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 359 毫秒
1.
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   

2.
The interaction of Escherichia coli RNA polymerase with poly[d(A-T)] and poly[d-(I-C)] was studied by difference absorption spectroscopy at temperatures, from 5 to 45°C in the absence and presence of Mg2+. The effect of KCl concentration, at a fixed temperature, was studied from 12.5 to 400 mM. Difference absorption experiments permitted calculation of the extent of DNA opening induced by RNA polymerase and estimation of the equilibrium constant associated with the isomerization from a closed to an open RNA polymerase-DNA complex. ΔH0 and ΔS0 for the closed-to-open transition with poly[d(A-T)] or poly[d(I-C)] complexed with RNA polymerase are significantly lower than the values associated with the helix-to-coil transition for the free polynucleotides. For the RNA polymerase complexes with poly[d(A-T)] and poly[d(I-C)] in 50 mM KCl, ΔH0 ≈ 15–16 kcal/mol (63–67 kJ/mol) and ΔS0 ≈ 50–57 cal/K per mol (209–239 J/K per mol). The presence of Mg2+ does not change these parameters appreciably for the RNA polymerase-poly[d(A-T)] complex, but for the RNA polymerase-poly[d(I-C)] complex in the presence of Mg2+, the ΔH0 and ΔS0 values are larger and temperature-dependent, with ΔH0 ≈ 22 kcal/mol (92 kJ/mol) and ΔS0 ≈ 72 cal/K per mol (approx. 300 J/K per mol) at 25°C, and ΔCp0 2 kcal/K per mol (approx. 8.3 kJ/K per mol). The circular dichroism (CD) changes observed for helix opening induced by RNA polymerase are qualitatively consistent with the thermally induced changes observed for the free polynucleotides, supporting the difference absorption method. The salt-dependent studies indicate that two monovalent cations are released upon helix opening. For poly[d(A-T)], the temperature-dependence of enzyme activity correlates well with the helix opening, implying this step to be the rate-determining step. In the case of poly[d(I-C)], the same is not true, and so the rate-determining step must be a process subsequent to helix opening.  相似文献   

3.
M F Maestre  D M Gray  R B Cook 《Biopolymers》1971,10(12):2537-2553
The MCD (magnetic circular dichroism) spectra of Ap, ApA, ApApA, poly A, Up, UpU, poly U and double-stranded poly A:U alternating copoly A–U and alternating deoxyribopoly A–T were measured with a Cary 61 spectropolarimeter fitted with a Varian superconducting magnet at a field strength of 50 Kgauss. The MCD spectra of T2 and T5 DNA at various stages of heal denaturation were measured as a function of hyperchromicity of the sample. MCD spectra of the intact and degraded T2 and T5 phages were used to study the degree of alteration of the DNA inside the phages versus the DNA in vitro. The results for the adenine polymers show that the main MCD bands, B2u(271 nm), B1u(252 nm), and E1u(212 nm), show a decrease in specific magnitude as the length of the polymer is increased, reflecting the degree of stacking of the polymer. In contrast, the uridine series of polymers shows little change of the MCD bands, indicating that there is little interaction between the bases regardless of the length of the polymers. The MCD spectra of poly A:U, alternating poly r(A–U): (A–U), and alternating poly d(A–T):(A–T) show significant differences among themselves in the magnitude of the B2u band and when compared with the sum of the spectrum for the poly A plus poly U. This may indicate the selective effect of hydrogen bonding on the B2u band. Alternatively, the difference may be due to the absence of an n → π* transition in the double-stranded polymer. Measurements of denatured T2 and To DNA's show increases in all MCD bands. The T2 DNA internally packed in phage shows an increase of the B2u and E1ubands, the B2u remaining unchanged. The internal T5 DNA shows an increase of the B1u band only. Thus, the internal DNA structure is altered in a manner quite different from a simple denaturation caused by hydrogen bond breaking. Furthermore, different MCD bands indicate that different modes of DNA packing exist for T2 and T5 phages.  相似文献   

