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 506 毫秒
1.
G Schwarz  J Seelig 《Biopolymers》1968,6(9):1263-1277
Dielectric relaxation of poly(γ-benzyl L -glutamate) in solution has been studied in the 5 kcps-10 Mcps range for various values of the helix content. The results give first experimental evidence for three effects of major significance. (1) The system exhibits dielectric relaxation due to a chemical rate process (namely helix formation). This confirms recent theoretical predictions. (2) The mean relaxation time τ* of the helix–coil transition could be evaluated as a function of the degree of transition. The results are in excellent agreement with a previously developed theory. At the midpoint of transition it is found τ*max = 5 × 10?7 sec. The elementary process of helical growth turns out to be practically diffusion-controlled (with a rate constant of hydrogen bond formation of 1.3 × 1010 sec?1). (3) There is a considerable electric field effect of the helix–coil transition. This indicates that conformation changes in biological systems could be potentially caused by direct action of an electric field.  相似文献   

2.
The thermodynamics of the hairpin helix-single strand transition of A6C6U6 has been analyzed by a staggering zipper model with consideration of single strand stacking. This analysis yields an enthalpy change of +11 kcal/mole for the formation of a first, isolated base pair. The stability constant of a first (intramolecular) base pair in A6C6U6 is around 2 × 1O?5 at 25°C, whereas a first (intermoleciilar) base pair in an A6 · U6 helix is characterised by a stability constant of about 4 × 10?3M?1 (25°C, extrapolated from An · Vn oligomer measurements). These data indicate a destabilizing effect of the C6 loop.The rate constant of hairpin helix formation is 2 to 3 × 104 sec?1 associated with an activation enthalpy of +2.5 kcal/mote. The rate of helix dissociation of the A6C6U6 hairpin is in the range of 103 to lO5 sec?1 with an activation enthalpy of 21 kcalmole. A comparison with the kinetic parameters obtained for A · U oligomer helices shows a specific influence of the C6 loop due to the stacking tendency of the cytosine residues. This intluence is preferentially reflected in the relatively low value of the rate constant of helix formation.  相似文献   

3.
A polarimetric electric-field-jump relaxation apparatus is described and used to determine the relaxation spectrum for the helix–coil transition of poly(α,L -glutamic acid) in water at 24°C. A maximum relaxation time of 1.7 μc occurs at the transition midpoint (pH = 5.9) yielding a rate constant for helical growth of 6 × 107 sec?1.  相似文献   

4.
The rate of conformational change of aqueous poly(α-L -lysine) solutions was measured using the electric field pulse relaxation method with conductivity detection. The relaxation time as a function of pH exhibits two maxima. One is assigned to a proton transfer reaction and the other to the helix–coil conformational transition. The helix nucleation parameter and the maximum relaxation time yield the rate constant of helix growth process (kF) according to Schwarz's kinetic theory as kF = 2 × 107 sec?1, which is comparable to that of the poly(glutamic acid) solution. The thermodynamic parameters of the helix growth process are compared with those of poly(glutamic acid).  相似文献   

5.
The kinetics of the coil-to-helix transition of (dG-dC)3 in M NaCl, 45 mM sodium cacodylate, pH 7, were measured in H2O, D2O, 10 mol % ethanol, 10 mol % urea, and 10 mol % glycerol. At 43°C in H2O the recombination rate is 1.3 ± 0.2 × 107 M?1 s?1; the dissociation rate is 68 ± 10 s?1. The destabilization of the helix in 10 mol % ethanol and 10 mol % urea relative to water is primarily due to a large increase in the helix-dissociation rate. In 10 mol % glycerol, the destabilization of the helix is due to a decrease in the recombination rate and an increase in the dissociation rate. Above 20°C, two exponential decays longer than 1 μs are observed after a temperature jump. The slower relaxation time is 4–10 times faster than the bimolecular component and is independent of oligomer concentration. We attribute this relaxation to a rapid equilibrium between two helical states. At low temperatures and oligomer concentrations of 1 mM or greater, the helices aggregate in 1M NaCl. Experimental data are presented under conditions where aggregation is unimportant and evidence is given that the ΔH-determined spectroscopically is unaffected by aggregation.  相似文献   

