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 DNA helix–coil transition has been studied in the presence of high concentrations of manganese ions (about 10?3M), which corresponds to the conditions close to equal stability of the A+T and G+C pairs, at the ionic strengths of 10?1, 10?2, and 1.6 × 10?3M Na+. With the Mn2+ ion effect, the transition range is significantly reduced to not more than 0.2°C at 1.2 × 10?3M Mn2+ and 1.6 × 10?3M Na+. The melting curves display a sharp kink at the end of the helix–coil transition, which is interpreted as an indication of the second-order phase transition. It is shown that the melting curves obtained can be approximated by a simple analytical expression 1 – θ = exp[–a(tc - t)], where θ is the DNA helix fraction, tc is the phase transition temperature, and a is an empirical parameter characterizing the breadth of the melting range and responsible for the magnitude of a jump of the helicity derivative with respect to the temperature at the phase transition point.  相似文献   

2.
An algorithm for the calculation of ligand-induced helix–coil transitions of macromolecules of any specific sequence was derived. The probabilities of each residue to be in helix and ligand-free, helix and ligand-bound, coil and ligand-free, and coil and ligand-bound states can be calculated by extending the recursion relation formula proposed by D. Poland [(1974) Biopolymers 13 , 1859–1871]. Calculations using hypothetical DNAs having block heterogeneities for melting reactions showed that cooperative binding specific to single strands weakens the effect of the heterogeneity. Fine structures in the melting profiles or cooperatively melting regions, which have so far been found in thermal melting experiments, were predicted to exist in ligand-induced melting reactions for natural DNA fragments.  相似文献   

3.
Thermodynamics of the B to Z transition in poly(dGdC)   总被引:1,自引:0,他引:1  
The thermodynamics of the B to Z transition in poly(dGdC) was examined by differential scanning calorimetry, temperature-dependent absorbance spectroscopy, and CD spectroscopy. In a buffer containing 1 mM Na cacodylate, 1 mM MgCl2, pH 6.3, the B to Z transition is centered at 76.4°C, and is characterized by ΔHcal = 2.02 kcal (mol base pair)?1 and a cooperative unit of 150 base pairs (bp). The tm of this transition is independent of both polynucleotide and Mg2+ concentrations. A second transition, with ΔHcal = 2.90 cal (mol bp)?1, follows the B to Z conversion, the tm of which is dependent upon both the polynucleotide and the Mg2+ concentrations. Turbidity changes are concomitant with the second transition, indicative of DNA aggregation. CD spectra recorded at a temperature above the second transition are similar to those reported for ψ(–)-DNA. Both the B to Z transition and the aggregation reaction are fully and rapidly reversible in calorimetric experiments. The helix to coil transition under these solution conditions is centered at 126°C, and is characterized by ΔHcal = 12.4 kcal (mol bp)?1 and a cooperative unit of 290 bp. In 5 mM MgCl2, a single transition is seen centered at 75.5°C, characterized by ΔHcal = 2.82 kcal (mol bp)?1 and a cooperative unit of 430 bp. This transition is not readily reversible in calorimetric experiments. Changes in turbidity are coincident with the transition, and CD spectra at a temperature just above the transition are characteristic of ψ(–)-DNA. A transition at 124.9°C is seen under these solution conditions, with ΔHcal = 10.0 kcal (mol bp)?1 and which requires a complex three-step reaction mechanism to approximate the experimental excess heat capacity curve. Our results provide a direct measure of the thermodynamics of the B to Z transition, and indicate that Z-DNA is an intermediate in the formation of the ψ-(–) aggregate under these solution conditions.  相似文献   

4.
R L Ornstein  J R Fresco 《Biopolymers》1983,22(8):1979-2000
Tm values for 20 DNA duplexes with different repeating base sequences provided the data base for developing a rational and relatively simple methodology for computing apparent enthalpies for the helix → coil transitions of DNA helices, ΔH calc. The computational variables and their range of acceptable values were selected on the basis of physically plausible arguments. Over 350,000 different combinations of the variables were tested for degree of fit. It was therby possible to find a combination giving a high degree of linear fit between Tm and ΔH calc (correlation coefficient, 0.99), with Tm values deviating (on average) from the regression line by only ±2.17°C. Most of this uncertainty is attributed to experimental limitations, although computational approximations also contribute. With ΔH calc for the melting of each of the unique complementary dinucleotide fragments computed by the method developed, it is possible to estimate Tm and (relative) ΔH calc reliable for the melting of any particular DNA [with base pairs G(I)·C and A·T] given only its base sequence. The ΔHcalc values for the complementary dinucleotide fragments, together with statistical considerations, make it apparent that Tm of DNAs with repeating base sequence show only an approximate linear dependence on G·C content because A·T and G·G pairs do not contribute to helix stability independently of the base-pair sequence in which they occur. In fact, the nearestneighbor stacking interactions are so significant that certain complementary dinucleotide fragment sequences with 0,50, and 100% G·C content have the same stability.  相似文献   

