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 31 毫秒
1.
Binding to DNA of two synthetic peptides, Val-Thr-Thr-Val-Val-NH-NH-Dns and Thr-Val-Thr-Lys-Val-Gly-Thr-Lsy-Val-Gly-Thr-Val-Val-NH-NH-Dns (where Dns is a residue of 5-dimethylaminonaphthalene-1-sulfonic acid), has been studied by circular dichroism, electron microscopy and fluorescence methods. It has been found that these two peptides can self-associate in aqueous solution as follows from the fact that concentration-dependent changes are observed in the UV absorbance and fluorescence spectra. The two peptides can bind to DNA both in self-associated and monomeric forms. The pentapeptide in the beta-associated form binds more strongly to poly(dG).poly(dC) than to poly[d(A-C)].poly[d(G-T)] and poly(dA).poly(dT) whereas the tridecapeptide exhibits an opposite order of preferences binding more strongly to poly[d(A-C)].poly[d(G-T)] and poly(dA).poly(dT) than to poly(dG).poly(dC). Binding is a cooperative process which is accompanied by the DNA compaction at peptide/DNA base pair ratios greater than 1. At the initial stage of the compaction process, the coalescence of DNA segments covered by bound peptide molecules leads to the formation of DNA loops stabilized by the interaction between peptide molecules bound to different DNA segments. Further increase in the peptide/DNA ratio leads to the formation of rod-like structures each consisting of two or more double-stranded DNA segments. The final stage of the compaction process involves folding of fibrillar macromolecular complexes into a globular structure containing only one DNA molecule.  相似文献   

2.
Binding of synthetic pentapeptide Val-Thr-Thr-Val-Val-N2H2Dns (where Dns is a residue of 5-dimethylamino naphthyl-1-sulfonic acid) is studied by circular dichroism, electron microscopy and fluorescence methods. It is found that this peptide can self-associate in aqueous solution as revealed from the concentration-dependent changes in the UV absorbance and fluorescence spectra. At high peptide concentration (3.10(-4) M) massive peptide aggregates are formed in solution and can be visualized by electron microscopy. It is shown that pentapeptide binds to DNA predominantly in a self-associated form and exhibits preferences for certain nucleotide sequences. It binds more strongly to poly(dG).poly(dC) and poly[d(A-C)].poly[d(G-T)] than to poly(dA).poly(dT). The complex with poly(dA).poly(dT) dissociates in the presence of 0.05 M NaCl, whereas the complex with poly(dG).poly(dC) is stable even in the presence of 0.2 M NaCl. The binding is a cooperative process which is accompanied by compaction of DNA at peptide/DNA base pair ratios greater than 2. At the initial stage of the compaction process the coalescence of DNA segments covered by bound peptide molecules results in the formation of DNA loops stabilized by interaction between bound peptide molecules. Increasing peptide/DNA ratio leads to the formation of rod-like particles as revealed from electron microscopy studies. Further increase in the peptide concentration leads to folding of fibrillar macromolecular complexes into globula each containing a single DNA molecule.  相似文献   

3.
Jean Sturm 《Biopolymers》1982,21(6):1189-1206
A temperature-jump relaxation study of the interaction of tilorone with different polynucleotides and DNA has been performed. A single relaxation time, attributed to the intercalation step, has been observed in the case of poly[d(A-T)]·poly[d(A-T)], poly[d(A-C)]·poly[d(G-T)], poly[d(G-C)]·poly[d(G-C)], and poly(dG)·poly(dC). No intercalation into poly(dA)·poly(dT) occurs, and the interaction with poly(dG)·poly(dC) is different from what is observed with the other intercalating homopolymers. Refinement of the binding model is suggested from the analysis of the kinetic data. The relaxation curves obtained with DNA are well simulated based on a binding mechanism where DNA is considered a heterogeneous lattice and each type of site behaves as if it were located in the corresponding homopolymer. Poly(dA)·poly(dT) shows a unique behavior: studies of the effects of concentration and temperature indicate that tilorone acts as a probe of a process involving the polynucleotide alone. This process appears to be related to the dynamic structure of the nucleic acid and is detectable only when the bound dye is not intercalated.  相似文献   

