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 364 毫秒
1.
Cytochromes c' are pentacoordinate heme proteins with sterically hindered distal sites that bind NO and CO but do not form stable complexes with O(2). Removal of distal pocket steric hindrance via a Leu→Ala mutation yields favorable O(2) binding (K(d) ~49 nM) without apparent H-bond stabilization of the Fe-O(2) moiety, as well as an extremely high distal heme-NO affinity (K(d) ~70 fM). The native Leu residue inhibits distal coordination of diatomic ligands by decreasing k(on) as well as increasing k(off). The connection between distal steric constraints, k(off) values, and distal to proximal heme-NO conversion is discussed.  相似文献   

2.
Tryptophan 2,3-dioxygenase (TDO) is a heme-dependent enzyme that catalyzes the oxidative degradation of L-tryptophan (L-Trp) to N-formylkynurenine (NFK). A highly conserved histidine residue in the distal heme pocket has attracted great attention in the mechanistic studies of TDO. However, a consensus has not been reached regarding whether and how this distal histidine plays a catalytic role after substrate binding. In this study, three mutant proteins, H72S, H72N, and Q73F were generated to investigate the function of the distal histidine residue in Cupriavidus metallidurans TDO (cmTDO). Spectroscopic characterizations, enzymatic kinetic analysis, and chemical rescue assays were employed to study the biochemical properties of the wild-type enzyme and the mutant proteins. Rapid kinetic methods were utilized to explore the molecular basis for the observed stimulation of catalytic activity by 2-methylimidazole in the His72 variants. The results indicate that the distal histidine plays multiple roles in cmTDO. First, His72 contributes to but is not essential for substrate binding. In addition, it shields the heme center from nonproductive binding of exogenous small ligand molecules (i.e., imidazole and its analogs) via steric hindrance. Most importantly, His72 participates in the subsequent chemical catalytic steps after substrate binding possibly by providing H-bonding interactions to the heme-bound oxygen.  相似文献   

3.
A study of the H-bonding between poly(ethylene glycol) (PEG) and polyvinylpyrrolidone (PVP) in the presence of supercritical carbon dioxide at various temperatures, pressures, different M(w) of PEG and PVP and different PEG/PVP ratios is presented. In situ attenuated total reflection Fourier transform infrared (ATR-FTIR) spectroscopy was used to investigate H-bonding by examining changes in the relative intensities and positions of peak maxima of the 2nd derivative ν(C=O) bands associated with 'free' and H-bonded C=O groups. In general, relative intensities of bands associated with H-bonded CO groups decreased upon CO(2) sorption and was accompanied by an increase in intensity of bands associated with 'free' C=O groups. At the same time, these bands were shifted to higher wavenumbers. These shifts were attributed to the shielding effect of CO(2) molecules on H-bonding interactions between PEG and PVP. The magnitude of the effects of CO(2) shielding generally increased with decreasing polymer M(w) and increasing CO(2) content. However, upon CO(2) venting the extent of the H-bonding between PEG and PVP reappeared. The extent of H-bonding recovery was greatest for blends with low M(w) PEG (M(w): 4×10(2)) and PVP (M(w): 9×10(3)) and PEG content ≥0.54 wt% under mild conditions of pressure (80 bar) and temperature (35°C). For the same low M(w) blends, increasing pressure to 150 bar, or temperature to 50°C, showed poor H-bond recovery upon CO(2) venting. Overall, it was shown that supercritical CO(2)-induced shielding of H-bonding interactions in polymer blends is reversible upon CO(2) venting, and the magnitude of both was influenced by processing conditions and blend composition.  相似文献   

