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 156 毫秒
1.
A nonoriented hydrated film of poly(dG-dC) with ?20 water molecules per nucleotide (called B* by Loprete and Hartman (Biochem. 32, 4077-4082 (1993)) was studied by Fourier transform infrared (FT-IR) spectroscopy either as equilibrated sample between 290 and 270 K or, after quenching into the glassy state, as nonequilibrated film isothermally at 200 and 220 K. IR spectral changes on isothermal relaxation at 200 and 220 K, caused by interconversion of two conformer substates, are revealed by difference spectra. Comparison with difference curves obtained in the same manner from two classical B-DNA forms, namely the d(CGCGAATTCGCG)(2) dodecamer and polymeric NaDNA from salmon testes, revealed that the spectral changes on B(I)-to-B(II) interconversion in the classical B-DNA forms are very similar to those in the B*-form, and that the spectroscopic differences between the B(I) and B(II) features from classical B-DNA and those from the modified B*-form are minor. Nonexponential kinetics of the B(I)-->B(II) transition in the B*-form of poly(dG-dC) at 200 K showed that the structural relaxation time is about three times of that in the classical B-DNA forms (approximately equal to 30 versus approximately equal to 10 min at 200 K). The unexpected reversal of conformer substates interconversion (that is B(II)-->B(I) transition on cooling from 290 K and B(I)-->B(II) transition on isothermal relaxation at 200 K) observed for classical B-DNA occurs also in the modified B*-form. We therefore conclude that restructuring of hydration shells rules the low-temperature dynamics of the B*-form via its two conformer substates in the same manner reported for classical B-DNA by Pichler et al. (J. Phys. Chem. B 106, 3263-3274 (2002)).  相似文献   

2.
Conformational substates of B-DNA had been observed so far in synthetic oligonucleotides but not in naturally occurring highly polymeric B-DNA. Our low-temperature experiments show that native B-DNA from salmon testes and the d(CGCGAATTCGCG)2 dodecamer have the same BI and BII substates. Nonequilibrium distribution of conformer population was generated by quenching hydrated unoriented films to 200 K, and isothermal structural relaxation toward equilibrium by interconversion of substates was followed by Fourier transform infrared spectroscopy. BI interconverts into BII on isothermal relaxation at 200 K, whereas on slow cooling from ambient temperature, BII interconverts into BI. Our estimation of the dodecamer's BI-to-BII conformer substate population by curve resolution of the symmetrical stretching vibration of the ionic phosphate is 2.4 +/- 0.5 to 1 at 200 K, and it is 1.3 +/- 0.5 to 1 between 270 and 290 K. Pronounced spectral changes upon BI-to-BII interconversion are consistent with base destacking coupled with migration of water from ionic phosphate toward the phosphodiester and sugar moieties. Nonspecific interaction of proteins with the DNA backbone could become specific by induced-fit-type interactions with either BI or BII backbone conformations. This suggests that the BI-to-BII substate interconversion could be a major contributor to the protein recognition process.  相似文献   

3.
Abstract

A nonoriented hydrated film of poly(dG-dC) with ≈20 water molecules per nucleotide (called B by Loprete and Hartman (Biochem. 32, 4077–4082 (1993)) was studied by Fourier transform infrared (FT-IR) spectroscopy either as equilibrated sample between 290 and 270 K or, after quenching into the glassy state, as nonequilibrated film isothermally at 200 and 220 K. IR spectral changes on isothermal relaxation at 200 and 220 K, caused by interconversion of two conformer substates, are revealed by difference spectra. Comparison with difference curves obtained in the same manner from two classical B-DNA forms, namely the d(CGCGAATTCGCG)2 dodecamer and polymeric NaDNA from salmon testes, revealed that the spectral changes on BIto-BII interconversion in the classical B-DNA forms are very similar to those in the B-form, and that the spectroscopic differences between the BI and BII features from classical B-DNA and those from the modified B-form are minor. Nonexponential kinetics of the BI→BII transition in the B-form of poly(dG-dC) at 200 K showed that the structural relaxation time is about three times of that in the classical B-DNA forms (≈30 versus ≈10 min at 200 K). The unexpected reversal of conformer substates interconversion (that is BII→BI transition on cooling from 290 K and BI→BII transition on isothermal relaxation at 200 K) observed for classical B-DNA occurs also in the modified B-form. We therefore conclude that restructuring of hydration shells rules the low-temperature dynamics of the B-form via its two conformer substates in the same manner reported for classical B-DNA by Pichler et al. (J. Phys. Chem. B 106, 3263–3274 (2002)).  相似文献   