5.
A S Benight  R M Wartell 《Biopolymers》1983,22(5):1409-1425
Theoretical melting curves were calculated for four DNA restriction fragments, 157–257 base pairs (bp), and a series of hypothetical block DNAs with sequences d(C2xAxC2x). d(C2xTxG2x), 5 ? x ? 40. These DNAs provided a mixture of A·T/G·C sequence distributions with which to investigate the effects of parameters and base-pair changes on the melting of short DNAs. The sensitivity of DNA melting curves to changes in internal loop melting parameters σ and κ was examined. As Expected, theoretical melting curves of short DNAs with a quasirandom base-pair sequence vary little with changes in internal loop parameters. End melting dominates the transition behaviour of these moleucles. This was also observed for the block DNAs up to x = 22. Beyond this length, melting curves are highly sensitive to the internal loop parameters. Sensitivity is also predicted for a 157-bp fragment with a block distribution of A·T and G·C pairs. These results indicate that accurate evaluation of internal loop parameters is possible with short DNAs (100–200 bp) containing a G·C/A·T/G·C block distribution with at least 22 bp in each block. Duplex-to-single-strands dissociation parameters were reevaluated form experimental melting curve data of eight DNA fragments using a least squares fit approach. This analysis confirmed parameter values previously found with a simplified dissociation model. A Priori predictions are made on the effects of base-pair changes on the melting curves of three characterized DNA restriction fragments. Single base-pair changes are predicted to induce small but measurable changes in the melting curves. The characteristics of the altered melting curves depend on the location of the base-pair change.  相似文献   

6.
The Poland–Fixman–Freire formalism was adapted for modeling of calorimetric DNA melting profiles, and applied to plasmid pBR 322 and long random sequences. We studied the influence of the difference (HGC?HAT) between the helix‐coil transition enthalpies of AT and GC base pairs on the calorimetric melting profile and on normalized calorimetric melting profile. A strong alteration of DNA calorimetrical profile with HGC?HAT was demonstrated. In contrast, there is a relatively slight change in the normalized profiles and in corresponding ordinary (optical) normalized differential melting curves (DMCs). For fixed HGC?HAT, the average relative deviation (S) between DMC and normalized calorimetric profile, and the difference between their melting temperatures (Tcal?Tm) are weakly dependent on peculiarities of the multipeak fine structure of DMCs. At the same time, both the deviation S and difference (Tcal?Tm) enlarge with the temperature melting range of the helix‐coil transition. It is shown that the local deviation between DMC and normalized calorimetric profile increases in regions of narrow peaks distant from the melting temperature.  相似文献   

7.
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   

8.
Many factors that change the temperature position and interval of the DNA helix–coil transition often also alter the shape of multi-peak differential melting curves (DMCs). For DNAs with a multi-peak DMC, there is no agreement on the most useful definition for the melting temperature, Tm, and temperature melting width, ΔT, of the entire DNA transition. Changes in Tm and ΔT can reflect unstable variation of the shape of the DMC as well as alterations in DNA thermal stability and heterogeneity. Here, experiments and computer modeling for DNA multi-peak DMCs varying under different factors allowed testing of several methods of defining Tm and ΔT. Indeed, some of the methods give unreasonable “jagged” Tm and ΔT dependences on varying relative concentration of DNA chemical modifications (rb), [Na+], and GC content. At the same time, Tm determined as the helix–coil transition average temperature, and ΔT, which is proportional to the average absolute temperature deviation from this temperature, are suitable to characterize multi-peak DMCs. They give smoothly varying theoretical and experimental dependences of Tm and ΔT on rb, [Na+], and GC content. For multi-peak DMCs, Tm value determined in this way is the closest to the thermodynamic melting temperature (the helix–coil transition enthalpy/entropy ratio).  相似文献   

