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 140 毫秒
1.
S P Fodor  P A Starr  T G Spiro 《Biopolymers》1985,24(8):1493-1500
Raman spectra have been recorded for poly(dG-dT) · poly(dA-dC) and poly(dA-dT) · poly(dA-dT) in low salt and at high concentrations of CsF. Poly(dG-dT) · poly(dA-dC) shows no change in the 682-cm?1 guanine mode, demonstrating the absence of the Z-structure at high salt. The 790-cm?1 phosphodiester symmetric stretch, however, shifts up 5 cm?1 in 4.3M CsF, suggesting a slight conformational change, associated with ion binding or hydration changes. Poly(dA-dT) · poly(dA-dT) shows an additional broad band at 816 cm?1, attributed to the phosphodiester modes associated with the C3′-endo deoxyribose units in the alternating B-structure. In this case, both the 841- and the 816-cm?1 asymmetric phosphodiester stretches, associated with the C2′- and C3′-endo units, shift down on addition of CsF in a sequential manner. Correlation of this sequence with that previously observed for the two 31P-nmr resonances, establishes that the phosphodiester stretching frequencies depend on the conformation of the 5′-sugar, and not on the 3′-sugar.  相似文献   

2.
Molecular mechanical simulations on base-paired deoxyhexanucleoside phosphates, (dAdT)3 · (dAdT)3, (dTdA)3 · (dTdA)3, (dGdC)3 · (dGdC)3, and (dCdG)3 · (dCdG)3, have been carried out to assess their energetic stabilities in left- and right-handed forms. These hexamers have also been simulated with alternating sugar-puckering profiles with the combinations (purine : C2′-endo–pyrimidine : C3′-endo) and (purine : C3′-endo–pyrimidine C2′-endo). The right-handed models have been found to be the energetically most stable structures and the left-handed structures are significantly destabilized. This instability has been rationalized in terms of competition between stabilizing stacking interactions on one hand, and distortions in the bond angles and torsion angles in the sugar-phosphate backbone on the other.  相似文献   

3.
Changes in the 31P-nmr spectra of sonicated natural DNA fragments were investigated in ethanol solutions where the fragments underwent, as checked by CD, the B-to-A conformational transition. The study produced the following conclusions: (1) The high DNA concentrations used for the 31P-nmr measurements promote the transition compared to dilute solutions that are commonly used for CD measurements. (2) The B-to-A transition was reflected in a cooperative downfield shift of the DNA 31P-nmr resonance, consistent with unwinding of the double helix. (3) Prior to the transition, the changes in chemical shift of double-and single-stranded DNAs were almost identical. It thus appears that the effect of ethanol on the geometry and hydration of phosphodiester linkages does not depend heavily on DNA base–base interactions. (4) The A-form resonances were 30–40% narrower than the B-form resonances, which is attributed to marked sequence-dependent variations in the latter conformation and to their reduction in the former. (5) The B-form DNA aggregated in the concentrated 31P-nmr samples in the presence of ethanol, judged from a milky opalescence of the solution and a substantial broadening of its 31P-nmr resonance. The broadening abruptly disappeared as soon as DNA adopted the A-form so that DNA, in dependence on the secondary structure, showed different tendencies to condense in the presence of ethanol. The condensation increased cooperativity of the B-to-A interconversion.  相似文献   