4.
The interaction of tilorone with DNA and five synthetic polydeoxyribonucleotides [(I): poly[d(A-T)]·poly[d(A-T)]; (II): poly[d(A-C)]·poly[d(G-T)]; (III): poly[d(G-C)]·poly[d(G-C)]; (IV): poly(dG)·poly(dC); and (V): poly(dA)·poly(dT)] has been investigated. Binding isotherms for the homopolymers were obtained by microdialysis equilibria using 14C-labeled tilorone and interpreted with different models: exclusion effect, associated or not associated with cooperativity, or variable exclusion. Affinity appears to be related more to local structure than to base composition and decreases in the following order: (I) > (II) > (III) > (IV) > (V). Intercalation in circular DNA was demonstrated by electrophoresis migration and electron microscopy, which yielded an average unwinding angle of 7° per bound dye. The behavior observed in CD and UV spectroscopy shows a sequence similar to the affinities. Tilorone seems to be less intercalated in (IV) and not at all in (V). The experimental binding isotherm of tilorone to DNA was well fitted on the basis of a model where DNA acts as a heterogeneous lattice built with the six different possible couples of adjacent base pairs, each potential site behaving as if it were in the corresponding homopolymer. The results are discussed in terms of specificity of alternating Pyr-Pur sequences and related to theoretical calculations on intercalation energies of DNA.  相似文献   

5.
Melting measurements of the mono-base-pair DNA polymers showed that the melting temperature Tm of the B-DNA homopolymer poly (dA ) · poly (dT) is higher than that of the copolymer poly [d(A-T)]. On the other hand, the Tmof the B-DNA homopolymer poly (dG) · poly (dC) is lower than that of the copolymer poly [d (G-C)]. From a structural point of view, the cross-strand base-stacking interaction in a DNA homopolymer is weaker than that in a DNA copolymer with the same base pair. One would then expect that all the DNA homopolymers are less stable than the copolymer with the same base pair. We find that the inversion of the melting order seen in the AT mono-base-pair DNA polymers is caused by the enhanced thermal stability of poly (dA) · poly (dT) from a well-defined spine of hydration attached to its minor groove. In this paper we employ the modified self-consistent phonon theory to calculate base-pair opening probabilities of four B-DNA polymers: poly(dA)-poly(dT), poly(dG) · poly(dC), poly[d(A-T)], and poly[d(G-C)] at temperatures from room temperature through the melting regions. Our calculations show that the spine of hydration can give the inverted melting order of the AT polymers as compared to the GC polymers in fair agreement with experimental measurements. Our calculated hydration spine disruption behavior in poly(dA) · poly(dT) at premelting temperatures is also in agreement with experimentally observed premelting transitions in poly (dA) · poly (dT). The work is in a sense a test of the validity of our models of nonbonded interactions and spine of hydration interactions. We find we have to develop the concept of a strained bond to fit observations in poly (dA) · poly(dT). The strained-bond concept also explains the otherwise anomalous stability of the hydration chain. © 1993 John Wiley & Sons, Inc.  相似文献   

6.
7.
Three types of DNA: approximately 2700 bp polydeoxyguanylic olydeoxycytidylic acid [poly(dG)-poly(dC)], approximately 2700 bp polydeoxyadenylic polydeoxythymidylic acid [poly(dA)-poly(dT)] and 2686 bp linear plasmid pUC19 were deposited on a mica surface and imaged by atomic force microscopy. Contour length measurements show that the average length of poly(dG)-poly(dC) is approximately 30% shorter than that of poly(dA)-poly(dT) and the plasmid. This led us to suggest that individual poly(dG)-poly(dC) molecules are immobilized on mica under ambient conditions in a form which is likely related to the A-form of DNA in contrast to poly(dA)-poly(dT) and random sequence DNA which are immobilized in a form that is related to the DNA B-form.  相似文献   