4.
The template activity of Cancer pagurus DNA and its two components (poly d(A-T) and main component) in response to a DNA polymerase purified from regenerating rat liver has been studied and compared to the results previously obtained with synthetic templates. In the double-stranded native state, whole crab DNA and the main component were poor templates. Their replication was increased by thermal denaturation and inhibited by actinomycin. Like the synthetic copolymer poly[d(A-T)·d(T-A)], native crab poly d(A-T) could be copied and its duplication was not inhibited by actinomycin. The structural difference between native poly d(A-T) Form I, isolated on a density gradient, and partially renatured poly d(A-T) Form II, isolated on hydroxylapatite, resulted in a modification of their template activity. The kinetic studies of [3H] dGMP and [3H] dAMP incorporation confirmed the importance of single-stranded regions (particulary dC regions) in the initiation of the in vitro duplication.  相似文献   

5.
Summary The effect of Aclacinomycin B (ACM-B), an anthracycline antitumor antibiotic, on the DNA-dependent RNA synthesis using single- and double-stranded DNAs of known base content and sequence is studied. The data show that ACM-B effectively inhibits the double-stranded DNA-directed RNA synthesis with a preference of poly[d(A-T)] > poly[d(G-C)] > poly[d(I-C)]. In contrast, it has no inhibitory effect on the template function of single-stranded DNA (e.g. poly dA, poly dT, and poly dC). These results suggest that the mechanism of ACM-13 inhibition, like other anthracycline antibiotics, is by intercalation. In addition to the base specificity, there are also dramatic differences in inhibition depending on the base sequence in the DNA template. Thus, ACM-13 preferentially inhibits the alternating double-stranded copolymers over the double-stranded homopolymers; e.g. poly [d(A-T)] is inhibited to a greater extent than poly dA · poly dT and poly [d(G-C)] is inhibited more than poly dG · poly dC. Since the inhibition by ACM-13 can be totally abolished when assayed in excess amount of DNA, this result suggests that ACM-B inhibition of RNA synthesis is solely on the DNA template (which is in support of the intercalation model), and has ruled out the possibility that ACM-B may also exert an inhibitory effect on the activity of RNA polymerase per se.  相似文献   

6.
Double-stranded synthetic polydeoxynucleotides of the general form poly[d(GnCn)] · poly[d(GnCn)], poly[d(GnC)] · poly[d(GCn)], and poly[d(AnTn)] · poly[d(AnTn)] have been synthesized. When n = 4 or larger, the CD spectra of polymers of the form poly[d(GnCn)] · poly[d(GnCn)] or poly[d(GnC)] · poly[d(GCn)] closely resemble the spectrum of poly[dG] · poly[dC], suggesting that a string of four continguous guanosine residues is sufficient to induce a conformation resembling that of the polypurine · polypyrimidine. With polymers of the form poly[d(AnTn)] · poly[d(AnTn)], however, the CD spectrum only gradually approaches that of poly[dA] · poly[dT].  相似文献   

7.
In the present work we studied the initial part of thermal denaturation curves of B. subtilis DNA in normal and heavy water, observing, by electron microscopy, the “opening” of DNA as a function of temperature. The results support the hypothesis that D2O plays a stabilizing role on strands separation during thermal denaturation of DNA.  相似文献   

8.
Abstract

CD spectra were used to compare the acid-induced structural transitions of poly[d(A)] and poly[d(C)] with those of poly[r(A)] and poly[r(C)], respectively. The types of base pairing were probably the same in the acid self-complexes of both A-containing polymers and in the acid self-complexes of both C-containing polymers. Similar base pairings were indicated by similarities in the difference CD spectra showing the changes during the first major acid- induced transitions of the polymers. Information from the CD spectra and pKa values of the transitions suggested that the transitions for the RNA polymers involved similar structural changes. The two DNA polymers were markedly different. Single-stranded poly [d(A)] was in the most stacked structure and had the lowest pKa for forming an acid self-complex. Single-stranded poly[d(C)] was in the least stacked structure and had the highest pKa for forming a protonated duplex.  相似文献   

