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 125 毫秒
1.
Coenzyme B(12) serves as a cofactor for enzymatic radical reactions. The essential steps in all the coenzyme B(12)-dependent rearrangements are two hydrogen abstraction steps: hydrogen abstraction of the adenosyl radical from substrates, and hydrogen back-abstraction (recombination) of a product-derived radical from 5'-deoxyadenosine. The energetic feasibility of these hydrogen abstraction steps in the diol dehyratase reaction was examined by theoretical calculations with a protein-free, simplified model at the B3LYP/6-311G* level of density functional theory. Activation energies for the hydrogen abstraction and recombination with 1,2-propanediol as substrate are 9.0 and 15.1 kcal/mol, respectively, and essentially not affected by coordination of the substrate and the radical intermediate to K+. Since these energies can be considered to be supplied by the substrate-binding energy, the computational results with this simplified model indicate that the hydrogen abstraction and recombination in the coenzyme B(12)-dependent diol dehydratase reaction are energetically feasible.  相似文献   

2.
We have used molecular dynamics simulations to determine the stability in water of a model beta-sheet formed by two alanine dipeptide molecules with two intermolecular hydrogen bonds in the closely spaced antiparallel arrangement. In this paper we describe our computations of the binding free energy of the model sheet and a portion of the free energy surface as a function of a reaction co-ordinate for sheet formation. We used the free energy surface to identify stable conformations along the reaction co-ordinate. To determine whether or not the model sheet with two hydrogen bonds is more stable than a single amide hydrogen bond in water, we compared the results of the present calculations to results from our earlier study of linear hydrogen bond formation between two formamide molecules (the formamide "dimer"). The free energy surfaces for the sheet and formamide dimer each have two minima corresponding to locally stable hydrogen-bonded and solvent-separated configurations. The binding free energies of the model sheet and the formamide dimer are -5.5 and -0.34 kcal/mol, respectively. Thus, the model sheet with two hydrogen bonds is quite stable while the simple amide hydrogen bond is only marginally stable. To understand the relative stabilities of the model sheet and formamide dimer in terms of solute-solute and solute-water interactions, we decomposed the free energy differences between hydrogen-bonded and solvent-separated conformations into energetic and entropic contributions. The changes in the peptide-peptide energy and the entropy are roughly twice as large for the sheet as they are for the formamide dimer. The magnitude of the peptide-water energy difference for the sheet is less than twice (by about 3.5 kcal/mol) that for the formamide dimer, and this accounts for the stability of the sheet. The presence of the side-chains and/or blocking groups apparently prevents the amide groups in the sheet from being solvated as favorably in the separated arrangement as in the formamide dimer, where the amide groups are completely exposed to the solvent.  相似文献   

3.
Mechanism-based inhibitors of dopamine beta-hydroxylase   总被引:2,自引:0,他引:2  
The copper-containing monooxygenase dopamine beta-hydroxylase catalyzes the hydroxylation of dopamine at the benzylic position to form norepinephrine. Mechanism-based inhibitors for dopamine beta-hydroxylase have been used as probes of the mechanism of catalysis. The variety of such inhibitors that have been developed for this enzyme can be divided into three groups: (i) those in which the inactivating species is formed by abstraction of a hydrogen atom to form a radical intermediate; (ii) those in which the inactivating species is formed by abstraction of an electron to form an epoxide-like intermediate; and (iii) those in which the product is the inactivating species. A mechanism consistent with inactivation by all three groups of inhibitors which proposes that hydroxylation of dopamine by dopamine beta-hydroxylase involves formation of a benzylic radical has been developed. The benzylic radical is formed by abstraction of a hydrogen atom from the substrate by a high-potential copper-oxygen species.  相似文献   

