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.
Andersson J  Barth A 《Biopolymers》2006,82(4):353-357
As part of our work to determine the bond properties of the aspartyl phosphate moiety of the Ca(2+)-ATPase (SERCA1a) phosphoenzymes, we analyzed Morse potentials of the bridging P-O bond as well as C=O bond strengths for the model compound acetyl phosphate and the two phosphoenzyme intermediates Ca(2)E1P and E2P. Reaction-induced infrared difference spectroscopy was used and a carbonyl band of E2P at 1708 cm(-1) in the presence of mM Mg(2+) was tentatively assigned to the carbonyl group of phosphorylated Asp(351) because of its sensitivity to divalent cations. This band is found at 1716 cm(-1) with mM Ca(2+), for Ca(2)E1P at 1717 cm(-1) with Mg(2+), and at 1719 cm(-1) with Ca(2+) and at 1718 cm(-1) for acetyl phosphate in the absence of divalent cations. The similar band positions indicate similar strengths of interaction of the carbonyl oxygen in acetyl phosphate and the two phosphoenzymes. Together with information on the P-O bond strengths, this implies that the bridging oxygen exerts stronger interactions in the phosphoenzymes than in acetyl phosphate.  相似文献   

2.
Phosphate binding to the sarcoplasmic reticulum Ca2+-ATPase was studied by time-resolved Fourier transform infrared spectroscopy with ATP and isotopically labeled ATP ([beta-18O2, betagamma-18O]ATP and [gamma-18O3]ATP). Isotopic substitution identified several bands that can be assigned to phosphate groups of bound ATP: bands at 1260, 1207, 1145, 1110, and 1085 cm(-1) are affected by labeling of the beta-phosphate, bands likely near 1154, and 1098-1089 cm(-1) are affected by gamma-phosphate labeling. The findings indicate that the strength of interactions of beta- and gamma- phosphate with the protein are similar to those in aqueous solution. Two bands, at 1175 and 1113 cm(-1), were identified for the phosphate group of the ADP-sensitive phosphoenzyme Ca2E1P. They indicate terminal and bridging P-O bond strengths that are intermediate between those of ADP-insensitive phosphoenzyme E2P and the model compound acetyl phosphate in water. The bridging bond of Ca2E1P is weaker than for acetyl phosphate, which will facilitate phosphate transfer to ADP, but is stronger than for E2P, which will make the Ca2E1P phosphate less susceptible to attack by water.  相似文献   

3.
A general method to study the phosphate group of phosphoenzymes with infrared difference spectroscopy by helper enzyme-induced isotope exchange was developed. This allows the selective monitoring of the phosphate P-O vibrations in large proteins, which provides detailed information on several band parameters. Here, isotopic exchange was achieved at the oxygen atoms of the catalytically important phosphate group that transiently binds to the sarcoplasmic reticulum Ca(2+)-ATPase (SERCA1a). [gamma-(18)O(3)]ATP phosphorylated the ATPase, which produced phosphoenzyme that was initially isotopically labeled. The helper enzyme adenylate kinase regenerated the substrate ATP from ADP (added or generated upon ATP hydrolysis) with different isotopic composition than used initially. With time this produced the unlabeled phosphoenzyme. The method was tested on the ADP-insensitive phosphoenzyme state of the Ca(2+)-ATPase for which the vibrational frequencies of the phosphate group are known, and it was established that the helper enzyme is effective in mediating the isotope exchange process.  相似文献   

4.
Kinetics of hydrolysis of poly(alkylene amidophosphate)s with amino acids or dipeptides as the side groups was studied by 31P NMR at pH 1.5, 6.5, and 8.5. The direction of hydrolysis and the relative rate coefficients of breaking P-O bonds in the main chain and P-N bonds in the side groups depend strongly on the pH of the medium of hydrolysis. The P-N (amide) bond hydrolyzes much faster than the P-O (ester) bond in acidic and close to neutral conditions (negligible P-O hydrolysis), whereas above pH > or = 8.5 these differences are much smaller. For instance, for 4-Ala the rate coefficients of hydrolysis are equal (in H2O at 37 degrees C and pH 8.5) to 1.9 x 10(-8) s(-1) and 1.0 x 10(-9) s(-1) for the P-N and P-O bonds, respectively, quite different from the values found for the low molecular model 2 (0 and 1.4 x 10(-7) s(-1), respectively).  相似文献   

