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 15 毫秒
1.
A new computer program, DISCO, running under Windows, has been developed under the project CSA98P22 falling within the Competitive Support Activities initiative launched within the EU 4th Framework Programme. DISCO allows the calculation of the stepwise acid dissociation constants of polyprotic molecules in water and in complex media (i.e., biofluids, etc.) from nuclear magnetic resonance (NMR) data (chemical shifts) by means of two derivative-free methods: Pit-mapping and Simplex. DISCO performances were tested using simulated-unaffected by experimental error-data sets, for systems having up to seven equilibrium constants and experimental NMR data of spermine, 6-monofluorospermine, and 6,6-difluorospermine, dissolved in D(2)O and in physiological solution (D(2)O/NaCl). Results demonstrated that (i) DISCO enables the determination of pK(A) values with high precision even when small-sized raw data sets are employed, when chemical shifts are measured with low precision (the usual condition in biofluids due to the impossibility to obtain narrow line shape), and when the guess solution, necessary as an initial step of the mathematical iterative process, is fixed within a large interval of variation; (ii) DISCO always converges to the root; (iii) DISCO permits the calculation of pK(A) values which lie within the observed pH range, independent of the narrowness of the pH range.  相似文献   

2.
Paper electrophoresis has been used over the pH range 1.2 to 10.4 to measure apparent pK(a) values for malvidin-3-O-glucoside of pK(a(1)) 1.76+/-0.07, pK(a(2)) 5.36+/-0.04, and pK(a(3)) 8.39+/-0.07. Using solvent partitioning between buffered aqueous solutions and n-octanol, several micro-pK(a) constants for malvidin-3-O-glucoside were also identified, highlighting the complex nature of malvidin-3-glucoside equilibria. As a nonspectrophotometric procedure, the charge-dependent electrophoretic mobility method provided independent information on the net charge and color of anthocyanin species at wine pH (ca. 3.6). At this pH, the color of malvidin-3-glucoside in red wines is consistent only with the uncharged quinonoidal base as a major colored component of the equilibria.  相似文献   

3.
The pK values of the titratable groups in ribonuclease Sa (RNase Sa) (pI=3.5), and a charge-reversed variant with five carboxyl to lysine substitutions, 5K RNase Sa (pI=10.2), have been determined by NMR at 20 degrees C in 0.1M NaCl. In RNase Sa, 18 pK values and in 5K, 11 pK values were measured. The carboxyl group of Asp33, which is buried and forms three intramolecular hydrogen bonds in RNase Sa, has the lowest pK (2.4), whereas Asp79, which is also buried but does not form hydrogen bonds, has the most elevated pK (7.4). These results highlight the importance of desolvation and charge-dipole interactions in perturbing pK values of buried groups. Alkaline titration revealed that the terminal amine of RNase Sa and all eight tyrosine residues have significantly increased pK values relative to model compounds.A primary objective in this study was to investigate the influence of charge-charge interactions on the pK values by comparing results from RNase Sa with those from the 5K variant. The solution structures of the two proteins are very similar as revealed by NMR and other spectroscopic data, with only small changes at the N terminus and in the alpha-helix. Consequently, the ionizable groups will have similar environments in the two variants and desolvation and charge-dipole interactions will have comparable effects on the pK values of both. Their pK differences, therefore, are expected to be chiefly due to the different charge-charge interactions. As anticipated from its higher net charge, all measured pK values in 5K RNase are lowered relative to wild-type RNase Sa, with the largest decrease being 2.2 pH units for Glu14. The pK differences (pK(Sa)-pK(5K)) calculated using a simple model based on Coulomb's Law and a dielectric constant of 45 agree well with the experimental values. This demonstrates that the pK differences between wild-type and 5K RNase Sa are mainly due to changes in the electrostatic interactions between the ionizable groups. pK values calculated using Coulomb's Law also showed a good correlation (R=0.83) with experimental values. The more complex model based on a finite-difference solution to the Poisson-Boltzmann equation, which considers desolvation and charge-dipole interactions in addition to charge-charge interactions, was also used to calculate pK values. Surprisingly, these values are more poorly correlated (R=0.65) with the values from experiment. Taken together, the results are evidence that charge-charge interactions are the chief perturbant of the pK values of ionizable groups on the protein surface, which is where the majority of the ionizable groups are positioned in proteins.  相似文献   