8.
We have undertaken a search for mammalian DNA-binding proteins that enhance the activity of DNA polymerases in a template sequence-specific fashion. In this paper, we report the extensive purification and characterization of a new DNA-binding protein from rabbit liver that selectively stimulates DNA polymerases to copy synthetic poly[d(G-C)] and the poly(dC) strand of poly(dC).poly(dG) as well as single-stranded natural DNA that contains stretches of oligo(dC). The enhancing protein, a polypeptide of 65 kDa designated factor C, stimulates the copying of the two synthetic templates by Escherichia coli DNA polymerase I, Micrococcus luteus polymerase, and eukaryotic DNA polymerases alpha and beta, but not by avian myeloblastosis virus polymerase. Factor C, however, does not affect utilization by these polymerases of the poly(dG) strand of poly(dC).poly(dG), of poly(dC) primed by oligo(dG), or of poly(dA).poly(dT) and poly[d(A-T)]. With polymerase I, Michaelis constants (Km) of poly[d(G-C)] and of the poly(dC) strand of poly(dC).poly(dG) are decreased by factor C 37- and 4.7-fold, respectively, whereas maximum velocity (Vmax) remains unchanged. By contrast, neither the Km value of the poly(dG) strand of poly(dC).poly(dG) nor the Vmax value with this template is altered by factor C. Rates of copying of activated DNA, denatured DNA, or singly primed M13 DNA are not affected significantly by factor C. However, primer extension analysis of the copying of recombinant M13N4 DNA that contains runs of oligo(dC) within an inserted thymidine kinase gene shows that factor C increases processivity by specifically augmenting the efficiency at which polymerase I traverses the oligo(dC) stretches. Direct binding of factor C to denatured DNA is indicated by retention of the protein-DNA complex on columns of DEAE-cellulose. Binding of factor C to poly[d(G-C)] is demonstrated by the specific adsorption of the enhancing protein to columns of poly[d(G-C)]-Sepharose. We propose that by binding to poly[d(G-C)] and to poly(dC).poly(dG), factor C enables tighter binding of some DNA polymerases to these templates and facilitates enzymatic activity.  相似文献   

9.
The interaction of the oligopeptides Ala-Gln-GIn-Leu-Ala-Gly-OH and Gln-Leu-Ala-Gly-OMe corresponding, respectively, to the sequence 53–58 and 55–58 oflac repressor protein with four polynucleotides was studied. The two peptides did not interact with poly dA. poly dT, poly d(A-T)·poly d(A-T) or poly d(A-G)·poly d(C-T). But they interacted in a characteristic way with poly d(A-C). poly d (G-T), the sequences of which are in abundance in thelac operator region. Both the peptides stabilised the melting of poly d (A-C). poly d (G-T) at a peptide to nucleotide ratio (P/N) of 4; at lower ratios, they destabilised the DNA slightly. The circular dichroism of the alternating polynucleotide with CAC/GTG sequences was perturbed by both the oligopeptides. The hexapeptide at a P/N of 4 caused the transformation of the B-form circular dichroism spectrum to a new state, characterised by strong 220 and 240 nm bands, and a rather weak long wavelength spectrum.  相似文献   

10.
Cystine peptide dimer (Lys-Gly-Val-Cys-Val-N2H2Dns)2 with S-S bridge was synthesized and its interactions with DNA and synthetic polynucleotides have been studied by optical spectroscopy methods. By recording fluorescent titration curves we have shown that the affinity of the peptide to different synthetic polynucleotides decreases in the order: poly(dG).poly(dC) greater than poly(dA).poly(dT) greater than poly(dGC).poly(dGC). The stability of complexes to increasing concentrations of NaCl diminishes in the same order. The association constant is about 20-fold greater for peptide binding to poly(dG).poly(dC) than to poly(dA).poly(dT). By using circular dichroism and fluorescence measurements we have shown that the peptide competes for the binding sites on DNA with two minor-groove binding antibiotics--distamycin A and sybiromycin. These results have suggested that the peptide also binds in the DNA minor groove. Investigation of the interactions between such peptides and DNA may be useful for constructing ligands with combined specificity to DNA.  相似文献   