4.
We use a recently developed formalism (1) to calculate the salt dependent part of the free energy determining DNA conformational stability in 1:1 electrolytes. The conformations studied are the A, B, C and alternating-B right-handed forms and the ZI, ZII left-handed forms of DNA. In the case of the B-ZI transition of d(G-C).d(G-C) helices in NaCl solution, the free energy contribution considered suffices to describe the transition in a quantitative manner. The theory also predicts the occurrence of salt-induced B-A transitions which have been recently observed with poly[d(n2 A-T)] and poly[d(G-C)]. In other cases, additional terms in the free energy balance, particularly due to hydration effects, must be at least as important as salt effects in determining conformational stability and structural transitions in solution. If diffuse ionic cloud electrostatic effects alone would dominate in all cases, the relative helical stabilities at 0.2 M monovalent salt would decrease in the order C greater than B greater than A greater than ZII greater than ZI greater than alternating-B. At high salt concentrations (2.0 M-5.0 M), the order would be alternating-B greater than ZI greater than A greater than ZII greater than B greater than C.  相似文献   

5.
2H NMR spin-lattice relaxation and line-shape analyses are performed to study the temperature-dependent dynamics of water in the hydration shells of myoglobin, elastin, and collagen. The results show that the dynamical behaviors of the hydration waters are similar for these proteins when using comparable hydration levels of h = 0.25–0.43. Since water dynamics is characterized by strongly nonexponential correlation functions, we use a Cole–Cole spectral density for spin-lattice relaxation analysis, leading to correlation times, which are in nice agreement with results for the main dielectric relaxation process observed for various proteins in the literature. The temperature dependence can roughly be described by an Arrhenius law, with the possibility of a weak crossover in the vicinity of 220 K. Near ambient temperatures, the results substantially depend on the exact shape of the spectral density so that deviations from an Arrhenius behavior cannot be excluded in the high-temperature regime. However, for the studied proteins, the data give no evidence for the existence of a sharp fragile-to-strong transition reported for lysozyme at about 220 K. Line-shape analysis reveals that the mechanism for the rotational motion of hydration waters changes in the vicinity of 220 K. For myoglobin, we observe an isotropic motion at high temperatures and an anisotropic large-amplitude motion at low temperatures. Both mechanisms coexist in the vicinity of 220 K. 13C CP MAS spectra show that hydration results in enhanced elastin dynamics at ambient temperatures, where the enhancement varies among different amino acids. Upon cooling, the enhanced mobility decreases. Comparison of 2H and 13C NMR data reveals that the observed protein dynamics is slower than the water dynamics.  相似文献   