9.
The Pressure Dependence of the Helix-Coil Transition Temperature (Tm) of Poly[d(G-C)] was studied as a function of sodium ion concentration in phosphate buffer. The molar volume change of the transition (ΔV) was calculated using the Clapeyron equation and calorimetrically determined enthalpies. The ΔV of the transition increased from +4.80 (±0.56) to +6.03 (±0.76) mL mol?1 as the sodium ion concentration changed from 0.052 to 1.0M. The van't Hoff enthalpy of the transition calculated from the half-width of the differentiated transition displayed negligible pressure dependence: however, the value of this parameter decreased with increasing sodium ion concentration, indicating a decrease in the size of the cooperative unit. The volume change of the transition exhibits the largest magnitude of any double-stranded DNA polymer measured using this technique. For poly[d(G-C)] the magnitude of the change in ΔV with sodium ion concentration (0.94 ± 0.05 mL mol?1) is approximately one-half that observed for either poly[d(A-T)] or poly (dA)·poly(dT). The ΔV values are interpreted as arising from changes in the hydration of the polymer due to the release of counterions and changes in the stacking of the bases of the coil form. As a consequence of solvent electrostriction, the release of counterions makes a net negative contribution to the total ΔV, implying that disruption of the slacking interactions contributes a positive volume change to the total ΔV. The larger magnitude of the ΔV compared with that of other double-stranded polymers may be due in part to the high helix-coil transition temperature of poly[d(G-C)], which will attenuate the contribution of electrostriction to the total volume change. The data in addition show that in the absence of other cellular components, the covalent structure of DNA is stabile under conditions of temperature and pressure more extreme than those experienced by any known organism. © 1995 John Wiley & Sons, Inc.  相似文献   

10.
The complexes of Hoechst 33258 with poly[d(A-T)2], poly[d(I-C)2], poly[d(G-C)2], and poly[d(G-m5C)2] were studied using linear dichroism, CD, and fluorescence spectroscopies. The Hoechst-poly[d(I-C)2] complex, in which there is no guanine amino group protruding in the minor groove, exhibits spectroscopic properties that are very similar to those of the Hoechst-poly[d(A-T)2] complex. When bound to both of these polynucleotides, Hoechst exhibits an average orientation angle of near 45° relative to the DNA helix axis for the long-axis polarized low-energy transition, a relatively strong positive induced CD, and a strong increase in fluorescence intensity—leading us to conclude that this molecule also binds in the minor groove of poly[d(I-C)2]. By contrast, when bound to poly[d(G-C)2] and poly[d(G-m5C)2], Hoechst shows a distinctively different behavior. The strongly negative reduced linear dichroism in the ligand absorption region is consistent with a model in which part of the Hoechst chromophore is intercalculated between DNA bases. From the low drug:base ratio onset of excitonic effects in the CD and fluorescence emission spectra, it is inferred that another part of the Hoechst molecule may sit in the major groove of poly[d(G-C)2] and poly[d(G-m5C)2] and preferentially stacks into dimers, though this tendency is strongly reduced for the latter polynucleotide. Based on these results, the importance of the interactions of Hoechst with the exocyclic amino group of guanine and the methyl group of cytosine in determining the binding modes are discussed. © 1996 John Wiley & Sons, Inc.  相似文献   

11.
Abstract

Circular dichroism measurements were used to study the binding of fd gene 5 protein to fd DNA, to six polydeoxynucleotides (poly(d(A)], poly[d(T)], poly[d(I)], poly[d(C)], poly[d(A-T)], and the random copolymer poly[d(A,T)]), and to three oligodeoxynucleotides (d(pA)20, d(pA)7, and d(pT)7). Titrations of these DNAs with fd gene 5 protein were generally done in a low ionic strength buffer (5 mM Tris-HCl, pH 7.0 or 7.8) to insure tight binding, needed to obtain stoichiometric endpoints. By monitoring the CD of the nucleic acids above 250 nm, where the protein has no significant intrinsic optical activity, we found that there were two modes of binding, with the number of nucleotides covered by a gene 5 protein monomer (n) being close to either 4 or 3. These stoichiometrics depended upon which polymer was titrated as well as upon the protein concentration. Single endpoints at nucleotide/protein molar ratios close to 3 were found during titrations of poly[d(T)] and fd DNA (giving n = 3.1 and 2.8 ± 0.2, respectively), while CD changes with two apparent endpoints at nucleotide/protein molar ratios close to 4 and approximately 3 were found during titrations of poly[d(A)], poly[d(I)], poly[d(A-T)], and poly[d(A,T)) (with the first endpoints giving n = 4.1, 4.0, 4.0, and 4.1 ± 0.3, respectively). Calculations showed that the CD changes we observed during these latter titrations were consistent with a switch between two non- interacting binding modes of n = 4 and n = 3. We found no evidence for an n = 5 binding mode. One implication of our results is that the Brayer and McPherson model for the helical gene 5 protein-DNA complex, which has 5 nucleotides bound per protein monomer (G. Brayer and A. McPherson, J. Biomol Struct, and Dyn. 2, 495-510, 1984), cannot be correct for the detailed solution structure of the complex.