11.
Native DNA duplexes in fibers exist usually in one of three well-known (A, B and C) forms depending on relative humidity, type of cations and the amount of retained salt. To determine the precise influence of these factors and the effect of base composition, as well as base sequence, on DNA secondary structure, X-ray diffraction methods have been used to study all four synthetic DNA duplexes with repeated dinucleotide sequences, eight of the 12 with repeated trinucleotide sequences and seven analogues in which guanine was replaced with hypoxanthine. The results indicate that there are at least six additional allomorphs denoted by B′, C′, C″, D, E and S.The B′ form (h = 0.329 nm) observed for poly(dA) · poly(dT), poly(dI) · poly(dC) and poly[d(A-I)] · poly[d(C-T)] is a minor variant of the traditional B form (h = 0.338 nm) of native DNA. The two C-like forms C′ for poly[d(A-G-C)] · poly-[d(G-C-T)] and poly[d(G-G-T)] · poly[d(A-C-C)] and C″ for poly[d(A-G)] · poly-[d(C-T)] have, respectively, 91 and 92 symmetries which reflect repetition of trinucleotide and dinucleotide sequences, respectively. Although isocompositional with poly(dA) · poly(dT), the existence of the rather different D form (81) for poly[d(A-T)] · poly[d(A-T)] or for poly[d(A-A-T)] · poly[d(A-T-T)] is a clear demonstration of the sequence effect. The I · C pair generally mimics an A · T pair, but poly[d(I-I-T)] · poly[d(A-C-C)] provides a new (E) form with approximately 152 screw symmetry and with 〈h〉 = 0.325 nm and 〈t〉 = 48 dg per nucleotide. The S form (65) observed for poly[d(G-C)] · poly[d(G-C)] and poly[d(A-C)] · poly[d(G-T)] is an unusual left-handed polydinucleotide helix and is accessible to any alternating purine-pyrimidine sequence. In it the two nucleotides have quite different conformations and involve syn purine and anti pyrimidine nucleosides.  相似文献   

12.
Abstract

In the present work, we employed UV-VIS spectroscopy, fluorescence methods, and circular dichroism spectroscopy (CD) to study the interaction of dye Hoechst 33258, Hoechst 33342, and their derivatives to poly[d(AT)]·poly[d(AT)], poly(dA)·poly(dT), and DNA dodecamer with the sequence 5′-CGTATATATACG-3′. We identified three types of complexes formed by Hoechst 33258, Hoechst 33342, and methylproamine with DNA, corresponding to the binding of each drug in monomer, dimer, and tetramer forms. In a dimer complex, two dye molecules are sandwiched in the same place of the minor DNA groove. Our data show that Hoechst 33258, Hoechst 33342, and methylproamine also form complexes of the third type that reflects binding of dye associates (probably tetramers) to DNA. Substitution of a hydrogen atom in the ortho position of the phenyl ring by a methyl group has a little effect on binding of monomers to DNA. However it reduces strength of binding of tetramers to DNA. In contrast, a Hoechst derivative containing the ortho-isopropyl group in the phenyl ring exhibits a low affinity to poly(dA)·poly(dT) and poly[d(AT)]·poly[d(AT)] and binds to DNA only in the monomer form. This can be attributed to a sterical hindrance caused by the ortho-isopropyl group for side-by-side accommodation of two dye molecules in the minor groove. Our experiments show that mode of binding of Hoechst 33258 derivatives and their affinity for DNA depend on substituents in the ortho position of the phenyl ring of the dye molecule. A statistical mechanical treatment of binding of Hoechst 33258 and its derivatives to a polynucleotide lattice is described and used for determination of binding parameters of Hoechst 33258 and its derivatives to poly[d(AT)]·poly[d(AT)] and poly(dA)·poly(dT).  相似文献   