4.
We report density functional theory calculations at the B3LYP/D95(d,p) level on the hydrogen-bonding complexes of tetraazaanthracenedione, 1, with N-(pyridin-2-yl)urea, 2H, or N-(6-aminopyridin-2-yl)urea, 2N. The interaction energy of the 1-2H complex exceeds that of 1-2N, despite the fact that 1-2N contains a strong N-H...O interaction in place of a weak C-H...O interaction in 1-2H. We show that the 1-2N interaction is weaker than the sum of the four normal individual H-bonding interactions because the steric constraints of the complex prevent the H-bonding donors and acceptors from optimally approaching each other to form the two central H-bonds. This steric phenomenon, which we call attractive strain, is likely present to some extent in most H-bonding systems that contain more than two H-bonds between rigid monomers. Attractive strain is unusually important in 1-2N. Attractive strain can be conceived of as an enthalpic cost for the entropic benefits of freezing the dihedral angles of the multiple H-bond donors and acceptors by designing rigid systems.  相似文献   

5.
In spite of many theoretical and experimental attempts for understanding intramolecular hydrogen bonding (H-bonding) in carbohydrates, a direct quantification of individual intramolecular H-bond energies and the cooperativity among the H-bonded networks has not been reported in the literature. The present work attempts, for the first time, a direct estimation of individual intramolecular O-H...O interaction energies in sugar molecules using the recently developed molecular tailoring approach (MTA). The estimated H-bond energies are in the range of 1.2-4.1 kcal mol(-1). It is seen that the OH...O equatorial-equatorial interaction energies lie between 1.8 and 2.5 kcal mol(-1), with axial-equatorial ones being stronger (2.0-3.5 kcal mol(-1)). The strongest bonds are nonvicinal axial-axial H-bonds (3.0-4.1 kcal mol(-1)). This trend in H-bond energies is in agreement with the earlier reports based on the water-water H-bond angle, solvent-accessible surface area (SASA), and (1)H NMR analysis. The contribution to the H-bond energy from the cooperativity is also estimated using MTA. This contribution is seen to be typically between 0.1 and 0.6 kcal mol(-1) when H-bonds are a part of a relatively weak equatorial-equatorial H-bond network and is much higher (0.5-1.1 kcal mol(-1)) when H-bonds participate in an axial-axial H-bond network.  相似文献   

6.
Heme proteins are found in all living organisms and are capable of performing a wide variety of tasks, requiring in many cases the binding of diatomic ligands, namely, O(2), CO, and/or NO. Therefore, subtle regulation of these diatomic ligands' affinity is one of the key issues for determining a heme protein's function. This regulation is achieved through direct H-bond interactions between the bound ligand and the protein, and by subtle tuning of the intrinsic heme group reactivity. In this work, we present an investigation of the proximal regulation of oxygen affinity in Fe(II) histidine coordinated heme proteins by means of computer simulation. Density functional theory calculations on heme model systems are used to analyze three proximal effects: charge donation, rotational position, and distance to the heme porphyrin plane of the proximal histidine. In addition, hybrid quantum-classical (QM-MM) calculations were performed in two representative proteins: myoglobin and leghemoglobin. Our results show that all three effects are capable of tuning the Fe-O(2) bond strength in a cooperative way, consistently with the experimental data on oxygen affinity. The proximal effects described herein could operate in a large variety of O(2)-binding heme proteins-in combination with distal effects-and are essential to understand the factors determining a heme protein's O(2) affinity.  相似文献   

7.
The nature of interactions of phenol with various molecules (Y = HF, HCl, H2O, H2S, NH3, PH3, MeOH, MeSH) is investigated using ab initio calculations. The optimized geometrical parameters and spectra for the global energy minima of the complexes match the available experimental data. The contribution of attractive (electrostatic, inductive, dispersive) and repulsive (exchange) components to the binding energy is analyzed. HF favors sigma O-type H-bonding, while H2O, NH3, and MeOH favor sigma H-type H-bonding, where sigma O-/sigma H-type is the case when a H-bond forms between the phenolic O/H atom and its interacting molecule. On the other hand, HCl, H2S, and PH3 favor pi-type H-bonding, which are slightly favored over sigma O-, sigma H-, sigma H-type bonding, respectively. MeSH favors chi H-type bonding, which has characteristics of both pi and sigma H. The origin of these conformational preferences depending on the type of molecules is elucidated. Finally, phenol-Y complexes are compared with water-Y complexes. In the water-Y complexes where sigma O/sigma H-type involves the H-bond by the water O/H atom, HF and HCl favor sigma O-type, H2O involves both sigma O-/sigma H-type, and H2S, NH3, PH3, MeOH, and MeSH favor sigma H-type bonding. Except for HF, seven other species have larger binding energies with a phenol molecule than a water molecule.  相似文献   

