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.
To evaluate the length and sequence dependence of the unusual interaction properties observed for nonalternating A/T sequences in deoxyribonucleic acid (DNA) [Wilson, W. D., Wang, Y. H., Krishnamoorthy, C. R., & Smith, J. C. (1985) Biochemistry 24, 3991-3999], we have synthesized the oligomers d(A-T)6, dA10 X dT10, and d(A6-T6) and evaluated their interaction with the intercalator propidium. Propidium visible spectral shifts on adding all three oligomers are quite similar. Low-temperature spectrophotometric binding measurements indicate that d(A-T)6 has a significantly larger binding constant for propidium than dA10.dT10, as with the analogous alternating and nonalternating DNA polymers. The oligomer dA10.dT10 displays positive cooperativity in its propidium binding isotherm, and its binding constant increases with increasing temperature while d(A-T)6 does not display positive cooperativity, and its binding constant decreases with temperature, again as with the analogous polymers. van't Hoff plots indicate that the propidium binding enthalpies are approximately -9 and +6 kcal/mol for the alternating and nonalternating DNA samples, respectively. The mixed-sequence self-complementary oligomer d(A6-T6) has an unusual low-temperature binding isotherm which suggests a single strong binding site and a larger number of weaker binding sites which bind propidium cooperatively. A van't Hoff plot indicates that the cooperative sites d(A-T)6 have binding constants and binding enthalpies similar to dA10.dT10. Similar rate constants are observed in the sodium dodecyl sulfate driven dissociation reaction of propidium from d(A-T)6 and d(A6-T6), but the association reaction of propidium is significantly slower with d(A6-T6) than with d(A-T)6.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Thermal denaturation of four oligonucleotides, viz. 3'-d(AT)5pO(CH2)6Opd(AT)5-3'(par(AT], 3'-d(AT)5pO(CH2)6Opd(AT)5-5'(anti(AT],3'-d(A)10pO(CH2) 6Op(T)10-3'(par(A-T], and 3'-d(A)10pO(CH2)6Opd(T)10-5' (anti(A-T], was studied in 0.01 M phosphate buffer, pH 7, in the presence of 0.1, 0.25, 0.5 and 1.0 M NaCl. All the oligomers were found to exist at a lower temperature (0 to 20 degrees C) as complexes composed either of two oligomer molecules (a canonical duplex) or of more oligomer molecules whereas, at a higher temperature (30 to 70 degrees C), they formed hairpins with a parallel (par(AT) and par(A-T] or antiparallel (anti(AT) and anti(A-T) orientation of the chains. Melting curves (A260(T] were used to calculate thermodynamic parameters for the formation of hairpins and "low-temperature" duplexes. Experiments on ethidium bromide binding to the oligonucleotides have shown that the oligomer anti(A-T) exists, at a low ionic strength, as a four stranded complex ("quadruplex") contains two antiparallel helices, d(A).d(T), which have a parallel orientation and are bound to one another owing to the formation of additional hydrogen bonds between nucleic acid bases. The possible biological function of quadruplexes is discussed.  相似文献   

3.