13.
The binding of quinolones, nalidixic acid (Nal), oxolinic acid (Oxo) with double stranded polynucleotides was undertaken by using UV-melting, UV-Vis absorption, fluorescence and CD spectroscopic techniques. The binding of Nal or Oxo to the polynucleotides under low-salt buffer conditions were determined for poly (dA).(dT), poly [d(A-T)], poly (dG).(dC), poly [d(G-C)] and E. coli DNA. The fluorescence data were analyzed using a previously established two step mechanism with two different DNA-Drug complexes [Rajeswari et al., Biochemistry 26, 6825-31 (1987)]. The first complex [DN](1) with a binding constant K(1), is formed where the interactions are 'nonspecific' and complex [DN](2) with a binding constant K(2), is formed where the interactions are "specific" which involve (additional) hydrophobic type of interactions like 'stacking' of the drug and the overall association constant is represented as K(=K(1)K(2)). The order of binding for Nal and Oxo is: poly [d(G-C)] > poly [d(A- T)] > E. coli > poly (dG).(dC) > poly (dA).(dT). Interaction of quinolones seems to be preferential in the alternating G, C or A, T stretches of DNA than those of non-alternating. Within any alternating or non-alternating in DNA sequences the G, C rich sequences have distinctly greater binding than A, T sequences. The overall association constant data (K) reveal higher binding of Oxo to DNA compared to Nal to any given polynucleotide investigated; which also explains the higher antibacterial potency of Oxo. Changes in the absorption difference spectra and in circular dichroic spectra also manifest these results. As the melting temperatures of the polynucleotides were only marginally raised in presence of the quinolone, we rule out the possibility of 'classical intercalation' of the drug. Amino group of guanine facilitates the binding of quinolones and therefore has the greater binding with the DNA. However, poly (dG).(dC) is known to exist in 'A' conformation which is not adopted by quinolones as in the case of poly (dA).(dT). Present results suggest that Nal or Oxo bind to DNA in a non-classical fashion which is partially stacking in nature.  相似文献   

14.
The rate constants of 1H----3H exchange between water and C8H-groups of purinic residues of alternating polynucleotides: poly[d(A-T)].poly[d(A-T)] (I), poly[d(G-C)].poly[d(G-C)] (II), poly[d(A-C)].poly[d(G-T)] (III) and homopolynucleotides: poly(dA).poly(dt) (IV), poly(dG).poly(dC) (V), as well as DNA E. coli, was determined in 0.15 M NaCl at 25 degrees C. The retardation of exchange observed at these conditions (compared to that of the B-form DNA) is in agreement with the model of B-alternating structure for the (I) and is attributed to the co-existence of B- and A-conformers for the (V) in solution. Absence of distinguishable differences in exchange rate constants for purinic residues of the (II), (III) and (IV) (compared to that of the B-form DNA) evidences that conformations of these polynucleotides in solution are similar to "canonical" B-form DNA and don't correlate with the model of "heteronomous" DNA which was proposed for (IV).  相似文献   

15.
Propidium iodide is used as a structural probe for alternating and non-alternating DNA polymers containing guanine and the results are compared to experiments with poly[d(A-T)2], poly(dA . dT) and random DNA sequences. Viscometric titrations indicate that propidium binds to all polymers and to DNA by intercalation. The binding constant and binding site size are quite similar for all alternating polymers, non-alternating polymers containing guanine and natural DNA. Poly(dA . dT) is unusual with a lower binding constant and positive cooperativity in its propidium binding isotherms. Poly(dA . dT) and poly(dG . dC) have similar salt effects but quite different temperature effects in propidium binding equilibria. Polymers and natural DNA have similar rate constants in their SDS driven dissociation reactions. The association rate constants are similar for the alternating polymers and poly(dG . dC) but are significantly reduced for poly(dA . dT). These results suggest that natural DNA, the alternating polymers, and non-alternating polymers containing guanine convert to an intercalated conformation with bound propidium in a very similar manner.  相似文献   