6.
The structures of ZI- and ZII-form RNA and DNA oligonucleotides were energy minimized in vacuum using the AMBER molecular mechanics force field. Alternating C-G sequences were studied containing either unmodified nucleotides, 8-bromoguanosine in place of all guanosine residues, 5-bromocytidine in place of all cytidine residues, or all modified residues. Some molecules were also energy minimized in the presence of H2O and cations. Free energy perturbation calculations were done in which G8 and C5 hydrogen atoms in one or two residues of Z-form RNAs and DNAs were replaced in a stepwise manner by bromines. Bromination had little effect on the structures of the energy-minimized molecules. Both the minimized molecular energies and the results of the perturbation calculations indicate that bromination of guanosine at C8 will stabilize the Z forms of RNA and DNA relative to the nonbrominated Z form, while bromination of cytidine at C5 stabilizes Z-DNA and destabilizes Z-RNA. These results are in agreement with experimental data. The destabilizing effect of br5C in Z-RNAs is apparently due to an unfavorable interaction between the negatively charged C5 bromine atom and the guanosine hydroxyl group. The vacuum-minimized energies of the ZII-form oligonucleotides are lower than those of the corresponding ZI-form molecules for both RNA and DNA. Previous x-ray diffraction, nmr, and molecular mechanics studies indicate that hydration effects may favor the ZI conformation over the ZII form in DNA. Molecular mechanics calculations show that the ZII-ZI energy differences for the RNAs are greater than three times those obtained for the DNAs. This is due to structurally reinforcing hydrogen-bonding interactions involving the hydroxyl groups in the ZII form, especially between the guanosine hydroxyl hydrogen atom and the 3'-adjacent phosphate oxygen. In addition, the cytidine hydroxyl oxygen forms a hydrogen bond with the 5'-adjacent guanosine amino group in the ZII-form molecule. Both of these interactions are less likely in the ZI-form molecule: the former due to the orientation of the GpC phosphate away from the guanosine ribose in the ZI form, and the latter apparently due to competitive hydrogen bonding of the cytidine 2'-hydroxyl hydrogen with the cytosine carbonyl oxygen in the ZI form. The hydrogen-bonding interaction between the cytidine hydroxyl oxygen and the 5'-adjacent guanosine amino group in Z-RNA twists the amino group out of the plane of the base. This may be responsible for differences in the CD and Raman spectra of Z-RNA and Z-DNA.  相似文献   

7.
8.
The product of the Epstein-Barr virus BZLF1 gene encodes a protein which is related to c-fos, it has been shown to bind specifically to a consensus AP-1 site, and its expression in latently Epstein-Barr virus-infected lymphocytes is sufficient to trigger the viral lytic cycle. We identified several elements within the BZLF1 promoter (Zp) which are responsive to the phorbol ester 12-O-tetradecanoylphorbol-13-acetate (TPA), an inducer of the viral lytic cycle. These elements fall into two classes based on the factors which bind to these sequences and their resulting functional behavior. Four of the elements are homologous (ZI elements) and share homology to a protein-binding domain in the promoter region of the coordinately expressed BRLF1 gene. When cloned upstream of heterologous promoters, the ZI elements function as silencers which exhibit TPA-inducible enhancer activity. A distinct TPA-responsive element (ZII) is located near the TATA box and shares homology with the AP-1-binding site in the c-jun promoter. A synthetic oligonucleotide with a sequence corresponding to the ZII element effectively competes for binding of nuclear factors to the c-jun AP-1 site. Furthermore, we found that a complex of c-jun and c-fos bound to the ZII domain.  相似文献   

9.
A numerical counterion condensation analysis of the B-Z transition of DNA   总被引:1,自引:0,他引:1  
M O Fenley  G S Manning  W K Olson 《Biopolymers》1990,30(13-14):1205-1213
  相似文献   

10.
The rates at which a peptide hexamer and a peptide octamer interconvert between left- and right-handed helical forms in CD2Cl2 solution have been characterized by 13C dynamic NMR (DNMR) spectroscopy. The peptide esters studied are Fmoc-(Aib)n-OtBu (n = 6 and 8), where Fmoc is 9-fluorenylmethyoxycarbonyl and Aib is the strongly helix-forming residue alpha-aminoisobutyric acid. Because the Aib residue is itself achiral, homooligomers of this residue form a 50/50 mixture of enantiomeric 3(10)-helices in solution. It has been demonstrated (R.-P. Hummel, C. Toniolo, and G. Jung, Angewandte Chemie International Edition, 1987, Vol. 26, pp. 1150-1152) that oligomers of Aib interconvert on the millisecond timescale. We have performed lineshape analysis of 13C-NMR spectra collected for our peptides enriched with 13C at a single residue. Rate constants for the octamer range from 6 s(-1) at 196 K to about 56,500 s(-1) at 320 K. At all temperatures, the hexamer interconverts about three times faster than the octamer. Eyring plots of the data reveal experimentally indistinguishable DeltaH++ values for the hexamer and octamer of 37.8 +/- 0.6 and 37.6 +/- 0.4 kJ mol(-1) respectively. The difference in the rates of interconversion is dictated by entropic factors. The hexamer and octamer exhibit negative DeltaS++ values of -29.0(-1) +/- 2.5 and -37.3 +/- 1.7 J K(-1) mol(-1), respectively. A mechanism for the helix-helix interconversion is proposed. and calculated DeltaG++ values are compared to the estimate for a decamer undergoing a helix-helix interconversion.  相似文献   

