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.
The changes in polymer-solvent interactions that occur when native calf thymus DNA is dialyzed against Na2SO4 solutions of a given ionic strength and buffer concentration but of varying concentrations in methylmercuric hydroxide have been investigated with the help of solution density measurements at 25 °C and pH 6.8–7.0. From measurements executed under equilibrium dialysis conditions at the three salt levels 5 mm, 0.05 m, and 0.5 m Na2SO4 (m refers to molality) and in the presence of 5 mm cacodylic acid buffer, the density increments (???c2)μ0 for native calf thymus DNA were determined as a function of CH3HgOH concentration. (???c2)μ0 was found not to vary with organomercurial concentration, irrespective of the concentration of supporting electrolyte, until a certain CH3HgOH concentration level has been reached, viz., pM1 ? 3.5 (pM1 = ?log mCH3HgOH), beyond which (???c2)μ0 increases strongly with increasing concentration of CH3HgOH. As is shown by optical melting, (???c2)μ0 becomes a function of organomercurial concentration the moment DNA undergoes denaturation brought about by the complexing of CH3HgOH with the various N-binding sites of the base residues in the DNA double helix.Polymer-solvent interactions, expressed in terms of preferential water interactions (“net hydration”) and preferential salt interactions (“salt solvation”), were derived from the (???c2)μ0 data in combination with data obtained on the preferential interaction of CH3HgOH with denatured DNA and data on the partial specific volumes of all major solution components, gathered from density measurements on solutions with fixed concentrations of diffusible components. Evidence is presented which shows that denaturation in general decreases the net hydration while salt becomes preferentially associated with the polyelectrolyte. This process is further amplified by the interaction of CH3HgOH with denatured DNA: Methylmercurated DNA alters the redistribution of diffusible components at dialysis equilibrium to such an extent that in a formal sense large amounts of water are rejected from the immediate vicinity of the polymer. The molecular implications of these findings are explored. The results are further discussed in the light of previous findings where the methylmercury-induced denaturation of DNA had been studied with the help of buoyant density measurements in a Cs2SO4 density gradient and by velocity-sedimentation in a variety of sulfate media.  相似文献   

2.
Using guanidinium and n-butylammonium cations (C+) as models for the positively charged side chains in arginine and lysine, we have determined the association constants with various oxyanions by potentiometric titration. For a dibasic acid, H2A, three association complexes may exist: K1M = [CHA][C+] [HA?]; K1D = [CA?][C+] [A2?]; K2D = [C2A][C+] [CA?]. For guanidinium ion and phosphate, K1M = 1.4, K1D = 2.6, and K2D = 5.1. The data for carboxylates indicate that the basicity of the oxyanion does not affect the association constant: acetate, pKa = 4.8, K1M = 0.37; formate, pKa = 3.8, K1M = 0.32; and chloroacetate, pKa = 2.9, K1M = 0.43, all with guanidinium ion. Association constants are also reported for carbonate, dimethylphosphinate, benzylphosphonate, and adenylate anions.  相似文献   

3.
Alain Boussac  Anne Lise Etienne 《BBA》1984,766(3):576-581
In Tris-washed Photosystem-II particles we are able to induce an EPR signal in the dark by addition of an iridium salt (K2IrCl6). This signal is attributed to signal IIs (slow) (D+) and the redox titration gives an Em value of 760 mV for the couple D+D. On the basis of our previous studies on the equilibrium between D+Z and DZ+ (K = 104) (Boussac, A. and Etienne, A.L. (1982) Biochem. Biophys. Res. Commun. 109, 1200–1205), we therefore attribute a value of 1 V for the Em of the Z+Z couple. A second effect of K2IrCl6 is to modify the spectral characteristics of signal II. We conclude that K2IrCl6 is able to change the environment of the species from which signal IIs and signal IIf originate.  相似文献   