5.
Native solium and potassium adenosine triphosphatase from guinea pig kidney accepted a phosphate group from radioactive inorganic phosphate to form an acyl phosphate bond at the active site in the presence or absence of sodium ion. Magnesium ion was always required. In the presence of sodium ion and absence of adenosine triphosphate, there was no phosphorylation by inorganic phosphate. Addition of unlabeled adenosine triphosphate produced a potassium-sensitive phosphoenzyme which exchanged its phosphate-group with radioactive inorganic phosphate. The dephosphoenzyme was an intermediate in this exchange. The rate constant for dephosphorylation was about 0.05 per second. Addition of rubidium ion, a congener of potassium ion, to the potassium-sensitive phosphoenzyme produced a phosphoenzyme labeled from inorganic phosphate with a corresponding rate constant of 0.26 per s. This was a rubidium-complexed phosphoenzyme. Addition of magnesium ion to potassium-sensitive phosphoenzyme converted it into insensitive phosphoenzyme, the splitting of which was not accelerated by potassium ion or by adenosine diphosphate. Its rate constant was 0.07 per s. In the absence of sodium ion and adenosine triphosphate, inorganic phosphate was incorporated directly into a similar insensitive phosphoenzyme. In the presence of potassium ion or rubidium ion, inorganic phosphate was incorporated into a potassium-complexed or rubidium-complexed phosphoenzyme which exchanged 32-P with inorganic phosphate completely in less than 3 s. Incorporation of inorganic phosphate into a complex of the enzyme with the inhibitor, ouabain, is already described in the literature. Its rate constant was about 0.02 per s. Thus there appear to be at least four reactive states of the phosphoenzyme which equilibrate measurably with inorganic phosphate, namely, potassium-sensitive phosphoenzyme, potassium-complexed phosphoenzyme, insensitive phosphoenzyme, and ouabain phosphoenzyme. Two of these reactive states are functional intermediates in native sodium and potassium ion transport adenosine triphosphatase. The results are compatible with control of the reactivity of the active site by conformational changes in the surrounding active center and with regulation of the energy level of the phosphate group according to the kind of monovalent cation bound to the enzyme.  相似文献   

6.
Y Kuriki  E Racker 《Biochemistry》1976,15(23):4951-4956
The bioflavonoid, quercetin, inhibited the (Na+, K+)adenosine triphosphatase purified from the electric organ of electric eel (Electrophorus electricus) or from lamb kidney. An analysis of its mode of action revealed that the formation of phosphoenzyme from Pi but not from ATP was inhibited. Quercetin increased the amount of ADP-sensitive phosphoenzyme (E1--P), indicating an inhibition of the conversion of E1--P to the ADP-insensitive form (E2--P). The rate of dephosphorylation of the phosphoenzyme formed from ATP was slowed by quercetin. These results suggest that quercetin inhibits the formation of E2--P from either Pi or E1-P as well as the hydrolysis of the phosphoenzyme. Its mode of action is therefore different from that of ouabain and other inhibitors of the Na+, K+)adenosine triphosphatase.  相似文献   

7.
Ca(2+)-ATPase belongs to the family of P-type ATPases and maintains low concentrations of intracellular Ca(2+). Its reaction cycle consists of four main intermediates that alternate ion binding in the transmembrane domain with phosphorylation of an aspartate residue in a cytoplasmic domain. Previous work characterized an ultrastable phosphoenzyme produced first by labeling with fluorescein isothiocyanate, then by allowing this labeled enzyme to establish a maximal Ca(2+) gradient, and finally by removing Ca(2+) from the solution. This phosphoenzyme is characterized by very low fluorescence and has specific enzymatic properties suggesting the existence of a high energy phosphoryl bond. To study the structural properties of this phosphoenzyme, we used cryoelectron microscopy of two-dimensional crystals formed in the presence of decavanadate and determined the structure at 8-A resolution. To our surprise we found that at this resolution the low fluorescence phosphoenzyme had a structure similar to that of the native enzyme crystallized under equivalent conditions. We went on to use glutaraldehyde cross-linking and proteolysis for independent structural assessment and concluded that, like the unphosphorylated native enzyme, Ca(2+) and vanadate exert a strong influence over the global structure of this low fluorescence phosphoenzyme. Based on a structural model with fluorescein isothiocyanate bound at the ATP site, we suggest that the stability as well as the low fluorescence of this phosphoenzyme is due to a fluorescein-mediated cross-link between two cytoplasmic domains that prevents hydrolysis of the aspartyl phosphate. Finally, we consider the alternative possibility that phosphate transfer to fluorescein itself could explain the properties of this low fluorescence species.  相似文献   

