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.
Bovine β-casein (β-CN) with its C-terminal truncated by chymosin digestion, β-CN-(f1-192), was examined and characterized using circular dichroism (CD) under various temperature conditions. CONTIN/LL analysis of the CD data revealed significant secondary structure disruption in β-CN-(f1-192) relative to its parent protein,β-CN, in the temperature range (5° to 70°C) studied. Near-UV CD spectra indicated significant temperature dependent structural changes. Analytical ultracentrifugation results showed significant reduction but not complete abolishment of self-association in β-CN-(f1-192) compared to whole β-casein at 2°–37°C. Furthermore, binding experiments with the common hydrophobic probe – 8-anilino-1- naphthalene sulfonic acid (ANS) illustrated that β-CN-(f1-192) is nearly incapable of binding to ANS relative to whole β-CN, suggesting a nearly complete open overall tertiary structure brought about by the C-terminal truncation. It has been demonstrated clearly that the tail peptide β-CN-(f193-209) is important in maintaining the hydrophobic core of β-CN but the residual association observed argues for a minor role for other sites as well.  相似文献   

2.
Studies of the binding of Ni2+ to adenylyl-3',5'-adenosine (ApA) at pH 6-0 by ultraviolet spectrophotometry indicate the formation of a 1:1 complex in the presence of a large excess of metal ion. At 25 °C. and ionic strength μ = 0.5 M, the stability constant of Ni(ApA) is evaluated to be K = 2.6 (±0.6) M?1. The low stability is taken as evidence that the predominant complex species is one in which the ApA acts as a monodentate ligand, mainly through the adenine group. The rate constants for complex formation and dissociation, kf = 1430 M?1 s?1 and kb = 665 s?1 (25°C. μ = 0.5M). determined by the temperature-jump relaxation technique, are consistent with this interpretation. The binding strength of Ni2+ to poly(adenylic acid) [poly(A)] has been studied at pH 7.0 using murexide as an indicator of the concentration of free Ni2+. Within the concentration range [Ni2+ = 1 × 10?5 × 10?3 M the data can be represented in the form of a linear Scatchard plot. i.e., the process can be described as the binding of Ni2+ to one class of independent binding sites. The number of binding sites per monomer is 0.26, and the stability constant K = 8.2×103 M?1 (25°C μ = 0.1 M). In kinetic studies of the reaction of Ni2+ with poly(A), two relaxation effects due to complex formation were detected, one with a concentration-independent time constant of about 0.4 ms, the other with a concentration-dependent time constant in the millisecond range. The concentration dependence of the longer relaxation time can be accounted for by a three-step mechanism which consists of a fast second-order association reaction followed by two first-order steps. There is evidence, however, that the overall process is more complicated than expressed by the three-step mechanism.  相似文献   

3.
Binding of ethidium to bacteriophage T7 and T7 deletion mutants   总被引:1,自引:0,他引:1  
Equilibrium binding of ethidium, quantitated by fluorescence enhancement, to DNA packaged in bacteriophage T7 and T7 deletion mutants has been compared with the binding of this dye to DNA released from its capsid (free DNA). During achievement of apparent equilibrium binding, no change in bacteriophage T7 structure occurred, by the criterion of agarose gel electrophoresis. However, excessive incubation with ethidium bromide caused detectable changes in bacteriophage structure, a possible explanation of disagreements in similar studies previously performed with T-even bacteriophages. Scatchard plots for packaged DNA had a curvature greater than the previously demonstrated [Bresloff, J. L. & Crothers, D. M. (1981) Biochemistry 20 , 3547–3553] curvature for free DNA. By treating plots for packaged DNA as though they were biphasic, it was found that binding to most sites occurred with an apparent association constant (Kap) 3.3–4.3 times lower than the Kap of free DNA. The number of these sites increased significantly as the density of packaged DNA was decreased by use of the deletion mutants. Values of ΔH° for these sites were negative and equal to the ΔH° for free DNA; values of ΔS° were positive and about half the ΔS° for free DNA. A second class of sites, roughly 1.2% of the total, had a significantly higher Kap and more negative ΔH° than those of the majority of sites.  相似文献   

