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 31 毫秒
1.
It was established for the first time by DFT and MP2 quantum-mechanical (QM) methods either in vacuum, so in the continuum with a low dielectric constant (ε = 4), typical for hydrophobic interfaces of specific protein-nucleic acid interactions, that the repertoire for the tautomerisation of the biologically important adenine·cytosine* (A·C*) mismatched DNA base pair, formed by the amino tautomer of the A and the imino mutagenic tautomer of the C, into the A*·C base mispair (?G = 2.72 kcal?mol?1 obtained at the MP2 level of QM theory in the continuum with ε = 4), formed by the imino mutagenic tautomer of the A and the amino tautomer of the C, proceeds via the asynchronous concerted double proton transfer along two antiparallel H-bonds through the transition state (TSA·C*?A*·C). The limiting stage of the A·C*→A*·C tautomerisation is the final proton transfer along the intermolecular N6H···N4 H-bond. It was found that the A·C*/A*·C DNA base mispairs with Watson–Crick geometry are associated by the N6H?N4/N4H?N6, N3H?N1/N1H?N3 and C2H?O2 H-bonds, respectively, while the TSA·C*?A*·C is joined by the N6–H–N4 covalent bridge and the N1H?N3 and C2H?O2 H-bonds. It was revealed that the A·C*?A*·C tautomerisation is assisted by the true C2H?O2 H-bond, that in contrast to the two others conventional H-bonds exists along the entire intrinsic reaction coordinate (IRC) range herewith becoming stronger at the transition from vacuum to the continuum with ε = 4. To better understand the behavior of the intermolecular H-bonds and base mispairs along the IRC of the A·C*?A*·C tautomerisation, the profiles of their electron-topological, energetical, geometrical, polar and charge characteristics are reported in this study. It was established based on the profiles of the H-bond energies that all three H-bonds are cooperative, mutually strengthening each other. The nine key points, providing a detailed physicochemical picture of the A·C*?A*·C tautomerisation, were revealed and thoroughly examined along the IRC. It was shown that the A*·C base mispair with the population ~1 % obtained at the MP2 level of QM theory in the continuum with ε = 4 is thermodynamically and dynamically stable structure. Its lifetime was calculated to be 5.76·10?10 s at the MP2 level of QM theory in the continuum with ε = 4. This lifetime, from the one side, enables all six low-frequency intermolecular vibrations to develop, but, from the other side, it is by order less than the time (several ns) required for the replication machinery to forcibly dissociate a base pair into the monomers during DNA replication. This means that the A*·C base mispair “slips away from the hands” of the replication machinery into the A·C* mismatched base pair. Consequently, the authors came to the conclusion that exactly the A·C* base mispair is an active player of the point mutational events and is effectively dissociated by the replication machinery into the A and C* monomers in contrast to the A*·C base mispair, playing the mediated role of a provider of the A·C* base mispair in DNA that is synthesised.  相似文献   

2.
3.
This work is devoted to the careful QM/QTAIM analysis of the evolution of the basic physico-chemical parameters along the intrinsic reaction coordinate (IRC) of the biologically important 2AP·T(WC)?2AP·T*(w) and 2AP·C*(WC)?2AP·C(w) Watson–Crick(WC)?wobble(w) tautomeric transformations obtained at each point of the IRC using original authors’ methodology. Established profiles reflect the high similarity between the courses of these processes. Basing on the scrupulous analysis of the profiles of their geometric and electron-topological parameters, it was established that the dipole-active WC?w tautomerizations of the Watson–Crick-like 2AP·T(WC)/2AP·C*(WC) mispairs, stabilized by the two classical N3H?N1, N2H?O2 and one weak C6H?O4/N4 H-bonds, into the wobble 2AP·T*(w)/2AP·C(w) base pairs, respectively, joined by the two classical N2H?N3 and O4/N4H?N1 H-bonds, proceed via the concerted stepwise mechanism through the sequential intrapair proton transfer and subsequent large-scale shifting of the bases relative each other, through the planar, highly stable, zwitterionic transition states stabilized by the participation of the four H-bonds – N1+H?O4/N4, N1+H?N3, N2+H?N3, and N2+H?O2. Moreover, it was found out that the 2AP·T(WC)?2AP·T*(w)/2AP·C*(WC)?2AP·C(w) tautomerization reactions occur non-dissociatively and are accompanied by the consequent replacement of the 10 unique patterns of the specific intermolecular interactions along the IRC. Obtained data are of paramount importance in view of their possible application for the control and management of the proton transfer, e.g. by external electric or laser fields.  相似文献   