8.
Time-resolved Fourier transform infrared difference spectra of the phosphoenzyme conversion and Ca(2+) release reaction (Ca(2)E(1)-P --> E(2)-P) of the sarcoplasmic reticulum Ca(2+)-ATPase were recorded at pH 7 and 1 degrees C in H(2)O and (2)H(2)O. In the amide I spectral region, the spectra indicate backbone conformational changes preserving conformational changes of the preceding phosphorylation step. beta-sheet or turn structures (band at 1685 cm(-1)) and alpha-helical structures (band at 1653 cm(-1)) seem to be involved. Spectra of the model compound EDTA for Ca(2+) chelation indicate the assignment of bands at 1570, 1554, 1411 and 1399 cm(-1) to Ca(2+) chelating Asp and Glu carboxylate groups partially shielded from the aqueous environment. In addition, an E(2)-P band at 1638 cm(-1) has been tentatively assigned to a carboxylate group in a special environment. A Tyr residue seems to be involved in the reaction (band at 1517 cm(-1) in H(2)O and 1515 cm(-1) in (2)H(2)O). A band at 1192 cm(-1) was shown by isotopic replacement in the gamma-phosphate of ATP to originate from the E(2)-P phosphate group. This is a clear indication that the immediate environment of the phosphoenzyme phosphate group changes in the conversion reaction, altering phosphate geometry and/or electron distribution.  相似文献   

9.
Using inside-out vesicles of human red cell membranes, the side-specific effects of Na+ on phosphorylation of (Na,K)-ATPase have been studied using low concentrations of [gamma-32P]ATP (less than or equal to 0.1 microM). Phosphorylation is stimulated by Na+ at the cytoplasmic membrane surface (extravesicular Na+) alone and not by Na+ at the external surface (intravesicular Na+). At 37 degrees C, external Na+ (less than or equal to 10 mM) does, however, increase the steady state level (approximately 2 1/2-fold) of phosphoenzyme above that observed with cytoplasmic Na+ alone; hydrolysis is increased to only a small extent. Little stimulation by external Na+ is observed at 0 degrees C. As Na+ at the cytoplasmic side is decreased to very low levels (less than or equal to 0.2 mM) several kinetic changes are observed: (i) the apparent turnover of phosphoenzyme (ratio Na+-ATP-ase/phosphoenzyme level) is markedly increased (approximately 3-fold, (ii) Rbext sensitivity (inhibition of (Na)-ATPase at low ATP levels) is reduced, and (iii) the ratio of Na+ ions transported per molecule of ATP hydrolyzed is decreased. These results are compatible with a reaction pathway involving a transition from one form of phosphoenzyme, E1-P, to another, E2-P of which the hydrolysis is decreased by moderate levels of external Na+. It is suggested also that an alternate reaction pathway for Na+-ATPase occurs at very low cytoplasmic Na+, one via hydrolysis of E1-P and not associated with Na+ translocation.  相似文献   