4.
The Rotational Isomeric States model is applied to calculate dipole moments of polypeptides of the twenty natural α-amino acids in the random coil state. Dipole moments of each repeat unit (μi), are evaluated using a quantum mechanics procedure. Dipole moment ratios (Dx = 〈μ2xμi2, x = number of repeat units) of homopolypeptides are calculated and extrapolated to x →?. With a few exceptions, D? = 0.36 ± 0.1. Ten actual proteins and three enzymes are also studied; their dipole ratios (Dx′ =〈μ〉/x) range from 7.34 to 10.57 in 10?59 C2 m2 (6.6–9.5 D2). Diffferences in the values of Dx′ are due mainly to the different contributions, μi, of the amino acid residues contained in each polymer, whereas the sequence of amino acids has a very minor effect.  相似文献   

5.
A class of indices that may be applied to quantitative data on nuclear families and that can help to assess degrees of mode of inheritance is developed. Given phenotype values of spouses x(1) and x(2) and offspring y, the deviation of an offspring value from the midparent is ¦y ? 12(x(1) + x(2), and those from the separate parents are ¦y ? x(1)¦ and ¦y ? x(2)¦. The indices called major-gene indices (MGI) investigated are functions of the deviations from midparental values compared to corresponding symmetric functions of the deviations from separate parents. Major-gene indices exceeding 1 may indicate some extent of major-gene inheritance, whereas an MGI less than 1 is suggestive of relatively more polygenic inheritance. Superposition of assortative mating and environmental effects will tend not to shift the MGI greater than 1 for polygenic inheritance, nor will they shift the MGI less than 1 for major-gene factors. The reliance on the proposed indices is reinforced on the basis of a hierarchy of representative models of monogenic and multifactorial inheritance. Extensions of the method to deal with multigenerational pedigrees are briefly discussed.  相似文献   

6.
7.
(1) H+/electron acceptor ratios have been determined with the oxidant pulse method for cells of denitrifying Paracoccus denitrificans oxidizing endogenous substrates during reduction of O2, NO?2 or N2O. Under optimal H+-translocation conditions, the ratios H+O, H+N2O, H+NO?2 for reduction to N2 and H+NO?2 for reduction to N2O were 6.0–6.3, 4.02, 5.79 and 3.37, respectively. (2) With ascorbate/N,N,N′,N′-tetramethyl-p-phenylenediamine as exogenous substrate, addition of NO?2 or N2O to an anaerobic cell suspension resulted in rapid alkalinization of the outer bulk medium. H+N2O, H+NO?2 for reduction to N2 and H+NO?2 for reduction to N2O were ?0.84, ?2.33 and ?1.90, respectively. (3) The H+oxidant ratios, mentioned in item 2, were not altered in the presence of valinomycinK+ and the triphenylmethylphosphonium cation. (4) A simplified scheme of electron transport to O2, NO?2 and N2O is presented which shows a periplasmic orientation of the nitrite reductase as well as the nitrous oxide reductase. Electrons destined for NO?2, N2O or O2 pass two H+-translocating sites. The H+electron acceptor ratios predicted by this scheme are in good agreement with the experimental values.  相似文献   

8.
5-hydroxylysine, an analogue of glutamate and lysine, causes NH4+ production by N2-fixing A. cylindrica; it also reversibly inhibits GS activity in vitro but has no effect on alanine dehydrogenase or GOGAT. On adding 5-hydroxylysine intracellular pools of glutamine, glutamate and aspartate decrease; those of alanine and serine increase. 5-hydroxylysine alleviates the inhibitory effect of NH4+ on heterocyst production and C2H2 reduction and in NH4+-grown cultures results in heterocyst synthesis and in C2H2 reduction. The data suggest that the GS-GOGAT pathway is the sole route of importance in primary NH4+ assimilation in A. cylindrica, that NH4+ alone does not inhibit nitrogenase and heterocyst production, and that GS and/or a product is involved in regulating the production of both.  相似文献   