4.
The self-assembled supramolecular structures of diacylphosphatidylcholine (diC(n)PC), diacylphosphatidylethanolamine (diC(n)PE), diacylphosphatidyglycerol (diC(n)PG), and diacylphosphatidylserine (diC(n)PS) were investigated by (31)P nuclear magnetic resonance (NMR) spectroscopy as a function of the hydrophobic acyl chain length. Short-chain homologs of these lipids formed micelles, and longer-chain homologs formed bilayers. The shortest acyl chain lengths that supported bilayer structures depended on the headgroup of the lipids. They increased in the order PE (C(6)) < PC (C(9)) < or = PS (C(9) or C(10)) < PG (C(11) or C(12)). This order correlated with the effective headgroup area, which is a function of the physical size, charge, hydration, and hydrogen-bonding capacity of the four headgroups. Electrostatic screening of the headgroup charge with NaCl reduced the effective headgroup area of PS and PG and thereby decreased the micelle-to-bilayer transition of these lipid classes to shorter chain lengths. The experimentally determined supramolecular structures were compared to the assembly states predicted by packing constraints that were calculated from the hydrocarbon-chain volume and effective headgroup area of each lipid. The model accurately predicted the chain-length threshold for bilayer formation if the relative displacement of the acyl chains of the phospholipid were taken into account. The model also predicted cylindrical rather than spherical micelles for all four diacylphospholipid classes and the (31)P-NMR spectra provided evidence for a tubular network that appeared as an intermediate phase at the micelle-to-bilayer transition. The free energy of micellization per methylene group was independent of the structure of the supramolecular assembly, but was -0.95 kJ/mol (-0.23 kcal/mol) for the PGs compared to -2.5 kJ/mol (-0.60 kcal/mol) for the PCs. The integral membrane protein OmpA did not change the bilayer structure of thin (diC(10)PC) bilayers.  相似文献   

5.
A general method for formulating complex thermodynamic systems in terms of hierarchical interactions has been developed, and has been applied in a previous analysis to hemoglobin oxygen binding data. Polyprotic acids can be considered a simple chemical model of thermodynamic interaction between ligand binding events. To further illustrate the hierarchical interaction approach it is applied to the analysis of the thermodynamic interactions between proton binding events in inorganic polyprotic acids. pK values for arsenate, carbonate, chromate, phosphate, phosphite, selenite, sulfide and sulfite were recast into hierarchical interaction terms. The intrinsic K(d,h) for protonation ranged from 8.8 x 10(-13) (M) for phosphate to 1.3 x 10(-6) (M) for chromate. Pairwise interactions (K(d,hh)) between protonation events ranged from 1.3 x 10(4) for phosphite to 9.4 x 10(5) for carbonate. Third order interactions (K(d,hhh)) were 0.91 and 0.51 for arsenate and phosphate, respectively, values relatively close to the no interaction value of 1. A principle feature of systems described by hierarchical interactions is that higher order interactions, representing more complex interactions, are less likely to be significant than lower order interactions, and this is further illustrated by these observations from polyprotic acids. The set of significant hierarchical interaction values can be used to predict values for as yet unobserved events, and projected pK values are made for all the polyprotic acids included in this study. Finally, application of this method to the protonation equilibria of water demonstrates a profound pairwise interaction between protonation events (K(d,hh) = 1.3 x 10(17)), which is attributed to oxygen's small size and lack of polarizability.  相似文献   