4.
This study provides the first accurate investigation of the tautomerization of the biologically important guanine*·thymine (G*·T) DNA base mispair with Watson–Crick geometry, involving the enol mutagenic tautomer of the G and the keto tautomer of the T, into the G·T* mispair (?G?=?.99?kcal?mol?1, population?=?15.8% obtained at the MP2 level of quantum-mechanical theory in the continuum with ε?=?4), formed by the keto tautomer of the G and the enol mutagenic tautomer of the T base, using DFT and MP2 methods in vacuum and in the weakly polar medium (ε?=?4), characteristic for the hydrophobic interfaces of specific protein–nucleic acid interactions. We were first able to show that the G*·T?G·T* tautomerization occurs through the asynchronous concerted double proton transfer along two antiparallel O6H···O4 and N1···HN3 H-bonds and is assisted by the third N2H···O2 H-bond, that exists along the entire reaction pathway. The obtained results indicate that the G·T* base mispair is stable from the thermodynamic point of view complex, while it is dynamically unstable structure in vacuum and dynamically stable structure in the continuum with ε?=?4 with lifetime of 6.4·10?12?s, that, on the one side, makes it possible to develop all six low-frequency intermolecular vibrations, but, on the other side, it is by three orders less than the time (several ns) required for the replication machinery to forcibly dissociate a base pair into the monomers during DNA replication. One of the more significant findings to emerge from this study is that the short-lived G·T* base mispair, which electronic interaction energy between the bases (?23.76?kcal?mol?1) exceeds the analogical value for the G·C Watson–Crick nucleobase pair (?20.38?kcal?mol?1), “escapes from the hands” of the DNA replication machinery by fast transforming into the G*·T mismatch playing an indirect role of its supplier during the DNA replication. So, exactly the G*·T mismatch was established to play the crucial role in the spontaneous point mutagenesis.  相似文献   

5.
Abstract

2′-Deoxycytidine hemidihydrogenphosphate has been crystallized in the hexagonal space group P62 with α=25.839(3), c = 12.529(1) Å. The structure has been solved using the Patterson search method. The asymmetric unit contains two protonated, base-paired 2′-deoxycytidine dimers and two H2PO4 ? anions. The C+·C base pairs are composed of a protonated and a neutral species each and are triple H-bonded, the central N(3)…N(3) bonds being 2.850(7) and 2.884(5) Å. The conformations of the four nucleosides fall in the same category (sugar puckers 2·-endo, glycosidic links anti) but in one of them the glycosidic torsion angle is quite low with consequences in other geometrical parameters. The H2PO4 ? anions are located on twofold axes and form two types of tight columns with P…P separations about 4.18 Å The neighboring units along a column are linked via two very short O…H…O hydrogen bonds (O…O about 2.49 Å) leading to effective equalization of the P-O bonds. The base pairs of the two dC+·dC cations are coplanar and form layers perpendicular to the phosphate columns repeating every c/3. Within the layers, the dimers form a network through 0(5′)…O(2) hydrogen bonds but their primary intermolecular interactions have the form of H-bond anchors [N(4)-H…O-P and 0(3′)-H…O-P] to the phosphate groups.  相似文献   