Factors influencing the binding of tetracationic porphyrin derivatives to DNA have been comprehensively evaluated by equilibrium dialysis, stopped-flow kinetics, etc., for mesotetrakis (4-N-methylpyridiniumyl)porphyrin [TMpyP (4)]. Technical difficulties have previously precluded a comprehensive study of metalloporphyrins. Since electrostatic interactions with the DNA and metal derivatization of the porphyrins have important consequences, we have investigated in greater detail two isomers of TMpyP (4) (meso-tetrakis(3-N-methylpyridiniumyl)porphyrin, [TMpyP (3)] and meso-tetrakis(2-N-methylpyridiniumyl)porphyrin [TMpyP (2)]) in which the position of the charged centers has been varied. A comprehensive study of the Cu(II) derivatives, e.g., CuTMpyP (4), was possible since the difficulties encountered previously with Ni(II) compounds were not a problem with Cu(II) porphyrins [J. A. Strickland, L. G. Marzilli, M. K. Gay, and W. D. Wilson (1988) Biochemistry 27, 8870-8878]. At 25 degrees C, the apparent equilibrium constants [Kobs] decreased with increasing [Na+] for all porphyrins. The Kobs values were comparable for TMpyP (4) and TMpyP (3) binding to either polyd(G-C).polyd(G-C) [poly[d(G-C)2]] or poly[d(A-T)].poly[d(A-T)] [poly[d(A-T)2]]. For the copper(II) porphyrins, the Kobs values were about fivefold greater. The Kobs value for CuTMpyP (2) binding to poly[d(G-C)2] was too small to measure under typical salt conditions; however, Kobs for binding to poly[d(A-T)2] was about two orders of magnitude smaller than those found for CuTMpyP (4) or CuTMpyP (3). Application of the condensation theory for polyelectrolytes suggests about three charge interactions when CuTMpyP (4), CuTMpyP (3), and TMpyP (3) bind to poly[d(G-C)2] or poly[d(A-T)2], a result comparable to that reported for TMpyP (4). At 20 degrees C and 0.115 M [Na+], incorporation of copper decreased the rates of dissociation from poly[d(A-T)2] by a 100-fold compared to those reported for TMpyP (4) but had little effect on the rates of dissociation from poly[d(G-C)2]. Also, movement of the H3CN+ group from the fourth to the third position of the pyridinium ring enhanced the rates of dissociation from poly[d(A-T)2] but decreased the rates of dissociation from poly[d(G-C)2]. From polyelectrolyte theory, the [Na+] dependence of the dissociation rates from poly[d(G-C)2] is consistent with intercalative binding, while that for poly[d(A-T)2] is consistent with an outside binding model. For calf thymus [CT] DNA at 20 degrees C, a greater decrease in the AT than in the GC imino 1H-nmr signal was observed upon addition of CuTMpyP (2), suggesting selective outside binding to the AT regions.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
The interactions of DAPI with natural DNA and synthetic polymers have been investigated by hydrodynamic, DNase I footprinting, spectroscopic, binding, and kinetic methods. Footprinting results at low ratios (compound to base pair) are similar for DAPI and distamycin. At high ratios, however, GC regions are blocked from enzyme cleavage by DAPI but not by distamycin. Both poly[d(G-C)]2 and poly[d(A-T)]2 induce hypochromism and shifts of the DAPI absorption band to longer wavelengths, but the effects are larger with the GC polymer. NMR shifts of DAPI protons in the presence of excess AT and GC polymers are significantly different, upfield for GC and mixed small shifts for AT. The dissociation rate constants and effects of salt concentration on the rate constants are also quite different for the AT and the GC polymer complexes. The DAPI dissociation rate constant is larger with the GC polymer but is less sensitive to changes in salt concentration than with the AT complex. Binding of DAPI to the GC polymer and to poly[d(A-C)].poly[d(G-T)] exhibits slight negative cooperativity, characteristic of a neighbor-exclusion binding mode. DAPI binding to the AT polymer is unusually strong and exhibits significant positive cooperativity. DAPI has very different effects on the bleomycin-catalyzed cleavage of the AT and GC polymers, a strong inhibition with the AT polymer but enhanced cleavage with the GC polymer. All of these results are consistent with two totally different DNA binding modes for DAPI in regions containing consecutive AT base pairs versus regions containing GC or mixed GC and AT base pair sequences. The binding mode at AT sites has characteristics which are similar to those of the distamycin-AT complex, and all results are consistent with a cooperative, very strong minor groove binding mode. In GC and mixed-sequence regions the results are very similar to those observed with classical intercalators such as ethidium and indicate that DAPI intercalates in DNA sequences which do not contain at least three consecutive AT base pairs.  相似文献   

5.
The antibiotic AT2433-B1 belongs to a therapeutically important class of antitumor agents. This natural product contains an indolocarbazole aglycone connected to a unique disaccharide consisting of a methoxyglucose and an amino sugar subunit, 2,4-dideoxy-4-methylamino-L-xylose. The configuration of the amino sugar distinguishes AT2433-B1 from its diastereoisomer iso-AT2433-B1. Here we have investigated the interaction of these two disaccharide indolocarbazole derivatives with different DNA sequences by means of DNase I footprinting and surface plasmon resonance (SPR). Accurate binding measurements performed at 4 and 25 degrees C using the BIAcore SPR method revealed that AT2433-B1 binds considerably more tightly to a hairpin oligomer containing a [CG](4) block than to an oligomer with a central [AT](4) tract. The kinetic analysis shows that the antibiotic dissociates much more slowly from the GC sequence compared to the AT one. Preferential binding of AT2433-B1 to GC-rich sequences in DNA was independently confirmed by DNase I footprinting experiments performed with a 117 bp DNA restriction fragment. The specific binding sequence 5'-AACGCCAG identified from the footprints was then converted into a biotin-labeled DNA hairpin duplex and compound interactions with this specific sequence were characterized by high resolution BIAcore SPR experiments. Such a combined approach provided a detailed understanding of the molecular basis of DNA recognition. The discovery that the glycosyl antibiotic AT2433-B1 preferentially recognizes defined sequences offers novel opportunities for the future design of sequence-specific DNA-reading small molecules.  相似文献   

6.