6.
This work presents a study aimed at the theoretical prediction of pK(a) values of aminopyridines, as a factor responsible for the activity of these compounds as blockers of the voltage-dependent K(+) channels. To cover a large range of pK(a) values, a total of seven substituted pyridines is considered as a calibration set: pyridine, 2-aminopyridine, 3-aminopyridine, 4-aminopyridine, 2-chloropyridine, 3-chloropyridine, and 4-methylpirydine. Using ab initio G1, G2 and G3 extrapolation methods, and the CPCM variant of the Polarizable Continuum Model for solvation, we calculate gas phase and solvation free energies. pK(a) values are obtained from these data using a thermodynamic cycle for describing protonation in aqueous and gas phases. The results show that the relatively inexpensive G1 level of theory is the most accurate at predicting pK(a) values in aminopyridines. The highest standard deviation with respect to the experimental data is 0.69 pK(a) units for absolute values calculations. The difference increases slightly to 0.74 pK(a) units when the pK(a) is computed relative to the pyridine molecule. Considering only compounds at least as basic as pyridine (the values of interest for bioactive aminopyridines) the error falls to 0.10 and 0.12 pK(a) units for the absolute and relative computations, respectively. The technique can be used to predict the effect of electronegative substituents in the pK(a) of 4-AP, the most active aminopyridine considered in this work. Thus, 2-chloro and 3-chloro-4-aminopyridine are taken into account. The results show a decrease of the pK(a), suggesting that these compounds are less active than 4-AP at blocking the K(+) channel.  相似文献   

7.
Menyhárd DK  Keserü GM 《FEBS letters》2005,579(24):5392-5398
pK(a) values of ionizable residues were calculated for the crystal structures describing the pH and NO binding dependant conformations of nitrophorin 4, a pH sensitive NO carrier heme protein. Comparison of resultant H-bonding patterns allowed the identification of the amino acids that take part in signaling pH change. We carried out MD simulations to show that the protonation state of Asp30, buried in the closed conformation, is crucial for maintaining the tight packed conformation of the closed form of the complex - presenting a model for the functional decrease of NO binding affinity of nitrophorins at physiological pH.  相似文献   

8.
Prostaglandin H synthase-1 of ram vesicular glands metabolises 5,8,11-eicosatrienoic (Mead) acid to 13R-hydroxy-5,8,11-eicosatrienoic and to 11R-hydroxy-5,8,12-eicosatrienoic in a 5:1 ration. We wanted to determine the metabolism of this fatty acid by prostaglandin H synthase-2. Western blot showed that microsomes of sheep and rabbit placental cotyledons contained prostaglandin H synthase-2, while prostaglandin H synthase-1 could not be detected. Microsomes of sheep cotyledons metabolised [1-14C]5,8,11-eicosatrienoic acid to many polar metabolites and diclofenac (0.05 mM) inhibited the biosynthesis. The two major metabolites were identified as 13-hydroxy-5,8,11-eicosatrienoic and 11-hydroxy-5,8,12-eicosatrienoic acids. They were formed in a ratio of 3:2, which was not changed by aspirin (2 mM). 5,8,11-Eicosatrienoic acid is likely oxygenated by removal of the pro-S hydrogen at C-13 and insertion of molecular oxygen at either C-13 or C-11, which is followed by reduction of the peroxy derivatives to 13-hydroxy-5,8,11-eicosatrienoic and 11-hydroxy-5,8,12-eicosatrienoic acids, respectively. Prostaglandin H synthase-1 and -2 oxygenate 5,8,11-eicosatrienoic acid only slowly compared with arachidonic acid.  相似文献   