8.
Complexes of the type syn-(R,S)-Me(3)dienPtL (Me(3)dien = N,N',N' '-trimethyldiethylenetriamine; L = guanine or hypoxanthine derivative) have two rotamers, a feature useful for assessing hydrogen-bond interactions between a Me(3)dien NH group and either the O6 or the phosphate group of the coordinated L. The two rotamers are defined as endo and exo for the rotamer with the six-membered ring of the purine on the same side and on the opposite side, respectively, of the coordination plane as the N-Me's. For L = 5'-GMP and 5'-IMP the endo rotamer is the exclusive form (at neutral and basic pH) or is present at 90% and more (low pH where 5'-phosphate group is protonated). A 5'-phosphate group can be positioned to form a direct H-bond with a Me(3)dien NH group only in the endo form; such an H-bond explains this high endo preference. Such a direct phosphate-NH H-bond is not possible for other complexes used in this study because either L has no phosphate group (9-EtG, Guo) or the phosphate is at the 3'-position (3'-GMP and 3'-IMP), too far for H-bonding. Nevertheless, a preference for the endo rotamer was observed for these L also. This result is opposite to that expected both from potential steric repulsion of the L O6 with the N-Me groups and also from the lack of a potential favorable H-bond interaction between L O6 and a Me(3)dien NH. For the 9-EtG adduct, the temperature dependence of the endo/exo equilibrium and the activation parameters for endo/exo interconversion suggest that the preference for the endo rotamer arises from the hydration of the Me(3)dien NH groups; such hydration is favorable in the endo rotamer. At basic pH, N1H deprotonation increases the H-bond capacity of O6, and the exo rotamer increases in stability, becoming the dominant rotamer for the 9-EtG and Guo adducts. For L = 3'-GMP and 3'-IMP, stabilization of the endo form upon phosphate deprotonation at neutral pH was observed. This result is attributed to an H-bonding network involving water, the 3'-phosphate, and the Me(3)dien NH groups.  相似文献   

9.
As an extension study, FTIR and molecular simulation methods were combined in the present paper to analyze the H-bond interactions resulting from multiple donors and acceptors that have led to self-assembly based on segmented polyurethane with carboxyl (PUc) and poly(4-vinylpyridine) (P4VP) in our previous work. Of them, FTIR was used to analyze the H-bonding types and interactions as well as their changes before and after self-assembly; molecular mechanics (MM/COMPASS) was used to study the effect of possible conformations on the H-bonds involved and analyze the most probable H-bond patterns; quantum mechanics (QM/B3LYP) was used to help confirm the experimental FTIR band assignments and calculate the H-bond energy. It was found that two types of H-bonds exist, namely, COOH...P4VP (type I) and (OCO)NH...P4VP (type II), based on OH and NH as the strong donors in the interaction between PUc and P4VP. Strong evidence has been obtained for a type II H-bond, which is the specialty in PUc/P4VP assembly. The type I and type II H-bonding energies are -11.293 and -7.150 kcal/mol, respectively. The forming probability of the type I H-bond accounts for 95.87%, while that of the type II H-bond is 4.13%, showing the primary driving force for the assembly based on PUc and P4VP is still the H-bond between COOH and P4VP, yet the H-bonds based on NH and pyridyl in P4VP cannot be ignored.  相似文献   

10.
Glycoluril derivative--whose bulky Ph-C[triple bond]C- substituents prevent formation of H-bonded tapes--undergoes solvent dependent assembly in the crystal; a tetrameric molecular bowl is formed by R(24) H-bonding interactions from CH(2)Cl(2) whereas DMF results in H-bond dimerization followed by oligomerization via C-H...pi interactions.  相似文献   