4.
D J Patel 《Biopolymers》1976,15(3):533-558
The Watson–Crick imino and amino exchangeable protons, the nonexchangeable base and sugar protons, and the backbone phosphates for d-CpG(pCpG)n, n = 1 and 2, have been monitored by high-resolution nmr spectroscopy in aqueous solution over the temperature range 0°–90°C. The temperature dependence of the chemical shifts of the tetramer and hexamer resonances is consistent with the formation of stable duplexes at low temperature in solution. Comparison of the spectral characteristics of the tetranucleotide with those of the hexanucleotide with temperature permits the differentiation and assignment of the cytosine proton resonances on base pairs located at the end of the helix from those in an interior position. There is fraying at the terminal base pairs in the tetranucleotide and hexanucleotide duplexes. The Watson–Crick ring imino protons exchange at a faster rate than the Watson–Crick side-chain amino protons, with exchange occurring by transient opening of the double helix. The structure of the d-CpG(pCpG)n double helices has been probed by proton relaxation time measurements, sugar proton coupling constants, and the proton chemical shift changes associated with the helix–coil transition. The experimental data support a structural model in solution, which incorporates an anti conformation about the glycosyl bonds, C(3) exo sugar ring pucker, and base overlap geometries similar to the B-DNA helix. Rotational correlation times of 1.7 and 0.9 × 10?9 sec have been computed for the hexanucleotide and tetranucleotide duplexes in 0.1 M salt, D2O, pH 6.25 at 27°C. The well-resolved 31P resonances for the internucleotide phosphates of the tetramer and hexamer sequences at superconducting fields shift upfield by 0.2–0.5 ppm on helix formation. These shifts reflect a conformational change about the ω,ω′ phosphodiester bonds from gauche-gauche in the duplex structure to a distribution of gauche-trans states in the coil structure. Significant differences are observed in the transition width and midpoint of the chemical shift versus temperature profiles plotted in differentiated form for the various base and sugar proton and internucleotide phosphorous resonances monitoring the d-CpG(pCpG)n helix–coil transition. The twofold symmetry of the d-CpGpCpG duplex is removed on complex formation with the antibiotic actinomycin-D. Two phosphorous resonances are shifted downfield by ~2.6 ppm and ~1.6 ppm on formation of the 1:2 Act-D:d-CpGpCpG complex in solution. Model studies on binding of the antibiotic to dinucleotides of varying sequence indicate that intercalation of the actinomycin-D occurs at the GpC site in the d-CpGpCpG duplex and that the magnitude of the downfield shifts reflects strain at the O-P-O backbone angles and hydrogen bonding between the phenoxazone and the phosphate oxygens. Actinomycin-D is known to bind to nucleic acids that exhibit a B-DNA conformation; this suggests that the d-CpG(pCpG)n duplexes exhibit a B-DNA conformation in solution.  相似文献   

5.
Formation of psi (+) and psi (-) DNA   总被引:4,自引:0,他引:4  
Y A Shin  G L Eichhorn 《Biopolymers》1984,23(2):325-335
DNA molecules can be organized into ordered aggregates of opposite handedness by complexation with polylysine and other polypeptides; we have investigated the conditions under which ψ(+) and ψ(?) structures are produced with double-helical synthetic polynucleotides. Both poly(dGdC)·poly(dGdC) and poly(dAdT)·poly(dAdT) readily form ψ(?) structures with polylysine, although the method of preparation can alter the CD spectra. The GC copolymer, which is more susceptible to conversion into A or Z conformers, forms ψ(+) structures with lysine–alanine copolypeptides more readily than the AT copolymer. Nucleotide base modifications that favor the Z structure, such as bromination and methylation, also favor ψ(+) formation, and the Co(NH3)6Cl3 reagent that readily induces the Z structure also leads to ψ(+). Thus, the production of the ψ(+) structure seems to be frequently correlated with susceptibility to A or Z formation, although there are some cases in which the B conformer also leads to ψ(+). Polyethylene glycol generally produces a ψ(?) structure; the differentiation between ψ(+) and ψ(?) structures seems to require electrically charged polymers.  相似文献   

6.
Abstract

Methylated lysine, arginine and histidine residues are found in a number of proteins (for example, histones, non-histone chromosomal proteins, ribosomal proteins, calmodulin, cytochrome C, etc.). We are studying the effects of methylation on the conformations of poly(lysine) and of the effects of methylation of poly(lysine) and poly(arginine) on interactions with polynucleotides. The conformational properties of e-amino-methylated poly(lysine) differ from those of unmodified poly(lysine). Methylation increases resistance to thermally- induced and NaCl-induced changes in the CD spectrum. Guanidinium chloride increases (proportional to the degree of methylation) the extent of approach to the conformation in dispute as to its being a random coil or an extended helix. Methylation enhances aggregation in the helix-inducing solvent 0.5 M Ca(ClO4)2. With increasing methylation of poly(lysine), the conformation in dodecyl sulfate changes from β, to 50% α, to random coil at the maximum methylation.

Increasing methylation of poly(lysine) weakens the interaction with polynucleotides in respect to dissociation by salt, linearly with methyl content. Complexes of (dAdT)n·(dAdT)n with the polypeptides are increasingly stabilized to heat denaturation by progressive methylation. However, with a series of synthetic double-stranded RNA's and DNA's a more complex situation exists, Tm increasing or decreasing, depending on the base composition, sequence and type of sugar. Methylation of poly(lysine) and poly(arginine) can have opposite effects on Tm based on results with complexes with (dI)n·(dC)n. Methylated poly(lysine) affects the CD spectrum of polynucleotides, in a manner dependent on base composition and sequence. In some cases large positive or negative ψ-spectra are induced, which, in the case of (dGdC)n·(dGdC)n, can be positive or negative depending on the degree of methylation of the polypeptide and the salt concentration.