9.
2-Deoxystreptamine (2-DOS) aminoglycoside antibiotics bind specifically to the central region of the 16S rRNA A site and interfere with protein synthesis. Recently, we have shown that the binding of 2-DOS aminoglycosides to an A site model RNA oligonucleotide is linked to the protonation of drug amino groups. Here, we extend these studies to define the number of amino groups involved as well as their identities. Specifically, we use pH-dependent 15N NMR spectroscopy to determine the pK(a) values of the amino groups in neomycin B, paromomycin I, and lividomycin A sulfate, with the resulting pK(a) values ranging from 6.92 to 9.51. For each drug, the 3-amino group was associated with the lowest pK(a), with this value being 6.92 in neomycin B, 7.07 in paromomycin I, and 7.24 in lividomycin A. In addition, we use buffer-dependent isothermal titration calorimetry (ITC) to determine the number of protons linked to the complexation of the three drugs with the A site model RNA oligomer at pH 5.5, 8.8, or 9.0. At pH 5.5, the binding of the three drugs to the host RNA is independent of drug protonation effects. By contrast, at pH 9.0, the RNA binding of paromomycin I and neomycin B is coupled to the uptake of 3.25 and 3.80 protons, respectively, with the RNA binding of lividomycin A at pH 8.8 being coupled to the uptake of 3.25 protons. A comparison of these values with the protonation states of the drugs predicted by our NMR-derived pK(a) values allows us to identify the specific drug amino groups whose protonation is linked to complexation with the host RNA. These determinations reveal that the binding of lividomycin A to the host RNA is coupled to the protonation of all five of its amino groups, with the RNA binding of paromomycin I and neomycin B being linked to the protonation of four and at least five amino groups, respectively. For paromomycin I, the protonation reactions involve the 1-, 3-, 2'-, and 2"'-amino groups, while, for neomycin B, the binding-linked protonation reactions involve at least the 1-, 3-, 2', 6'-, and 2"'-amino groups. Our results clearly identify drug protonation reactions as important thermodynamic participants in the specific binding of 2-DOS aminoglycosides to the A site of 16S rRNA.  相似文献   

10.
A general method for formulating complex thermodynamic systems in terms of hierarchical interactions has been developed, and has been applied in a previous analyses to the theoretical analysis of cooperativity in a dimeric protein, to the statistical analysis of hemoglobin oxygen binding data, and to the protonation equilibria of inorganic polyprotic acids. Organic polyprotic acids have served as a demonstration system for the development of concepts and methods for treating complex biochemical equilibria. Glutamic acid is the classic test case for understanding proton-proton interactions in organic polyprotic acids, and this system is analyzed using the concept of hierarchical interactions. Second order interactions were apparent between all three possible proton interactions, as has been established previously. The third order interaction between the three protons was found to be insignificant, indicating that protonation of one site on glutamate has no effect on the interaction between the other two protonation sites. This further reinforces the premise that higher order terms, representing more complex interactions, are less likely to be significant than lower order terms. To allow correlation of the interaction values from glutamate with other organic acids, pairwise interaction values between protonation events were then calculated from known pKd values for a number of diprotic acids and bases. For simple straight chain acids and bases a linear log-log relationship was apparent between the number of intervening atoms between the protons and the pK(d,hh) (pKd of interaction). This relationship extended from three atoms (carbonate) up to 11 atoms (azelaic acid) and applied to both dicarboxylic acids and diamine bases. The pairwise interactions in glutamate also followed this simple relationship.  相似文献   

11.
The mechanism of excretion into bile of hepatospecific magnetic resonance imaging (MRI) contrast media employed labeled Gd-reagents EOB.DTPA, BOPTA, B 20790 (iopanoate-linked), and B 21690 (glycocholate-linked) for measurement in rat liver canalicular plasma membrane vesicles and yeast vacuoles. The presence of ATP gave threefold greater transport of B 20790 and B 21690 than of EOB.DTPA and BOPTA. In yeast vacuoles the ATP stimulatory effect was eightfold with B 20790 and fivefold greater for B 21690, whereas in YCF1- or YLLO115w-deleted yeast cells the transport was significantly reduced and absent from double mutants, YCF1 and YLLO15w. The transport was similar in wild-type and deletant cells for B 21690; taurocholate gave 85% inhibition. These data suggest that bilary secretion of structurally related MRI agents depend on molecular structure. The findings are suggestive as of possible value for clinical diagnosis of inherited hyperbilirubinemias and other liver disorders.  相似文献   