Interactions of meso-tetrakis(4-N-methylpyridiniumyl)porphyrin [TMpyP(4)] with poly[d(G-C)].poly[d(G-C)] [poly[d(G-C)2] and poly[d(A-T)].poly[d(A-T)] [poly[d(A-T)2] were studied by equilibrium dialysis and stopped-flow dissociation kinetics as a function of [Na+]. Metalloderivatives of TMpyP(4), NiTMpyP(4), and ZnTMpyP(4) were also investigated. The apparent equilibrium binding constants (Kobs) were approximately the same for TMpyP(4) binding to either poly[d(G-C)2] or poly[d(A-T)2] and decreased with increasing [Na+]. The slopes of the plots of log Kobs vs log [Na+] were similar, with values close to -2.7. Contrary to implications in previously reported studies, these data do not indicate that TMpyP(4) prefers to bind to GC sites at low ionic strength and to AT sites at high ionic strength. In contrast, binding of ZnTMpyP(4) to these two polymers is very different. Comparisons of Kobs values at 0.065 M [Na+] indicate that ZnTMpyP(4) binding to AT sites is approximately 200 times more favorable than binding to GC sites, a finding in agreement with previous qualitative observations. Although the binding of the Zn species to the GC polymer was too weak for us to assess the salt effect, the plot of log Kobs vs log [Na+] gave a slope of -2.0 for ZnTMpyP(4) binding to poly[d(A-T)2]. Application of condensation theory for polyelectrolytes suggests similar charge interactions for ZnTMpyP(4) and for TMpyP(4) binding to poly[d(A-T)2]. Likewise, the rates of dissociation from poly[d(A-T)2] were similar for TMpyP(4) and ZnTMpyP(4) [and also NiTMpyP(4)]. However, whereas TMpyP(4) [and NiTMpyP(4)] dissociation from poly[d(G-C)2] was measurable, that for ZnTMpyP(4) was too fast to measure.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Apparent molar adiabatic compressibilities and apparent molar volumes of poly[d(A-T)].poly[d(A-T)], poly(dA).poly(dT), DNA and poly(A).poly(U) in aqueous solutions were determined at 1 degree C. The change of concentration increment of the ultrasonic velocity upon replacing counter ion Cs+ by the Mg2+ ion was also determined for these polymers. The following conclusions have been made: (1) the hydration of the double helix of poly(dA).poly(dT) is remarkably larger than that of other polynucleotides; (2) the hydration of the AT pair in the B-form DNA is larger than that of the GC pair; (3) the substitution of Cs+ for Mg2+ ions as counter ions results in a decrease of hydration of the system polynucleotide plus Mg2+, and (4) the magnitude of this dehydration depends on the nucleotide sequence; the following rule is true: the lesser is a polynucleotide hydration, the larger dehydration upon changing Cs+ for Mg2+ ions in the ionic atmosphere of polynucleotide.  相似文献   

8.