9.
Y C Fu  H V Wart  H A Scheraga 《Biopolymers》1976,15(9):1795-1813
The enthalpy change associated with the isothermal pH-induced uncharged coil-to-helix transition ΔHh° in poly(L -ornithine) in 0.1 N KCl has been determnined calorimetrically to be ?1530 ± 210 and ?1270 ± 530 cal/mol at 10° and 25°C, respectively. Titration data provided information about the state of charge of the polymer in the calorimetric experiments, and optical rotatory dispersion data about its conformation. In order to compute ΔHh°, the observed calorimetric heat was corrected for the heat of breaking the sample cell, the heat of dilution of HCl, the heat of neutralization of the OH? ion, and the heat of ionization of the δ-amino group in the random coil. The latter was obtained from similar calorimetric measurements on poly(D ,L -ornithine). Since it was discovered that poly(L -ornithine) undergoes chain cleavage at high pH, the calorimetric measurements were carried out under conditions where no degradation occurred. From the thermally induced uncharged helix–coil transition curve for poly(L -ornithine) at pH 11.68 in 0.1 N KCl in the 0°–40°C region, the transition temperature Ttr and the quantity (?θh/?T)Ttr have been obtained. From these values, together with the measured values of ΔHh°, the changes in the standard free energy ΔGh° and entropy ΔGh°, associated with the uncharged coil-to-helix transition at 10°C have been calculated to be ?33 cal/mol and ?5.3 cal/mol deg, respectively. The value of the Zimm–Bragg helix–coil stability constant σ has been calculated to be 1.4 × 10?2 and the value of s calculated to be 1.06 at 10°C, and between 0.60 and 0.92 at 25°C.  相似文献   

10.
The vacuum-ultraviolet circular dichroism (VUCD) of chondroitin and chontroitin-6-sulfate has been measured to 160 nm for films and to 170 nm for D2O solutions. The pD-dependent dichroic behavior of these glycosaminoglycans in D2O is similar above 200 nm and is in agreement with previous studies. Near 190 nm, the CD band sign is also dependent on pD. VUCD spectra were recorded for films and solutions of poly(L -arginine). In trifluoroethanol the polypeptide is α-helical, while in D2O it exists as a random coil. The well-characterized coil–helix transition of poly(L -arginine) during complexation with chondroitin-6-sulfate was observed by VUCD, including the previously inaccessible entire π → π* band. By construction of difference spectra it was also possible to monitor the VUCD of the polysaccharide component during complexation.  相似文献   

11.
12.
The binding of neomycin sulfate (NS)/paromomycin sulfate (PS) with DNA was investigated by fluorescence quenching using acridine orange (AO) as a fluorescence probe. Fluorescence lifetime, FT-IR, circular dichroism (CD), relative viscosity, ionic strength, DNA melting temperature, and molecular docking were performed to explore the binding mechanism. The binding constant of NS/PS and DNA was 6.70 × 103/1.44 × 103 L mol?1 at 291 K. The values of ΔHθ, ΔSθ, and ΔGθ suggested that van der Waals force or hydrogen bond might be the main binding force between NS/PS and DNA. The results of Stern–Volmer plots and fluorescence lifetime measurements all revealed that NS/PS quenching the fluorescence of DNA–AO was static in nature. FT-IR indicated that the interaction between DNA and NS/PS did occur. The relative viscosity and melting temperature of DNA were almost unchanged when NS/PS was introduced to the solution. The fluorescence intensity of NS/PS–DNA–AO was decreased with the increase in the ionic strength. For CD spectra of DNA, the intensity of positive band at nearly 275 nm was decreased and that of negative band at nearly 245 nm was increased with the increase in the concentration of NS/PS. The binding constant of NS/PS with double-stranded DNA (dsDNA) was larger than that of NS/PS with single-stranded DNA (ssDNA). From these studies, the binding mode of NS/PS with DNA was evaluated to be groove binding. The results of molecular docking further indicated that NS/PS could enter into the minor groove in the A–T rich region of DNA.  相似文献   