4.
The basic theoretical groundwork for the use of derivative binding isotherms in the analysis of ligand binding is presented. The derivative binding isotherm is defined as Γ (Y) = df/dy where f = fractional degree of saturation and y = natural logarithm of the free ligand concentration. Since Γ (y) is a positive function, which goes to zero as y → ±∞, the mean value of y, 〈y〉, and the second and third moments, μ2 and μ3 about 〈y〉 are well defined. For a macromolecular system consisting of N equivalent and independent binding sites, Γ (y) is a symmetrical bell-shaped function with one maximum. The maximum occurs when y = ?ln Kassoc; μ2 = π2/3, and μ3 = 0. For multiple sets of independent binding sites, Γ (y) is a superposition of Γ-type functions. If the sets are sufficiently well separated in binding free energy, multiple extrema may be seen at positions corresponding to the logarithms of the dissociation constants for the individual sets. In any case, 〈y〉 is equal to the mean value of the logarithms of the dissociation constants for the sets; μ2 > π2/3 and equal to π2/3 plus the variance of the logarithms of the dissociation constants about their mean value; and μ3 is, except by coincidence, not equal to zero and equals the third moment of the distribution of logarithms of the dissociation constants about their mean value. Analysis of Γ(y) for the case of cooperative interactions within a set of binding sites was investigated by examining (1) the Hill model (whose mathematical representation is equivalent to that used to describe antibody heterogeneity except that in the latter case the parameter a, the Sips, constant, is constrained (0 < a ≤1);(2) a common model for cooperativity in which the cooperative free energy is a linear function of the fraction bound; and (3) a general representation of cooperative interactions within a set of sites in terms of ?(f), a smooth function that gives the interaction free energy in units of RT. For the Hill model (or Sips model) Γ(y) is a symmetrical function with one maximum at y = (?1)/a)lnK, μ2 = π2/3a2; and μ3 = 0. For the case in which the cooperative free energy is a linear function of f [?(f) = cf], 〈y〉 = ?ln K0 + (c/2); μ2 = (π2/3) + c[(c/12) + 1] where c > ?4; and μ3 = 0. General expressions for the moments in terms of ?(f) are derived. In general, μ2 < (π2/3) for positive cooperativity and μ2 > (π2/3) for negative for negative cooperativity. Γ(y) will be symmetrical if and only if the cooperative free energy is introduced symmetrically about f = 0.5.  相似文献   

5.
T W Sturgill 《Biopolymers》1978,17(7):1793-1810
A self-consistent thermodynamic characterization of the binding of ethidium to yeast phenylalanine-specific tRNA at 25°C, pH 7.0, in 11 nM MgCl2, 375 nM NaCl, and 25 mM sodium phosphate has been obtained. Two ethidium molecules bind per tRNA under these conditions. The stronger site has a dissociation constant equal to 1.9 ± 0.5 μM and ΔHdis°′ = 12 ± 1 Kcal/mol, and the weaker sites has a dissociation constant equal to 24 ± 9 μM and ΔHdis°′ = 8.9 ± 1.5 Kcal/mol. The average calorimetric ΔHdis°′ for the to sites 10.6 ± 0.4 kcal/mol. The thermodynamics of binding to the stranger sites are most probably the thermodynamics of interaction between A·U (6) and A·U (7), the unique site identified by Jones and Kearns. The binding is enthalpically driven and classical hydrophobic interactions do not appear to be important in the binding reaction.  相似文献   