Hemoglobin can be specifically carboxymethylated at its NH2-terminal amino groups (i.e. HbNHCH2COO-) to form the derivatives alpha 2Cm beta 2, alpha 2 beta 2Cm, and alpha 2Cm beta 2Cm, where Cm represents carboxymethyl. Previous studies (DiDonato, A., Fantl, W. J., Acharya, A. S., and Manning, J. M. (1983) J. Biol. Chem. 258, 11890-11895) suggested that these derivatives could be used as stable analogues of the corresponding carbamino (Hb-NHCOO-) forms of hemoglobin, adducts that are generated reversibly in vivo when CO2 combines with alpha-amino groups. In this paper we present x-ray diffraction studies of both carbamino hemoglobin and carboxymethylated hemoglobin that verify this proposal and we use the carboxymethylated derivatives to study the functional consequences of placing a covalently bound carboxyl group at the NH2 terminus of each hemoglobin subunit. Our studies also provide additional information concerning the oxygen-linked binding of anions and protons to Val-1 alpha. Difference electron density analysis of deoxy alpha 2Cm beta 2Cm versus the unmodified deoxyhemoglobin tetramer (deoxy alpha 2 beta 2) shows that the covalently bound carboxyl moieties replace inorganic anions that are normally bound to the free NH2-terminal amino groups in crystals of native deoxyhemoglobin grown from solutions of concentrated (2.3 M) ammonium sulfate. In the case of the beta-subunits, the carboxymethyl group replaces an inorganic anion normally bound between the alpha-amino group of Val-1 beta, the epsilon-amino group of Lys-82 beta, and backbone NH groups at the NH2-terminal end of the F'-helix. In the case of the alpha-subunits, the carboxymethyl group replaces an anion that is normally bound between the alpha-amino group of Val-1 alpha and the beta-OH group of Ser-131 alpha. A corresponding difference electron map of carbamino deoxyhemoglobin in low-salt (50 mM KCl) crystals shows that CO2 bound in the form of carbamate occupies the same two anion binding sites. The alkaline Bohr effect of alpha 2Cm beta 2 is only marginally lower (approximately 7%) than that of alpha 2 beta 2. Previous studies (Kilmartin, J. V., 1977) have shown that about 30% of the alkaline Bohr effect is the result of an oxygen-linked change in the pK alpha of Val-1 alpha, and O'Donnell et al., 1979, found that this portion of the Bohr effect is the result of the oxygen-linked binding of chloride to Val-1 alpha.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

9.
The conventional wisdom argues that DNA intercalators possess a condensed polyaromatic ring whereas DNA minor groove binders generally contain unfused aromatic heterocycles, frequently separated by amide bonds. Recently, this view has been challenged with the discovery of powerful intercalating agents formed by unfused aromatic molecules and groove binders containing a polyaromatic nucleus. Bis-amidinocarbazoles belong to this later category of drugs having a planar chromophore and capable of reading the genetic information accessible within the minor groove of AT-rich sequences [Tanious, F.A., Ding, D., Patrick, D.A., Bailly, C., Tidwell, R.R. & Wilson, W.D. (2000) Biochemistry 39, 12091-12101]. But in addition to the tight binding to AT sites, we show here that bis-amidinocarbazoles can also interact with GC sites. The extent and mode of binding of 2,7 and 3,6 substituted amidinocarbazoles to AT and GC sequences were investigated by complementary biochemical and biophysical methods. Absorption, fluorescence, melting temperature and surface plasmon resonance (SPR) measurements indicate that the position of the two amidine groups on the carbazole ring influences significantly the drug-DNA interaction. SPR and DNase I footprinting data confirm the AT-preference of the compounds and provide useful information on their additional interaction with GC sequences. The 3,6-carbazole binds approximately twice as strongly to the GC-containing hairpin oligomer than the 2,7-regioisomer. The high tendency of the 3,6 compound to intercalate into different types of DNA containing G.C base pairs is shown by electric linear dichroism. This work completes our understanding of the sequence-dependent DNA binding properties of carbazole dications.  相似文献   

10.
A new series of phenanthridinone derivatives, and diketo acid analogs, as well as related phenanthrene and anthracene diketo acids have been synthesized and evaluated as HIV integrase (IN) inhibitors. Several new beta-diketo acid analogs with the phenanthridinone scaffold replaced by phenanthrene, anthracene or pyrene exhibited the highest IN inhibitory potency. There is a general selectivity against the integrase strand transfer step. The most potent IN was 2,4-dioxo-4-phenanthren-9-yl-butyric acid (27f) with an IC(50) of 0.38microM against integrase strand transfer. The phenanthrene diketo acids 27d-f were more potent (IC(50)=2.7-0.38microM) than the corresponding phenanthridinone diketo acid 16 (IC(50)=65microM), suggesting that the polar amide bridge in the phenanthridinone system decreases inhibitory activity relative to the more lipophilic phenanthrene system. This might have to do with the possible binding of the aryl group of the compounds binding to a lipophilic pocket at the integrase active site as suggested by the docking simulations. Molecular modeling also suggested that effectiveness of chelation of the active site Mg(2+) contributes to IN inhibitory potency. Finally, some of the potent compounds inhibited HIV-1 replication in human peripheral blood mononuclear cells (PBMC) with EC(50) down to 8microM for phenanthrene-3-(2,4-dioxo)butyric acid (27d), with a selectivity index of 10 against PBMCs.  相似文献   

11.