10.
We re-examined the kinetics of the bisphosphatase reaction of rat hepatic 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase after depleting the enzyme of bound fructose 6-phosphate and found a hyperbolic dependence on fructose 2,6-bisphosphate at concentrations below 100 nM. The Michaelis constant was 4 nM, the Vmax was about 12 nmol X mg-1 X min-1 at 22 degrees C but the substrate inhibited at concentrations above 100 nM. Both phosphate and alpha-glycerol phosphate strongly inhibited phosphoenzyme formation and hydrolytic rate below 100 nM, but relieved the inhibition by substrate at higher concentrations probably by antagonizing substrate binding. A number of observations support the proposition that the phosphoenzyme is a necessary participant in catalysis. 1) The amount of phosphoenzyme measured during steady-state hydrolysis as a function of substrate concentration correlated with the velocity profile. 2) Rapid mixing experiments demonstrated that over a broad range of substrate concentrations phosphoenzyme formation was faster than the net rate of hydrolysis. 3) Both phosphate and alpha-glycerol phosphate inhibited the rate of phosphoenzyme formation and, at low substrate concentrations, reduced the steady-state phosphoenzyme levels. The latter correlated with inhibition of substrate hydrolysis. 4) Both phosphate and alpha-glycerol phosphate stimulate the rate of phosphoenzyme breakdown, consistent with their stimulation of substrate hydrolysis at high substrate concentrations. 5) The fractional rate of phosphoenzyme breakdown, which was pH and substrate dependent, multiplied by the amount of phosphoenzyme obtained in the steady state at that pH and substrate concentration approximated the observed rate of hydrolysis. We conclude that the phosphoenzyme is a reaction intermediate in the hepatic fructose-2,6-bisphosphatase reaction.  相似文献   

11.
The intrinsic fluorescence of sarcoplasmic reticulum vesicles was measured under conditions allowing ATPase phosphorylation from inorganic phosphate. Significant fluorescence enhancement of up to 4% resulted from gradient-independent enzyme phosphorylation at pH 6, in the absence of KCl. The equilibrium fluorescence data obtained at various magnesium and phosphate concentrations agree with a reaction scheme in which Mg2+, as direct activator, and free phosphate, as the true substrate, bind to the enzyme in random order to give a noncovalent ternary complex (Mg.*E.Pi), in equilibrium with the covalent phosphoenzyme (Mg.*E-P). The transient kinetics of the fluorescence rise was also studied, and the resulting data were generally consistent with the above scheme, assuming that binding reactions are fast compared to covalent phosphoenzyme formation. This, however, might be valid only as a first approximation. At 20 degrees C and pH 6, the phosphate concentration for half-maximum phosphorylation rate constant, at 20 mM magnesium, was higher than 20 mM. Similarly, the magnesium concentration for half-maximum phosphorylation rate constant, at 20 mM phosphate, was also higher than 20 mM. The maximum phosphorylation rate was faster than 25 s-1, and the phosphoenzyme hydrolysis rate constant was 1.5-2 s-1 under these conditions, so that the equilibrium constant between Mg.*E.Pi and Mg.*E-P largely favors the phosphoenzyme.  相似文献   

12.
In order to examine the energetics in base-catalyzed hydrolysis of RNA, a tentative pentacoordinated intermediate (3) has been characterized by molecular orbital calculations. Ab initio studies at the level of 3-21G* indicate that, under the Cs symmetry restricted conditions, the P-O(2) bond possessing antiperiplanar (app) lone pair electrons (Ip) on the equatorial oxygen (O(3)) can be cleaved with almost no barrier (TS1 transition state; 0.08 kcal mol-1), from the pentacoordinated intermediate (3) of base-catalyzed hydrolysis of phosphate, compared to the P-O(5) bond (TS2 transition state; 28.9 kcal mol-1) which lacks app lp assistance from O(3). The dianionic intermediate, however, loses the TS1 transition state thus its property as an intermediate when the Cs restriction is removed. The analysis of the entire potential energy surface enables us to conclude that, in a related system examined by Lim and Karplus [1990) J. Am. Chem. Soc., 112, 5872-5873) for attack by OH- on ethylene phosphate monoanion, the TS1 transition state had also been lost and thus no intermediate had been found. These results further support our earlier conclusions (Taira et al. (1990) Protein Engineering, 3, 691-701) of rate limiting transition state possessing extended P-O(5') bond breaking character (the TS2 transition state) in the base-catalyzed hydrolysis of RNA. Finally, although the lack of 2',3' -migration of phosphate moieties in basic condition appears to be in accord with the short-lived intermediate, it really does not prove the absence of the intermediate. The detail will be discussed in the text.  相似文献   