6.
The thermodynamics of ethidium ion binding to the double strands formed by the ribooligonucleotides rCA5G + rCU5G and the analogous deoxyribo-oligonucleotides dCA5G + dCT5G were determined by monitoring the absorbance versus temperature at 260 and 283 nm at several concentrations of oligonucleotides and ethidium bromide. A maximum of three ethidium ions bind to the oligonucleotides, which is consistent with intercalation and nearest-neighbor exclusion. For the ribo-oligonucleotide the binding mechanism is complex. Either two sites (assumed to be the intercalation sites at the two ends of the oligonucleotide) bind more strongly by a factor of 140 than the third site, or all sites are identical, but there is strong anticooperativity on binding (cooperativity parameter, 0.1). In sharp contrast, the binding to the same sequence (with thymine substituted for uracil) in the deoxyribo-oligonucleotide showed all sites equivalent and no cooperativity. For the ribo-oligonucleotides the enthalpy for ethidium binding is ?14 kcal/mol. The equilibrium constants at 25°C depend on the model; either K = 6 × 105M?1 for the two strong sites (4 × 103M?1 for the weak site) or K = 2.5 × 105M?1 for the intrinsic constant of the anticooperative model. For the equivalent deoxyribo-oligonucleotide the enthalpy of binding is -9 kcal/mol and the equilibrium constant at 25°C is a factor of 10 smaller (K = 2.5 × 104M?1).  相似文献   

7.
Bovine brain hexokinase enhances the effect of Mn(II) on the longitudinal relaxation rate of water protons. Direct interaction of Mn(II) with the enzyme has been studied using electron spin resonance and proton relaxation rate enhancement methods. The results indicate that brain hexokinase has 1.05 ± 0.13 tight binding sites and 7 ± 2 weak binding sites with a dissociation constant, KD = 25 ± 4 μM and KD = 1050 ± 290 μM, respectively, at pH 8.0, 23 °C. The characteristic enhancement ?b) for hexokinase-Mn(II) complex evaluated from proton relaxation rate enhancement studies, gave ?b = 3.5 ± 0.4 for tight binding sites and an average ?b = 2.3 ± 0.5 per site for weak binding sites at 9 MHZ. The dissociation constant of Mn(II) for tight binding sites on the enzyme exhibits strong temperature dependence. In the low-temperature region (5–12 °C) brain hexokinase probably undergoes a conformational change. Frequency dependence of the normalized relaxation rate for bound water at various temperatures has shown that the number of exchangeable water molecules left in the first coordination sphere of bound Mn(II) is about one at 30 °C and about two at 18 °C. Binding of glucose 6-phosphate to hexokinase results in large-line broadening of the resonances of anomeric protons of the sugar. However, no such effect was observed in the case of glucose binding. These results suggest different modes of interaction of these two sugars to hexokinase. Line broadening of the C-(1) hydrogen resonances of glucose caused by Mn(II) in the presence of hexokinase suggests the proximity of the Mn(II) binding site to that of glucose. A lower limit of 1330 ± 170 s?1 for the rate of dissociation of glucose from enzyme-Mn(II)-glucose complex has been obtained from these studies.  相似文献   

8.
Variations in incubation temperature can markedly differentiate opiate receptor binding of agonists and antagonists. In the presence of sodium increasing incubation temperatures from 0° to 30° reduces receptor binding of 3H-naloxone by 50% while tripling the binding of the agonist 3H-dihydromorphine. Lowering incubation temperature from 25° to 0° reduces the potency of morphine in inhibiting 3H-naloxone binding by 9-fold while not affecting the potency of the antagonist nalorphine. At temperatures of 25° and higher the number of binding sites for opiate antagonists is increased by sodium and the number of sites for agonists is decreased by sodium with no changes in affinity. By contrast, in the presence of sodium lowering of incubation temperature to 0° increases opiate receptor binding of the antagonist naloxone by enhancing its affinity for binding sites even though the total number of binding sites are not changed.  相似文献   

9.
Common killifish Fundulus heteroclitus were acclimated to ecologically relevant temperatures (5, 15 and 33°C) and their maximum heart rate (fHmax) was measured at each acclimation temperature during an acute warming protocol. Acclimation to 33°C increased peak fHmax by up to 32% and allowed the heart to beat rhythmically at a temperature 10°C higher when compared with acclimation to 5°C. Independent of acclimation temperature, peak fHmax occurred about 3°C cooler than the temperature that first produced cardiac arrhythmias. Thus, when compared with previously published values for the critical thermal maximum of F. heteroclitus, the temperature for peak fHmax was cooler and the temperature that first produced cardiac arrhythmias was similar to these critical thermal maxima. The considerable thermal plasticity of fHmax demonstrated in the present study is entirely consistent with eurythermal ecology of killifish, as shown previously for another eurythermal fish Gillichthys mirabilis.  相似文献   