N Sugimoto  R Kierzek  D H Turner 《Biochemistry》1988,27(17):6384-6392
The self-splicing intervening sequence from the rRNA precursor in Tetrahymena thermophila produces a covalently closed, circularized form (C IVS). Reaction rates for reverse cyclization (linearization) of C IVS by the covalent addition of the oligoncleotides CU, UCU, CUCU, and CUCUCU have been measured. The dependence of the observed rates on oligomer and Mg2+ concentrations indicates the presence of intermediates that are generated by separate binding steps for both oligomer and Mg2+. Linearization of C IVS by OH- hydrolysis is suppressed in the presence of oligomer, suggesting oligomer binds near the active site. The binding constants derived for CU at 30 degrees C in 1 and 10 mM Mg2+ are 5 X 10(3) and 2.5 X 10(4) M-1, respectively. These are roughly 4 orders of magnitude larger than expected for simple Watson-Crick base pairing. The binding constants derived for UCU, CUCU, and CUCUCU at 30 degrees C in 10 mM Mg2+ are 1.2 X 10(5), 4 X 10(5), and greater than 10(7) M-1, respectively. The free energy increments for binding of UCU and CUCU relative to CU are similar to those expected from a nearest-neighbor model for addition of base pairs. This indicates the factors responsible for the unusually strong binding of CU to C IVS are restricted to two nucleotides.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
K X Chen  N Gresh    B Pullman 《Nucleic acids research》1987,15(5):2175-2189
Theoretical computations are performed of the intercalative binding of the neocarzinostatin chromophore (NCS) with the double-stranded oligonucleotides d(CGCG)2, d(GCGC)2, d(TATA)2 and d(ATAT)2. Minor groove binding is preferred over major groove binding. It is found that the long axis of the stacked naphtoate ring lies approximately parallel to the long axis of the base pairs of the intercalation site. The galactosamine ammonium group interacts with specific sites of the groove (O2/N3 of bases 2 and O1' of sugar S3), whereas the dodecadyine ring system wraps around the groove towards the backbone. An overall AT versus GC preference is derived. Intercalation in a central purine-(3', 5')-pyrimidine sequence appears to be preferred over that in a central pyrimidine-(3', 5')-purine sequence.  相似文献   

13.
F A Tanious  S F Yen  W D Wilson 《Biochemistry》1991,30(7):1813-1819
The interaction of a symmetric naphthalene diimide with alkylamino substituents at each imide position was investigated with the alternating sequence polymers, poly[d(A-T)]2 and poly[d(G-C)]2. Spectrophotometric binding studies indicate strong binding of the diimide to both sequences although the GC binding constant is 20-25 times larger than the AT binding constant. Analysis of the effects of salt concentration on the binding equilibria shows that the diimide forms two ion pairs in its complex with both polymers as expected for a simple dication. Stopped-flow kinetics experiments demonstrate that the diimide both associates and dissociates from DNA more slowly than classical intercalators with similar binding constants. Analysis of salt concentration effects on dissociation kinetics rate constants (kd) reveals that slopes in log kd versus log [Na+] plots are only approximately half the value obtained for classical dicationic intercalators that have both charged groups in the same groove. These kinetics results support a threading intercalation model, with one charged diimide substituent in each of the DNA grooves rather than with both side chains in the same groove, for the diimide complex with DNA. In the rate-determining step of the mechanism for dissociation of a threading complex only one ion pair is broken; the free side chain can then slide between base pairs to put both diimide side chains in the same groove, and this is followed by rapid full dissociation of the diimide. This sequential release of ion pairs makes the dissociation slope for dicationic threading intercalators more similar to the slope for classical monocationic intercalating ligands.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
We designed and synthesized the peptide nucleic acid (PNA)-peptide conjugates having anthracene chromophores and investigated their interactions with calf thymus DNA, [d(AT)(10)](2), [d(GC)(10)](2), and [d(AT)(10)dA(6)](2). Considering the synthesis compatibility and expecting that a novel DNA analogue, PNA, can improve DNA binding properties of alpha-helix peptides, we attempted to attach thymine PNA oligomers at the C-terminus of a 14 amino acid alpha-helix peptide that contained a pair of artificial intercalators, anthracene, as a probe, and to examine their interactions with DNA using anthracene UV, fluorescence and circular dichroism properties. The results observed in this study showed that the designed peptide folded in an alpha-helix structure in the presence of calf thymus DNA, [d(AT)(10)](2), and [d(AT)(10)dA(6)](2) with the chromophores at the side-chain being fixed with a left-handed chiral-sense orientation. The alpha-helix and the anthracene signals were not observed for [d(GC)(10)](2). Incorporation of thymine PNA oligomers into the designed alpha-helix peptide increased the DNA binding ability to [d(AT)(10)dA(6)](2) with increasing the length of the PNA without changing the conformations of the peptide backbone and the anthracene side-chains.  相似文献   

15.