11.
Investigation of the intrinsic H-bonding pattern of the guanine complex with a sizable segment (from Asn43 to Glu46) of the primary recognition site (PRS) in RNase T1 at the B3LYP/6-311G(d,p) level of theory enables the electronic density characteristics of the H-bonding patterns of the guanine-PRS complexes to be identified. The perfect H-bonding pattern in the guanine recognition site is achieved through the guanine complex interactions with the large segment of the PRS. Two significant short H-bonds, O epsilon 1...HN1 and O epsilon 2...HN2, have been identified. The similar short H-bond distances found in the anionic GC- base pair and in this study suggest that the short hydrogen-bond distances may be characteristic of the multiple H-bonded anionic nucleobases. The H-bonding energy distribution, the geometric analysis of the H-bonding pattern, and the electron structure characteristics of the H-bonds in the guanine PRS of RNase T1 all suggest that the O epsilon 1...HN1 and O epsilon 2...HN2 side-chain H-bonds dominate the binding at the guanine recognition site of RNase T1. Also, the geometry evidence, the electron structure characteristics, and the properties of the bond critical points of the H-bonds reveal that the side-chain H-bonding and the main-chain H-bonding are mutually intensifying. Thus the positive cooperativity between Asn43 to Tyr45 and Glu46 is proposed.  相似文献   

12.
In heme-based sensor proteins, ligand binding to heme in a sensor domain induces conformational changes that eventually lead to changes in enzymatic activity of an associated catalytic domain. The bacterial oxygen sensor FixL is the best-studied example of these proteins and displays marked differences in dynamic behavior with respect to model globin proteins. We report a mid-IR study of the configuration and ultrafast dynamics of CO in the distal heme pocket site of the sensor PAS domain FixLH, employing a recently developed method that provides a unique combination of high spectral resolution and range and high sensitivity. Anisotropy measurements indicate that CO rotates toward the heme plane upon dissociation, as is the case in globins. Remarkably, CO bound to the heme iron is tilted by ~30° with respect to the heme normal, which contrasts to the situation in myoglobin and in present FixLH-CO X-ray crystal structure models. This implies protein-environment-induced strain on the ligand, which is possibly at the origin of a very rapid docking-site population in a single conformation. Our observations likely explain the unusually low affinity of FixL for CO that is at the origin of the weak ligand discrimination between CO and O(2). Moreover, we observe orders of magnitude faster vibrational relaxation of dissociated CO in FixL than in globins, implying strong interactions of the ligand with the distal heme pocket environment. Finally, in the R220H FixLH mutant protein, where CO is H-bonded to a distal histidine, we demonstrate that the H-bond is maintained during photolysis. Comparison with extensively studied globin proteins unveils a surprisingly rich variety in both structural and dynamic properties of the interaction of a diatomic ligand with the ubiquitous b-type heme-proximal histidine system in different distal pockets.  相似文献   

