtained, respectively. The differences between these two motional models are discussed based on a simple simulation of J-coupling constants.  相似文献   

14.
The hematopoietic cellular kinase (Hck) is a member of the Srcfamily of non-receptor protein-tyrosine kinases that is expressedpredominantly in granulocytes, monocytes and macrophages. Recentobservations suggest that Hck may be activated in HIV-infected macrophagesand in chronic myelogenous leukemia cells that express Bcr-Abl. In order toincrease our understanding of the structural basis for regulation of Hckactivity under normal and pathological conditions, we have solved thesolution structure of the uncomplexed Hck SH2 domain using NMR spectroscopy.A novel procedure that uses intraresidueHNH distances as references forconverting NOE intensities into distance restraints has been described. Atotal of 1757 significant experimental restraints were derived from NMRspectroscopic data including 238 medium-range and 487 long-range distancerestraints and 177 torsion angle restraints. These restraints were used in asimulated annealing procedure to generate 20 structures with the programDYANA. Superimposition of residues 5–104 upon the mean coordinate setyielded an average atomic rmsd values of 0.42 ± 0.08 Å for theN,C,C atoms and 0.81 ± 0.08 Å forall heavy atoms. Rmsd values for those residues in the regions of orderedsecondary structure were 0.27 ± 0.04 Å for theN,C,C atoms and 0.73 ± 0.06 Å forall heavy atoms.  相似文献   

15.
By means of reaction calorimetry we measured the apparent enthalpy change, Happ, of the binding of Mn2+-ions to goat -lactalbumin as a function of temperature. The observed Happ can be written as the sum of contributions resulting from a conformational and a binding process. In combination with the thermal unfolding curve of goat -lactalbumin, we succeeded in separating the complete set of thermodynamic parameters (H, G, S, Cp) into the binding and conformational contributions. By circular dichroism we showed that NH 4 + -ions, upon binding to bovine a-lactalbumin, induce the same conformational change as do Na+ and K+: the binding constant equals 98 ± 9 M–1.Abbreviations BLA bovine -lactalbumin - GLA goat -lactalbumin - HLA human -lactalbumin - CD circular dichroism Offprint requests to: H. Van DaelDeceased  相似文献   

16.
We calculate the thermodynamic properties of a two-dimensional fluid of hard disks with embedded dipoles. Our attention is centered on the isotherms in the neighborhood of the critical point. Evaluating the canonical partition function by the "factor cluster expansion", we exhibit the Van der Waals loops obtained considering the exact two-body clusters and the "hard core" contribution of the three-body clusters. The Van der Waals isotherms can be scaled as universal functions of the parameter =p2/4r 0 3 kT, where p, r0, , are the dipole moment, hard core radius, and permittivity which characterize the interaction. The model is applied to the lipid phase transition found in natural and synthetic membranes. The typical critical parameters (Tc300K, C50 dyne/cm) reflect a physically reasonable value for the dipole moment of a polar head group of a lipid but a much-too-small value for the hard core radius.  相似文献   

17.
Micropropagated plantlets are fragile and often lack sufficient vigour to survive the acclimatization shock during transplantation to the soil. Effects of photosynthetic photon flux densities (PPFDs) on growth, photosynthesis and anatomy of micropropagated Doritaenopsis were studied after 4 months of acclimatization in a greenhouse at 25 °C. The plantlets were transferred to three different PPFDs for four months, i.e. low light (175), intermediate light (270) and high light (450 mol m–2 s–1). For most of the growth parameters measured i.e. leaf length, leaf area, leaf width, fresh weight, dry weight, chlorophyll (Chl) a/b ratio, were greater for the intermediate light levels after 4 months of acclimatization. The only exception was leaf thickness, which was increased more under high light levels. Results showed that the survival of Doritaenopsis plantlets was greatest (90%) in low light and intermediate light (89%) compared with only (73%) at high light. However, at low light levels, pigment concentrations (chlorophyll a, b and total chlorophyll) were higher. Net CO2 assimilation (A), stomatal conductance (g) and transpiration (E) were higher in plantlets grown at high level PPFD than at low after 4 months of acclimatization. Photosynthetic efficiency (Fv/Fm) decreased insignificantly; only at mid day for the high light treatment whereas leaf temperature and stomatal closure increased compared to low light. Scanning electron microscopic (SEM) images of leaves from acclimatized plantlets showed an increase in wax formation for the high light grown plantlets compared to those at low light. Microscopic analysis of acclimatized root sections showed highly developed multiseriate-velamen layers and higher root cell activity; while shoots had larger leaf air spaces than those of in vitro grown plantlets. These results suggest that physiological acclimation occurs at the intermediate PPFD (270 mol m–2 s–1) in Doritaenopsis compared to treatment at the high light level.  相似文献   