12.
The method of proton magnetic resonance was used to obtain information on the active site of the guanyl-specific ribonuclease from Penicillium chrysogenum, strain 152A. Four pH-dependent signals in the aromatic region of the proton NMR spectrum of the enzyme were assigned to the C-2 and C-4 protons of the two histidine residues. To determine the pK values and the environment of the histidine residues the pH dependence of their chemical shifts was studied and experimental curves thus obtained were analyzed taking into account the effect of other dissociating groups of the enzyme. The pK values of the histidine residues were found to be equal to 7.92 +/- 0.04 and 7.86 +/- 0.09. The results of the calculations indicate that each histidine residue should interact with an acidic group (carboxylic) of the protein (pK 4.33 and 3.48) and the distance between two histidine residues does not exceed 0.85 nm. The rate constants for the quasi-first order reaction of deuterium exchange of the histidine residues (11.2 s-1 and 3.7 x-1) suggest that both residues are accessible, though to a different degree to solvent. Formation of a complex between the enzyme and guanosine 3'-phosphate (Guo3'P) is accompanied by the shift of the histidine pK toward the alkaline region by 0.5. The existence of the complex is controlled by dissociation of a histidine residue with pK 8.7 in alkaline medium and by protonation of the N-7 of Guo3'P (pK 2.4) in acid medium. Nuclear Overhauser effect measurements were used to determine the glycosidic torsion angle for the Guo3'P in the complex and to estimate the distances between the histidine residues of the enzyme and ribose ring of Guo-3'P. The results obtained suggest that the nucleotide in the complex has an anti conformation and the least exposed histidine is spaced not more than 0.5 nm from the C-1' proton of the nucleotide ribose ring. A model for the enzyme-nucleotide complex is presented.  相似文献   

13.
Atypical antipsychotic properties of 4-(4-fluorobenzylidene)-1-[2-[5-(4-fluorophenyl)-1H-pyrazol-4-yl]ethyl] piperidine (NRA0161) were investigated by in vitro receptor affinities, in vivo receptor occupancies and findings were compared with those of risperidone and haloperidol in rodent behavioral studies. In in vitro receptor binding studies, NRA0161 has a high affinity for human cloned dopamine D(4) and 5-HT(2A) receptor with Ki values of 1.00 and 2.52 nM, respectively. NRA0161 had a relatively high affinity for the alpha(1) adrenoceptor (Ki; 10.44 nM) and a low affinity for the dopamine D(2) receptor (Ki; 95.80 nM). In in vivo receptor binding studies, NRA0161 highly occupied the 5-HT(2A) receptor in rat frontal cortex. In contrast, NRA0161 did not occupy the striatal D(2) receptor. In behavioral studies, NRA0161, risperidone and haloperidol antagonized the locomotor hyperactivity in mice, as induced by methamphetamine (MAP). At a higher dosage, NRA0161, risperidone and haloperidol dose-dependently antagonized the MAP-induced stereotyped behavior in mice and NRA0161 dose-dependently and significantly induced catalepsy in rats. The ED(50) value in inhibiting the MAP-induced locomotor hyperactivity was 30 times lower than that inhibiting the MAP-induced stereotyped behavior and 50 times lower than that which induced catalepsy.These findings suggest that NRA0161 may have atypical antipsychotic activities yet without producing extrapyramidal side effects.  相似文献   