It is suggested that the biological effects of methylated proteins may be evoked by salt changes in the cell cycle, and that methylation can affect local interactions with nucleic acids and larger scale structure, and interactions with lipids.  相似文献   

7.
Induced CD of DNA intercalators: electric dipole allowed transitions   总被引:3,自引:0,他引:3  
The induced CD of an electric dipole allowed transition of a DNA intercalator has been calculated using the “matrix method” and a set of DNA ππ* transitions recently adopted for calculating the CD of DNA by Rizzo and Schellman [(1984) Biopolymers 23 , 435–470]. The induced CD is strongly dependent on the angular orientation of the intercalator and only moderately dependent on its location within the intercalation pocket. The dependence of the CD on the orientation and location of the intercalator was studied for some representative conformations of di- and tetranucleotide duplexes of (dGdC) and (dAdT). The effect of alternative DNA transition moment directions was also tested. The orientation dependence compares well with the previously predicted 1-2 cos2 γ dependence [B. Nordén and F. Tjerneld (1982) Biopolymers 21 , 1713–1734]. Measured induced CD spectra of methylene blue (MB) intercalated in double-stranded poly(dAdT), poly(dGdC), and calf-thymus DNA are discussed on the basis of the results of the calculations. Major differences between the induced CD spectra are likely to reflect different modes of intercalation for the different nucleotide sequences. In particular, the concluded geometry in solution for MB intercalated in poly(dAdT) differs significantly from the corresponding geometry found in dinucleotide–intercalator crystals.  相似文献   

8.
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   

9.
At low temperature and low salt concentration, both imino proton and 31p-nmr spectra of DNA complexes with the intercalators ethidium and propidium are in the slow-exchange region. Increasing temperature and/or increasing salt concentration results in an increase in the site exchange rate. Ring-current effects from the intercalated phenanthridinium ring of ethidium and propidium cause upfield shifts of the imino protons of A · T and G · C base pairs, which are quite similar for the two intercalators. The limiting induced chemical shifts for propidium and ethidium at saturation of DNA binding sites are approximately 0.9 ppm for A · T and 1.1 ppm for G · C base pairs. The similarity of the shifts for ethidium and propidium, in both the slow- and fast-exchange regions over the entire titration of DNA, shows that a binding model for propidium with neighbor-exclusion binding and negative ligand cooperativity is correct. The fact that a unique chemical shift is obtained for imino protons at intercalated sites over the entire titration and that no unshifted imino proton peaks remain at saturation binding of ethidium and propidium supports a neighbor-exclusion binding model with intercalators bound at alternating sites rather than in clusters on the double helix. Addition of ethidium and propidium to DNA results in downfield shifts in 31P-nmr spectra. At saturation ratios of intercalator to DNA base pairs in the titration, a downfield shoulder (approximately ?2.7 ppm) is apparent, which accounts for approximately 15% of the spectral area. The main peak is at ?3.9 to ?4.0 ppm relative to ?4.35 in uncomplexed DNA. The simplest neighbor-binding model predicts a downfield peak with approximately 50% of the spectral area and an upfield peak, near the chemical shift for uncomplexed DNA, with 50% of the area. This is definitely not the case with these intercalators. The observed chemical shifts and areas for the DNA complexes can be explained by models, for example, that involve spreading the intercalation-induced unwinding of the double helix over several base pairs and/or a DNA sequence- and conformation-dependent heterogeneity in intercalation-induced chemical shifts and resulting exchange rates.  相似文献   