11.
The (1)H- and (13)C-NMR spectra of antifreeze glycoprotein fractions 1-5 from Antarctic cod have been assigned, and the dynamics have been measured using (13)C relaxation at two temperatures. The chemical shifts and absence of non-sequential (1)H-(1)H NOEs are inconsistent with a folded, compact structure. (13)C relaxation measurements show that the protein has no significant long-range order, and that the local correlation times are adequately described by a random coil model. Hydroxyl protons of the sugar residues were observed at low temperature, and the presence of exchange-mediated ROEs to the sugar indicate extensive hydration. The conformational properties of AFGP1-5 are compared with those of the previously examined 14-mer analog AFGP8, which contains proline residues in place of some alanine residues (Lane, A. N., L. M. Hays, R. E. Feeney, L. M. Crowe, and J. H. Crowe. 1998. Protein Sci. 7:1555-1563). The infrared (IR) spectra of AFGP8 and AFGP1-5 in the amide I region are quite different. The presence of a wide distribution of backbone torsion angles in AFGP1-5 leads to a rich spectrum of frequencies in the IR spectrum, as interconversion among conformational states is slow on the IR frequency time scale. However, these transitions are fast on the NMR chemical shift time scales. The restricted motions for AFGP8 may imply a narrower distribution of possible o, psi angles, as is observed in the IR spectrum. This has significance for attempts to quantify secondary structures of proteins by IR in the presence of extensive loops.  相似文献   

12.
1. Three zymogens have been isolated from human gastric mucosae and two from the stomachs of the toad Caudiverbera caudiverbera. 2. Human zymogens I and III were immunologically related and cross-reacted with antisera prepared against porcine pepsinogen. The third, (II), showed no cross-reactivity. 3. Human zymogens I and III and toad zymogen ZII gave rise to two human pepsins and to a pepsin-like enzyme, respectively. 4. Human zymogen II (gastricsinogen) and toad zymogen ZI gave rise to human gastricsin and to a gastricsin-like enzyme respectively. 5. The toad enzymes showed much greater stability at neutral and alkaline pH values than the human enzymes.  相似文献   

13.
Induction of the Epstein-Barr virus lytic cycle is mediated through the immediate-early BZLF1 gene and the coordinately regulated BRLF1 gene. The BZLF1 gene product, Zta, transactivates its own promoter, as well as the promoters of a number of lytic genes, thereby initiating a cascade of viral gene expression. Previous work identified four related elements (ZIA, ZIB, ZIC, and ZID) and a cyclic AMP response element binding-AP-1 element (ZII) that are involved in the induction of the BZLF1 promoter (Zp) by the phorbol ester 12-O-tetradecanoylphorbol-13-acetate (TPA) (E. Flemington and S. H. Speck, J. Virol. 64:1217-1226, 1990). Here we report a detailed characterization of TPA induction mediated by the ZI domains. Mutation of individual ZI domains within the context of the intact promoter significantly diminished TPA induction. Cloning of individual ZI domains upstream of a minimal promoter demonstrated that the ZIA, ZIC, and ZID domains, but not the ZIB domain, are TPA responsive. Furthermore, cloning of the ZII domain downstream of the ZI domains significantly augmented TPA induction. The critical regions within the ZIA and ZIC elements involved in binding of cellular factors were identified by using methylation interference and electrophoretic mobility shift analyses of ZI domain mutants. Four specific complexes were observed with the ZIA and ZID domains, all of which could be specifically competed for by either the ZIA or ZID domain. Methylation interference analyses of bound complexes revealed the presence of two overlapping binding sites for cellular factors in the ZIA domain, and functional studies provided evidence that both of these sites are involved in TPA induction. Functional analyses of the ZIC domain revealed that the 5' region of this domain is largely responsible for mediating TPA induction. Binding data correlated well with functional activity and revealed that the ZIC domain binds only a subset of the cellular factors that bind to the ZIA and ZID domains. Analysis of factor binding to the ZIB domain revealed only a single shifted complex, which correlated with the most slowly migrating complex observed with the ZIA and ZID domains. These data provide a direct demonstration of TPA induction mediated by the ZIA, ZIC, and ZID domains and also provide the first evidence that the ZI domains exhibit distinct functional characteristics.  相似文献   