13.
Hydrolysis of acetyl phosphate is inhibited by high concentrations of Pi and MgCl2, probably due to an increase in the steady-state level of phosphoenzyme formed from Pi in the medium. A dual effect of ADP during steady-state hydrolysis of acetyl phosphate was observed. ADP inhibited hydrolysis in the presence of 5 mM MgCl2 and no added Pi, whereas it stimulated hydrolysis when phosphoenzyme formation by Pi was favored by including 6 mM Pi and 20 mM MgCl2 in the assay medium. ATP inhibited acetyl phosphate hydrolysis in both of these assay media. When phosphoenzyme formation by Pi in the presence of acetyl phosphate was stimulated at Ca2+ concentrations sufficient to saturate the low-affinity Ca2+-binding sites, ADP stimulated acetyl phosphate hydrolysis and also promoted ATP synthesis by reversal of the catalytic cycle. The rate of ATP synthesis was dependent on ADP, Pi and Ca2+. Phosphoenzyme formation by Pi and MgCl2, whether in the absence of Ca2+ and acetyl phosphate, or during acetyl phosphate hydrolysis, was inhibited by ADP and ATP. These results suggest that ADP interacts with different intermediates of the catalytic cycle and that expression of inhibition or activation of acetyl phosphate hydrolysis depends on the steady-state level of phosphoenzyme formed by Pi.  相似文献   

14.
Inesi G  Lewis D  Ma H  Prasad A  Toyoshima C 《Biochemistry》2006,45(46):13769-13778
We relate solution behavior to the crystal structure of the Ca2+ ATPase (SERCA). We find that nucleotide binding occurs with high affinity through interaction of the adenosine moiety with the N domain, even in the absence of Ca2+ and Mg2+, or to the closed conformation stabilized by thapsigargin (TG). Why then is Ca2+ crucial for ATP utilization? The influence of adenosine 5'-(beta,gamma-methylene) triphosphate (AMPPCP), Ca2+, and Mg2+ on proteolytic digestion patterns, interpreted in the light of known crystal structures, indicates that a Ca2+-dependent conformation of the ATPase headpiece is required for a further transition induced by nucleotide binding. This includes opening of the headpiece, which in turn allows inclination of the "A" domain and bending of the "P" domain. Thereby, the phosphate chain of bound ATP acquires an extended configuration allowing the gamma-phosphate to reach Asp351 to form a complex including Mg2+. We demonstrate by Asp351 mutation that this "productive" conformation of the substrate-enzyme complex is unstable because of electrostatic repulsion at the phosphorylation site. However, this conformation is subsequently stabilized by covalent engagement of the -phosphate yielding the phosphoenzyme intermediate. We also demonstrate that the ADP product remains bound with high affinity to the transition state complex but dissociates with lower affinity as the phosphoenzyme undergoes a further conformational change (i.e., E1-P to E2-P transition). Finally, we measured low-affinity ATP binding to stable phosphoenzyme analogues, demonstrating that the E1-P to E2-P transition and the enzyme turnover are accelerated by ATP binding to the phosphoenzyme in exchange for ADP.  相似文献   

15.
The molecular structure and relative stability of north and south conformers of 2'-deoxyribonucleotides containing pyrimidine nucleic acid bases ( 2'-deoxythymidilic (pdT), 2'-deoxycytidilic (pdC) acids and their mono- and dianions) have been obtained and analyzed at the DFT/B3LYP level using the standard 6-31G(d) basis set. We have revealed that, when the nucleobase moiety is incorporated into the nucleotides, it maintains a nonplanar and nonrigid conformation due to out-of-plane deformation of the amino group and pyrimidine ring. It has been demonstrated that an increase of negative charge of the phosphate group results in increase of amino group pyramidalization, discrimination between conformers with syn and anti orientation of base with respect to sugar, strengthening of intramolecular C-H.O hydrogen bonds leading to deformation and fixation of geometry of nucleotides, and weakening of phosphodiester bond. These results allow to make suggestions about sources of twist and buckle deformations of base pairs, mechanisms of repaire of DNA via change of base orientation, and conditions for breakage of the P-O bonds during hydrolysis.  相似文献   