18.
Summary In this paper, the results of the preceding electrophysiological study of sodium-alanine cotransport in pancreatic acinar cells are compared with kinetic models. Two different types of transport mechanisms are considered. In the simultaneous mechanism the cotransporterC forms a ternary complexNCS with Na+ and the substrateS; coupled transport of Na+ andS involves a conformational transition between statesNCS andNCS with inward- and outward-facing binding sites. In the consecutive (or ping-pong) mechanism, formation of a ternary complex is not required; coupled transport occurs by an alternating sequence of association-dissociation steps and conformational transitions. It is shown that the experimentally observed alanine- and sodium-concentration dependence of transport rates is consistent with the predictions of the simultaneous model, but incompatible with the consecutive mechanism. Assuming that the association-dissociation reactions are not rate-limiting, a number of kinetic parameters of the simultaneous model can be estimated from the experimental results. The equilibrium dissociation constants of Na+ and alanine at the extracellular side are determined to beK N <-64mm andK S <-18mm. Furthermore, the ratioK N /K N S of the dissociation constants of Na+ from the binary (NC) and the ternary complex (NCS) at the extracellular side is estimated to be <-6. This indicates that the binding sequence of Na+ andS to the transporter is not ordered. The current-voltage behavior of the transporter is analyzed in terms of charge translocations associated with the single-reaction steps. The observed voltage-dependence of the half-saturation concentration of sodium is consistent with the assumption that a Na+ ion that migrates from the extracellular medium to the binding site has to traverse part of the transmembrane voltage.  相似文献   

19.
Summary The angular dependence of1JC,H in model compounds related to -linked oligosaccharides has been established by FPT INDO quantum chemical calculations. Values calculated for models of (1 1)-, (1 2)-, (1 3)- and (1 4)-linked disaccharides were compared, and the effect of the orientation of HO-2 elucidated. The angular dependence of1JC,H on the torsional angles H and H and the solvent dielectric constant (s) was characterized in the form:1JC,H = A cos2+B cos + C sin2 + D since + E + Fe. The1JC,H values, measured by DEPT methods for C-1-H-1 and C-X-H-X in cellobiose, cyclic trisaccharide and hexopyranoses were used to adjust the calculated angular dependences. Based on the occurrence of the conformers for agarobiose, neoagarobiose, mannobiose and methyl -xylobioside, the thermodynamically averaged <1JC,H > values were calculated. The results obtained (<1JC-1,H-1 > 162.4, <1JC-4, H-4 > 147.6 Hz for methyl -xylobioside; <1JC-1,H-1 > 162.4 and <1JC-4,H-4] > 147.6 Hz for mannobiose; <1JC-1,H-1 > 162.8 Hz for neo agarobiose and <1JC-1,H-1 > 163.2 Hz for agarobiose) agree well with the experimental values of 162.7, 147.5, 160.4, 147.2, 160.9 and 165.7 Hz, respectively.  相似文献   

20.
Summary Previous experiments have shown that during prey-catching behavior (orienting, snapping) in response to a worm-like moving stripe common toads.Bufo bufo (L.) exhibit a contrast-and direction-dependent edge preference. To a black (b) stripe moving against a white (w) background (b/w), they respond (R*) preferably toward the leading (l) rather the trailing (t) edge (R l * > R t * ), thus displaying head preference. If the contrastdirection is reversed (w/b), the stripe's trailing edge is preferred (R l * < R t * ), hence showing tail preference. In the present study, neuronal activities of retinal classes R2 and R3 and tectal classes T5(2) and T7 have been extracellularly recorded in response to leading and trailing edges of a 3 ° × 30 ° stripe simulating a worm and traversing the centers of their excitatory receptive fields (ERF) horizontally at a constant angular velocity in variable movement direction (temporo-nasal or naso-temporal).The behavioral contrast-direction dependent edge preferences are best resembled by the responses (R) of prey-selective class T5(2) neurons (Rl Rt=101 for b/w, 0.31 for w/b) and T7 neurons (RlRt=61 for b/w, 0.41 for w/b); the T7 responses may be dendritic spikes. This property can be traced back to off-responses dominated retinal class R3 neurons (RlRt=61 for b/w, 0.51 for w/b), but not to class R2 (RlRt =1.21 for b/w and 0.91 for w/b). The respective edge preference phenomena are independent of the direction of movement.When stimuli were moved against a stationary black-white structured background, the head preference to the black stripe and the tail preference to the white stripe were maintained in class R3, T5(2), and T7 neurons. If the stripe traversed the ERF together with the structured background in the same direction at the same velocity, the responses of tectal class T5(2) and T7 neurons were strongly inhibited, particularly in the former. Responses of retinal R2 neurons in comparable situations could be reduced by about 50%, while class R3 neurons responded to both the stimulus and the moving background structure.The results support the concept that the prey feature analyzing system in toads applies principles of (i) parallel and (ii) hierarchial information processing. These are (i) divergence of retinal R3 neuronal output contributes to stimulus edge positioning and (in combination with R2 output) area evaluation intectal neurons and to stimulus area evaluation and (in combination with R4 output) sensitivity for moving background structures inpre tectal neurons; (ii) convergence of tectal excitatory and pretectal inhibitory inputs specify the property of prey-selective tectal T5(2) neurons which are known to project to bulbar/spinal motor systems.Abbreviations ERF excitatory receptive field - IRF inhibitory receptive field - N nasal - T temporal - R w response to a worm-like stripe moving in the direction of its longer axis - R A response to an antiworm-like stripe whose longer axis is oriented perpendicular to the direction of movement - R l response to the leading edge of a worm-like moving stripe - R t response to the trailing edge of a worm-like moving stripe - b/w black stimulus against a white background - w/b white stimulus against a black background - sm structured moving background - ss structured stationary background - u minimal structure width of a structured background consisting of rectangular black and white patches in random distribution - HRP horseradish peroxidase  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