14.
Cyclic peptide disulfides of the general formula H-Cys-(Gly)n-Cys-OH (n = 0-4) were synthesized from the corresponding peptide derivatives [Boc-Cys(Trt)(Gly)-n-Cys(Trt)-OBut] by oxidation with iodine in methanol and by subsequent removal of the terminal groups with trifluoroacetic acid. Acid ionization constants of the obtained peptides were determined by potentiometric titration in aqueous KCl (0.1 mol/L) medium. All compounds have two dissociable hydrogens, corresponding to carboxyl (pK1 = 2.35-2.84) and to terminal amino group (pK2 = 5.61-6.93); pK1 values show first an upward and then a downward trend with the increase in ring size; the opposite is true for pK2 values. These trends could be tentatively attributed to the intramolecular salt bridge (-COO- ----NH+3-) formation.  相似文献   

15.
The equilibrium behavior of cationic iron(III) meso-tetrakis(4-N-methyl-pyridiniumyl) porphyrin, Fe(III)TMPyP, in aqueous solution was studied as a function of pH by optical absorption, EPR and (1)H NMR spectroscopies. The presence of several Fe(III)TMPyP species in solution was unequivocally demonstrated: monomeric porphyrin species (a monoaqueous five-coordinated complex, a diaaqueous six-coordinated complex and a monoaqueous-hydroxo six-coordinated complex), a micro-oxo dimer and a bis-hydroxo complex. The addition of salt to the porphyrin solution leads to a simplification of the equilibrium as a function of pH. In this case, only three species were observed in solution: a monomeric porphyrin species, a micro-oxo dimer and a bis-hydroxo complex. Optical absorption, EPR and (1)H NMR spectra contributed to the characterization of these species. Four critical pH values (pK) for Fe(III)TMPyP were obtained in pure buffer and only three pK values were observed in the presence of NaCl. The addition of salt favors the presence of the dimeric species in solution and simplifies the equilibrium in the acidic pH range.  相似文献   

16.
The dark incubation at room temperature of photosystem II (PS II) membrane fragments in a chloride-free medium at pH 6.3 slowly leads to large chloride-restorable and non-restorable O2 evolution activity losses with time as compared with control samples incubated in the presence of 10 mM NaCl. The chloride requirement in O2 evolution generated under these conditions reveals a complex interplay among various experimental parameters, including the source of the plant material, the times of incubation, the sample concentration, the chloride concentration, as well as those treatments which are believed to specifically displace chloride from PS II such as alkaline pH pretreatment and Na2SO4 addition. The results indicate that secondary, structural changes within the PS II complex are an important factor in determining the influence of chloride on the O2 evolution activity and raise the question whether or not chloride ions actually play a direct cofactor role in the water-oxidizing reactions leading to O2 evolution.Abbreviations Chl chlorophyll - EPR electron paramagnetic resonance - MES 2-(N-morpholino) ethanesulfonic acid - NMR nuclear magnetic resonance - PS II photosystem II  相似文献   

17.
Deprotonation of D-mannitol was studied in aqueous basic solutions by means of potentiometry and (13)C NMR spectroscopy. Two-step dissociation in the pH range from 12 to 13.8 was shown, and successive dissociation constants K(a1) and K(a2) were determined. In a solution with ionic strength I = 1.0 M (NaOH + NaNO(3)) pK(a1) = 13.1 +/- 0.1 and pK(a2) = 13.8 +/- 0.2. With increasing ionic strength from 0.75 to 3.0 M, both pK(a1) and pK(a2) values decrease. Deprotonation-induced chemical shifts in pH-variable (13)C NMR spectra show that the OH-groups next to internal carbon atoms C-3 and C-4 dissociate to a greater extent compared to OH-groups next to external carbon atoms C-1 and C-6.  相似文献   