6.
D J Patel 《Biopolymers》1976,15(3):533-558
The Watson–Crick imino and amino exchangeable protons, the nonexchangeable base and sugar protons, and the backbone phosphates for d-CpG(pCpG)n, n = 1 and 2, have been monitored by high-resolution nmr spectroscopy in aqueous solution over the temperature range 0°–90°C. The temperature dependence of the chemical shifts of the tetramer and hexamer resonances is consistent with the formation of stable duplexes at low temperature in solution. Comparison of the spectral characteristics of the tetranucleotide with those of the hexanucleotide with temperature permits the differentiation and assignment of the cytosine proton resonances on base pairs located at the end of the helix from those in an interior position. There is fraying at the terminal base pairs in the tetranucleotide and hexanucleotide duplexes. The Watson–Crick ring imino protons exchange at a faster rate than the Watson–Crick side-chain amino protons, with exchange occurring by transient opening of the double helix. The structure of the d-CpG(pCpG)n double helices has been probed by proton relaxation time measurements, sugar proton coupling constants, and the proton chemical shift changes associated with the helix–coil transition. The experimental data support a structural model in solution, which incorporates an anti conformation about the glycosyl bonds, C(3) exo sugar ring pucker, and base overlap geometries similar to the B-DNA helix. Rotational correlation times of 1.7 and 0.9 × 10?9 sec have been computed for the hexanucleotide and tetranucleotide duplexes in 0.1 M salt, D2O, pH 6.25 at 27°C. The well-resolved 31P resonances for the internucleotide phosphates of the tetramer and hexamer sequences at superconducting fields shift upfield by 0.2–0.5 ppm on helix formation. These shifts reflect a conformational change about the ω,ω′ phosphodiester bonds from gauche-gauche in the duplex structure to a distribution of gauche-trans states in the coil structure. Significant differences are observed in the transition width and midpoint of the chemical shift versus temperature profiles plotted in differentiated form for the various base and sugar proton and internucleotide phosphorous resonances monitoring the d-CpG(pCpG)n helix–coil transition. The twofold symmetry of the d-CpGpCpG duplex is removed on complex formation with the antibiotic actinomycin-D. Two phosphorous resonances are shifted downfield by ~2.6 ppm and ~1.6 ppm on formation of the 1:2 Act-D:d-CpGpCpG complex in solution. Model studies on binding of the antibiotic to dinucleotides of varying sequence indicate that intercalation of the actinomycin-D occurs at the GpC site in the d-CpGpCpG duplex and that the magnitude of the downfield shifts reflects strain at the O-P-O backbone angles and hydrogen bonding between the phenoxazone and the phosphate oxygens. Actinomycin-D is known to bind to nucleic acids that exhibit a B-DNA conformation; this suggests that the d-CpG(pCpG)n duplexes exhibit a B-DNA conformation in solution.  相似文献   

7.
A direct measurement of the unzippering rate of a nucleic acid double helix   总被引:1,自引:0,他引:1  
The rate of double helix unzippering was determined directly by application of a fast temperature jump method to a nucleotide system of partly unzippered helices formed from oligoriboadenylates and oligoribouridylates of equal chain lengths (14 and 18 nucleotide residues). These helices showed a relaxation process in the time range of 0.1 to 0.3 μsec, that is assigned to the unzippering reaction. Measurements at 0.05 M and 0.1 M [Na+] demonstrated a rather small dependence upon the ionic strength. Increase of temperature increases the rate of unzippering. Simulation of the unzippering relaxation by a zipper model yielded a rate constant of base pair formation adjacent to a helix sequence of 8 × 106 sec?1 at 25°C associated with an activation enthalpy of 4 kcalmole. This elementary rate constant is higher than that obtained from a simulation of the overall recombination and dissociation rates of entire helices. The difference is attributed to reduced electrostatic and steric hindrance effects for base pair equilibration at helix ends.  相似文献   