6.
The ground-state tautomerization of the G·C Watson–Crick base pair by the double proton transfer (DPT) was comprehensively studied in vacuo and in the continuum with a low dielectric constant (??=?4), corresponding to a hydrophobic interface of protein–nucleic acid interactions, using DFT and MP2 levels of quantum-mechanical (QM) theory and quantum theory “Atoms in molecules” (QTAIM). Based on the sweeps of the electron-topological, geometric, polar, and energetic parameters, which describe the course of the G·C???G?·C? tautomerization (mutagenic tautomers of the G and C bases are marked with an asterisk) through the DPT along the intrinsic reaction coordinate (IRC), it was proved that it is, strictly speaking, a concerted asynchronous process both at the DFT and MP2 levels of theory, in which protons move with a small time gap in vacuum, while this time delay noticeably increases in the continuum with ??=?4. It was demonstrated using the conductor-like polarizable continuum model (CPCM) that the continuum with ??=?4 does not qualitatively affect the course of the tautomerization reaction. The DPT in the G·C Watson–Crick base pair occurs without any intermediates both in vacuum and in the continuum with ??=?4 at the DFT/MP2 levels of theory. The nine key points along the IRC of the G·C base pair tautomerization, which could be considered as electron-topological “fingerprints” of a concerted asynchronous process of the tautomerization via the DPT, have been identified and fully characterized. These key points have been used to define the reactant, transition state, and product regions of the DPT reaction in the G·C base pair. Analysis of the energetic characteristics of the H-bonds allows us to arrive at a definite conclusion that the middle N1H?N3/N3H?N1 and the lower N2H?O2/N2H?O2 parallel H-bonds in the G·C/G?·C? base pairs, respectively, are anticooperative, that is, the strengthening of the middle H-bond is accompanied by the weakening of the lower H-bond. At that point, the upper N4H?O6 and O6H?N4 H-bonds in the G·C and G?·C? base pairs, respectively, remain constant at the changes of the middle and the lower H-bonds at the beginning and at the ending of the G·C???G?·C? tautomerization. Aiming to answer the question posed in the title of the article, we established that the G?·C? Löwdin’s base pair satisfies all the requirements necessary to cause point mutations in DNA except its lifetime, which is much less than the period of time required for the replication machinery to forcibly dissociate a base pair into the monomers (several ns) during DNA replication. So, from the physicochemical point of view, the G?·C? Löwdin’s base pair cannot be considered as a source of point mutations arising during DNA replication.  相似文献   

7.
The biologically important tautomerization of the Hyp·Cyt, Hyp*·Thy and Hyp·Hyp base pairs to the Hyp*·Cyt*, Hyp·Thy* and Hyp*·Hyp* base pairs, respectively, by the double proton transfer (DPT) was comprehensively studied in vacuo and in the continuum with a low dielectric constant (ε?=?4) corresponding to hydrophobic interfaces of protein–nucleic acid interactions by combining theoretical investigations at the B3LYP/6-311++G(d,p) level of QM theory with QTAIM topological analysis. Based on the sweeps of the energetic, electron-topological, geometric and polar parameters, which describe the course of the tautomerization along the intrinsic reaction coordinate (IRC), it was proved that the tautomerization through the DPT is concerted and asynchronous process for the Hyp·Cyt and Hyp*·Thy base pairs, while concerted and synchronous for the Hyp·Hyp homodimer. The continuum with ε?=?4 does not affect qualitatively the course of the tautomerization reaction for all studied complexes. The nine key points along the IRC of the Hyp·Cyt?Hyp*·Cyt* and Hyp*·Thy?Hyp·Thy* tautomerizations and the six key points of the Hyp·Hyp?Hyp*·Hyp* tautomerization have been identified and fully characterized. These key points could be considered as electron-topological “fingerprints” of concerted asynchronous (for Hyp·Cyt and Hyp*·Thy) or synchronous (for Hyp·Hyp) tautomerization process via the DPT. It was found, that in the Hyp*·Cyt*, Hyp·Thy*, Hyp·Hyp and Hyp*·Hyp* base pairs all H-bonds are significantly cooperative and mutually reinforce each other, while the C2H…O2 H-bond in the Hyp·Cyt base pair and the O6H…O4 H-bond in the Hyp*·Thy base pair behave anti-cooperatively, i.e., they become weakened, while two others become strengthened.  相似文献   

8.
The molecular and crystal structures of one derivative and three homopeptides (from the di-to the tetrapeptide level) of the chiral, Cα, α-disubstituted glycine Cα-methyl, Cα-benzylglycine [(αMe)Phe], have been determined by x-ray diffraction. The derivative is mClAc-D -(αMe)Phe-OH, and the peptides are pBrBz-[D -(αMe)Phe]2-NHMe, pBrBz-[D -(αMe)Phe]3-OH hemihydrate, and pBrBz-[D -(αMe)Phe]4-OtBu sesquihydrate. All (αMe)Phe residues prefer ?,ψ torsion angles in the helical region of the conformational map. The dipeptide methylamide and the tripeptide carboxylic acid adopt a β-turn conformation with a 1 ← 4 C?O…?H? N intramolecular H bond. The structure of the tripeptide carboxylic acid is further stabilized by a 1 ← 4 C?O…?H? O intramolecular H bond, forming an “oxy-analogue” of a β-turn. The tetrapeptide ester is folded in a regular (incipient) 310-helix. In general, the relationship between (αMe)Phe chirality and helix screw sense is opposite to that exhibited by protein amino acids. A comparison is made with the conclusions extracted from published work on homopeptides from other Cα-methylated α-amino acids. © 1993 John Wiley & Sons, Inc.  相似文献   