4.
A parallel study of the radical copper enzyme galactose oxidase (GOase) and a low molecular weight analog of the active site was performed with dynamical density functional and mixed quantum-classical calculations. This combined approach enables a direct comparison of the properties of the biomimetic and the natural systems throughout the course of the catalytic reaction. In both cases, five essential forms of the catalytic cycle have been investigated: the resting state in its semi-reduced (catalytically inactive) and its oxidized (catalytically active) form, A semi and A ox, respectively; a protonated intermediate B; the transition state for the rate-determining hydrogen abstraction step C, and its product D. For A and B the electronic properties of the biomimetic compound are qualitatively very similar to the ones of the natural target. However, in agreement with the experimentally observed difference in catalytic activity, the calculated activation energy for the hydrogen abstraction step is distinctly lower for GOase (16 kcal/mol) than for the mimetic compound (21 kcal/mol). The enzymatic transition state is stabilized by a delocalization of the unpaired spin density over the sulfur-modified equatorial tyrosine Tyr272, an effect that for geometric reasons is essentially absent in the biomimetic compound. Further differences between the mimic and its natural target concern the structure of the product of the abstraction step, which is characterized by a weakly coordinated aldehyde complex for the latter and a tightly bound linear complex for the former. Received 14 October 1999 · Accepted: 19 January 2000  相似文献   

5.
The seemingly simple proton abstraction reactions underpin many chemical transformations, including isomerization reactions, and are thus of immense biological significance. Despite the energetic cost, enzyme-catalyzed proton abstraction reactions show remarkable rate enhancements. The pathways leading to these accelerated rates are numerous and on occasion partly enigmatic. The isomerization of the steroid Δ5-androstene-3,17-dione by the glutathione transferase A3-3 in mammals was investigated to gain insight into the mechanism. Particular emphasis was placed on the nature of the transition state, the intermediate suspected of aiding this process, and the hydrogen bonds postulated to be the stabilizing forces of these transient species. The UV-visible detection of the intermediate places this species in the catalytic pathway, whereas fluorescence spectroscopy is used to obtain the binding constant of the analog intermediate, equilenin. Solvent isotope exchange reveals that proton abstraction from the substrate to form the intermediate is rate-limiting. Analysis of the data in terms of the Marcus formalism indicates that the human glutathione transferase A3-3 lowers the intrinsic kinetic barrier by 3 kcal/mol. The results lead to the conclusion that this reaction proceeds through an enforced concerted mechanism in which the barrier to product formation is kinetically insignificant.  相似文献   

6.
A DFT study with QST3 approach method is used to calculate kinetic, thermodynamic, spectral and structural data of tautomers and transition state structures of some N-hydroxy amidines. All tautomers and transition states are optimized at the B3LYP/6-311++g** and B3LYP/aug-cc-pvtz level, with good agreement in energetic result with energies obtained from CBS-QB3, a complete basis set composite energy method. The result shows that the tautomer a (amide oxime) is more stable than the tautomer b (imino hydroxylamine) as is reported in the literature. In addition, our finding shows that, the energy difference between two tautomers is only in about 4–10 kcal/mol but the barrier energy found in traversing each tautomer to another one is in the range of 33–71 kcal/mol. Therefore, it is impossible to convert these two tautomers to each other at room temperature. Additionally, transition state theory is applied to estimate the barrier energy and reaction rate constants of the hydrogen exchange between tautomers in presence of 1–3 molecules of water. The computed activation barrier shows us that the barrier energy of solvent assisted tautomerism is about 9–20 kcal/mol and lower than simple tautomerism and this water-assisted tautomerism is much faster than simple tautomerism, especially with the assisting two molecules of water.  相似文献   

7.
Aromatic rings act as hydrogen bond acceptors   总被引:24,自引:0,他引:24  
Simple energy calculations show that there is a significant interaction between a hydrogen bond donor (like the greater than NH group) and the centre of a benzene ring, which acts as a hydrogen bond acceptor. This interaction, which is about half as strong as a normal hydrogen bond, contributes approximately 3 kcal/mol (1 cal = 4.184 J) of stabilizing enthalpy and is expected to play a significant role in molecular associations. It is of interest that the aromatic hydrogen bond arises from small partial charges centred on the ring carbon and hydrogen atoms: there is no need to consider delocalized electrons. Although some energy calculations have included such partial charges, their role in forming such a strong interaction was not appreciated until after aromatic hydrogen bonds had been observed in protein-drug complexes.  相似文献   