9.
A method for calculating the rate constant (KA1A2) for the oxidation of the primary electron acceptor (A1) by the secondary one (A2) in the photosynthetic electron transport chain of purple bacteria is proposed.The method is based on the analysis of the dark recovery kinetics of reaction centre bacteriochlorophyll (P) following its oxidation by a short single laser pulse at a high oxidation-reduction potential of the medium. It is shown that in Ectothiorhodospira shaposhnikovii there is little difference in the value of KA1A2 obtained by this method from that measured by the method of Parson ((1969) Biochim. Biophys. Acta 189, 384–396), namely: (4.5±1.4) · 103s?1 and (6.9±1.2) · 103 s?1, respectively.The proposed method has also been used for the estimation of the KA1A2 value in chromatophores of Rhodospirillum rubrum deprived of constitutive electron donors which are capable of reducing P+ at a rate exceeding this for the transfer of electron from A1 to A2. The method of Parson cannot be used in this case. The value of KA1A2 has been found to be (2.7±0.8) · 103 s?1.The activation energies for the A1 to A2 electron transfer have also been determined. They are 12.4 kcal/mol and 9.9 kcal/mol for E. shaposhnikovii and R. rubrum, respectively.  相似文献   

10.
An electrostatic calculation suggests that when an ion is bound near the mouth of a channel penetrating a low-dielectric membrane, a counter ion may form an ion pair with this ion. The tendency towards ion-pair formation is remarkably enhanced at channel mouths by forces (image forces) arising from the charges induced on the boundaries between different dielectrics. The binding constant for the formation of ion-pairs of monovalent ions is estimated under the assumption that local interactions between the counter ion and the channel wall are negligibly small. It is of the order of 1–10 molal?1 or more for the binding of a Cl? (F?) counter ion to an Na+ (Li+) ion if appropriate conditions are fulfilled. The binding constant depends on the position of the binding site, the dimensions and geometries of the channel and channel mouth, and the state of ion loading of the channel, as well as the ionic species. The present results also indicate that when cation (anion) channels have anionic (cationic) groups as integrant parts of their channel walls, interactions between these charged groups and permeant ions are markedly enhanced by the image forces.  相似文献   

11.
Charles F Fowler  Bessel Kok 《BBA》1976,423(3):510-523
Using a rapid pH electrode, measurements were made of the flash-induced proton transport in isolated spinach chloroplasts. To calibrate the system, we assumed that in the presence of ferricyanide and in steady-state flashing light, each flash liberates from water one proton per reaction chain. We concluded that with both ferricyanide and methylviologen as acceptors two protons per electron are translocated by the electron transport chain connecting Photosystem II and I. With methyl viologen but not with ferricyanide as an acceptor, two additional protons per electron are taken up due to Photosystem I activity. One of these latter protons is translocated to the inside of the thylakoid while the other is taken up in H2O2 formation. Assuming that the proton released during water splitting remains inside the thylakoid, we compute H+e? ratios of 3 and 4 for ferricyanide and methyl viologen, respectively.In continuous light of low intensity, we obtained the same H+e? ratios. However, with higher intensities where electron transport becomes rate limited by the internal pH, the H+e? ratio approached 2 as a limit for both acceptors.A working model is presented which includes two sites of proton translocation, one between the photoacts, the other connected to Photosystem I, each of which translocates two protons per electron. Each site presents a ≈ 30 ms diffusion barrier to proton passage which can be lowered by uncouplers to 6–10 ms.  相似文献   

12.
Microelectrophoretic studies of the binding of a number of commonly used hydrophobic amine drugs to liposomes demonstrated the existence of relatively large surface potentials associated with binding of the protonated forms of the drugs. A theoretical treatment based on Langmuir adsorption isotherms and the Gouy-Chapman theory of the diffuse double layer allows estimation of drug-binding constants from electrophoretic mobility data. Such constants allow calculation of the charge effects arising from drug binding in more complex membrane systems, and it is shown that shifts in the apparent Ca2+ affinity of the (Ca2+ + Mg2+)-ATPase of sarcoplasmic reticulum in the presence of hydrophobic amine drugs are readily explicable in terms of the electrostatic effects of drug binding.  相似文献   