9.
Combining quantum-mechanical (QM) calculations with quantum theory of atoms in molecules (QTAIM) and using the methodology of sweeps of the energetic, electron-topological, geometric and polar parameters, which describe the course of the tautomerization along the intrinsic reaction coordinate (IRC), we showed for the first time that the biologically important A?A* base pair (Cs symmetry) formed by the amino and imino tautomers of adenine (A) tautomerizes via asynchronous concerted double proton transfer (DPT) through a transition state (TS), which is the A+?A? zwitterion with the separated charge, with Cs symmetry. The nine key points, which can be considered as electron-topological “fingerprints” of the asynchronous concerted A?A*?A*?A tautomerization process via the DPT, were detected and completely investigated along the IRC of the A?A*?A*?A tautomerization. Based on the sweeps of the H-bond energies, it was found that intermolecular antiparallel N6Н?N6 (7.01 kcal mol?1) and N1H?N1 (6.88 kcal mol?1) H-bonds are significantly cooperative and mutually reinforce each other. It was shown for the first time that the A?A*?A*?A tautomerization is assisted by the third C2H?HC2 dihydrogen bond (DHB), which, in contrast to the two others N6H?N6 and N1H?N1 H-bonds, exists within the IRC range from ?2.92 to 2.92 Å. The DHB cooperatively strengthens, reaching its maximum energy 0.42 kcal mol?1 at IRC?=??0.52 Å and minimum energy 0.25 kcal mol?1 at IRC?=??2.92 Å, and is accompanied by strengthening of the two other aforementioned classical H-bonds. We established that the C2H?HC2 DHB completely satisfies the electron-topological criteria for H-bonding, in particular Bader’s and all eight “two-molecule” Koch and Popelier’s criteria. The positive value of the Grunenberg’s compliance constant (5.203 Å/mdyn) at the TSA?A*?A*?A proves that the C2H?HC2 DHB is a stabilizing interaction. NBO analysis predicts transfer of charge from σ(C2–H) bonding orbital to σ*(H–C2) anti-bonding orbital; at this point, the stabilization energy E(2) is equal to 0.19 kcal mol?1 at the TSA?A*?A*?A.  相似文献   

10.
In this paper, we consider the mutagenic properties of the 2-aminopurine (2AP), which has intrigued molecular biologists, biophysicists and physical chemists for a long time and been widely studied by both experimentalists and theorists. We have shown for the first time using QM calculations, that 2AP very effectively produces incorporation errors binding with cytosine (C) into the wobble (w) C·2AP(w) mispair, which is supported by the N4H?N1 and N2H?N3 H-bonds and is tautomerized into the Watson–Crick (WC)-like base mispair C*·2AP(WC) (asterisk denotes the mutagenic tautomer of the base), that quite easily in the process of the thermal fluctuations acquires enzymatically competent conformation. 2AP less effectively produces transversions forming the wobble mispair with A base – A·2AP(w), stabilized by the participation of the N6H?N1 and N2H?N1 H-bonds, followed by further tautomerization A·2AP(w) → A*·2AP(WC) and subsequent conformational transition A*·2AP(WC) → A*·2APsyn thus acquiring enzymatically competent structure. In this case, incorporation errors occur only in those case, when 2AP belongs to the incoming nucleotide. Thus, answering the question posed in the title of the article, we affirm for certain that 2AP induces incorporation errors at the DNA replication. Obtained results are consistent well with numerous experimental data.  相似文献   