8.
The RNA recognition motif (RRM) binds to single-stranded RNA target sites of diverse sequences and structures. A conserved mode of base recognition by the RRM involves the simultaneous formation of a network of hydrogen bonds with the base functional groups and a stacking interaction between the base and a highly conserved aromatic amino acid. We have investigated the energetic contribution of the functional groups involved in the recognition of an essential adenine, A6, in stem–loop 2 of U1 snRNA by the N-terminal RRM of the U1A protein. Previously, we found that elimination of individual hydrogen bond donors and acceptors on A6 destabilized the complex by 0.8–1.9 kcal/mol, while mutation of the aromatic amino acid (Phe56) that stacks with A6 to Ala destabilized the complex by 5.5 kcal/mol. Here we continue to probe the contribution of A6 to complex stability through mutation of both the RNA and protein. We have removed two hydrogen-bonding functional groups by introducing a U1A mutation, Ser91Ala, and replacing A6 with tubercidin, purine, or 1-deazaadenine. We find that the complex is destabilized an additional 1.2–2.6 kcal/mol by the elimination of the second hydrogen bond donor or acceptor. Surprisingly, deletion of all of the functional groups involved in hydrogen bonds with the U1A protein by substituting adenine with 4-methylindole reduced the binding free energy by only 2.0 kcal/mol. Experiments with U1A proteins containing mutations of Phe56 suggested that improved stacking interactions due to the greater hydrophobicity of 4-methylindole than adenine may be partly responsible for the small destabilization of the complex upon substitution of 4-methylindole for A6. The data imply that hydrophobic interactions can compensate energetically for the disruption of the complex hydrogen-bonding network between nucleotide and protein.  相似文献   

9.
Cooperative hydrogen bond interactions in the streptavidin-biotin system   总被引:1,自引:0,他引:1  
The thermodynamic and structural cooperativity between the Ser45- and D128-biotin hydrogen bonds was measured by calorimetric and X-ray crystallographic studies of the S45A/D128A double mutant of streptavidin. The double mutant exhibits a binding affinity approximately 2x10(7) times lower than that of wild-type streptavidin at 25 degrees C. The corresponding reduction in binding free energy (DeltaDeltaG) of 10.1 kcal/mol was nearly completely due to binding enthalpy losses at this temperature. The loss of binding affinity is 11-fold greater than that predicted by a linear combination of the single-mutant energetic perturbations (8.7 kcal/mol), indicating that these two mutations interact cooperatively. Crystallographic characterization of the double mutant and comparison with the two single mutant structures suggest that structural rearrangements at the S45 position, when the D128 carboxylate is removed, mask the true energetic contribution of the D128-biotin interaction. Taken together, the thermodynamic and structural analyses support the conclusion that the wild-type hydrogen bond between D128-OD and biotin-N2 is thermodynamically stronger than that between S45-OG and biotin-N1.  相似文献   

10.
Calculations predict that cation- interactions make an important contribution to protein stability. While there have been some attempts to experimentally measure strengths of cation-pi interactions using peptide model systems, much less experimental data are available for globular proteins. We have attempted to determine the magnitude of cation-pi interactions of Lys with aromatic amino acids in four different proteins (LIVBP, MBP, RBP, and Trx). In each case, Lys was replaced with Gln and Met. In a separate series of experiments, the aromatic amino acid in each cation-pi pair was replaced by Leu. Stabilities of wild-type (WT) and mutant proteins were characterized by both thermal and chemical denaturation. Gln and aromatic --> Leu mutants were consistently less stable than corresponding Met mutants, reflecting the nonisosteric nature of these substitutions. The strength of the cation-pi interaction was assessed by the value of the change in the free energy of unfolding [DeltaDeltaG(degrees) = DeltaG(degrees)(Met) - DeltaG(degrees)(WT)]. This ranged from +1.1 to -1.9 kcal/mol (average value -0.4 kcal/mol) at 298 K and +0.7 to -2.6 kcal/mol (average value -1.1 kcal/mol) at the Tm of each WT. It therefore appears that the strength of cation-pi interactions increases with temperature. In addition, the experimentally measured values are appreciably smaller in magnitude than calculated values with an average difference /DeltaG(degrees)expt - DeltaG(degrees)calc/av of 2.9 kcal/mol. At room temperature, the data indicate that cation-pi interactions are at best weakly stabilizing and in some cases are clearly destabilizing. However, at elevated temperatures, close to typical Tm's, cation-pi interactions are generally stabilizing.  相似文献   