We interpreted the CD changes above 250 nm upon binding of the gene 5 protein to single-stranded DNAs to be the result of a slight unstacking of the bases, along with a significant alteration of the CD contributions of the individual nucleotides in the case of A- and/or T-containing DNAs, Interestingly, CD contributions attributed to nearest-neighbor interactions in free poly[d(A-T)], poly[d(A,T)], poly[d(A)], and poly[d(T)] were partially maintained in the CD spectra of the protein-saturated polymers, so that neighboring nucleotides, when bound to the protein at 20°C, appeared to interact with one another in much the same manner as in the free polymers at 50°C. Finally, we found that the protein tyrosyl CD band at 228.5 nm decreased 39-42% when the protein bound to poly[d(A)] or poly[d(T)], but this band decreased no more than 9% when the gene 5 protein bound to short A- or T-containing oligomers. Thus, at least one tyrosyl residue has a significantly altered optical activity only when the DNA substrate is long enough either to cause a transition to a different protein conformation or to allow additional protein-protein contacts between adjacent helical turns of the DNA-protein complex.  相似文献   

12.
In the present work, we measured survival and the platinum on the genome after treatment of repair-proficient or repair-deficient Escherichia coli strains with trans-[PtCl2(E-iminoether)2] and compared these results with the effects of “classical” cisplatin. We found that toxicity of antitumor trans-[PtCl2(E-iminoether)2] in repair-deficient trains was much less than that of cisplatin. This markedly reduced toxicity was not a consequence of the reduced uptake or low levels of DNA binding in the bacteria cells but rather appeared to reflect DNA binding mode of this trans-platinum drug different from that of cisplatin.  相似文献   

13.
P A Mirau  D R Kearns 《Biopolymers》1985,24(4):711-724
1H-nmr relaxation has been used to study the effect of sequence and conformation on imino proton exchange in adenine–thymine (A · T) and adenine–uracil (A · U) containing DNA and RNA duplexes. At low temperature, relaxation is caused by dipolar interactions between the imino and the adenine amino and AH2 protons, and at higher temperature, by exchange with the solvent protons. Although room temperature exchange rates vary between 3 and 12s?1, the exchange activation energies (Eα) are insensitive to changes in the duplex sequence (alternating vs homopolymer duplexes), the conformation (B-form DNA vs A-form RNA), and the identity of the pyrimidine base (thymine vs uracil). The average value of the activation energy for the five duplexes studied, poly[d(A-T)], poly[d(A) · d(T)], poly[d(A-U)], Poly[d(A) · d(U)], and poly[r(A) · r(U)], was 16.8 ± 1.3 kcal/mol. In addition, we find that the average Eα for the A.T base pairs in a 43-base-pair restriction fragment is 16.4 ± 1.0 kcal/mol. This result is to be contrasted with the observation that the Eα of cytosine-containing duplexes depends on the sequence, conformation, and substituent groups on the purine and pyrimidine bases. Taken together, the data indicate that there is a common low-energy pathway for the escape of the thymine (uracil) imino protons from the double helix. The absolute values of the exchange rates in the simple sequence polymers are typically 3–10 times faster than in DNAs containing both A · T and G · C base pairs.  相似文献   