13.
The entropy-driven polymerization of tobacco mosaic virus protein is favored by an increase in ionic strength, μ, and by a decrease in pH. The effect of ionic strength is interpreted in terms of salting-out and electrical work, a function of charge and, therefore, of pH as well as of μ. The extent of polymerization is measured in terms of a characteristic temperature, T1, corresponding to a characteristic value of the equilibrium constant, KcT1 is measured at an early stage in the polymerization process where the optical density increment from light scatter is 0.01. The theory developed encompassing both salting-out and electrical work terms relates 1T1 to μ approximately according to the equation, 1T1 = C + Bμ ? Aμ12, where C is the ratio of entropy to enthalpy, B is proportional to the salting-out constant divided by enthalpy, and Aμ12 depends upon the square of the charge and is proportional to the electrical work contribution divided by the enthalpy. Data in which μ varied from 0.025 to 0.150 at three pH values, 5.95, 6.35, and 6.50, were fitted to this equation and the parameters C, B, and A were evaluated. Experiments were also carried out at a constant μ of 0.10 at pH values in increments of 0.1 between 5.9 and 6.8. The theory predicts that, at constant μ, 1T1, corrected for the electrical work contribution, is a linear function of pH with a negative slope proportional to the number of hydrogen ions bound per protein unit during polymerization, divided by the enthalpy. The data obtained fit two straight lines with different slopes above and below pH 6.3. Independent experiments carried out by the method of Stevens and Loga show that the number of hydrogen ions bound per protein unit also differs above and below pH 6.3 and the ratio of these is the same as the ratio of the above mentioned slopes. The data, therefore, make it possible to evaluate the enthalpy to be 24.8 kcal/mol of associating A protein and, with this value, the parameters C, B, and A can be interpreted. Standard entropies range from 86 e.u. at pH 6.5 to 88.5 at pH 5.95 and the salting-out constant, KS, is 2.2 at all pH values studied. At μ = 0.10, the values of the electrical work contribution at pH 5.95, 6.35, and 6.50 are +0.298, +0.455, and +0.534 kcal/mol, respectively. Theoretical calculations from models predict values in agreement within a factor of less than two.  相似文献   

14.
Proton inventory investigations of the hydrolysis N-acetylbenzotriazole at pH 3.0 (or the equivalent point on the pD rate profile) have been conducted at two different temperatures and at ionic strengths ranging from 0 to 3.0 M. The solvent deuterium isotope effects and proton inventories are remarkably similar over this wide range of conditions. The proton inventories suggest a cyclic transition state involving four protons contributing to the solvent deuterium isotope effect for the water-catalyzed hydrolysis. The hydrolysis data are described by the equation kn = ko (1 ? n + nπa1)4 with πa1 ~ 0.74, where ko is the observed first-order rate constant in protium oxide, n is the atom fraction of deuterium in the solvent, kn is the rate constant in a protium oxide-deuterium oxide mixture, and πa1 is the isotopic fractionation factor.  相似文献   

15.
The structural changes accompanying the recently described sub-transition of hydrated dipalmitoylphosphatidylcholine (Chen, S.C., Sturtevant, J.M. and Gaffney, B.J. (1980) Proc. Natl. Acad. Sci. USA 77, 5060–5063) have been defined using X-ray diffraction methods. Following prolonged storage at ?4°C the usual Lβ′ gel form of hydrated dipalmitoylphosphatidylcholine (DPPC) is converted into a more ordered stable ‘crystal’ form. The bilayer periodicity is 59.1 Å and the most striking feature is the presence of a number of X-ray reflections in the wide angle region. The most prominent of these are a sharp reflection at 14.4A??1 and a broader reflection at 13.9A??1. This diffraction pattern is indicative of more ordered molecular and hydrocarbon chain packing modes in this low temperature ‘crystal’ bilayer form. At the sub-transition (Trmsub = 15–20°C) an increase in the bilayer periodicity occurs (d=63.6 A?) and a strong reflection at approx. 14.2A??1 with a shoulder at approx. 14.1A??1 is observed. This diffraction pattern is identical to that of the bilayer gel (Lβ′) form of hydrated DPPC. Thus, the sub-transition corresponds to a bilayer ‘crystal’ → bilayer Lβ′ gel structural rearrangement accompanied by a decrease in the lateral hydrocarbon chain interactions. Differential scanning calorimetry and X-ray diffraction show that on further heating the usual structural changes Lβ′ → Pβ′ and Pβ′ → Lα occur at the pre- and main transitions, at approx. 35°C and 41°C, respectively.  相似文献   