11.
It is difficult to increase protein stability by adding hydrogen bonds or burying nonpolar surface. The results described here show that reversing the charge on a side chain on the surface of a protein is a useful way of increasing stability. Ribonuclease T1 is an acidic protein with a pI approximately 3.5 and a net charge of approximately -6 at pH 7. The side chain of Asp49 is hyperexposed, not hydrogen bonded, and 8 A from the nearest charged group. The stability of Asp49Ala is 0.5 kcal/mol greater than wild-type at pH 7 and 0.4 kcal/mol less at pH 2.5. The stability of Asp49His is 1.1 kcal/mol greater than wild-type at pH 6, where the histidine 49 side chain (pKa = 7.2) is positively charged. Similar results were obtained with ribonuclease Sa where Asp25Lys is 0.9 kcal/mol and Glu74Lys is 1.1 kcal/mol more stable than the wild-type enzyme. These results suggest that protein stability can be increased by improving the coulombic interactions among charged groups on the protein surface. In addition, the stability of RNase T1 decreases as more hydrophobic aromatic residues are substituted for Ala49, indicating a reverse hydrophobic effect.  相似文献   

12.
Equilibrium H/D fractionation factors have been extensively employed to qualitatively assess hydrogen bond strengths in protein structure, enzyme active sites, and DNA. It remains unclear how fractionation factors correlate with hydrogen bond free energies, however. Here we develop an empirical relationship between fractionation factors and free energy, allowing for the simple and quantitative measurement of hydrogen bond free energies. Applying our empirical relationship to prior fractionation factor studies in proteins, we find: [1] Within the folded state, backbone hydrogen bonds are only marginally stronger on average in α‐helices compared to β‐sheets by ~0.2 kcal/mol. [2] Charge‐stabilized hydrogen bonds are stronger than neutral hydrogen bonds by ~2 kcal/mol on average, and can be as strong as –7 kcal/mol. [3] Changes in a few hydrogen bonds during an enzyme catalytic cycle can stabilize an intermediate state by –4.2 kcal/mol. [4] Backbone hydrogen bonds can make a large overall contribution to the energetics of conformational changes, possibly playing an important role in directing conformational changes. [5] Backbone hydrogen bonding becomes more uniform overall upon ligand binding, which may facilitate participation of the entire protein structure in events at the active site. Our energetic scale provides a simple method for further exploration of hydrogen bond free energies.  相似文献   

13.
Evidence from proteins and peptides supports the conclusion that intrapeptide hydrogen bonds stabilize the folded form of proteins. Paradoxically, evidence from small molecules supports the opposite conclusion, that intrapeptide hydrogen bonds are less favorable than peptide-water hydrogen bonds. A related issue-often lost in this debate about comparing peptide-peptide to peptide- water hydrogen bonds-involves the energetic cost of an unsatisfied hydrogen bond. Here, experiment and theory agree that breaking a hydrogen bond costs between 5 and 6 kcal/mol. Accordingly, the likelihood of finding an unsatisfied hydrogen bond in a protein is insignificant. This realization establishes a powerful rule for evaluating protein conformations.  相似文献   

14.
Purified manganese peroxidase (MnP) from Phanerochaete chrysosporium oxidizes nonphenolic beta-1 diarylpropane lignin model compounds in the presence of Tween 80, and in three- to fourfold lower yield in its absence. In the presence of Tween 80, 1-(3',4'-diethoxyphenyl)-1-hydroxy-2-(4'-methoxyphenyl)propane (I) was oxidized to 3,4-diethoxybenzaldehyde (II), 4-methoxyacetophenone (III) and 1-(3',4'-diethoxyphenyl)-1-oxo-2-(4'-methoxyphenyl)propane (IV), while only 3,4-diethoxybenzaldehyde (II) and 4-methoxyacetophenone (III) were detected when the reaction was conducted in the absence of Tween 80. In contrast to the oxidation of this substrate by lignin peroxidase (LiP), oxidation of substrates by MnP did not proceed under anaerobic conditions. When the dimer (I) was deuterated at the alpha position and subsequently oxidized by MnP in the presence of Tween 80, yields of 3,4-diethoxybenzaldehyde, 4-methoxyacetophenone remained constant, while the yield of the alpha-keto dimeric product (IV) decreased by approximately sixfold, suggesting the involvement of a hydrogen abstraction mechanism. MnP also oxidized the alpha-keto dimeric product (IV) to yield 3,4-diethoxybenzoic acid (V) and 4-methoxyacetophenone (III), in the presence and, in lower yield, in the absence of Tween 80. When the reaction was performed in the presence of 18O2, both products, 3,4-diethoxybenzoic acid and 4-methoxyacetophenone, contained one atom of 18O. Finally, MnP oxidized the substrate 1-(3',5'-dimethoxyphenyl)-1-hydroxy-2-(4'-methoxyphenyl)propane (IX) to yield 3,5-dimethoxybenzaldehyde (XI), 4-methoxyacetophenone (III) and 1-(3',5'-dimethoxyphenyl)-1-oxo-2-(4'-methoxyphenyl)propane (X). In sharp contrast, LiP was not able to oxidize IX. Based on these results, we propose a mechanism for the MnP-catalyzed oxidation of these dimers, involving hydrogen abstraction at a benzylic carbon, rather than electron abstraction from an aromatic ring.  相似文献   