13.
CD spectra and melting curves were collected for a 28 base-pair DNA fragment in the form of a DNA dumbbell (linked on both ends by T4 single-strand loops) and the same DNA sequence in the linear form (without end loops). The central 16 base pairs (bp) of the 28-bp duplex region is the poly(pu) sequence: 5′-AGGAAGGAGGAAAGAG-3′. Mixtures of the dumbbell and linear DNAs with the 16-base single-strand sequence 5′-TCCTTCCTCCTTTCTC-3′ were also prepared and studied. At 22°C, CD measurements of the mixtures in 950 mM NaCl, 10 mM sodium acetate, 1 mM EDTA, pH 5.5, at a duplex concentration of 1.8 μM, provided evidence for triplex formation. Spectroscopic features of the triplexes formed with either a dumbbell or linear substrate were quite similar. Melting curves of the duplex molecules alone and in mixtures with the third strand were collected as a function of duplex concentration from 0.16 to 2.15 μM. Melting curves of the dumbbell alone and mixtures with the third strand were entirely independent of DNA concentration. In contrast, melting curves of the linear duplex alone or mixed with the third strand were concentration dependent. At identical duplex concentrations, the dumbbell alone melts ~20°C higher than the linear duplex. The curve of the linear duplex displayed a significant pretransition probably due to end fraying. On melting curves of mixtures of the dumbbell or linear duplex with the third strand, a low temperature transition with much lower relative hyperchromicity change (~ 5%) was observed. This transition was attributed to the melting of a new molecular species, e.g., the triplex formed between the duplex and single-strand DNA molecules. In the case of the dumbbell/single-strand mixture, these melting transitions of the triplex and the dumbbell were entirely resolvable. In contrast, the melting transitions of the linear duplex and the triplex overlapped, thereby preventing their clear distinction. To analyze the data, a three-state equilibrium model is presented. The analysis utilizes differences in relative absorbance vs temperature curves of dumbbells (or linear molecules) alone and in mixtures with the third strand. From the model analysis a straightforward derivation of fT(T), the fraction of triplex as a function of temperature, was obtained. Analysis of fT vs temperature curves, in effect melting curves of the triplexes, provided evaluation of thermodynamic parameters of the melting transition. For the triplex formed with the dumbbell substrate, the total transition enthalpy is ΔHT = 118.4 ± 12.8 kcal/mol (7.4 ± 0.8 kcal/mol per triplet unit) and the total transition entropy is ΔST = 344 ± 36.8 cal/K · mol (eu) (21.5 ± 2.3 eu per triple unit). The transition curves of the triplex formed with the linear duplex substrate displayed two distinct regions. A broad pretransition region from fT = 0 to 0.55 and a higher, sharper transition above fT = 0.55. The transition parameters derived from the lower temperature region of the curve are ΔHT = 44.8 ± 9.6 kcal/mol and ΔST = 112 ± 33.6 eu (or ΔH′ = 2.8 ± 0.6 kcal/mol and ΔS′ = 7.0 ± 2.1 eu per triplet). These values are probably too small to correspond to actual melting of the triplex but instead likely reveal effects of end fraying of the duplex substrate on triplex stability. Transition parameters of the upper transition are ΔHT = 128.0 ± 2.3 kcal/mol and ΔST = 379.2 ± 6.4 eu (ΔH′ = 8.0 ± 0.2 kcal/mol and ΔS′ = 23.7 ± 0.4 eu per triplet) in good agreement (within experimental error) with the transition parameters of the triplex formed with the dumbbell substrate. Supposing this upper transition reflects actual dissociation of the third strand from the linear duplex substrate this triplex is comparable in thermodynamic stability to the triplex formed with a dumbbell substrate. Even so, the biphasic melting character of the linear triplex obscures the whole analysis, casting doubt on its absolute reliability. Apparently triplexes formed with a dumbbell substrate offer technical advantages over triplexes formed from linear or hairpin duplex substrates for studies of DNA triplex stability. © 1993 John Wiley & Sons, Inc.  相似文献   