16.
Joël Lunardi  Pierre V. Vignais 《BBA》1982,682(1):124-134
(1) N-4-Azido-2-nitrophenyl-γ-[3H]aminobutyryl-AdoPP[NH]P(NAP4-AdoPP[NH]P) a photoactivable derivative of 5-adenylyl imidodiphosphate (AdoPP[NH]P), was synthesized. (2) Binding of 3H]NAP4-AdoPP[NH]P to soluble ATPase from beef heart mitochrondria (F1) was studied in the absence of photoirradiation, and compared to that of [3H]AdoPP[NH]P. The photoactivable derivative of AdoPP[NH]P was found to bind to F1 with high affinity, like AdoPP[NH]P. Once [3H]NAP4-AdoPP[NH]P had bound to F1 in the dark, it could be released by AdoPP[NH]P, ADP and ATP, but not at all by NAP4 or AMP. Furthermore, preincubation of F1 with unlabeled AdoPP[NH]P, ADP, or ATP prevented the covalent labeling of the enzyme by [3H]NAP4-AdoPP[NH]P upon photoirradiation. (3) Photoirradiation of F1 by [3H]NAP4-AdoPP[NH]P resulted in covalent photolabeling and concomitant inactivation of the enzyme. Full inactivation corresponded to the binding of about 2 mol [3H]NAP4-AdoPP[NH]Pmol F1. Photolabeling by NAP4-AdoPP[NH]P was much more efficient in the presence than in the absence of MgCl2. (4) Bound [3H]NAP4-AdoPP[NH]P was localized on the α- and β-subunits of F1. At low concentrations (less than 10 μM), bound [3H]NAP4-AdoPP[NH]P was predominantly localized on the α-subunit; at concentrations equal to, or greater than 75 μM, both α- and β-subunits were equally labeled. (5) The extent of inactivation was independent of the nature of the photolabeled subunit (α or β), suggesting that each of the two subunits, α and β, is required for the activity of F1. (6) The covalently photolabeled F1 was able to form a complex with aurovertin, as does native F1. The ADP-induced fluorescence enhancement was more severely inhibited than the fluorescence quenching caused by ATP. The percentage of inactivation of F1 was virtually the same as the percentage of inhibition of the ATP-induced fluorescence quenching, suggesting that fluorescence quenching is related to the binding of ATP to the catalytic site of F1.  相似文献   