The interaction of naphthothiophene, phenanthrene and anthracene ring systems, which have amide and ester side chains with cationic groups (synthesized from the aromatic acid chlorides and appropriate amines and alcohols), with calf thymus DNA has been investigated by using viscometric titrations, spectrophotometric binding experiments and 1H-, 31P- and 17O-NMR methods. The viscosity and NMR experiments suggest that all of these compounds bind to DNA by intercalation. These experiments and spectrophotometric binding studies, however, indicate that there is considerable variation in the interaction of these compounds with DNA. These variations can all be explained by the geometry of the ring systems, the position of protons adjacent to the side chains, and the relative sizes of the amide and ester side chains. With the naphthothiophene ester and amide, for example, the planar amide cannot rotate into the plane of the naphthothiophene ring whereas the smaller planar ester can. With this ring system the ester has a significantly higher binding constant than the amide derivative. Additional binding studies with poly[d(A-T)2] and poly[d(G-C)2] have shown that all of these compounds bind more strongly to the A-T- than the G-C-containing polymer. Since the ester compounds do not have hydrogen bond donating groups proximate to the aromatic ring, these results suggest a model for the A-T specificity of these compounds that involves a solvent-mediated hydrogen bond between the C-2 carbonyl of thymine and the carbonyl group of the intercalators.  相似文献   

16.
The binding of propidium to poly(dA).poly(dT) [poly(dA.dT)] and to poly[d(A-T)].poly[d(A-T)] [poly[d(A-T)2]] has been compared under a variety of solution conditions by viscometric titrations, binding studies, and kinetic experiments. The binding of propidium to poly[d(A-T)2] is quite similar to its binding to calf thymus deoxyribonucleic acid (DNA). The interaction with poly(dA.dT), however, is quite unusual. The viscosity of a poly(dA.dT) solution first decreases and then increases in a titration with propidium at 18 degrees C. The viscosity of poly[d(A-T)2] shows no decrease in a similar titration. Scatchard plots for the interaction of propidium with poly(dA.dT) show the classical upward curvature for positive cooperativity. The curvature decreases as the temperature is increased in binding experiments. A van't Hoff plot of the observed binding constants yields an apparent positive enthalpy of approximately +6 kcal/mol for the propidium-poly(dA.dT) interaction. Propidium binding to poly[d(A-T)2] shows no evidence for positive cooperativity, and the enthalpy change for the reaction is approximately -9 kcal/mol. Both the magnitude of the dissociation constants and the effects of ionic strength are quite similar for the dissociation of propidium from poly(dA-T)2] and from poly[d(A-T)2], suggesting that the intercalated states are similar for the two complexes. The observed association reactions, under pseudo-first-order conditions, are quite different. Plots of the observed pseudo-first-order association rate constant vs. polymer concentration have much larger slopes for propidium binding to poly[d(A-T)2] than to poly(dA.dT).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
The complex formation of S- and R-ofloxacin with the self-complementary oligonucleotides, namely d[ATAGCGCTAT](2), d[GCGATATCGC](2) and d[ATAICICTAT](2), were investigated by the molecular dynamics (MD) simulation. Four starting positions, including two intercalation positions with different insertion directions and two minor groove binding positions, were considered. The total energy of both S- and R-ofloxacin-d[ATAGCGCTAT](2) complex, in which ofloxacin binds in the minor groove of the oligonucleotide, were lower than any intercalation binding mode. For both enantiomers, formation of the complex with GC oligonucleotide is more favorable than AT and IC oligonucleotides. When S- and R-ofloxacin are compared, the S-enantiomer exhibits more favorable total energy and torsion angles in the complex formation. This result is in agreement with the experimental observation [Hwangbo et al., Eur J Pharm Sci 18, 197 (2003)]. In the complex, both enantiomers form two hydrogen bonds: one between the carbonyl group of ofloxacin and the amine group of G16 and the other between the fluorine group and the G6 amine for S-ofloxacin. However, only one hydrogen bond is formed between endocyclic hydrogen atom at the C2 position of adenine and inosine base and carbonyl group of ofloxacin, which may be the reason for the GC preferentiality of ofloxacin.  相似文献   

18.