14.
C Zimmer  H Triebel 《Biopolymers》1969,8(5):573-593
Reversible and irreversible conformational changes in the acid-induced denaturation of DNA were studied by spectrophotometric titration, sedimentation, and melting measurements. A GC-rich DNA (72 mole-%) shows complete or partial reversibility of the titration profiles within the pH region of transition from helix to coil, while AT-rich DNA (29 mole-%) is irreversible in its titration behavior at each acid pH below the onset of the transition. The results for GC-rich DNA further indicate distinct differences in the titration behavior, which can be attributed to differences in the frequency of GC clusters along the DNA molecule. Plots of the sedimentation coefficient and the parameter asapp against pH lead to the conclusion that conformational changes occur before the onset of the acid-induced helix–coil transition. These alterations are more pronounced upon protonation of larger GC-rich domains than of smaller ones, as concluded from very marked differences observed in the sedimentation–pH behavior of two GC-rich DNA's. An acid denaturation scheme for a GC-rich DNA segment is suggested. Reversibility of the acid denaturation is explained by the existence of stable, protonated, single GC base pairs in nonprotonated stacked single-stranded domains formed in the acid-induced transition region.  相似文献   

15.
The melting transition of DNA in alkaline CsCl can be followed in the analytical ultracentrifuge. Equilibrium partially denatured states can be observed. These partially denatured DNA bands have bandwidths of up to several times those of native DNA. Less stable molecules melt early and are found at heavier densities in the melting region. An idealized ultracentrifuge melting transition is described. The melting transition of singly nicked PM-2 DNA resembles the idealized curve. The DNA profile is a Gaussian band at all points in the melt. DNA's from mouse, D. Melanogaster, M. lysodeikticus, T4, and T7 also show equilibrium bands at partially denatured densities, some of which are highly asymmetric. Simple sequence satellite DNA shows an all-or-none transition with no equilibrium bands at partially denatured densities. The temperature at which a DNA denatures is an increasing function of the (G + C) content of the DNA. The Tm does not show a molecular-weight dependence in the range 1.2 × 106–1.5 × 107 daltons (single strand) for mouse, M. lysodeikticus, or T4 DNA. The mouse DNA partially denatured bands do not change shape as a function of molecular weight. The T4 DNA intermediate band develops a late-melting tail at low molecular weight. M. lysodeikticus DNA bands at partially denatured densities become broader as the molecular weight is decreased. Mouse DNA is resolved into six Gaussian components at each point in the melting transition.  相似文献   

16.
The CD spectra of the peptides Boc-X-(Aib-X)n-OMe (n = 1, 2, 3) and Boc-(Aib-X)5-OMe, where X = L -Ala or L -Val have been examined in several solvents. The X = Ala and Val peptides behave similarly in all solvents, suggesting that the Aib residues dominate the folding preferences of these peptides. The decapeptides adopt helical conformations in methanol and trifluoroethanol, with characteristic negative CD bands at 222 and 205 nm. In the heptapeptides, similar spectra with reduced intensities are observed. Comparison with nmr studies suggest that estimates of helical content in oligopeptides by CD methods may lead to erroneous conclusions. The pentapeptides yield solvent-dependent spectra indicative of conformational perturbations. Peptide association in dioxane results in an unusual spectrum with a single negative band at 210 nm for the decapeptides. Disaggregation is induced by the addition of methanol or water to dioxane solutions. Aggregation of the heptapeptides is less pronounced in dioxane, suggesting that a critical helix length may be necessary to promote association stabilized by helix dipole–dipole interactions.  相似文献   

17.
Interaction between polylysine and DNA's of varied G + C contents was studied using thermal denaturation and circular dichroism (CD). For each complex there is one melting band at a lower temperature tm, corresponding to the helix–coil transition of free base pairs, and another band at a higher temperature tm, corresponding to the transition of polylysine-bound base pairs. For free base pairs, with natural DNA's and poly(dA-dT) a linear relation is observed between the tm and the G + C content of the particular DNA used. This is not true with poly(dG)·poly(dC), which has a tm about 20°C lower than the extrapolated value for DNA of 100% G + C. For polylysine-bound base pairs, a linear relation is also observed between the tm and the G + C content of natural DNA's but neither poly(dA-dT) nor poly(dG)·poly(dC) complexes follow this relationship. The dependence of melting temperature on composition, expressed as dtm/dXG·C, where XG·C is the fraction of G·C pairs, is 60°C for free base pairs and only 21°C for polylysine-bound base pairs. This reduction in compositional dependence of Tm is similar to that observed for pure DNA in high ionic strength. Although the tm of polylysine-poly(dA-dT) is 9°C lower than the extrapolated value for 0% G + C in EDTA buffer, it is independent of ionic strength in the medium and is equal to the tm0 extrapolated from the linear plot of tm against log Na+. There is also a noticeable similarity in the CD spectra of polylysine· and polyarginine·DNA complexes, except for complexes with poly(dA-dT). The calculated CD spectrum of polylysine-bound poly(dA-dT) is substantially different from that of polyarginine-bound poly(dA-dT).  相似文献   