15.
Dopamine beta-hydroxylase is inactivated by phenyl-, phenethyl-, benzyl-, and methylhydrazine, but not by hydrazine itself. With phenyl-, methyl-, and phenethylhydrazine, the rate of inactivation decreases in the presence of ascorbate and increases in the presence of tyramine. Reduction of the enzyme-bound copper occurs with all of the hydrazines tested. In the presence of the spin trap alpha-(4-pyridyl-1-oxide)-N-tert-butylnitrone the carbon-centered radicals generated from each compound are trapped. This is consistent with reduction of the enzyme-bound copper by the hydrazine-containing compounds, resulting in formation of the hydrazine cation radical. Homolytic cleavage of the carbon-nitrogen bond then generates a carbon-centered radical which reacts with the enzyme, resulting in inactivation. Inactivation with [14C]phenylhydrazine results in the incorporation of 0.94 molecule of label per enzyme subunit. Benzylhydrazine behaves as a mechanism-based inhibitor of the enzyme. Both benzyl- and phenethylhydrazine are substrates for dopamine beta-hydroxylase. The second-order rate constant for inactivation of dopamine beta-hydroxylase by benzylhydrazine in the presence of ascorbate is increased about 4-fold when the benzylic hydrogens are replaced with deuterium. The apparent Vmax shows an observed deuterium kinetic isotope effect of 13 +/- 2. The partition ratio for product formation versus inactivation is 11-fold less for alpha,alpha-d2-benzylhydrazine. These results are interpreted in terms of a model where inactivation is due to abstraction of an electron from nitrogen instead of abstraction of a hydrogen atom from the benzylic carbon.  相似文献   

16.
The classical Zn finger contains a phenylalanine at the crux of its three architectural elements: a beta-hairpin, an alpha-helix, and a Zn(2+)-binding site. Surprisingly, phenylalanine is not required for high-affinity Zn2+ binding, but instead contributes to the specification of a precise DNA-binding surface. Substitution of phenylalanine by leucine leads to a floppy but native-like structure whose Zn affinity is maintained by marked entropy-enthalpy compensation (DeltaDeltaH -8.3 kcal/mol and -TDeltaDeltaS 7.7 kcal/mol). Phenylalanine and leucine differ in shape, size, and aromaticity. To distinguish which features correlate with dynamic stability, we have investigated a nonstandard finger containing cyclohexanylalanine at this site. The structure of the nonstandard finger is similar to that of the native domain. The cyclohexanyl ring assumes a chair conformation, and conformational fluctuations characteristic of the leucine variant are damped. Although the nonstandard finger exhibits a lower affinity for Zn2+ than does the native domain (DeltaDeltaG -1.2 kcal/mol), leucine-associated perturbations in enthalpy and entropy are almost completely attenuated (DeltaDeltaH -0.7 kcal/mol and -TDeltaDeltaS -0.5 kcal/mol). Strikingly, global changes in entropy (as inferred from calorimetry) are in each case opposite in sign from changes in configurational entropy (as inferred from NMR). This seeming paradox suggests that enthalpy-entropy compensation is dominated by solvent reorganization rather than nominal molecular properties. Together, these results demonstrate that dynamic and thermodynamic perturbations correlate with formation or repair of a solvated packing defect rather than type of physical interaction (aromatic or aliphatic) within the core.  相似文献   