8.
Absorbance-temperature profiles have been determined for the following self-complementary oligonucleotides or equimolar paris of complementary oligonucleotides containing GC base pairs: A2GCU2, A3GCU3, A4GCU4, A6CG + CGU6, A8CG + CGU8, A4G2 + C2U4, A5G2 + C2U5, A4G3 + C3U4, and A5G3 + C3U5. In all cases cooperative melting transitions indicate double-helix formation. As was found previously, the stability of GC containing oligomer helices is much higher than that of AU helices of corresponding length. Moreover, helices with the same length and base composition but different sequences also have quite different stabilites. The melting curves were andlyzed using a zipper model and the thermodynamic parameters for the AU pairs determined previously. The effect of single-strand stacking was considered separately. According to this model, the formation of a GC pair from unstacked single strands is associated with an ethalpy change of ?15 kcal/mole. Due to the high degree of single-strand stacking at room temperature the enthalpy change for the formation of GC pairs from unstacked single strands is only ?5 to ?6 kcal/mole. (The corresponding parameters for AU pairs are ?10.7 kcal/mole and ?5 to ?6 kcal/mole.) The sequence dependence of helix stability seems to be primarily entropic since no differences in ΔH were seen among the sequence isomers. The kinetics of helix formation was investigated for the same molecules using the temperature jump technique. Recombination of strands is second order with rate constants in the range of 105 to 107M?1 sec?1 depending on the chain length and the nucleotide sequence. Within a series of oligomers of a given type, the rates of recombination decrease with increasing chain length. Oligomers with the sequence AnGCUn recombine six to eight times slower than the other oligomers of corresponding chain length. The experimental enthalpies of activation of 6 to 9 kcal/mole suggest a nucleation length of one or two GC base pairs. The helix dissociation process has rate constants between 0.5 and 500 sec?1 and enthalpies of activation of 25 to 50 kcal/mole. An increase of chain length within a given nucleotide series leads to decreased rates of dissociation and increased enthalpies of activation. An investigation of the effect of ionic strength on AnGCUn helix formation showed that the rates of recombination increase considerably with increased ionic strength.  相似文献   

9.
The collagen-like peptides (L -Pro-L -Pro-Gly)n and (L -Pro-L -Hyp-Gly)n with n = 5 and 10, were examined in terms of their triple helix ? coil transitions in aqueous and nonaqueous solvents. The peptides were soluble in 1,2-propanediol containing 3% acetic acid and they were found to form triple-helical structures in this solvent system. The water content of the solvent system and the amount of water bound to the peptides were assayed by equilibrating the solvent with molecular sieves and carrying out Karl Fischer titrations on the solvent phase. After the solvent was dehydrated, much less than one molecule of water per tripeptide unit was bound to the peptides. Since the peptides remained in a triple-helical conformation, the results indicated that water was not an essential component of the triple-helical structure. Comparison of peptides with the same chain length demonstrated that the presence of hydroxyproline increased the thermal stability of the triple helix even under anhydrous conditions. The results, therefore, did not support recent hypotheses that hydroxyproline stabilizes the triple helix of collagen and collagen-like peptides by a specific interaction with water molecules. Analysis of the thermal transition curves in several solvent systems showed that although the peptides containing hydroxyproline had tm values which were 18.6° to 32.7°C higher, the effect of hydroxyproline on ΔG was only 0.1 to 0.3 kcal per tripeptide unit at 25°C. The results suggested, therefore, that the influence of hydroxyproline on helical stability may be explained by intrinsic effects such as dipole–dipole interactions or by changes in the solvation of the peptides by alcohol, acetic acid, and water. A direct calorimetric measurement of the transition enthalpy for (L -Pro-L -Pro-Gly)n in 3% or 10% acetic acid gave a value of ?1.84 kcal per tripeptide unit for the coil-to-helix transition. From the value for enthalpy and from data on the effects of different chain lengths on the thermal transition, it was calculated that the apparent free energy for nucleation was +5 kcal/mol at 25°C (apparent nucleation parameter = 2 × 10?4 M?2). The value was dependent on solvent and on chemical modification of end groups.  相似文献   

10.
K Suto  H Noda 《Biopolymers》1974,13(11):2391-2404
Measurements of the molecular weight of (Pro-Pro-Gly)n and (Pro-Pro-Gly)n(Ala-Pro-Gly)m(Pro-Pro-Gly)n, which were synthesized by the solid-phase method, revealed that they formed a trimer in an aqueous solution, and dissociated into single-stranded chains on warming. Accompanying the transition, a large decrease of optical rotation was observed, like the collagen–gelatin transition. The shape of the trimeric molecule was rodlike, and the dimensions were 12 Å in diameter and 2.8 Å per residue in length, regardless of the length of Ala-Pro-Gly sequences in a peptide chain. The data indicate that both Pro-Pro-Gly sequences and Ala-Pro-Gly sequences from the triple-helical structure similar to that of collagen in aqueous solution. All optical rotational dispersion (ORD) curves of solutions of the peptides were represented by a single-term Drude equation, and the Drude constant λc was 200 nm for all peptides regardless of the length of Ala-Pro-Gly sequences. The resemblance between the helical structure formed by Pro-Pro-Gly sequences and that by Ala-Pro-Gly sequences was also suggested by the formation of the hybrid triple helix from two kinds of peptide chains with different lengths of Ala-Pro-Gly sequences.  相似文献   