10.
We report the relative molar sound velocity increments, [U], partial molar volumes, V(o), and partial molar adiabatic compressibilities, K(S)(o), of the Li(+), Na(+), K(+), Rb(+), Cs(+), NH(4)(+), and N(CH(3))(4)(+) salts of poly(dAdT)poly(dAdT), poly(dGdC)poly(dGdC), poly(dIdC)poly(dIdC), poly(rA)poly(rU), poly(rG)poly(rC), poly(rI)poly(rC), and poly(rU) at 25 degrees C. When analyzing these data, we take into account the Donnan membrane equilibrium effect. Comparison between the values of [U], V(o), and K(S)(o) exhibited by the nucleic acid salts and respective chlorides (LiCl, NaCl, KCl, RbCl, CsCl, NH(4)Cl, and N(CH(3))(4)Cl) yields information about the state of counterion hydration in the vicinity of each nucleic acid structure studied here. Our analysis reveals that the poly(dGdC)poly(dGdC), poly(dIdC)poly(dIdC), and poly(rI)poly(rC) duplexes and single-stranded poly(rU) do not significantly influence the hydration properties of their condensed counterions. In the vicinity of these polymers, counterions retain their full hydration shells (within +/-15%). By contrast, counterions condensed around the poly(dAdT)poly(dAdT), poly(rA)poly(rU), and poly(rG)poly(rC) duplexes are significantly dehydrated and retain, respectively, only 65(+/-18)%, 34(+/-21)%, and 33(+/-9)% of their original hydration shells. Taken together, the volumetric data reported here provide important new information that ultimately may help us understand the central role that hydration and counterions play in modulating the conformational preferences of nucleic acids and the energetics of DNA recognition events.  相似文献   

11.
B P Cross  T Schleich 《Biopolymers》1973,12(10):2381-2389
The solution conformation of β-D -O2,2′-cyclouridine has been determined at 27 and 88°C in D2O by proton magnetic resonance spectroscopy. The conformation is described in terms of a fixed syn-like sugar-base torsional angle, a type S furanose ring conformation (similar to 2′-endo), and a temperature-dependent exocyclic C(4)′–C(5′) rotamer population containing approximately 50% of the gauche-gauche form at 27°C. β-D -O2,2′-Cyclouridine 5′-phosphate likewise possesses a type S furanose ring conformation.  相似文献   

12.
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   

13.
In the course of investigation of 8-alkylamino substituted adenosines, the title compounds were synthesized as potential partial agonists for adenosine receptors. The structure determination of these compounds was carried out with the X-ray crystallography study. Crystals of 8-(2-hydroxyethylamino)adenosine are monoclinic, space group P 21; a = 7.0422(2), b = 11.2635(3), c = 8.9215(2) Å, β = 92.261(1)°, V = 707.10(3) Å3, Z = 2; R-factor is 0.0339. The nucleoside is characterized by the anti conformation; the ribose ring has the C(2′)-endo conformation and gauchegauche form across C(4′)–C(5′) bond. The molecular structure is stabilized by intramolecular hydrogen bond of N–H·O type. Crystals of 8-(pyrrolidin-1-yl)adenosine are monoclinic, space group C 2; a = 19.271(1), b = 7.3572(4), c = 11.0465(7) Å, β = 103.254(2)°, V = 1524.4(2) Å3, Z = 4; R-factor is 0.0498. In this compound, there is syn conformation of the nucleoside; the ribose has the C(2′)-endo conformation and gauchegauche form across C(4′)–C(5′) bond. The molecular structure is stabilized by intramolecular hydrogen bond of O–H·N type. For both compounds, the branching net of intermolecular hydrogen bonds occur in the crystal structures.  相似文献   

14.
《Biophysical journal》2020,118(12):2997-3007
We aimed to investigate insulin amyloid fibril polymorphism caused by salt effects and heating temperature and to visualize the structural differences of the polymorphisms in situ using Raman imaging without labeling. The time course monitoring for amyloid formation was carried out in an acidic condition without any salts and with two species of salts (NaCl and Na2SO4) by heating at 60, 70, 80, and 90°C. The intensity ratio of two Raman bands at 1672 and 1657 cm−1 due to antiparallel β-sheet and α-helix structures, respectively, was revealed to be an indicator of amyloid fibril formation, and the relative proportion of the β-sheet structure was higher in the case with salts, especially at a higher temperature with Na2SO4. In conjunction with the secondary structural changes of proteins, the S-S stretching vibrational mode of a disulfide bond (∼514 cm−1) and the ratio of the tyrosine doublet I850/I826 were also found to be markers distinguishing polymorphisms of insulin amyloid fibrils by principal component analysis. Especially, amyloid fibrils with Na2SO4 media formed the gauche-gauche-gauche conformation of disulfide bond at a higher rate, but without any salts, the gauche-gauche-gauche conformation was partially transformed into the gauche-gauche-trans conformation at higher temperatures. The different environments of the hydroxyl groups of the tyrosine residue were assumed to be caused by fibril polymorphism. Raman imaging using these marker bands also successfully visualized the two- and three- dimensional structural differences of amyloid polymorphisms. These results demonstrate the potential of Raman imaging as a diagnostic tool for polymorphisms in tissues of amyloid-related diseases.  相似文献   