17.
The reactivities of anionic nitroalkanes with 2-nitropropane dioxygenase of Hansenula mrakii, glucose oxidase of Aspergillus niger, and mammalian d-amino acid oxidase have been compared kinetically. 2-Nitropropane dioxygenase is 1200 and 4800 times more active with anionic 2-nitropropane than d-amino acid oxidase and glucose oxidase, respectively. The apparent Km values for anionic 2-nitropropane are as follows: 2-nitropropane dioxygenase, 1.61 mm; glucose oxidase, 16.7 mm; and d-amino acid oxidase, 11.1 mm. Anionic 2-nitropropane undergoes an oxygenase reaction with 2-nitropropane dioxygenase and glucose oxidase, and an oxidase reaction with d-amino acid oxidase. In contrast, anionic nitroethane is oxidized through an oxygenase reaction by 2-nitropropane dioxygenase, and through an oxidase reaction by glucose oxidase. All nitroalkane oxidations by these three flavoenzymes are inhibited by Cu and Zn-superoxide dismutase of bovine blood, Mn-superoxide dismutases of bacilli, Fe-superoxide dismutase of Serratia marcescens, and other O2? scavengers such as cytochrome c and NADH, but are not affected by hydroxyl radical scavengers such as mannitol. None of the O2? scavengers tested affected the inherent substrate oxidation by glucose oxidase and d-amino acid oxidase. Furthermore, the generation of O2? in the oxidation of anionic 2-nitropropane by 2-nitropropane dioxygenase was revealed by ESR spectroscoy. The ESR spectrum of anionic 2-nitropropane plus 2-nitropropane dioxygenase shows signals at g1 = 2.007 and g11 = 2.051, which are characteristic of O2?. The O2? generated is a catalytically essential intermediate in the oxidation of anionic nitroalkanes by the enzymes.  相似文献   

18.
The ultimate rate of approach to equilibrium in the infinite stepping-stone model is calculated. The analysis is restricted to a single locus in the absence of selection, and every mutant is assumed to be new to the population. Let f(t, x) be the probability that two homologous genes separated by the vector x in generation t are the same allele. It is supposed that f(0, x) = O(x?2?η), η > 0, as x ≡ ¦ x ¦ → ∞. In the absence of mutation, f(t, x) tends to unity at the rate t?12 in one dimension and (ln t)?1 in two dimensions. Thus, the loss of genetic variability in two dimensions is so slow that evolutionary forces not considered in this model would supervene long before a two-dimensional natural population became completely homogeneous. If the mutation rate, u, is not zero f(t, x) asymptotically approaches equilibrium at the rate (1 ? u)2tt?32 in one dimension and (1 ? u)2tt?1(lnt)?2 in two dimensions. Integral formulas are presented for the spatial dependence of the deviation of f(t, x) from its stationary value as t → ∞, and for large separations this dependence is shown to be (const + x) in one dimension and (const + ln x) in two dimensions. All the results are the same for the Malécot model of a continuously distributed population provided the number of individuals per colony is replaced by the population density. The relatively slow algebraic and logarithmic rates of convergence for the infinite habitat contrast sharply with the exponential one for a finite habitat.  相似文献   

19.
The redox potentials for cytochrome c-552 at different ionic strengths, pH 7, have been determined, together with the thermodynamic parameters of the redox reaction. The effects of the electrostatic media on the redox potential of cytochrome c-552 do not depend on the nature of the ions employed. At 25 °C and pH 7 the observed potentials depend on the ionic strength, I, according to the equation: Eobso = 0.280 + .525 (I12(I + I12)). The significance of the ionic strength dependence of the redox potentials and their derived thermodynamic parameters are discussed and compared to those of mammalian cytochrome c. It is concluded that the redox potentials for ionic strength approaching zero are not affected by the overall net charge of the proteins; at finite ionic strengths, the protein charges play a very important role in determining the observed redox potentials.  相似文献   

20.
In an accompanying publication by Duckwitz-Peterlein, Eilenberger and Overath ((1977) Biochim. Biophys. Acta 469, 311–325) it is shown that the exchange of lipid molecules between negatively charged vesicles consisting of total phospholipid extracts from Escherichia coli occurs by the transfer of single lipid monomers or small micelles through the water. Here a kinetic interpretation is presented in terms of a rate constant, k?, for the escape of lipid molecules from the vesicle bilayer into the water. The evaluated rate constants are k?P = (0.86 ± 0.05) · 10?5s?1 and k?E = (1.09 ± 0.13) · 10?6s?1 for phospholipid molecules with trans-Δ9-hexadecenoate and trans-Δ9-octadecenoate, respectively, as the predominant acyl chain component. The rate constants are discussed in terms of the acyl chain and polar head group composition of the lipids.  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