13.
A method is described to evaluate backbone interactions in proteins via computational unnatural amino acid mutagenesis. Several N-acetyl polyalanyl amides (AcA(n)NH(2)) were optimized in the representative helical (3(10)-, 4(13)-, and a "hybrid" kappa-helix, n = 7, 9, 10, 14) and hairpin (two- and three-stranded antiparallel beta-sheets with type I turns betaalphaalphaepsilon, n = 6, 9, 10) conformations, and extended conformers of N-acetyl polyalanyl methylamides (n = 2, 3) were used to derive multistranded beta-sheet fragments. Subsequently, each residue of every model structure was substituted, one at a time, with l-lactic acid. The resulting mutant structures were again optimized, and group-transfer energies DeltaE(GT) were obtained as heats of the isodesmic reactions: AcA(n)NHR + AcOMe --> AcA(x)LacA(y)NHR + AcNHMe (R = H, CH(3)). These group-transfer energies correlate with the degree of charge polarization of the substituted peptide linkages as measured by the difference Deltae in H and O Mulliken populations in HN-C=O and with the H-bond distances in the "wild-type" structures. A good correlation obtains for the HF/3-21G and B3LYP/6-31G* group-transfer energies. The destabilization effects are interpreted in terms of loss of interstrand and intrastrand H-bonds, decrease in Lewis basicity of the C=O group, and O...O repulsion. On the basis of several comparisons of Ala --> Lac DeltaE(GT)'s with heats of the NH --> CH(2) substitutions, the latter contribution is estimated (B3LYP/6-31G*) to range between 1.5 and 2.4 kcal mol(-1), a figure close to the recent experimental DeltaDeltaG(o) value of 2.6 kcal mol(-1) (McComas, C. C.; Crowley, B. M.; Boger, D. L. J. Am.Chem. Soc. 2003, 125, 9314). The partitioning yields the following maximum values of the electronic association energy of H-bonds in the examined sample of model structures (B3LYP/6-31G* estimates): 3(10)-helix D(e) = -1.7 kcal mol(-1), alpha-helix D(e) = -3.8 kcal mol(-1), beta-sheet D(e) = -6.1 kcal mol(-1). The premise of experimental evaluations of the backbone-backbone H-bonding that Ala --> Lac substitution in proteins is isosteric (e.g., Koh, J. T.; Cornish, V. W.; Schultz, P. G. Biochemistry 1997, 36, 11314) is often but not always corroborated. Examination of the integrity of H-bonding pattern and phi(i), psi(i) distribution identified several mutants with significant distortions of the "wild-type" structure resulting inter alia from the transitions between i, i + 3 and i, i + 4 H-bonding in helices, observed previously in the crystallographic studies of depsipeptides (Ohyama, T.; Oku, H.; Hiroki, A.; Maekawa, Y.; Yoshida, M.; Katakai, R. Biopolymers 2000, 54, 375; Karle, I. L.; Das, C.; Balaram, P. Biopolymers 2001, 59, 276). Thus, the isodesmic reaction approach provides a simple way to gauge how conformation of the polypeptide chain and dimensions of the H-bonding network affect the strength of backbone-backbone C=O...HN bonds. The results indicate that the stabilization provided by such interactions increases on going from 3(10)-helix to alpha-helix to beta-sheet.  相似文献   

14.
The bonding of mandelic acid enantiomers has been studied on benzene-leucine, dinitrobenzene-leucine and dinitrobenzene-phenylalanine type chiral stationary phases connected to zeolite A supports. The pi-donor, pi-acceptor and H-bonding interactions responsible for diastereomer pair formations can be studied under quasi in situ chromatographic conditions by Fourier transform Raman and surface enhanced Raman spectroscopic techniques. Structural differences between diastereomer pairs result in observable spectral differences at a phase load of approx. 50%. It was shown that the decreasing pi-acceptor character of the phase is associated with its increasing capability of H-bond formation. Correlating spectral data to chromatographic results it can be concluded that, in addition to H-bonding as well as to pi-donor-pi-acceptor interactions, steric hindrances due to bulky moieties of either the stationary phase or the analyte molecules are of importance in successful separations.  相似文献   

15.
Complexes containing pyridine-2-carboxaldehyde (pyca) ligand acting as κ(2)-(N,O) chelates in [MX(CO)(3)(pyca)] (M = Mn, Re; X = Cl, Br), or [MoX(methallyl)(CO)(2)(pyca)] (X = Cl, Br), are good precursors for iminopyridine complexes derived from amino esters and peptides of formula [MX(CO)(3)(py-2-C(H)═NCHX-COOY)] or [MoX(methallyl)(CO)(2)(py-2-C(H)═NCHX-COOY)], via Schiff condensation of the aldehyde function of pyca with the terminal NH(2) group of the amino ester or peptide. X-ray determinations confirm the structures and show that in solid phase the peptide chains assemble through H-bonds adopting different patterns which depend on the geometry of the metal-ligand fragments. The H-bonding patterns have been analyzed in detail and described by using graph set methods. In most cases, Mo complexes show intramolecular arrangement involving the halogen (Cl or Br) and an NH group of the side chain. For the Mn and Re complexes, the peptide side arms form infinite chains, helices, and rings. In many cases, the terminal carboxylic O-H function is engaged in a "terminal" H-bond with a polar molecule of solvent (THF or acetone), instead of forming the usual head-to-head arrangement found in simple carboxylic acids. For the longer tripeptide Gly-Gly-Gly, a discrete, dimeric association is observed, in which the peptide chains show antiparallel arrangement with a complementary disposition of the internal N-H and C═O functions. DOSY experiments in solution show significant changes in the diffusion rates upon addition of OPBu(3), which indicate H-bonding interaction of OPBu(3) with the peptide hydrogens.  相似文献   