14.
31P- and 1H-nmr and laser Raman spectra have been obtained for poly[d(G-T)]·[d(C-A)] and poly[d(A-T)] as a function of both temperature and salt. The 31P spectrum of poly[d(G-T)]·[d(C-A)] appears as a quadruplet whose resonances undergo separation upon addition of CsCl to 5.5M. 1H-nmr measurements are assigned and reported as a function of temperature and CsCl concentration. One dimensional nuclear Overhauser effect (NOE) difference spectra are also reported for poly[d(G-T)]·[d(C-A)] at low salt. NOE enhancements between the H8 protons of the purines and the C5 protons of the pyrimidines, (H and CH3) and between the base and H-2′,2″ protons indicate a right-handed B-DNA conformation for this polymer. The NOE patterns for the TH3 and GH1 protons in H2O indicate a Watson–Crick hydrogen-bonding scheme. At high CsCl concentrations there are upfield shifts for selected sugar protons and the AH2 proton. In addition, laser Raman spectra for poly[d(A-T)] and poly[d(G-T)]·[d(C-A)] indicate B-type conformations in low and high CsCl, with predominantly C2′-endo sugar conformations for both polymers. Also, changes in base-ring vibrations indicate that Cs+ binds to O2 of thymine and possibly N3 of adenine in poly[d(G-T)]·[d(C-A)] but not in poly[d(A-T)]. Further, 1H measurements are reported for poly[d(A-T)] as a function of temperature in high CsCl concentrations. On going to high CsCl there are selective upfield shifts, with the most dramatic being observed for TH1′. At high temperature some of the protons undergo severe changes in linewidths. Those protons that undergo the largest upfield shifts also undergo the most dramatic changes in linewidths. In particular TH1′, TCH3, AH1′, AH2, and TH6 all undergo large changes in linewidths, whereas AH8 and all the H-2′,2″ protons remain essentially constant. The maximum linewidth occurs at the same temperature for all protons (65°C). This transition does not occur for d(G-T)·d(C-A) at 65°C or at any other temperature studied. These changes are cooperative in nature and can be rationalized as a temperature-induced equilibrium between bound and unbound Cs+, with duplex and single-stranded DNA. NOE measurements for poly[d(A-T)] indicate that at high Cs+ the polymer is in a right-handed B-conformation. Assignments and NOE effects for the low-salt 1H spectra of poly[d(A-T)] agree with those of Assa-Munt and Kearns [(1984) Biochemistry 23 , 791–796] and provide a basis for analysis of the high Cs+ spectra. These results indicate that both polymers adopt a B-type conformation in both low and high salt. However, a significant variation is the ability of the phosphate backbone to adopt a repeat dependent upon the base sequence. This feature is common to poly[d(G-T)]·[d(C-A)], poly[d(A-T)], and some other pyr–pur polymers [J. S. Cohen, J. B. Wouten & C. L Chatterjee (1981) Biochemistry 20 , 3049–3055] but not poly[d(G-C)].  相似文献   

15.
The effect of negatively charged dilauroylphosphatidic acid (DLPA) vesicles on the conformation of poly( -lysine) was investigated by circular dichroism measurements. DLPA vesicles induced a confomiational change Of poly( -lysine) from the random coil to β-structure in 5 mM Tes, pH 7.0. The fraction of induced β-structure (Fβ) was determined via a procedure of curve fit the observed spectra to the reference spectra. Fβ increased linearly with the molar ratio, r, of DLPA to lysine residues up to r 0.7, and reached a saturation value of 1 at r > 1. Within the range 0.7 r 1, precipitation occurred. The effect of dilution of the negative charge on vesicle membranes was examined by mixing DLPA with dilauroylphosphatidylcholine (DLPC). Although the β-structure Of poly -lysine) was also induced by mixed vesicles, the saturation value of Fβ decreased with decreasing DLPA content in mixed vesicles. The variation in saturation value of Fβ with the composition of mixed vesicles was interpreted in terms of the change in average distance between DLPA head groups in mixed vesicles.  相似文献   

16.
The complex [Mn(L)(NO3)2(H2O)2] (1) (L=2H-5-hydroxy-1,2,5-oxadiazo[3,4-f]1,10-phenanthroline) was synthesized and characterized by elemental analysis, IR and UV. The crystal and molecular structure of 1 was determined by single-crystal X-ray diffraction; crystal data: light yellow, monoclinic, space group P21/n, Z=4, a=7.432(2) Å, b=9.582(3) Å, c=23.445(7) Å, β=90.519(5)°. The Mn atom in 1 is hexa-coordinated in a distorted octahedral arrangement by two N atoms of the ligand L and four O atoms of two water molecules and two nitrate anions. Biological tests in vitro showed that 1 has significant antitumor activity against HL-60, KB, Hela and BGC-823 cells. The interaction of 1 with calf thymus DNA was investigated by absorption titration, thermal denaturation and viscosity measurements. The results suggest that 1 binds with DNA by intercalating via the ligand L.  相似文献   