17.
We present the initial findings of a theoretical study of hydrogen bond formation between two formamide molecules in water and in carbon tetrachloride. These systems were chosen as the simplest models for secondary structure formation in the polar environment near the protein surface and the apolar environment of the protein interior. We have employed thermodynamic simulation methods to obtain absolute binding free energies and free energy profiles for the formation of peptide hydrogen bonds in the two solvents. We find that the amide hydrogen bond is stable by 8.4 kcal/mol in CCl4, and by 0.3 kcal/mol in water. Our results indicate also that the hydrogen-bonded dimer is 2.2 kcal/mol more stable in water than it is in CCl4. We compare our results with those from experiment, and discuss their use in interpreting mechanisms of protein folding.  相似文献   

18.
Density functional theory analysis was performed to elucidate the impact of one-electron reduction upon the initial step of adenosylcobalamin-dependent enzymatic catalysis. The transition state (TS) corresponding to the Co–C bond cleavage and subsequent hydrogen abstraction from the substrate was located. The intrinsic reaction coordinate calculations predicted that the reaction consisting of Co–C5′ bond cleavage in [CoIII(corrin)]–Rib (where Rib is ribosyl) and hydrogen-atom abstraction from the CH3–CH2–CHO substrate occurs in a concerted fashion. The computed activation energy barrier of the reaction (15.0 kcal/mol) was lowered by approximately 54.5% in comparison with the reaction involving the positively charged cofactor model (Im–[CoIII(corrin)]–Rib+, where Im is imidazole; energy barrier = 33.0 kcal/mol). The Im base was detached during the TS search in the reaction involving the one-electron-reduced analogue. Thus, to compare the energetics of the two reactions, the axial Im ligand detachment energy for the Im–[CoIII(corrin)]–Rib model was computed [7.6 kcal/mol (gas phase); 4.6 kcal/mol (water)]. Consequently, the effective activation energy barrier for the reaction mediated by the Im-off [CoIII(corrin)]–Rib was estimated to be 22.6 kcal/mol, which implied an overall 31.5% reduction in the energetic demands of the reaction. Considering that the lengthened Co–Naxial bond has been observed in X-ray crystal structure studies of B12-dependent mutases, the catalytic impact induced by one-electron reduction of the cofactor is expected to be higher in the presence of the enzymatic environment.  相似文献   

19.
Docking results have enabled us to propose how resveratrol could act as a selective PGHS-1 peroxidase site inhibitor. The docking model has predicted a slightly less favorable DeltaG(bind) (-17.9 kcal/mol) of the resveratrol to the PGHS-2 peroxidase site in comparison with its corresponding binding to the PGHS-1 (-20.4 kcal/mol). The formation of hydrogen bonds among the hydroxyl groups of the resveratrol phenyl rings, the backbone of Fe-heme and the carbonyl group of Leu294 inside the PGHS-1 peroxidase site, associated with the absence of His214 in the backbone of PGHS-1, are essential features that are required to maintain the aromatic rings of the natural product parallel to the Fe-heme group and transverse to the peroxidase access channel promoting a large steric hindrance at this site and its consequent selective inhibition.  相似文献   

20.
Every AE  Russu IM 《Biopolymers》2007,87(2-3):165-173
Aromatic stacking and hydrogen bonding between nucleobases are two of the key interactions responsible for stabilization of DNA double-helical structures. The present work aims at defining the specific contributions of these interactions to the stability of individual base pairs in DNA. The two DNA double helices investigated are formed, respectively, by the palindromic base sequences 5'-dCCAACGTTGG-3' and 5'-dCGCAGATCTGCG-3'. The strength of the N==H...N inter-base hydrogen bond in each base pair is characterized from the measurement of the protium-deuterium fractionation factor of the corresponding imino proton using NMR spectroscopy. The structural stability of each base pair is evaluated from the exchange rate of the imino proton, measured by NMR. The results reveal that the fractionation factors of the imino protons in the two DNA double helices investigated fall within a narrow range of values, between 0.92 and 1.0. In contrast, the free energies of structural stabilization for individual base pairs span 3.5 kcal/mol, from 5.2 to 8.7 kcal/mol (at 15 degrees C). These findings indicate that, in the two DNA double helices investigated, the strength of N==H...N inter-base hydrogen bonds does not change significantly depending on the nature or the sequence context of the base pair. Hence, the variations in structural stability detected by proton exchange do not involve changes in the strength of inter-base hydrogen bonds. Instead, the results suggest that the energetic identity of a base pair is determined by the number of inter-base hydrogen bonds, and by the stacking interactions with neighboring base pairs.  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