11.
Extracellular invertase (EC 3.2.1.26) of Saccharomyces cerevisiae was stabilized against thermal denaturation by intermolecular and intramolecular crosslinking of the surface nucleophilic functional groups with diisocyanate homobifunctional reagents (O?C?N(CH2)nN?C?O) of various lengths (n = 4, 6, 8). Crosslinking with 1,4‐diisocyanatobutane (n = 4) proved most effective in enhancing thermostability. Stability was improved dramatically by crosslinking 0.5 mg/mL of protein with 30 μmol/mL of the reagent. Molecular engineering by crosslinking reduced the first‐order thermal denaturation constant at 60°C from 1.567 min?1 (for the native enzyme) to 0.437 min?1 (for the stabilized enzyme). Similarly, the best crosslinking treatment increased the activation energy for denaturation from 391 kJ mol?1 (for the native protein) to 466 kJ mol?1 (for the stabilized enzyme). Crosslinking was confirmed by sodium dodecyl sulfate polyacrylamide gel electrophoresis. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

12.
Binary diastereomeric (-) (1R,2S)-ephedrine salts of various mandelic acids obtained from 95% ethanol show considerable differences in solubility. Structures and some properties of the less-soluble (L) and more-soluble (M) solid phases of (-)-ephedrine with unsubstituted mandelic acid, 2-, 3-, and 4-monosubstituted halo (F, Cl, Br) mandelic acids, and 3- and 4-methylmandelic acids have been determined. Salts were found to be binary, without solvent of crystallization, and composed of double-layered arrays of alternating anions and cations linked by H-bonds normal to the layers. H-bonding links charged donors and acceptors usually along a crystallographic 2-fold screw axis. A striking discrimination is evident in that the (2R)-mandelate salts typically display a compact four-atom chain as the H-bonding repeating unit [+N—H…O(—C?—O)…H-N′, C21(4)] while the (2S)-mandelate salts adopt a more dimensionally variable six-atom chain repeating unit [+N—H…O—C?—O…H—N′, C22(6)]. Two distinct packing schemes display the shorter H-bonding chain of the (2R)-mandelates which always occurs with ephedrinium ions in the fully extended conformation. Slightly greater packing efficiency and H-bonding energies of the (2R)-mandelate salts correlates with increased fusion points, lower solubilities (95% ethanol), and higher heats of fusion relative to the phase adopted by their diastereoisomers. In contrast, (2S)-mandelate salts exhibit considerably more structural variability involving all three major ephedrinium conformations, and at least four distinct packing motifs. Mandelates with larger 3′-substituents (Cl, Br, methyl) show similar property discriminations, but these occur with an opposing trend, that is, between phases in which the less-soluble salts contain (2S)-mandelates. Salts with 2-bromomandelate do not show property disparities and their structures are dissimilar to the other phases. © 1995 Wiley-Liss, Inc.  相似文献   

13.
The kinetics of the light-driven Cl? uptake pump of Synechococcus R-2 (PCC 7942) were investigated. The kinetics of Cl? uptake were measured in BG-11 medium (pHo, 7·5; [K+]o, 0·35 mol m?3; [Na+]o, 18 mol m?3; [Cl?]o, 0·508 mol m?3) or modified media based on the above. Net36Cl? fluxes (?Cl?o,i) followed Michaelis-Menten kinetics and were stimulated by Na+ [18 mol m?3 Na+ BG-11 ?Cl?max= 3·29±0·60 (49) nmol m?2 s?1 versus Na+-free BG-11 ?Cl?max= 1·02±0·13 (54) nmol m?2 s?1] but the Km was not significantly different in the presence or absence of Na+ at pHo 10; the Km was lower, but not affected by the presence or absence of Na+ [Km = 22·3±3·54 (20) mmol m?3]. Na+ is a non-competitive activator of net ?Cl?o,i. High [K+]o (18 mol m?3) did not stimulate net ?Cl?o,i or change the Km in Na+-free medium. High [K+]o (18 mol m?3) added to Na+ BG-11 medium decreased net ?Cl?o,i [18 mol m?3K+ BG-11; ?Cl?max= 2·50±0·32 (20) nmol m?2 s?1 versus BG-11 medium; ?Cl?max= 3·35±0·56 (20) nmol m?2 s?1] but did not affect the Km 55·8±8·100 (40) mmol m?3]. Na+-stimulation of net ?Cl?o,i followed Michaelis-Menten kinetics up to 2–5 mol m?3 [Na+]o but higher concentrations were inhibitory. The Km for Na+-stimulation of net ?Cl?o,i [K1/2(Na+)] was different at 47 mmol m?3 [Cl?]o (K1/2[Na+] = 123±27 (37) mmol m?3]. Li+ was only about one-third as effective as Na+ in stimulating Cl? uptake but the activation constant was similar [K1/2(Li+) = 88±46 (16) mmol m?3]. Br? was a competitive inhibitor of Cl? uptake. The inhibition constant (Ki) was not significantly different in the presence and absence of Na+. The overall Ki was 297±23 (45) mmol m?3. The discrimination ratio of Cl? over Br? (δCl?/δBr?) was 6·38±0·92 (df = 147). Synechococcus has a single Na+-stimulated Cl? pump because the Km of the Cl? transporter and its discrimination between Cl? and Br? are not significantly different in the presence and absence of Na+. The Cl? pump is probably driven by ATP.  相似文献   