14.
The poly(dG-dC) helical duplex forms a modified, B-family structure (B*) at very high hydration and a normal B structure at slightly lower hydration. The B* structure is slightly different in sugar-phosphate and base-stacking conformations than the B structure. Increasing the hydration or decreasing the NaCl content stabilizes B* with respect to B. Poly(dG-dC) forms the Z structure at low NaCl contents when the hydration is sufficiently reduced. At moderate NaCl content, the B to Z transition is sharp and cooperative for hydration with D2O. Hydration with H2O broadens the transition which occurs at lower hydration. This suggests that hydrogen bonding is stronger in the Z structure and helps stabilize Z over B. IR spectra may be used to quantitatively estimate the fractions of B and Z structures present in a sample. Some new indicator bands are described.  相似文献   

15.
From experimental data on the hydration energetics of nucleic acids obtained by differential scanning calorimetry under isothermal conditions, dielectric relaxation time tau d and "free volume" Vf occupied by water molecules in hydration shells of natural DNA and model polyribonucleotides were calculated. In addition, systems consisting of dinucleotides ApA, TpT, UpU, TpU, UpT and water clusters of various sizes (from 20 to 400 water molecules) were studied by Monte Carlo computer simulation. It was shown that, as water content in systems increases, the dynamic characteristics of bound water obtained with both methods approached the values for bulk water.  相似文献   

16.
The dynamic properties of water in the hydration shell of hemoglobin have been studied by means of dielectric permittivity measurements and nuclear magnetic resonance spectroscopy. The temperature behavior of the complex permittivity of hemoglobin solutions has been measured at 3.02, 3.98, 8.59, and 10.80 GHz. At a temperature of 298 K the average rotational correlation time tau of water within a hydration shell of 0.5-nm thickness is determined from the activation parameters to be 68 +/- 10 ps, which is 8-fold the corresponding value of bulk water. Solvent proton magnetic relaxation induced by electron-nuclear dipole interaction between hemoglobin bound nitroxide spin labels and water protons is used to determine the translational diffusion coefficient D(T) of the hydration water. The temperature dependent relaxation behavior for Lamor frequencies between 3 and 90 MHz yields an average value D(298K) = (5 +/- 2) x 10(-10)m2 s-1, which is about one-fifth of the corresponding value of bulk water. The decrease of the water mobility in the hydration shell compared to the bulk is mainly due to an enhanced activation enthalpy.  相似文献   