18.
Spectroscopic and calorimetric melting studies of 28 DNA hairpins were performed. These hairpins form by intramolecular folding of 16 base self‐complementary DNA oligomer sequences. Sequence design dictated that the hairpin structures have a six base pair duplex linked by a four base loop and that the first five base pairs in the stem are the same in every molecule. Only loop sequence and identity of the duplex base pair closing the loop vary for the set of hairpins. For these DNA samples, melting studies were carried out to investigate effects of the variables on hairpin stability. Stability of the 28 oligomers was ascertained from their temperature‐induced melting transitions in buffered 115 mM Na+ solvent, monitored by ultraviolet absorbance and differential scanning calorimetry (DSC). Experiments revealed the melting temperatures of these molecules range from 32.4 to 60.5°C and are concentration independent over strand concentrations of 0.5 to 260 μM; thus, as expected for hairpins, the melting transitions are apparently unimolecular. Model independent thermodynamic transition parameters, ΔHcal, ΔScal, and ΔGcal, were determined from DSC measurements. Model dependent transition parameters, ΔHvH, ΔSvH, and ΔGvH were estimated from a van't Hoff (two‐state) analysis of optical melting transitions. Results of these studies reveal a significant sequence dependence to DNA hairpin stability. Thermodynamic parameters evaluated by either procedure reveal the transition enthalpy, ΔHcalHvH) can differ by as much as 20 kcal/mol depending on sequence. Similarly, values of the transition entropy ΔScalSvH) can differ by as much as 60 cal/Kmol (eu) for different molecules. Differences in free energies ΔGcalGvH) are as large as 4 kcal/mol for hairpins with different sequences. Comparisons between the model independent calorimetric values and the thermodynamic parameters evaluated assuming a two‐state model reveal that 10 of the 28 hairpins display non‐two‐state melting behavior. The database of sequence‐dependent melting free energies obtained for the hairpins was employed to extract a set of n‐n (nearest‐neighbor) sequence dependent loop parameters that were able to reproduce the input data within error (with only two exceptions). Surprisingly, this suggests that the thermodynamic stability of the DNA hairpins can in large part be reasonably represented in terms of sums of appropriate nearest‐neighbor loop sequence parameters. © 1999 John Wiley & Sons, Inc. Biopoly 50: 425–442, 1999  相似文献   

19.
Sharon S. Yu  Hsueh Jei Li 《Biopolymers》1973,12(12):2777-2788
Protamine–DNA complexes prepared by the method of direct and slow mixing in 2.5 × 10?4M EDTA, pH 8.0, have been studied by thermal denaturation and circular dichroism. The complexes show biphasic melting with Tm at about 50 °C corresponding to the melting of free DNA regions and Tm′ at about 92 °C corresponding to the melting of protamine-bound regions. In protamine-bound regions there are 1.38 amino acid residues per nucleotide, indicating a nearly completely charge neutralization. Tm is increased but Tm′ is not when the ionic strength of the buffer is raised. This also supports a full charge neutralization in protamine-bound regions. The circular dichroism of the complexes can be decomposed into two components, Δε0 of free DNA regions in B-form conformation and Δεb of protamine-bound regions in a characteristic conformation neither that of B- nor C-form but somewhere between them.  相似文献   

20.
Kunio Takeda 《Biopolymers》1985,24(4):683-694
Conformational changes of poly(L-ornithine) [(Orn)n] were studied in a sodium dodecyl sulfate (NaDodSO4) solution by CD. (Orn)n adopted an unstable and a stable helical structure below and above the NaDodSO4 concentration range where β-structure was favored, respectively. CD stopped-flow was used to monitor the transitions from coil to the unstable helix, from the helix to β-structure, and from coil to β-structure. Only the rate of the helix to β-structure transition was accelerated by an increase in NaDodSO4 concentration, whereas the rates of the others were independent of NaDodSO4 concentration. The fractions of coil, α-helix, and β-structure in each conformation of (Orn)n caused by NaDodSO4 were computed by simulating a mixed spectrum of typical CD spectra for these structures to the experimentally obtained spectrum. The contents of the unstable and stable helical structures were less than 50 and 73%, respectively.  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