10.
We report for the first time the presence of a sex steroid-binding protein in the plasma of green sea turtles Chelonia mydas, which provides an insight into reproductive status. A high affinity, low capacity sex hormone steroid-binding protein was identified in nesting C. mydas and its thermal profile was established. In nesting C. mydas testosterone and oestradiol bind at 4°C with high affinity (K a = 1.49 ± 0.09 × 109 M−1; 0.17 ± 0.02 × 107 M−1) and low binding capacity (B max = 3.24 ± 0.84 × 10−5 M; 0.33 ± 0.06 × 10−4 M). The binding affinity and capacity of testosterone at 23 and 36°C, respectively were similar to those determined at 4°C. However, oestradiol showed no binding activity at 36°C. With competition studies we showed that oestradiol and oestrone do not compete for binding sites. Furthermore, in nesting C. mydas plasma no high-affinity binding was observed for adrenocortical steroids (cortisol and corticosterone) and progesterone. Our results indicate that in nesting C. mydas plasma temperature has a minimal effect on the high-affinity binding of testosterone to sex steroid-binding protein, however, the high affinity binding of oestradiol to sex steroid-binding protein is abolished at a hypothetically high (36°C) sea/ambient/body temperature. This suggests that at high core body temperatures most of the oestradiol becomes biologically available to the tissues rather than remaining bound to a high-affinity carrier.  相似文献   

11.
The coupling between the quaternary structure, stability and function of streptavidin makes it difficult to engineer a stable, high affinity monomer for biotechnology applications. For example, the binding pocket of streptavidin tetramer is comprised of residues from multiple subunits, which cannot be replicated in a single domain protein. However, rhizavidin from Rhizobium etli was recently shown to bind biotin with high affinity as a dimer without the hydrophobic tryptophan lid donated by an adjacent subunit. In particular, the binding site of rhizavidin uses residues from a single subunit to interact with bound biotin. We therefore postulated that replacing the binding site residues of streptavidin monomer with corresponding rhizavidin residues would lead to the design of a high affinity monomer useful for biotechnology applications. Here, we report the construction and characterization of a structural monomer, mSA, which combines the streptavidin and rhizavidin sequences to achieve optimized biophysical properties. First, the biotin affinity of mSA (Kd = 2.8 nM) is the highest among nontetrameric streptavidin, allowing sensitive monovalent detection of biotinylated ligands. The monomer also has significantly higher stability (Tm = 59.8°C) and solubility than all other previously engineered monomers to ensure the molecule remains folded and functional during its application. Using fluorescence correlation spectroscopy, we show that mSA binds biotinylated targets as a monomer. We also show that the molecule can be used as a genetic tag to introduce biotin binding capability to a heterologous protein. For example, recombinantly fusing the monomer to a cell surface receptor allows direct labeling and imaging of transfected cells using biotinylated fluorophores. A stable and functional streptavidin monomer, such as mSA, should be a useful reagent for designing novel detection systems based on monovalent biotin interaction. Biotechnol. Bioeng. 2013; 110: 57–67. © 2012 Wiley Periodicals, Inc.  相似文献   