15.
Interactions of meso-tetra(4-N-methylpyridyl)porphyrin [TMpyP(4)], meso-tetra(2-N-methylpyridyl)porphyrin [TMpyP(2)], and meso-tetra(para-N-trimethylanilinium)porphyrin (TMAP) with several native and synthetic DNAs were studied by a variety of physical techniques: nmr (31P and 1H), absorption spectroscopy, viscosity, and flow dichroism (FD). Of the three porphyrins studied, only the interaction of TMpyP(4) with poly [d(G-C)2] was fully consistent with intercalation. In particular, a large increase in viscosity, a downfield 31P-nmr signal (ca. -1 ppm), and upfield imino proton signals (11 to 12 ppm range) were observed. Comparison of the effects of TMpyP(4) on DNAs of different GC contents revealed larger changes in solution viscosity with increased GC content. However, the characteristic changes in 31P- and 1H-nmr spectra were not observed. The viscosity increases observed in studies with poly[d(A-C)(G-T)] and C. Perf. DNA were much lower than with poly[d(G-C)2], M. Lys. DNA, and calf thymus DNA. Thus, GC sequence and content are clearly important. The principal change in the 31P-nmr signal of native DNA is the appearance of a very broad shoulder centered at ca. -2.0 ppm, which is larger in M. Lys. DNA than in C. Perf. DNA. FD studies indicate highly ordered TMpyP(4) cations arranged perpendicular to the DNA axis of calf thymus DNA. Together, these results suggest the major effects of TMpyP(4) on DNA properties are due to strong GC-binding interactions that influence DNA structure. The data are consistent with combined intercalative and outside binding interactions of TMpyP(4) with GC regions of DNA. In contrast, similar studies with TMAP suggest that it influences AT regions of DNA by an outside binding mode. On the other hand, TMpyP(2) effects on DNA properties are consistent with nonselective outside binding.  相似文献   

16.
Abstract

Equilibrium binding is believed to play an important role in directing the subsequent covalent attachment of many carcinogens to DNA. We have utilized UV spectroscopy to examine the non-covalent interactions of aflatoxin B1 and B2 with calf thymus DNA, poly(dAdT):poly(dAdT), and poly(dGdC):poly(dGdC), and have utilized NMR spectroscopy to examine non-covalent interactions of aflatoxin B2 with the oligodeoxynucleotide d(ATGCAT)2. UV-VIS binding isotherms suggest a greater binding affinity for calf thymus DNA and poly(dAdT):poly(dAdT) than for poly(dGdC):poly(dGdC). Scatchard analysis of aflatoxin B1 binding to calf thymus DNA in 0.1 M NaCl buffer indicates that binding of the carcinogen at levels of bound aflatoxin ? 1 carcinogen per 200 base pairs occurs with positive cooperativity. The cooperative binding effect is dependent on the ionic strength of the medium; when the NaCl concentration is reduced to 0.01 M, positive cooperativity is observed at carcinogen levels ? 1 carcinogen per 500 base pairs. The Scatchard data may be fit using a “two-site” binding model [L.S. Rosenberg, M J. Carvlin, and T.R. Krugh, Biochemistry 25, 1002–1008 (1986)]. This model assumes two independent sets of binding sites on the DNA lattice, one a high affinity site which binds the carcinogen with positive cooperativity, the second consisting of lower affinity binding sites to which non-specific binding occurs. NMR analysis of aflatoxin B2 binding to d(ATGCAT)2 indicates that the aflatoxin B2/oligodeoxynucleotide complex is in fast exchange on the NMR time scale. Upfield chemical shifts of 0.1–0.5 ppm are observed for the aflatoxin B2 4-OCH3, H5, and H6a protons. Much smaller chemical shift changes ? 0.06 ppm) are observed for the oligodeoxynucleotide protons. The greatest effect for the oligodeoxynucleotide protons is observed for the adenine H2 protons, located in the minor groove. Nonselective T1 experiments demonstrate a 15–25 % decrease in the relaxation time for the adenine H2 protons when aflatoxin B2 is added to the solution. This result suggests that aflatoxin B2 protons in the bound state may be in close proximity to these protons, providing a source of dipolar relaxation. Further experiments are in progress to probe the nature of the aflatoxin B1 and B2 complexes with polymeric DNA and oligodeoxynucleotides, and to establish the relationship between the non-covalent DNA-carcinogen complexes observed in these experiments, and covalent aflatoxin B1,-guanine N7 DNA adducts.  相似文献   