17.
Zhao P  Xu LC  Huang JW  Fu B  Yu HC  Zhang WH  Chen J  Yao JH  Ji LN 《Bioorganic chemistry》2008,36(6):278-287
Four cationic porphyrin–anthraquinone (Por–AQ) hybrids differing in lengths of flexible alkyl linkage, 5-[4-(1-N-anthraquinonon-yl)-l-oxophenyl]-10,15,20-tris(N-methylpyridinium-4-yl)porphyrin triiodide, (l = acetyl, pentanoyl, octanoyl, undecanoyl, designed as [AQATMPyP]I3, [AQPTMPyP]I3, [AQOTMPyP]I3 and [AQUTMPyP]I3, respectively, see Fig. 1), were synthesized and their interactions with DNA were investigated. The results of spectroscopic, denaturation and viscosity measurements suggest that [AQATMPyP]I3 binds to DNA through non-intercalative mode while the other three hybrids with longer links bind via bis-intercalative mode. Ethidium bromide (EB) competition experiment was carried out to determine the binding constants (Kb) of these compounds for CT DNA, and [AQPTMPyP]I3 shows the largest Kb among these hybrids. The photocleavage mechanism and wavelength-dependent cleaving abilities of these hybrids to pBR322 plasmid DNA were also comparably investigated.  相似文献   

18.
Our previous results indicate that during protoplast isolation an oxidative burst occurs [A.K. Papadakis and KA Roubelakis-Angelakis (1999) Plant Physiol 127:197–205] and that suppression of totipotency is correlated with reduced antioxidant activity and low redox state [A.K. Papadakis et al. (2001b) Plant Physiol 126:434–444]. Polyamines are known to affect cell development and to act as antioxidants. Polyamines applied during isolation of tobacco (Nicotiana tabacum L.) protoplasts reduced the accumulation of O2· but not that of H2O2. This antioxidant effect is probably due to the inhibition of microsomal membrane NADPH oxidase, which occurred in a concentration-dependent manner, with spermine exerting the highest inhibitory effect. However, during protoplast culture, polyamine oxidase activity increased severalfold in spermidine- and spermine-treated protoplasts, concomitant with H2O2 titers. A cell death program was executed in untreated protoplasts, as documented by membrane malfunction, induced DNase activity, DNA fragmentation and a positive TUNEL reaction. Protoplast cell death was prevented in protoplasts treated with putrescine, but not by treatment with spermidine or spermine, which rather had the opposite effect. The data presented suggest that PAs may be implicated in the expression of plant protoplast totipotency.  相似文献   

19.
The interaction between DNA and ionen polymers, -[N+(CH3)2(CH2)mN+(CH3)2(CH2)n], with m-n of 3–3, 6–6, and 6–10 were examined in order to know how the binding behavior of cationic polymers with DNA depends on the charge density of polycation. The ionen polymer has no bulky side chain and the binding forces with DNA would be attributed mainly to electrostatic interaction. When 3–3 ionen polymers were added to DNA solution, precipitable complexes with the ratio of cationic residue to DNA phosphate (+/?) of 1/1 and the free DNA molecules were segregated, while 6–6 and 6–10 ionen polymers formed soluble complexes with DNA molecules up to (+/?) = 0.5. This suggests that 3–3 ionen polymers bind cooperatively with DNA while 6–6 and 6–10 ionen polymers bind noncooperatively. The cooperative binding of 3–3 ionen polymer and the noncooperative binding of 6–6 ionen polymer were also supported by the thermal melting and recooling profiles from the midpoint between first and second meltings. It was concluded that the charge density of DNA phosphate is a critical value determining whether the ionen polymers bind to DNA by a cooperative or by a noncooperative binding, since the distance between successive cationic charges of 3–3 ionen polymer is shorter than that between successive phosphate charges on DNA double helix and those of 6–6 and 6–10 ionen polymers are longer.  相似文献   

20.
The binding of the antibiotics netropsin and distamycin A to DNA has been studied by thermal melting, CD and sedimentation analysis. Netropsin binds strongly at antibiotic/nucleotide ratios up to at least 0.05. CD spectra obtained using DNA model polymers reveal that netropsin binds tightly to poly (dA) · poly (dT), poly (dA-dT) · poly(dA-dT) and poly (dI-dC) · poly (dI-dC) but poorly, if at all, to poly (dG) · poly (dC). Binding curves obtained with calf thymus DNA reveal one netropsin-binding site per 6.0 nucleotides (Ka=2.9 · 105 M−1); corresponding values for distamycin A are one site per 6.1 nucleotides with Ka= 11.6 · 105 M−1. Binding sites apparently involve predominantly A·T-rich sequences whose specific conformation determines their high affinity for the two antibiotics. It is suggested that the binding is stabilized primarily by hydrogen bonding and electrostatic interactions probably in the narrow groove of the DNA helix, but without intercalation. Any local structural deformation of the helix does not involve unwinding greater than approximately 3° per bound antibiotic molecule.  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