11.
Conformational analysis of triple helics of a type of collagen was performed with typical collagen tripeptide sequences based on Gly-Pro-Ala, Gly-Ala-Hyp, and Gly-Ala-Ala. During energy minimization, the possibility of continual deformation of the pyrrolidine cycle was taken into account in order to achieve better accuracy in the resulting structure. The (Gly-Pro-Ala)n structure is almost isomorphic to the (Gly-Pro-Hyp)n structure obtained in the previous work [Tumanyan, V. G. & Esipova, N.G. (1982) Biopolymers 21 , 475–497]. For a collagen-type structure, the optimal conformation of (Gly-Ala-Hyp)n tends to have a decreased unit twist (t = 15°), although the energy advantage with respect to the conformation with t = 45° is not so significant. A similar situation is observed for (Gly-Ala-Ala)n. In this case, the energy decrease during unwinding to t = 15° from t = 45° is quite small. The conformations of (Gly-Ala-Hyp)n and (Gly-Ala-Ala)n with t = 15° exhibit a similarity with a triple complex of polyproline II helices—a noncoiled coil such as (Gly-Pro-Hyp)n and (Gly-Pro-Ala)n. A similar structure may be postulated for subcomponent cq1 of the first component of a human complement containing substantial Gly-X-Pro and Gly-X-Y tripeptide derivatives in the primary structure (X, Y = any amino acid). The results suggest that the observed helical symmetry of collagen (t = 36°) is a consequence of superposition of diffraction patterns (for sufficiently long segments) from various helices (t varies from ~15° for Gly-X-Hyp and Gly-X-Y to ~56° for Gly-Pro-Ala). For short alternating segments, some unification of different helical structures is possible.  相似文献   

12.
The iodine which is added to an aqueous amylose solution is bound only partly by the amylose while forming the blue complex and partly remains free. The equilibrium normality of the free and the bound iodine at half-saturation of amylose by iodine is designated as [If]v and [Ib]w, respectively. The stability of the poly iodine chain formed within the axis of amylose helices depends on its length, i.e., indirectly on the DP of the amylose helices: the greater this stability, the lower the [If]v value. The amylose molecule consists of helical segments. Such a molecule may behave as a random coil. The average length of the helical segments in freshly prepared amylose-iodine complexes depends on temperature, pH, iodide concentration, the presence of other complex-forming agents, and the DP of the amylose. This latter factor is investigated in the present paper. By the aid of an automatically recording photometrictitrating device the coherent values of [Ib] and [If] were determined. Plotting these values against DP n for mechanochemically degraded as well as for periodateo-xidized amyloses resulted in curves consisting of two linear sections. The break of the curves occurred between DP n 110 and 130. It was concluded that below DP n = 100 the DP of helical segments (= sDP n) is identical to the DP n of the total molecule, i.e., the molecule consists of only a single, relatively stiff helix. Above this limit the molecule contains several helical segments. The DP of these helical segments can be calculated as follows: sDP n = 141.1 ? 10.2 × 105[If]v. This equation is considered to be valid for 0.5–0.6 mg. amylose in 100 ml. 0.1N HCl at 20°C., λ = 650 mμ, euuvet diameter 3.4 cm., the feed rate of the iodate-iodide titrating solution (in acid medium resulting in a 5 × 10?3N I2 solution with a molar iodide to iodine ratio of 1.5) is 0.4ml./min. Amylose molecules of, e.g., DP n = 1380 consist of an average of 11.4 segments having a DP of about 120 and consisting of an average of 15–18 helical turns.  相似文献   