17.
Chicken erythrocyte inner histones (H2A, H2B, H3 and H4) were associated with the two complementary homopolymeric polydeoxyribonucleotides and the two alternating copolymeric polydeoxyribonucleotides. No evidence for formation of chromatin-like structures was obtained for the complexes with poly(dG) . poly(dC) or poly(dA) . poly(dT). Both poly (dGdC) . poly(dGdC) and poly(dAdT) . poly(dAdT) could be folded by histones to yield material digested by DNAase I to multiples of about 10 and by staphylococcal nuclease to 146 bp core particles. Due to the lack of sequence heterogeniety in the complex of histones with poly(dAdT) . poly(dAdT), core particles with remarkable fine structural detail are obtained. The internal organization of DNA in the AT-containing and GC-containing core particles appears not to be identical.  相似文献   

18.
C. P. Beetz  G. Ascarelli 《Biopolymers》1982,21(8):1569-1586
We have measured the ir absorption of 5′CMP, 5′IMP, and poly(I)·poly(C) from ~25 to ~500 cm?1. From a comparison of the data with the previously measured absorption of the corresponding nucleosides and bases we can identify several “lines” associated with the deformation of the ribose ring. Out-of-plane deformation of the bases contributes strongly to vibrations near 200 cm?1. The same ribose vibrations observed in the nucleotides are found in poly(I)·poly(C). They sharpen with increasing water absorption. A study of the spectra of poly(I)·poly(C) as a function of the adsorbed water indicates that water does not contribute in a purely additive fashion to the polynucleotide spectrum but depends on the conformation of the helix. However, the only spectral feature that shifts drastically with conformation is near 45 cm?1. Measurements at cryogenic temperatures indicate some sharpening of the spectrum of poly(I)·poly(C). Instead, no sharpening is observed in the spectrum of the nucleotides. Shear degradation of poly(I)·poly(C) produces significant spectral changes in the 200-cm?1 region and sharpening of the features assigned to the low-frequency ribose-ring vibrations.  相似文献   

19.
Metal-induced sequential transitions among DNA conformations   总被引:1,自引:0,他引:1  
The action of [Co(NH3)6]Cl3 on poly(dGdC) · poly(dGdC) can lead to a series of consecutive reactions, in which B-DNA is first converted to Z-DNA [M. Behe and G. Felsenfeld (1981) Proc. Natl. Acad. Sci USA 78 , 1619–1623], which in turn is transformed into an unidentified conformer that we tentatively call “U”, and finally the highly associated anisotropic ψ-DNA is produced. Conditions are given under which the sequence B ? Z ? “U” ψ(+) can be stopped at any point in the direction from left to right. The reverse processes, from right to left, occur when ψ(+) or “U”-DNA are treated with various amounts of salt concentrations and lowering temperature. Thus it is demonstrated that four conformers of poly(dGdC) · poly(dGdC) are readily interconvertible, and that Z-DNA and “U” conformers are intermediates in the reversible transformations of B- and ψ-DNA.  相似文献   

20.
Abstract

2′-Deoxy-2′-methylidenecytidine (DMDC), a potent inhibitor of the growth of tumor cells, was crystallized with two different forms. One is dihydrated (DMDC·2H2O) and the other is its hydrochloride salt (DMDC·HCLl). Both crystal and molecular structures have been determined by the X-ray diffraction method. In both forms the glycosidic and sugar conformations are anti and C(4′)-exo, respectively, whereas the conformation about the exocyclic bond is trans for DMDC·2H2O and gauche + for DMDC·HCl. Proton nuclear magnetic resonance data of DMDC indicate a preference for the anti C(4′)-exo conformation found in the solid state. These molecular conformations were compared with the related pyrimidine nucleosides. When the cytosine bases are brought into coincidence, DMDC displays the exocyclic C(4′)-C(5′) bond located on the very close position to those of pyrimidine nucleosides with typical overall conformations. On the other hand, the hydroxyl O(3′)-H groups are separated by ca. 3 Å in the cases of DMDC and other pyrimidine nucleosides which have the C(2′)-endo sugar conformation. This result may be useful for the implication about the mechanism of the biological activity of DMDC.  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