16.
Coordinated N,N',N"-trimethyldiethylenetriamine (Me3dien) has several possible configurations: two have mirror symmetry (R,S configurations at the terminal nitrogens) and the terminal N-Me's anti or syn with respect to the central N-Me (anti-(R,S) and syn-(R,S) isomers, respectively), and two are nonsymmetrical (R,R and S,S configurations at terminal nitrogens, rac denotes a 1:1 mixture of the two isomers). For each configuration, two Me3dienPtG atropisomers can be formed (anti or syn orientation of central N-Me and G 06, G = guanine derivative), and these can be observed since the terminal N-Me's decrease the rate of G rotation about the Pt-N7 bond. In symmetrical syn-(R,S)-Me3dienPtG derivatives with G = 9-EtG and 3'-GMP, the anti rotamer, which can form O6-NH H-bonds, was slightly favored over the syn rotamer but never more than 2:1. This anti rotamer is also favored by lower steric repulsion between the terminal N-Me's and G O6; thus, the contribution of O6-NH H-bonding to the stability of the anti rotamer could be rather small. With G = 5'-GMP, an O6-NH H-bond in the anti rotamer and a phosphate-NH H-bond in the syn rotamer can form. Only the syn rotamer was detected in solution, indicating that NH H-bonds to 5'-phosphate are far more important than to O6, particularly since steric factors favor the anti rotamer. Interconversion between rotamers was faster for syn-(R,S)- than for rac-Me3dien derivatives. This appears to be determined by a smaller steric impediment to G rotation of two "quasi equatorial" N-Me's, both on one side of the platinum coordination plane (syn-(R,S) isomer), than one "quasi equatorial" and one "quasi axial" N-Me on either side of the coordination plane (rac isomer).  相似文献   

17.
The solution (1)H 1D and 2D NMR spectra of the high-spin ferric, resting-state, substrate-bound complex of heme oxygenase, HO, from the pathological bacterium N. meningitidis have been investigated to assess the prospects for definitive assignment of hyperfine shifted and relaxed residue protons and the interpretation of those shifts in terms of the anisotropy and orientation of the paramagnetic susceptibility tensor, chi. Appropriately tailored 1D/2D NMR data, together with analyses of paramagnetic relaxation and a preliminary estimate of the magnetic anisotropy, reveal a chi that is axially anisotropic and oriented along the Fe-His vector. Together with T(-)(2) dependence of the shifts, Deltachi(ax) yields a zero-field splitting constant, D = 9.1 cm(-)(1), which is expected to serve as a very sensitive probe of H-bond interactions between the iron-ligated water and a series of distal ordered water molecules implicated in the mechanism of HO action. The side chains, Gln49 and His53, involved in the stabilization of catalytically relevant water molecules, were found to exhibit orientations rotated by 180 degrees about the beta-gamma bonds in solution relative to those in the crystal. The implication of these reorientations on the details of the distal H-bond network is discussed. The H-bond donor strengths of Gln 49 and His53 were found to respond appropriately to H-bond donor (water) versus H-bond acceptor (cyanide) iron ligands. Very slow NH exchange for the N-terminal portion of the distal helix suggest that an intrinsically "unstable" distal helix may be valid only for the C-terminal portion.  相似文献   