16.
Abstract

We have performed a conformational analysis of DNA double helices with parallel directed backbone strands connected with the second order symmetry axis being at the same time the helix axis. The calculations were made for homopolymers poly(dA) · poly(dA), poly(dC) · poly(dC), poly(dG) poly(dG), and poly(dT) · poly(dT). All possible variants of hydrogen bonding of base pairs of the same name were studied for each polymer. The maps of backbone chain geometrical existence were constructed. Conformational and helical parameters corresponding to local minima of conformational energy of “parallel” DNA helices, calculated at atom-atom approximation, were determined. The dependence of conformational energy on the base pair and on the hydrogen bond type was analysed. Two major conformational advantageous for “parallel” DNA's do not depend much on the hydrogen-bonded base pair type were indicated. One of them coincided with the conformational region typical for “antiparallel” DNA in particular for the B-form DNA Conformational energy of “parallel” DNA depends on the base pair type and for the most part is similar to the conformational energy of “antiparallel” B-DNA.  相似文献   

17.
The interaction of the fluorescent dye thiazole orange (TO) with nucleic acids is characterized. It is found that TO binds with highest affinity to double-stranded (ds) DNA [log(K) ≈ 5.5 at 100 mM salt], about 5–10 times weaker to single-stranded polypurines, and further 10–1000 times weaker to single-stranded polypyrimidines. TO binds as a monomer to dsDNAs and poly(dA), both as a monomer and as a dimer to poly(dG) and mainly as a dimer to poly(dC) and poly(dT). The fluorescence quantum yield of TO free in solution is about 2 · 10−4, and it increases to about 0.1 when bound to dsDNA or to poly(dA), and to about 0.4 when bound to poly(dG). Estimated quantum yields of TO bound to poly(dC) and poly(dT) are about 0.06 and 0.01, respectively. The quantum yield of bound TO depends on temperature and decreases about threefold between 5 and 50°C. © 1998 John Wiley & Sons, Inc. Biopoly 46: 39–51, 1998  相似文献   

18.
Abstract

Thermodynamic parameters of melting process (δHm, Tm, δTm) of calf thymus DNA, poly(dA)poly(dT) and poly(d(A-C))·poly(d(G-T)) were determined in the presence of various concentrations of TOEPyP(4) and its Zn complex. The investigated porphyrins caused serious stabilization of calf thymus DNA and poly poly(dA)poly(dT), but not poly(d(A-C))poly(d(G-T)). It was shown that TOEpyp(4) revealed GC specificity, it increased Tm of satellite fraction by 24°C, but ZnTOEpyp(4), on the contrary, predominately bound with AT-rich sites and increased DNA main stage Tm by 18°C, and Tm of poly(dA)poly(dT) increased by 40 °C, in comparison with the same polymers without porphyrin. ZnTOEpyp(4) binds with DNA and poly(dA)poly(dT) in two modes—strong and weak ones. In the range of r from 0.005 to 0.08 both modes were fulfilled, and in the range of r from 0.165 to 0.25 only one mode—strong binding—took place. The weak binding is characterized with shifting of Tm by some grades, and for the strong binding Tm shifts by ~ 30–40°C. Invariability of ΔHm of DNA and poly(dA)poly(dT), and sharp increase of Tm in the range of r from 0.08 to 0.25 for thymus DNA and 0.01–0.2 for poly(dA)poly(dT) we interpret as entropic character of these complexes melting. It was suggested that this entropic character of melting is connected with forcing out of H2O molecules from AT sites by ZnTOEpyp(4) and with formation of outside stacking at the sites of binding. Four-fold decrease of calf thymus DNA melting range width ΔTm caused by increase of added ZnTO- Epyp(4) concentration is explained by rapprochement of AT and GC pairs thermal stability, and it is in agreement with a well-known dependence, according to which ΔT~TGC-TAT for DNA obtained from higher organisms (L. V. Berestetskaya, M. D. Frank-Kamenetskii, and Yu. S. Lazurkin. Biopolymers 13, 193–205 (1974)). Poly (d(A-C))poly(d(G-T)) in the presence of ZnTOEpyp(4) gives only one mode of weak binding. The conclusion is that binding of ZnTOEpyp(4) with DNA depends on its nucleotide sequence.  相似文献   