12.
The root‐lesion nematode Pratylenchus thornei is a major pathogen of wheat and other field crops, particularly in the northern grain region of sub‐tropical eastern Australia. Research was conducted into the temperature requirements of P. thornei for reproduction on wheat to increase the reliability of selection in resistance tests for wheat breeding. Final population densities (Pf) of P. thornei were determined on four wheat cultivars (Gatcher, GS50a, Potam and Suneca) at fortnightly intervals from 8 to 18 weeks at a range of six soil temperatures (15°C, 20°C, 22.5°C, 25°C, 27.5°C and 30°C) in a glasshouse experiment. Pratylenchus thornei had the highest Pf in the temperature range of 20–25°C on all wheat cultivars at all growth times after sowing, with no nematode reproduction measured at 30°C and very little at 15°C. The wheat cv. GS50a consistently produced lower Pf than cvs Gatcher, Potam and Suneca in the optimum temperature range of 20–25°C. In carrot disc cultures, P. thornei had an optimum temperature of 25°C with little reproduction at 17.5°C and none detectable at 30°C. A standard soil temperature of 23°C was chosen to maximise differences in nematode reproduction between resistant and susceptible wheat genotypes for selection in wheat breeding, and to improve reproducibility among successive experiments. The relationships derived from these experiments will be valuable for simulation of P. thornei reproduction in crop growth models. They also indicate that early sowing of wheat into cool soil (≤15°C) in farmers' fields of the northern grain region should favour wheat growth over nematode reproduction and increase grain yield.  相似文献   

13.
The anion exchange system of human red blood cells is highly inhibited and specifically labeled by isothiocyano derivatives of benzene sulfonate (BS) or stilbene disulfonate (DS). To learn about the site of action of these irreversibly binding probes we studied the mechanism of inhibition of anion exchange by the reversibly binding analogs p-nitrobenzene sulfonic acid (pNBS) and 4,4′-dinitrostilbene-disulfonic acid (DNDS). In the absence of inhibitor, the self-exchange flux of sulfate (pH 7.4, 25°C) at high substrate concentration displayed self-inhibitory properties, indicating the existence of two anion binding sites: one a high-affinity transport site and the other a low-affinity modifier site whose occupancy by anions results in a noncompetitive inhibition of transport. The maximal sulfate exchange flux per unit area was JA = (0.69 ± 0.11) × 10-10 moles · min-1 · cm-2 and the Michaelis-Menten constants were for the transport site KS = 41 ± 14 mM and for the modifier site KS' = 653 ± 242 mM. The addition to cells of either pNBS at millimolar concentrations or DNDS at micromolar concentrations led to reversible inhibition of sulfate exchange (pH 7.4, 25°C). The relationship between inhibitor concentration and fractional inhibition was linear over the full range of pNBS or DNDS concentrations (Hill coefficient n ? 1), indicating a single site of inhibition for the two probes. The kinetics of sul- fate exchange in the presence of either inhibitor was compatible with that of competitive inhibition. Using various analytical techniques it was possible to determine that the sulfate trans- port site was the target for the action of the inhibitors. The in- hibitory constants (Ki j for the transport sites were 0.45 ± 0.10 PM for DNDS and 0.21 ± 0.07 mM for pNBS. From the similarities between reversibly and irreversibly binding BS and DS inhibitors in structures, chemical properties, modus oper- andi, stoichiometry of interaction with inhibitory sites, and relative inhibitory potencies, we concluded that the anion trans- port sites are also the sites of inhibition and of labeling of co- valent binding analogs of BS and DS.  相似文献   

14.
The binding of high specific activity, radioactive Concanavalin A to cultured normal human fibroblasts was investigated. We report the presence of two classes of Concanavalin A binding sites on the plasma membranes of these cells. These classes of binding sites are distinguished by their affinities for the lectin. Scatchard analysis of the binding data indicates the presence of a class of high affinity sites which are saturated at about 0.25 μg/ml of Concanavalin A. The other, lower affinity binding sites are not saturated until 50–100 μg/ml Concanavalin A levels are achieved. At 4°C the Ka for the high affinity sites varies between 1.5 – 5 × 109 M?1 depending on the method used to label the Concanavalin A. For the lower affinity sites Ka varies between 1 – 4 × 106 M?1. The average number of high affinity sites per cell is 8 × 105 representing less than 1% of the total receptor sites for the lectin.  相似文献   