13.
Nongelling solutions of structurally regular chain segments of agarose sulphate show disorder–order and order–disorder transitions (as monitored by the temperature dependence of optical rotation) that are closely similar to the conformational changes that accompany the sol–gel and gel–sol transitions of the unsegmented polymer. The transition midpoint temperature (Tm) for formation of the ordered structure on cooling is ~25 K lower than Tm for melting. Salt-induced conformational ordering, monitored by polarimetric stopped-flow, occurs on a millisecond time scale, and follows the dynamics expected for the process 2 coil ? helix. The equilibrium constant for helix growth (s) was calculated as a function of temperature from the calorimetric enthalpy change for helix formation (ΔHcal = ?3.0 ± 0.3 kJ per mole of disaccharide pairs in the ordered state), measured by differential scanning calorimetry. The temperature dependence of the nucleation rate constant (knuc), calculated from the observed second-order rate constant (kobs) by the relationship kobs = knuc(1 ? 1/s) gave the following activation parameters for nucleation of the ordered structure of agarose sulphate (1 mg mL?1; 0.5M Me4NCl or KCl): ΔH* = 112 ± 5 kJ mol?1; ΔS* = 262 ± 20 J mol?1 K?1; ΔG*298 = 34 ± 6 kJ mol?1; (knuc)298 = (7.5 ± 0.5) × 106 dm3 mol?1 s?1. The endpoint of the fast relaxation process corresponds to the metastable optical rotation values observed on cooling from the fully disordered form. Subsequent slow relaxation to the true equilibrium values (i.e., coincident with those observed on heating from the fully ordered state) was monitored by conventional optical rotation measurements over several weeks and follows second-order kinetics, with rate constants of (2.25 ± 0.07) × 10?4 and (3.10 ± 0.10) × 10?4 dm3 mol?1 s?1 at 293.7 and 296.2 K, respectively. This relaxation is attributed to the sequential aggregation processes helix + helix → dimer, helix + dimer → trimer, etc., with depletion of isolated helix driving the much faster coil–helix equilibrium to completion. Light-scattering measurements above and below the temperature range of the conformational transitions indicate an average aggregate size of 2–3 helices.  相似文献   

14.
The binding behavior of lysozyme with Al(III) is described using luminol as a luminescence probe by flow injection–chemiluminescence (FI–CL) analysis. It was found that the CL intensity of the luminol–lysozyme reaction could be markedly enhanced by Al(III), and the increase in CL intensity was linear with the Al(III) concentration over the range 0.3–30.0 pg mL?1, with a detection limit of 0.1 pg mL?1 (3σ). Based on the interaction model of lysozyme with Al(III), lg[(I ? I0)/(2I0 ? I)] = lgK + nlg[M], the binding constant K = 6.84 × 106 L mol–1 and the number of binding sites (n) = 0.76. The relative standard deviations were 3.2, 2.4 and 2.0% for 10.0, 20.0 and 30.0 pg mL?1 Al(III) (n = 7), respectively. This new method was successfully applied to continuous, quantitative monitoring of picogram level Al(III) in human saliva following oral intake of compound aluminum hydroxide tablets. It was found that Al(III) in saliva reached a maximum of 101.2 ng mL?1 at 3.0 h. The absorption rate constant ka, elimination rate constant k and half‐life time t1/2 of Al(III) were 1.378 h?1, 0.264 h?1 and 2.624 h, respectively. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

15.
Y Suezaki  N Go 《Biopolymers》1974,13(5):919-929
A theoretical analysis is given of the triple-helix–random-coil transition in a mixed solution of poly(Pro-Pro-Gly)n with two different but defined degrees of polymerization n and n′. Because of the highly cooperative nature of this helix–coil transition, each polypeptide chain tends to form a triple helix with other polypeptide chains with the same degree of polymerization (size recognition). Occurrence of triple helices consisting of polypeptide chains with different degrees of polymerization (error in recognition) is studied in detail as a function of the cooperativity, and n and n′. Implication of this analysis for molecular recognition in general is discussed.  相似文献   

16.
A 13C-nmr study of the salt-induced helix–coil transition of the basic polypeptides poly(L -lysine) [(Lys)n], poly(L -arginine) [(Arg)n], and poly (L -ornithine) [(Orn)n] was performed to serve as a reference of the helical portion of histones and other proteins. As is the case with pH-induced helix–coil transition, the downfield displacement of the Cα and carbonyl carbon signals are observed in the helical state. The upfield shift of the Cβ signals, on the other hand, is noted in the salt-induced transition. Regardless of the differences in the side chains and also the salts used, very similar helix-induced chemical shifts are obtained for (Lys)n and (Arg)n. However, the displacement of the Cα, Cβ, and carbonyl carbons of (Orn)n in the presence of 4M NaClO4 is found to be almost 50% of that of (Lys)n and (Arg)n. This is explained by the fact that the maximum helical content is about 50%, consistent with the ORD result. Further, the motion of the backbone and side chains of the helical from was estimated by measuring the spin-lattice relaxation time (T1), nuclear Overhauser enhancement (NOE), and line width. In the case of (Lys)n, the motion of the side chains is charged very little in comparison with that of the random coil. Indicating that the aggregation of the salt-induced helix is small in contrast to that of the pH-induced helix. For (Arg)n, however, the precipitate of the helical polymers is mainly due to aggregation.  相似文献   