Imino proton and 31P NMR studies were conducted on the binding of actinomycin D (ActD) to self-complementary oligodeoxyribonucleotides with one GC binding site [d(ATATGCATAT) (1), d-(ATACGCGTAT) (2), and d(ATATACGCGTATAT) (3)] and with two GC sites [d(ATGCATGCAT) (4)]. At R = 1 (molar ratio of ActD to oligomer duplex) ActD caused a doubling of the number of imino proton signals at, and adjacent to, the GC binding site of 1. One of the G.C base pair signals shifted upfield while the other shifted downfield. Both of the signals for the A.T base pairs adjacent to the binding site shifted downfield. All imino proton signals of 2 and the longer sequence, 3, shifted upfield on binding of ActD to the GC site, indicating a sequence-dependent change in base stacking on complex formation. For both 1 and 2 addition of ActD resulted in a similar pattern of three downfield 31P NMR signals. The two most downfield signals have chemical shift and temperature dependence which are characteristic of phosphate groups at isolated intercalation sites. At R = 1 the ActD complex with 4 has very complex spectra with both upfield and downfield A.T and G.C imino signals. All these data were consistent with two 1:1 complexes with the unsymmetrical phenoxazone ring adopting both of the two possible orientations.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
4', 6-Diamidine-2-phenylindole forms fluorescent complexes with synthetic DNA duplexes containing AT, AU and IC base pairs; no fluorescent complexes were observed with duplexes containing GC base pairs or with duplexes containing a single AT base pair sandwiched between GC pairs. The binding site size is one molecule of dye per 3 base pairs. The intrinsic binding constants are higher for alternating sequence duplexes than for the corresponding homopolymer pairs. With the exception of the four-stranded helical poly rI which exhibits considerable fluorescence enhancement upon binding of the ligand, none of the single- or multi- stranded polyribonucleotides and ribo-deoxyribonucleotide hybrid structures form fluorescent complexes with the dye. Poly rI is the only RNA which forms a DNA B-like structure (Arnott et al. (1974) Biochem. J. 141, 537). The B conformation of the helix and the absence of guanine appear to be the major determinants of the specificity of the fluorescent binding mode of the dye. Nonfluorescent interactions of the dye with polynucleotides are nonspecific; UV absorption and circular dichroic spectra demonstrate binding to synthetic single- and double-stranded DNA and RNA analogs, including those containing GC base pairs.  相似文献   

20.
Flow cytometry, using propidium iodide and 4',6-diamidano-2-phenylindole staining, was used to estimate the nuclear DNA content (2C) and the proportion of A-T base pairs in 16 species of the Mediterranean genus Cistus. Genome sizes were shown to be constant within species, since no significant intraspecific variation in 2C DNA content was detected. At the genus level, up to about 1.5-fold differences in absolute DNA amounts were observed, ranging from 3.92 pg in C. crispus to 5.88 pg in C. monspeliensis. The (AT) : (GC) ratio was close to 1, and was similar for all species examined, ranging from 47.87% A-T content in C clusii, to 50.67% in C. populifolius. Pink-flowered species (subgenus Cistus) had lower DNA amounts than white-flowered species (subgenera Leucocistus and Halimioides). However, the distribution of DNA amounts in Cistus appeared to be continuous and did not permit a clear separation of infra-generic ranks in the genus.  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