15.
A study was made of the binding of a fluorescent probe K-35 (N-(carboxyphenyl)imide of 4-(dimethylamino)naphthalic acid), used as an indicator of albumin structural changes in pathology, to human serum albumin (HSA). Based on the data on the fluorescence decay of the probe, four types of site of K-35 binding to HSA have been recognized, which differ in fluorescence decay time (τ) and binding constant (K). Probe molecules bound to the first type of site have a decay time of 8–10 ns; this value corresponds to a high fluorescence quantum yield of about 0.7. These sites have a maximal binding constant, K 1 = 5 · 104 M−1. The τ2 of the second type of site is close to 3.6 ns and K 2 = 1 · 104 M−1, which is much lower than K 1; however, the number of these sites is several times greater. The number of sites of the third type and the binding constant are close to those of the second type, but the decay time τ3 is 1 ns, which is significantly lower than τ2. The binding of K-35 to sites of the second and the third types is characterized by a positive cooperativity. Their properties are similar but not completely identical. The total number of sites of these three types is about two per one HSA molecule. There are also one-two sites of the fourth type where bound K-35 molecules have a very short decay time τ4 ≪ 1, i.e., are virtually nonfluorescent, and K 4 = 1 · 104 M−1. The major contribution to the steady-state fluorescence is made by probe molecules bound to sites of the first and second types. As a rule, the concentration of albumin binding sites in blood is significantly higher than the concentration of metabolites and xenobiotics transferred by albumin. Therefore, the metabolite—or the probe in these experiments—is distributed among different sites in accordance with their K i n i values (n i is the number of sites of the i-th type per albumin molecule). The low occupancy of the sites results in an approximately equal number of K-35 molecules bound to different sites of types 1, 2, and 3. The competition of K-35 with phenylbutazone, a marker of the albumin drug-binding site I, allows one to suggest that the K-35 site of the first type is localized exactly in the drug site I region, while the sites of the second and third types are close to it.  相似文献   

16.
The kinetics of double-helix formation by poly U and the complementary monomer N-6,9-dimethyladenine (m6m9A) has been measured using a new fast temperature-jump apparatus. The cooperative binding kinetics are complicated by the extensive self-association of the monomers, but a satisfactory analysis using average relaxation times was possible in terms of three different models. Application of a model which considers only monomer binding yields the upper limit for the binding rate constant of an m6m9A monomer next to an already bound monomer on a poly U strand: (2 ± 0.4) × 108 M?1sec?1. A lower limit is found by using a model which allows for binding of all m6m9A stacks to poly U with equal rate constants: (3 ± 0.3) × 107 M?1sec?1. A third model with “weighted” rate constants consistent with the data: (7.5 ± 1.0) × 107 M?1sec?1. The rate of cooperative binding of m6m9A to the trimer UpUpU has also been measured. The rate constants obtained with the trimer agree with those obtained with the polymer for each of the three models within experimental error.  相似文献   

17.
The interaction of surfactant–cobalt(III) complexes [Co(bpy)(dien)TA](ClO4)3 · 3H2O (1) and [Co(dien)(phen)TA](ClO4)3 · 4H2O (2), where bpy = 2,2′‐bipyridine, dien = diethylenetriamine, phen = 1,10‐phenanthroline and TA = tetradecylamine with human serum albumin (HSA) under physiological conditions was analyzed using steady state, synchronous, 3D fluorescence, UV/visabsorption and circular dichroism spectroscopic techniques. The results show that these complexes cause the fluorescence quenching of HSA through a static mechanism. The binding constant (Kb) and number of binding‐sites (n) were obtained at different temperatures. The corresponding thermodynamic parameters (?G°, ?H° and ?S°) and Ea were also obtained. According to Förster's non‐radiation energy transfer theory, the binding distance (r) between the complexes and HSA were calculated. The results of synchronous and 3D fluorescence spectroscopy indicate that the binding process has changed considerably the polarity around the fluorophores, along with changes in the conformation of the protein. The antimicrobial and anticancer activities of the complexes were tested and the results show that the complexes have good activities against pathogenic microorganisms and cancer cells. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