17.
Electric birefringence measurements of suspensions of T3 and T7 bacteriophages in 10?2 M phosphate buffer, pH 6.9, show that there is a difference in their rotational diffusion coefficient. The values corrected to 25°C and water viscosity are D25,w = 4630 ± 130 sec?1 and D25,w = 5290 ± 260 sec?1 for T3 and T7, respectively. The value obtained from shell model calculations (according to Filson and Bloomfield) is D25,w = 4500 ± 600 sec?1. The apparent permanent dipole moments are 4.5 × 10?26 C·m and 1.7 × 10?26 C·m for T3 and T7, respectively. For both phage particles the intrinsic optical anisotropy is +7.2 × 10?3. It is shown that this anisotropy is mainly due to the DNA molecule inside the head of the phage. Its positive value means that there exists an excess orientation of the DNA helix perpendicular to the symmetry axis of the particle. For T7 an unexpectedly large increase of Δns and Ksp occurs at a glycerol concentration of about 30% (v/v). This increase is interpreted as being caused by a change of the shape of the particle and/or a change in the secondary structure of the DNA inside the head of the bacteriophage.  相似文献   

18.
A J Lomant  J R Fresco 《Biopolymers》1973,12(8):1889-1903
Stoichiometry and thermodynamic properties of polyadenylate–polyuridylate helices containing varying proportions of near-randomly distributed non-complementary uridine residues were charactrized from an analysis of their mixing curves and melting profiles measured at 259 nm and at appropriate longer wavelength isochromic points. The noncomplementary residues in this homopolymer–copolymer system (in which the homopolymer has the capacity to readjust with respect to the residues with which it is in opposition) show absolute preference for an extrahelical conformation even when situated in … AAUAA … sequences and must occur therefore as single loops. As the frequency of extrahelical residues in creases, the electrostatic energy of these complexes becomes greater, and is particularly severe for the three-stranded helices. Thus, an adenyl-ate-uridylate copolymer containing 35.2 mole percent uridine residues does not form a three-stranded complex with polyuridylate even in 0.7M Na+at O°C. The imperfections introduced into the helix lattice by extrahelical residues decrease the cooperativity of thermal denaturation as well as Tm. However, for the helices with extrahelical residues in low frequency (~1 per helix turn) only small increases in concentration of charge-neutralizing ions are required to bring Tm to the level of their perfect analogs. Two-stranded helices with a higher density of extra helical residues (~5 per helix turn) show [Na+] dependence of Tm characteristic of perfect three-stranded helices. These findings together with the absence of an effect of these imperfections on the hypochromicity per base-pair suggest only minimal disruption of helix continuity or distortion of stacking interactions that normally in volve the base pairs or triplets.  相似文献   

19.
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   

20.
Double-helical poly(dG-dC) and poly(dA-dT) are DNA analogs in which the interactions between the two strands of the helix are, respectively, either the stronger G/C type or the weaker A/T type along the entire length of macromolecules. Thus, these synthetic polynucleotides can be considered as representatives of the most stable and the least stable DNA. In the investigations presented here, potentiometric titrations and stopped-flow kinetic experiments were carried out in order to compare the pH-induced helix–coil conformations (10°C and 150mM [Na+]) the pH of the helix–coil transition (pHm) is 12.81 for poly(dG-dC) and 11.76 for poly(dA-dT). The unwinding of double-helical poly(dG-dC) initiated by a sudden change in pH was found to be a simple exponential process with rate constants in the range of 200–600 sec?1, depending on the final value of the pH jump. The intramolecular double-helix formation of poly(dG-dC) was studied by lowering the pH of the solutions from a value above pHm to that below pHm in dilute solutions (15.5 ug/ml [polymer]). Under these conditions, the observed rewinding reactions displayed a major and two exponential phases, all of which were independent of polymer concentration. From the comparison of the results of poly(dA-dT) and poly(dG-dT) would unwind faster than poly(dG-dC). However, if the pH jumps are such that they present the same perturbation of these polymers relative to their pHm values, no significant differences exist between the rates of helix–coil conformation changes of poly(dA-dT) and poly(dG-dC).  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