14.
Increase of the atmospheric concentration of halogenated organic compounds is partially responsible for a change of the global climate. In this work we have investigated the interaction between halogenated ether and water, which is one of the most important constituent of the atmosphere. The structures of the complexes formed by the two most stable conformers of enflurane (a volatile anaesthetic) with one and two water molecules were calculated by means of the counterpoise CP-corrected gradient optimization at the MP2/6–311++G(d,p) level. In these complexes the CH…Ow hydrogen bonds are formed, with the H…Ow distances varying between 2.23 and 2.32 Å. A small contraction of the CH bonds and the blue shifts of the ν(CH) stretching vibrations are predicted. There is also a weak interaction between one of the F atoms and the H atom of water, with the Hw…F distances between 2.41 and 2.87 Å. The CCSD(T)/CBS calculated stabilization energies in these complexes are between ?5.89 and ?4.66 kcal?mol?1, while the enthalpies of formation are between ?4.35 and ?3.22 kcal?mol?1. The Cl halogen bonding between enflurane and water has been found in two complexes. The intermolecular (Cl···O) distance is smaller than the sum of the corresponding van der Waals radii. The CCSD(T)/CBS stabilization energies for these complexes are about ?2 kcal?mol?1.
Figure
Complex between enflurane and water molecules  相似文献   

15.
Crystals of 5‐fluorouridine (5FUrd) have unit cell dimensions a = 7.716(1), b = 5.861(2), c = 13.041(1)Å, α = γ = 90°, β = 96.70° (1), space group P21, Z = 2, ρobs = 1.56 gm/c.c and ρcalc = 1574 gm/c.c The crystal structure was determined with diffractometric data and refined to a final reliability index of 0.042 for the observed 2205 reflections (I ≥ 3σ). The nucleoside has the anti conformation [χ = 53.1(4)°] with the furanose ring in the favorite C2′–endo conformation. The conformation across the sugar exocyclic bond is g+, with values of 49.1(4) and ? 69.3(4)° for Φθc and Φ respectively. The pseudorotational amplitude τm is 34.5 (2) with a phase angle of 171.6(4)°. The crystal structure is stabilized by a network of N–H…O and O–H…O involving the N3 of the uracil base and the sugar O3′ and O2′ as donors and the O2 and O4 of the uracil base and O3′ oxygen as acceptors respectively. Fluorine is neither involved in the hydrogen bonding nor in the stacking interactions. Our studies of several 5‐fluorinated nucleosides show the following preferred conformational features: 1) the most favored anti conformation for the nucleoside [χ varies from ? 20 to + 60°] 2) an inverse correlation between the glycosyl bond distance and the χ angle 3) a wide variation of conformations of the sugar ranging froni C2′–endo through C3′–endo to C4′–exo 4) the preferred g+ across the exocyclic C4′–C5′ bond and 5) no role for the fluorine atom in the hydrogen bonding or base stacking interactions.  相似文献   

16.
The influences of water solvent on the structures and stabilities of the complex ion conformers formed by the coordination of alanine dipeptide (AD) and Na+ have been investigated using supramolecular and polarizable continuum solvation models at the level of B3LYP/6-311++G**, respectively; 12 monohydrated and 12 dihydrated structures of Na+–AD complex ion were obtained after full geometrical optimization. The results showed that H2O molecules easily bind with Na+ of Na+–AD complex ion, forming an ion-lone pair interaction with the Na–O bond length of 2.1–2.3 Å. Besides, H2O molecules also can form hydrogen bonds OW–HW···O(1), OW–HW···O(2), N(1)–H(1)···OW or N(2)–H(2)···OW with O or N groups of the Na+–AD backbone. The most stable gaseous bidentate conformer C7AB of Na+–AD is still the most stable one in the solvent of water. However, the structure of the most unstable gaseous conformer α′B of Na+–AD collapses under the attack of H2O molecules and changes into C7AB conformation. Computations with IEFPCM solvation model of self-consistent reaction field theory give that aqueous C5A is more stable than C7eqB and that the stabilization energies of water solvent on monodentate conformers of Na+–AD complex ion (about 272–294 kJ/mol) are more than those on bidentate ones (about 243 kJ/mol).  相似文献   