18.
G Careri  A Giansanti  E Gratton 《Biopolymers》1979,18(5):1187-1203
By a combined gravimetric and ir technique, spectra of protein films are recorded during sorption isotherms at constant water content h (mg D2O/mg dry protein) in the range 0 les; h ? 0.35 at 27 and 38°C. Computer-aided differential analysis shows the effect of progressive hydration on some significant sites of the protein such as the ionizable acidic side chains and the backbone amide carbonyls, as well as the spectrum of the adsorbed water itself. In order to derive thermodynamic properties of these sites, the measured sorption isotherm is decomposed in terms of a model which postulates the existence of two classes of primary sorption sites only, and these two contributions are independently checked by the ir data. The free energy of binding of the strong and weak binding sites is found to be 2.0 ± 0.2 and 0.40 ± 0.1 kcal/mol, respectively. A water-induced transition region is clearly detected in all the observed properties at 0.06 < h < 0.10 at 38°C and is shown to be due to changes involving both the structure of the absorbed water and the coverage of the absorption sites. A detailed picture of the hydration events is offered, and the relevance of these findings to protein dynamics is discussed.  相似文献   

19.
N Sasaki 《Biopolymers》1984,23(9):1724-1734
The frequency dependences of the dielectric constant, ε′, and the loss factor, ε″, in collagen were measured at several water contents from 0.1 to 0.3 g/g collagen over a frequency range of 30 Hz to 100 kHz and at a temperature of 20°C. Remarkable dispersion was observed at the lower frequencies for higher water contents. According to accumulated results on the thermodynamic and structural investigations, the dispersion has some analogy to the surface conduction proposed by B. V. Hamon [(1953) Aust. J. Phys. 6 , 304–315]. An empirical relation bewteen ε″ and frequency, f, ε″ ∝? fn, where 0 < n < 1, suggests that the dielectric and conductive properties of hydrated collagen are related to carrier jumps between neighboring sites. For the polarization mechanism of this dispersion, we supposed a model of the transfer of protons between absorbed water molecules, and found that the time–water content superposition procedure is applicable to slightly hydrated collagen. The results derived from the superposition procedure show that the water content, ?, is related to the conductivity, σ, or the dielectric loss factor by the following equations: σ (?, f) = const. × ?nm?1f1?n and ε″ (?, f) = const. ?nmf?n, respectively, where m is a constant independent of ? and f. These results agree with that derived by another treatment of the same data. The role of water molecules in the conduction and polarization in slightly hydrated collagen is considered to be not far from that assumed in the model.  相似文献   

20.
Here, we show that heart rate in zebrafish Danio rerio is dependent upon two pacemaking mechanisms and it possesses a limited ability to reset the cardiac pacemaker with temperature acclimation. Electrocardiogram recordings, taken from individual, anaesthetised zebrafish that had been acclimated to 18, 23 or 28°C were used to follow the response of maximum heart rate (fHmax) to acute warming from 18°C until signs of cardiac failure appeared (up to c. 40°C). Because fHmax was similar across the acclimation groups at almost all equivalent test temperatures, warm acclimation was limited to one significant effect, the 23°C acclimated zebrafish had a significantly higher (21%) peak fHmax and reached a higher (3°C) test temperature than the 18°C acclimated zebrafish. Using zatebradine to block the membrane hyperpolarisation-activated cyclic nucleotide–gated channels (HCN) and examine the contribution of the membrane clock mechanisms to cardiac pacemaking, f Hmax was significantly reduced (by at least 40%) at all acute test temperatures and significantly more so at most test temperatures for zebrafish acclimated to 28°C vs. 23°C. Thus, HCN channels and the membrane clock were not only important, but could be modified by thermal acclimation. Using a combination of ryanodine (to block sarcoplasmic calcium release) and thapsigargin (to block sarcoplasmic calcium reuptake) to examine the contribution of sarcoplasmic reticular handling of calcium and the calcium clock, f Hmax was again consistently reduced independent of the test temperature and acclimation temperature, but to a significantly lesser degree than zatebradine for zebrafish acclimated to both 28 and 18°C. Thus, the calcium clock mechanism plays an additional role in setting pacemaker activity that was independent of temperature. In conclusion, the zebrafish cardiac pacemaker has a limited temperature acclimation ability compared with known effects for other fishes and involves two pacemaking mechanisms, one of which was independent of temperature.  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