18.
Inspired by the observation of polar interactions between CO and O(2) ligands and the peptide residues at the active site of hemoglobin and myoglobin, we synthesized two kinds of superstructured porphyrins: TCP-IM, which contains a linked imidazole ligand, and TCP-PY, which contains a linked pyridine ligand, and examined the thermodynamic, kinetic, and spectroscopic (UV/Vis, IR, NMR, and resonance Raman) properties of their CO and O(2) complexes. On both sides of each porphyrin plane, bulky binaphthyl bridges form hydrophobic cavities that are suitable for the binding of small molecules. In the proximal site, an imidazole or pyridine residue is covalently fixed and coordinates axially to the central iron atom. In the distal site, two naphtholic hydroxyl groups overhang toward the center above the heme. The CO affinities of TCPs are significantly lower than those of other heme models. In contrast, TCPs have moderate O(2) binding ability. Compared with reported model hemes, the binding selectivity of O(2) over CO in TCP-IM and TCP-PY complexes is greatly improved. The high O(2) selectivity of the TCPs is mainly attributable to a low CO affinity. The comparison of k(on)(CO) values of TCPs with those of unhindered hemes indicates the absence of steric hindrance to the intrinsically linear CO coordination to Fe(II) in TCP-IM and TCP-PY. The abnormally large k(off)(CO) values are responsible for the low CO affinities. In contrast, k(off)(O(2)) of TCP-PY is smaller than those of other pyridine-coordinated model hemes. For the CO adducts of TCPs, unusually low nu(Fe-CO) and unusually high nu(C-O) frequencies are observed. These results can be ascribed to decreased back-bonding from the iron atom to the bound CO. The lone pairs of the oxygen atoms of the hydroxyl groups prevent back-bonding by exertion of a strong negative electrostatic interaction. On the other hand, high nu(Fe-O(2)) frequencies are observed for the O(2) adducts of TCPs. In the resonance Raman (RR) spectrum of oxy-TCP-IM, we observed simultaneous enhancement of the Fe-O(2) and O-O stretching modes. Furthermore, direct evidence for hydrogen bonding between the hydroxyl groups and bound dioxygen was obtained by RR and IR spectroscopy. These spectroscopic data strongly suggest that O(2) and CO binding to TCPs is controlled mainly by the two different electrostatic effects exerted by the overhanging OH groups: destabilization of CO binding by decreasing back-bonding and stabilization of O(2) binding by hydrogen bonding.  相似文献   

19.
Non-covalent (H-bonding) interactions, either intramolecular or with the surrounding medium, have a major influence on the activity of natural and synthetic phenolic antioxidants, due to the modulation of their reactivity with radical species, such as peroxyl radicals. Different cases can be distinguished. (i) Intra- or inter-molecular H-bonding involving the reactive -OH moiety will depress the antioxidant activity if the -OH acts as H-bond donor, while the opposite will generally occur if it acts as H-bond acceptor. (ii) Remote intra- and inter-molecular H-bonding, involving a distant -OH group (in polyphenols) or a ring substituent, may increase or decrease the reactivity of an antioxidant toward free radicals, depending on whether the stabilization produced by the H-bond increases or decreases along the reaction coordinate, on proceeding from reactants to the transition state. In this Perspective, the role of non-covalent interactions in the complex chemistry of natural polyphenolic antioxidants is discussed with the aid of literature data on simplified model compounds, aiming at the composition of a clear picture that might guide future research.  相似文献   

20.
Proton spin relaxation induced by the triplet ground state of O(2) in the zinc-containing diamagnetic analogue of sperm whale deoxymyoglobin has been measured as a function of oxygen concentration. As no covalent binding of oxygen to the metal occurs in the zinc species, the relaxation effects of O(2) on the protein (1)H resonances arise exclusively via much weaker noncovalent interactions. The relaxation effects at the amide proton sites are found to be highly localized and are derived almost exclusively from O(2) binding at the four previously identified xenon binding sites. Relative binding constants of 1.0, 0.08, 0.07, and 0.23 were determined for the Xe 1, Xe 2, Xe 3, and Xe 4 sites, respectively. In combination with earlier measurements of the kinetics of the heme binding of oxygen, these equilibria measurements enable a more detailed analysis of models characterizing O(2) entry and egress. A correlation is established between the fraction of O(2) which enters the Fe(2+)-binding site via rotation of the distal histidine side chain (so-called "histidine gate") and the experimentally observable O(2) (or CO) lifetime in the Xe 1 site. A physiological role for these secondary oxygen binding sites is proposed in enhancing the efficiency of the O(2) association reaction by rendering more favorable its competition with water binding in the distal heme pocket.  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