17.
M A Eriksson  A Laaksonen 《Biopolymers》1992,32(8):1035-1059
Twelve dinucleotides (one complete turn) of left-handed, flexible, double-helix poly(dG-dC) Z-DNA have been simulated in aqueous solution with K+ counterions for 70 ps. Most of the d(GpC) phosphates have rotated in accordance with a ZI----ZII transition. The ZII conformation was probably partly stabilized by counterions, which coordinate one of the anionic oxygens and the guanine-N7 of the next (5'----3' direction) base. The presence of base-coordinating ions close to the helical axis rotated and pulled about half of the d(CpG) phosphates further into the groove. These ions also gave rise to rather large deviations from the crystal structure (ZI) with their tendency of pulling the bases closer toward the helical axis. A flipping of the orientation about the glycosyl bond from the +sc to the -sc region was observed for one guanosine, also leading to deviations from the crystal structure. Many bridges containing one or two water molecules were found, with a dominance for the latter. They essentially formed a network of intra- and interstrand bridges between anionic and esterified phosphate oxygens. A "spine" of water molecules could be distinguished as a dark zig-zag pattern in the water density map. The lifetime of a bridge containing one water was about twice as long as that of a two-water bridge and it lasted 5-15 times longer than a hydrogen bond in water. The lifetimes were also calculated for a selection of bridge types, in order of decreasing stability: O1P/O2P ... W ... O'4 much greater than O1P/O2P ... W ... guanine-N2 greater than O1P/O2P ... W ... O1P/O2P. The reorientational motion of water molecules in the first hydration shell around selected groups was slowed down considerably compared to bulk water and the decreasing order of correlation times was guanine-N2 greater than O'4 greater than O'3/O'5 greater than O1P/O2P.  相似文献   

18.
Quasielastic neutron and light-scattering techniques along with molecular dynamics simulations were employed to study the influence of hydration on the internal dynamics of lysozyme. We identified three major relaxation processes that contribute to the observed dynamics in the picosecond to nanosecond time range: 1), fluctuations of methyl groups; 2), fast picosecond relaxation; and 3), a slow relaxation process. A low-temperature onset of anharmonicity at T approximately 100 K is ascribed to methyl-group dynamics that is not sensitive to hydration level. The increase of hydration level seems to first increase the fast relaxation process and then activate the slow relaxation process at h approximately 0.2. The quasielastic scattering intensity associated with the slow process increases sharply with an increase of hydration to above h approximately 0.2. Activation of the slow process is responsible for the dynamical transition at T approximately 200 K. The dependence of the slow process on hydration correlates with the hydration dependence of the enzymatic activity of lysozyme, whereas the dependence of the fast process seems to correlate with the hydration dependence of hydrogen exchange of lysozyme.  相似文献   

19.
Molecular dynamics simulations of the Z-DNA hexamer 5BrdC-dG-5BrdC-dG-5BrdC-dG were performed at several temperatures between 100 K and 300 K. Above 250 K, a strong sequence-dependent flexibility in the nucleic acid is observed, with the guanine sugar and the phosphate of GpC sequences much more mobile than the cytosine sugar and phosphate of CpG sequences. At 300 K, the hexamer is in dynamic equilibrium between several Z forms, including the crystallographically determined ZI and ZII forms. The local base-pair geometry, however, is not very variable, except for the roll of the base-pairs. The hexamer molecular dynamics trajectories have been used to test the restrained parameter crystallographic refinement model for nucleic acids. X-ray diffraction intensities corresponding to observed diffraction data were computed. The average structures obtained from the simulations were then refined against the calculated intensities, using a restrained least-squares program developed for nucleic acids in order to analyse the effects of the refinement model on the derived quantities. In general, the temperature dependence of the atomic fluctuations determined directly from the refined Debye-Waller factors is in reasonably good agreement with the results obtained by calculating the atomic fluctuations directly from the Z-DNA molecular dynamics trajectories. The agreement is best for refinement of temperature factors without restraints. At the highest temperature studied (300 K), the effect of the refinement on the most mobile atoms (phosphates) is to significantly reduce the mean-square atomic fluctuations estimated from the refined Debye-Waller factors below the actual values (less than (delta r)2 greater than congruent to 0.5 A2). Analysis of the temperature-dependence of the mean-square atomic fluctuations provides information concerning the conformational potential within which the atoms move. The calculated temperature-dependence and anharmonicity of the Z-DNA helix are compared with the results observed for proteins. The average structures from the simulations were refined against the experimental X-ray intensities. It is found that low-temperature molecular dynamics simulations provide a useful tool for optimizing the refinement of X-ray structures.  相似文献   

20.
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