19.
The binding site and the geometry of Co(III)meso-tetrakis(N-methylpyridinium-4-yl)porphyrin (CoTMPyP) complexed with double helical poly(dA)·poly(dT) and poly(dG)·poly(dC), and with triple helical poly(dA)·[poly(dT)]2 and poly(dC)·poly(dG)·poly(dC)+ were investigated by circular and linear dichroism (CD and LD). The appearance of monomeric positive CD at a low [porphyrin]/[DNA] ratio and bisignate CD at a high ratio of the CoTMPyP-poly(dA)·poly(dT) complex is almost identical with its triplex counterpart. Similarity in the CD spectra was also observed for the CoTMPyP-poly(dG)·poly(dC) and -poly(dC)·poly(dG)·poly(dC)+ complex. This observation indicates that both monomeric binding and stacking of CoTMPyP to these polynucleotides occur at the minor groove. However, different binding geometry of CoTMPyP, when bind to AT- and GC-rich polynucleotide, was observed by LD spectrum. The difference in the binding geometry may be attributed to the difference in the interaction between polynucleotides and CoTMPyP: in the GC polynucleotide case, amine group protrude into the minor groove while it is not present in the AT polynucleotide.  相似文献   

20.
Thermodynamic parameters of melting process (DeltaHm, Tm, DeltaTm) of calf thymus DNA, poly(dA)poly(dT) and poly(d(A-C)).poly(d(G-T)) were determined in the presence of various concentrations of TOEPyP(4) and its Zn complex. The investigated porphyrins caused serious stabilization of calf thymus DNA and poly poly(dA)poly(dT), but not poly(d(A-C))poly(d(G-T)). It was shown that TOEpyp(4) revealed GC specificity, it increased Tm of satellite fraction by 24 degrees C, but ZnTOEpyp(4), on the contrary, predominantly bound with AT-rich sites and increased DNA main stage Tm by 18 degrees C, and Tm of poly(dA)poly(dT) increased by 40 degrees C, in comparison with the same polymers without porphyrin. ZnTOEpyp(4) binds with DNA and poly(dA)poly(dT) in two modes--strong and weak ones. In the range of r from 0.005 to 0.08 both modes were fulfilled, and in the range of r from 0.165 to 0.25 only one mode--strong binding--took place. The weak binding is characterized with shifting of Tm by some grades, and for the strong binding Tm shifts by approximately 30-40 degrees C. Invariability of DeltaHm of DNA and poly(dA)poly(dT), and sharp increase of Tm in the range of r from 0.08 to 0.25 for thymus DNA and 0.01-0.2 for poly(dA)poly(dT) we interpret as entropic character of these complexes melting. It was suggested that this entropic character of melting is connected with forcing out of H2O molecules from AT sites by ZnTOEpyp(4) and with formation of outside stacking at the sites of binding. Four-fold decrease of calf thymus DNA melting range width DeltaTm caused by increase of added ZnTOEpyp(4) concentration is explained by rapprochement of AT and GC pairs thermal stability, and it is in agreement with a well-known dependence, according to which DeltaT approximately TGC-TAT for DNA obtained from higher organisms (L. V. Berestetskaya, M. D. Frank-Kamenetskii, and Yu. S. Lazurkin. Biopolymers 13, 193-205 (1974)). Poly (d(A-C))poly(d(G-T)) in the presence of ZnTOEpyp(4) gives only one mode of weak binding. The conclusion is that binding of ZnTOEpyp(4) with DNA depends on its nucleotide sequence.  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