16.
Membrane-bound (H+ + K+)-ATPase purified from hog gastric mucosa was exposed to limited papain digestion. Such treatment resulted in a rapid inhibition of the K+-stimulated adenosine triphosphatase and p-nitrophenyl phosphatase activities, with about 90% of these activities lost after 3 min incubation at 37 degrees C with 0.1 units of papain per mg of enzyme protein. Parallel to the inhibition of the enzyme activities, there was a production of a 77 kDa membrane-bound fragment containing the aspartyl phosphate residue of the phospho-intermediate. This fragment accounted for about 45% of the total enzyme protein after the 3 min papain treatment. The digestion barely affected the steady-state level of phosphorylation, allowed the aspartyl phosphate of the 77 kDa fragment to undergo the transition to the E2P form, and did not significantly alter the fraction of ADP-sensitive phosphoenzyme. The presence of KCl, however, depressed the steady-state level of phosphoenzyme formed from [gamma-32P]ATP considerably less than that of the control enzyme. With further exposure to papain the 77 kDa peptide became fragmented into a 28 kDa soluble peptide that retained the phosphorylating site. Binding of fluorescein 5'-isothiocyanate (FITC) to the native enzyme did not affect the sites of papain hydrolysis because the same peptide fragments were obtained. The FITC reaction site was also in the 28 kDa soluble peptide fragment.  相似文献   

17.
Inhibitor and ion binding sites on the gastric H,K-ATPase   总被引:2,自引:0,他引:2  
Munson K  Garcia R  Sachs G 《Biochemistry》2005,44(14):5267-5284
The gastric H,K-ATPase catalyzes electroneutral exchange of H(+) for K(+) as a function of enzyme phosphorylation and dephosphorylation during transition between E(1)/E(1)-P (ion site in) and E(2)-P/E(2) (ion site out) conformations. Here we present homology modeling of the H,K-ATPase in the E(2)-P conformation as a means of predicting the interaction of the enzyme with two known classes of specific inhibitors. All known proton pump inhibitors, PPIs, form a disulfide bond with cysteine 813 that is accessible from the luminal surface. This allows allocation of the binding site to a luminal vestibule adjacent to Cys813 enclosed by part of TM4 and the loop between TM5 and TM6. K(+) competitive imidazo-1,2alpha-pyridines also bind to the luminal surface of the E(2)-P conformation, and their binding excludes PPI reaction. This overlap of the binding sites of the two classes of inhibitors combined with the results of site-directed mutagenesis and cysteine cross-linking allowed preliminary assignment of a docking mode for these reversible compounds in a position close to Glu795 that accounts for the detailed structure/activity relationships known for these compounds. The new E(2)-P model is able to assign a possible mechanism for acid secretion by this P(2)-type ATPase. Several ion binding side chains identified in the sr Ca-ATPase by crystallography are conserved in the Na,K- and H,K-ATPases. Poised in the middle of these, the H,K-ATPase substitutes lysine in place of a serine implicated in K(+) binding in the Na,K-ATPase. Molecular models for hydronium binding to E(1) versus E(2)-P predict outward displacement of the hydronium bound between Asp824, Glu820, and Glu795 by the R-NH(3)(+) of Lys791 during the conformational transition from E(1)P and E(2)P. The site for luminal K(+) binding at low pH is proposed to be between carbonyl oxygens in the nonhelical part of the fourth membrane span and carboxyl oxygens of Glu795 and Glu820. This site of K(+) binding is predicted to destabilize hydrogen bonds between these carboxylates and the -NH(3)(+) group of Lys791, allowing the Lys791 side chain to return to its E(1) position.  相似文献   