18.
An immobilized system was developed to detect interactions of human cytochromes P450 (P450) with the accessory proteins NADPH-P450 reductase and cytochrome b(5) (b(5)) using an enzyme-linked affinity approach. Purified enzymes were first bound to wells of a polystyrene plate, and biotinylated partner enzymes were added and bound. A streptavidin-peroxidase complex was added, and protein-protein binding was monitored by measuring peroxidase activity of the bound biotinylated proteins. In a model study, we examined protein-protein interactions of Pseudomonas putida putidaredoxin (Pdx) and putidaredoxin reductase (PdR). A linear relationship (r(2)=0.96) was observed for binding of PdR-biotin to immobilized Pdx compared with binding of Pdx-biotin to immobilized PdR (the estimated K(d) value for the Pdx.PdR complex was 0.054muM). Human P450 2A6 interacted strongly with NADPH-P450 reductase; the K(d) values (with the reductase) ranged between 0.005 and 0.1muM for P450s 2C19, 2D6, and 3A4. Relatively weak interaction was found between holo-b(5) or apo-b(5) (devoid of heme) with NADPH-P450 reductase. Among the rat, rabbit, and human P450 1A2 enzymes, the rat enzyme showed the tightest interaction with b(5), although no increases in 7-ethoxyresorufin O-deethylation activities were observed with any of the P450 1A2 enzymes. Human P450s 2A6, 2D6, 2E1, and 3A4 interacted well with b(5), with P450 3A4 yielding the lowest K(d) values followed by P450s 2A6 and 2D6. No appreciable increases in interaction between human P450s with b(5) or NADPH-P450 reductase were observed when typical substrates for the P450s were included. We also found that NADPH-P450 reductase did not cause changes in the P450.substrate K(d) values estimated from substrate-induced UV-visible spectral changes with rabbit P450 1A2 or human P450 2A6, 2D6, or 3A4. Collectively, the results show direct and tight interactions between P450 enzymes and the accessory proteins NADPH-P450 reductase and b(5), with different affinities, and that ligand binding to mammalian P450s did not lead to increased interaction between P450s and the reductase.  相似文献   

19.
Mead (5,8,11-icosatrienoic) acid was found to be metabolized by the cyclooxygenase enzyme system of ram seminal vesicle microsomes in a calcium-dependent manner. Although the enzyme converted Mead acid to products more slowly and less completely than the isomeric 8,11,14-icosatrienoic acid, both oxidations were inhibitable by indomethacin. Experiments using purified cyclooxygenase confirmed the participation of this enzyme system in the calcium-dependent oxidation. The products of the oxidation were separated by high performance liquid chromatography and analyzed by ultraviolet and gas chromatography-mass spectrometry. The spectra obtained were consistent with the products having the structures 13-hydroxy-5,8,11-icosatrienoate (the major product), 11-hydroxy-5,8,12-icosatrienoate, 9-hydroxy-5,7,11-icosatrienoate, and two isomeric 8,11-dihydroxy-5,9,12-icosatrienoates. No prostaglandin-like, cyclized products could be identified. This report is only the second to illustrate a calcium-dependent oxidation of a polyunsaturated fatty acid by a cyclooxygenase enzyme system and further extends the metabolic potential of Mead acid.  相似文献   

20.
The electron-transfer reactions and thermodynamic equilibria involving the quinone acceptor complex in bacterial reaction centers from R. sphaeroides were investigated. The reactions are described by the scheme: (Formula: see text). We found that the charge recombination pathway of D+QAQ(-)B proceeds via the intermediate state D+Q(-)AQB, the direct pathway contributing less than approx. 5% to the observed recombination rate. The method used to obtain this result was based on a comparison of the kinetics predicted for the indirect pathway (given by the product kAD-times the fraction of reaction centers in the Q-AQB state) with the observed recombination rate, kobsD+----D. The kinetic measurements were used to obtain the pH dependence (6.1 smaller than or equal to pH smaller than or equal to 11.7) of the free energy difference between the states Q(-)AQB and QAQ(-)B. At low pH (less than 9) QAQ(-)B is stabilized relative to Q(-)AQB by 67 meV, whereas at high pH Q(-)AQB is energetically favored. Both Q(-)A and Q(-)B associate with a proton, with pK values of 9.8 and 11.3, respectively. The stronger interaction of the proton with Q(-)B provides the driving force for the forward electron transfer.  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