17.
The nature of hydrogen bonds formed between carboxylic acid residues and histidine residues in proteins is studied by ir spectroscopy. Poly(glutamic acid) [(Glu)n] is investigated with various monomer N bases. The position of the proton transfer equilibrium OH…?N ? O?…?H+N is determined considering the bands of the carboxylic group. It is shown that largely symmetrical double minimum energy surfaces are present in the OH…?N ? O?…?H+N bonds when the pKa of the protonated N base is two values larger than that of the carboxylic groups of (Glu)n. Hence OH…?N ? O?…?H+N bonds between glutamic and aspartic acid residues and histidine residues in proteins may be easily polarizable proton transfer hydrogen bonds. The polarizability of these bonds is one to two orders of magnitude larger than usual electron polarizabilities; therefore, these bonds strongly interact with their environment. It is demonstrated that water molecules shift these proton transfer equilibria in favor of the polar proton boundary structure. The access of water molecules to such bonds in proteins and therefore the position of this proton transfer equilibrium is dependent on conformation. The amide bands show that (Glu)n is α-helical with all systems. The only exception is the (Glu)n-n-propylamine system. When this system is hydrated (Glu)n is α-helical, too. When it is dried, however, (Glu)n forms antiparallel β-structure. This conformational transition, dependent on degree of hydration, is reversible. An excess of n-propylamine has the same effect on conformation as hydration.  相似文献   

18.
Studies of uptake of ionic sources of N by two hydroponically grown rice (Oryza sativa L.) cultivars (paddy‐field‐adapted Koshihikari and dryland‐adapted Kanto 168) showed that the magnitude of the nitrogen isotope fractionation (?) for uptake of NH4+ depended on the concentrations of NH4+ and cultivar (averaging –6·1‰ for Koshihikari and –12·0‰ for Kanto 168 at concentrations from 40 to 200 mmol m?3 and, respectively, –13·4 and –28·9‰ for the two cultivars at concentrations from 0·5 to 4 mol m?3). In contrast, the ? for uptake of NO3? in similar experiments was almost insensitive to the N concentration, falling within a much narrower range (+3·2‰ to –0·9‰ for Koshihikari and –0·9‰ to –5·1‰ for Kanto 168 over NO3? concentrations from 0·04 to 2 mol m?3). From longer term experiments in which Norin 8 and its nitrate‐reductase deficient mutant M819 were grown with 2 or 8 mol m?3 NO3? for 30 d, it was concluded that the small concentration‐independent isotopic fractionation during absorption of this ion was not related to nitrate reductase activity.  相似文献   

19.
U Burget  G Zundel 《Biopolymers》1987,26(1):95-108
(L -His)n- dihydrogen phosphate systems are studied by ir spectroscopy in the presence of various cations and as a function of the degree of hydration. Ir continua indicate that (I) OH … N ? O?…H+N (IIR) hydrogen bonds are formed and that these bonds show high proton polarizability, which increases from the Li+ to the K+ system. In the K+?system, His-Pi-Pi chains are formed, showing particularly high proton polarizability due to collective proton motion within both hydrogen bonds. The OH N ? O?…H?N equilibria are determined from ir bands. With the Li+ system, 55% of the protons are present at the histidine residues; this percentage is smaller with the Na+ system (41%), and amounts to only 32% with the K+ system. With the increasing degree of hydration the removal of the degeneracy of νas?PO2?3 vanishes, indicating loosening of the cations from the phosphates. Nevertheless, the hydrogen bond acceptor O atom becomes more negative; a shift of the equilibrium to the right is observed in the OH… N ? O?…H+N bond. This is explained by the strong interaction of the dipole of the hydrogen bonds with the water molecules. All these results show that protons can be shifted easily in these hydrogen bonds due to their high proton polarizability. The transfer equilibria can be controlled easily by local electrical fields. In addition, these results may be of significance when phosphates interact with proteins.  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