18.
Y Guan  G J Thomas  Jr 《Biophysical journal》1996,71(5):2802-2814
A generalized valence force field is derived for the diethyl phosphate anion [(CH3CH2O)2PO2-] and its deuterium [(CH3CD2O)2PO2-, (CD3CH2O)2PO2- and (CD3CD2O)2PO2-] and carbon-13 [(CH3 13CH2O)2PO2-] derivatives in the stable trans-gauche-gauche-trans conformation. Normal coordinate analysis of the trans-gauche-gauche-trans conformer, which serves as a structural analog of the nucleic acid phosphodiester group, is based on comprehensive infrared and Raman spectroscopic data and vibrational assignments obtained for the diethyl phosphate anion. The generalized valence force field is in good agreement with the scaled ab initio force field of diethyl phosphate and represents significant improvement over earlier modeling of the phosphodiester moiety with dimethyl phosphate. The conformational dependence of skeletal C-C-O-P(O2-)-O-C-C stretching vibrations is also explored. Starting with the trans-gauche-gauche-trans conformation, the frequency dependence of skeletal stretching modes has been obtained by stepwise rotation of the torsion angles of the P-O and C-O bonds corresponding to nucleic acid torsions alpha (P-O5'), beta (O5'-C5'), epsilon (C3'-O3'), and zeta (O3'-P). Both symmetric and antisymmetric phosphoester stretching modes are highly sensitive to P-O and C-O torsions, whereas symmetric and antisymmetric phosphodioxy (PO2-) stretching modes are less sensitive. The present results provide an improved structural basis for understanding previously developed empirical correlations between vibrational marker bands and nucleic acid backbone conformation.  相似文献   

19.
Abstract

The molecular structure and relative stability of north and south conformers of 2′-deoxyribonucleotides containing pyrimidine nucleic acid bases (2′-deoxythymidilic (pdT), 2′- deoxycytidilic (pdC) acids and their mono- and dianions) have been obtained and analyzed at the DFT/B3LYP level using the standard 6–31G(d) basis set. We have revealed that, when the nucleobase moiety is incorporated into the nucleotides, it maintains a nonplanar and nonrigid conformation due to out-of-plane deformation of the amino group and pyrimidine ring. It has been demonstrated that an increase of negative charge of the phosphate group results in increase of amino group pyramidalization, discrimination between conformers with syn and anti orientation of base with respect to sugar, strengthening of intramolecular C-H…O hydrogen bonds leading to deformation and fixation of geometry of nucleotides, and weakening of phosphodiester bond. These results allow to make suggestions about sources of twist and buckle deformations of base pairs, mechanisms of repaire of DNA via change of base orientation, and conditions for breakage of the P-O bonds during hydrolysis.  相似文献   

20.
Grzyska PK  Kim Y  Jackson MD  Hengge AC  Denu JM 《Biochemistry》2004,43(27):8807-8814
Dual-specificity phosphatases (DSPs) belong to the large family of protein tyrosine phosphatases that contain the active-site motif (H/V)CxxGxxR(S/T), but unlike the tyrosine-specific enzymes, DSPs are able to catalyze the efficient hydrolysis of both phosphotyrosine and phosphoserine/threonine found on signaling proteins, as well as a variety of small-molecule aryl and alkyl phosphates. It is unclear how DSPs accomplish similar reaction rates for phosphoesters, whose reactivity (i.e., pK(a) of the leaving group) can vary by more than 10(8). Here, we utilize the alkyl phosphate m-nitrobenzyl phosphate (mNBP), leaving-group pK(a) = 14.9, as a physiological substrate mimic to probe the mechanism and transition state of the DSP, Vaccinia H1-related (VHR). Detailed pH and kinetic isotope effects of the V/K value for mNBP indicates that VHR reacts with the phosphate dianion of mNBP and that the nonbridge phosphate oxygen atoms are unprotonated in the transition state. (18)O and solvent isotope effects indicate differences in the respective timing of the proton transfer to the leaving group and P-O fission; with the alkyl ester substrate, protonation is ahead of P-O fission, while with the aryl substrate, the two processes are more synchronous. Kinetic analysis of the general-acid mutant D92N with mNBP was consistent with the requirement of Asp-92 in protonating the ester oxygen, either in a step prior to significant P-O bond cleavage or in a concerted but asynchronous mechanism in which protonation is ahead of P-O bond fission. Collectively, the data indicate that VHR and likely all DSPs can match leaving-group potential with the timing of the proton transfer to the ester oxygen, such that diverse aryl and alkyl phosphoesters are turned over with similar catalytic efficiency.  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