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.
Summary A new computational method for simultaneously estimating all the proton-proton coupling constants in a molecule from COSY spectra [Yang, J.-X. and Havel, T.F. (1994) J. Biomol. NMR, 4, 807–826] is applied to experimental data from two polypeptides. The first of these is a cyclic hexapeptide denoted as VDA (-d-Ala1-Phe2-Trp3-Lys(Z)4-Val5-Phe6-), in deuterated DMSO, while the second is a 39-residue protein, called decorsin, in aqueous solution. The effect of different data processing strategies and different initial parameter values on the accuracy of the coupling constants was explored. In the case of VDA, most of the coupling constants did not depend strongly on the initial values chosen for the optimization or on how the data were processed. This, together with our previous experience using simulated data, implies strongly that these values are accurate estimates of the coupling constants. They also differ by an average of only 0.36 Hz from the values of the 14 coupling constants that could be measured independently by established methods. In the case of decorsin, many of the coupling constants exhibited a moderate dependence on their initial values and a strong dependence on how the data were processed. With the most successful data processing strategy, the amide- coupling constants differed by an average of 1.11 Hz from the 21 values that could be measured by established methods, while two thirds of the three-bond coupling constants fell within 1.0 Hz of the ranges obtained by applying the Karplus relation to an independently computed ensemble of distance geometry structures. The averages of the coupling constants over multiple optimizations using random initial values were computed in order to obtain the best possible estimates of the coupling constants. Most clearly incorrect averages can be identified by large standard deviations in the coupling constants or the associated line widths and chemical shifts, and can be explained by strong coupling and/or overlap with the water signal, the diagonal peaks or other cross peaks.  相似文献   

2.
《BBA》2006,1757(9-10):1133-1143
In cytochrome c oxidase, oxido-reductions of heme a/CuA and heme a3/CuB are cooperatively linked to proton transfer at acid/base groups in the enzyme. H+/e cooperative linkage at Fea3/CuB is envisaged to be involved in proton pump mechanisms confined to the binuclear center. Models have also been proposed which involve a role in proton pumping of cooperative H+/e linkage at heme a (and CuA). Observations will be presented on: (i) proton consumption in the reduction of molecular oxygen to H2O in soluble bovine heart cytochrome c oxidase; (ii) proton release/uptake associated with anaerobic oxidation/reduction of heme a/CuA and heme a3/CuB in the soluble oxidase; (iii) H+ release in the external phase (i.e. H+ pumping) associated with the oxidative (R  O transition), reductive (O  R transition) and a full catalytic cycle (R  O  R transition) of membrane-reconstituted cytochrome c oxidase. A model is presented in which cooperative H+/e linkage at heme a/CuA and heme a3/CuB with acid/base clusters, C1 and C2 respectively, and protonmotive steps of the reduction of O2 to water are involved in proton pumping.  相似文献   

3.
T3 administration increases the extent of non-linearity in the flow-force relationship between pump proton conductance and protonmotive force. The effect is present also at the ATPase proton pump. These effects are not accompanied by changes in passive proton conductance. Incubation of mitochondria at 45 degrees C also causes an increased non-linearity, accompanied by a partial increase of proton conductance. It appears that the increase of respiratory activity following T3 administration is due to loss of thermodynamic control within or at the proton pumps, an effect which might be attributed to increased slipping.  相似文献   

4.
Neutron diffraction is used to localize water molecules and/or exchangeable hydrogen ions in the purple membrane by H2O/2H2O exchange experiments at different values of relative humidity. At 100% relative humidity, differences in the hydration between protein and lipid areas are observed, accounting for an excess amount of about 100 molecules of water in the lipid domains per unit cell. A pronounced isotope effect was observed, reproducibly showing an increase in the lamellar spacing from 60 A in 2H2O to 68 A in H2O. At 15% relative humidity, the positions of exchangeable protons became visible. A dominant difference density peak corresponding to 11 +/- 2 exchangeable protons was detected in the central part of the projected structure of bacteriorhodopsin at the Schiff's base end of the chromophore. A difference density map obtained from data on purple membrane films at 15% relative humidity in 2H2O, and the same sample after complete drying in vacuum, revealed that about eight of these protons belong to four water molecules. This is direct evidence for tightly bound water molecules close to the chromophore binding site of bacteriorhodopsin, which could participate in the active steps of H+ translocation as well as in the proton pathway across this membrane protein.  相似文献   

5.
The Fo membrane domain of FoF1-ATPase complex had been purified from porcine heart mitochondria. SDS-PAGE with silver staining indicated that the purity of Fo was about 85% and the sample contained no subunits of F1-ATPase. The purified Fo was reconstituted into liposomes with different phospholipid composition, and the effect of CL (cardiolipin), PA (phosphatidic acid), PI (phosphatidylinositol) and PS (phosphatidylserine) on the H+ translocation activity of Fo was investigated. The results demonstrated that CL, PA and PI could promote the proton translocation of Fo with the order of CL>PA>>PI, while PS inhibited it. Meanwhile ADM (adriamycin) severely impaired the proton translocation activity of Fo vesicles containing CL, which suggested that CL's stimulation of the activity of reconstituted Fo might correlate with its non-bilayer propensity. After Fo was incorporated into the liposomes containing PE (phosphatidylethanolamine), DOPE (dioleoylphosphatidylethanolamine) as well as DEPE (dielaidoylphospha  相似文献   

6.
AimThe feasibility of using 230 MeV proton cyclotrons in proton therapy centers as a spallation neutron source for Boron Neutron Capture Therapy (BNCT) was investigated.BackgroundBNCT is based on the neutron irradiation of a 10B-containing compound located selectively in tumor cells. Among various types of neutron generators, the spallation neutron source is a unique way to generate high-energy and high-flux neutrons.Materials and MethodsNeutron beam was generated by a proton accelerator via spallation reactions and then the produced neutron beam was shaped to be appropriate for BNCT. The proposed Beam Shaping Assembly (BSA) consists of different moderators, a reflector, a collimator, as well as thermal and gamma filters. In addition, the simulated Snyder head phantom was utilized to evaluate the dose distribution in tumor and normal tissue due to the irradiation by the designed beam. MCNPX2.6 Monte Carlo code was used to optimize BSA as well as evaluate dose evaluation.ResultsA BSA was designed. With the BSA configuration and a beam current of 104 nA, epithermal neutron flux of 3.94 × 106 [n/cm2] can be achieved, which is very low. Provided that we use the beam current of 5.75 μA, epithermal neutron flux of 2.18 × 108 [n/cm2] can be obtained and the maximum dose of 38.2 Gy-eq can be delivered to tumor tissue at 1.4 cm from the phantom surface.ConclusionsResults for 230 MeV protons show that with proposed BSA, proton beam current about 5.75 μA is required for this purpose.  相似文献   

7.
In an effort to develop DS02, a new radiation dosimetry system for the atomic bomb survivors of Hiroshima and Nagasaki, measurements of neutron-induced activities have provided valuable information to reconstruct the radiation situation at the time of the bombings. In Hiroshima, the depth profile of (152)Eu activity measured in a granite pillar of the Motoyasu Bridge (128 m from the hypocenter) was compared with that calculated using the DS02 methodology. For calculation of the (152)Eu production due to the thermal-neutron activation reaction, (151)Eu(n,gamma)(152)Eu, information on the hydrogen content in granite is important because the transport and slowing-down process of neutrons penetrating into the pillar is strongly affected by collisions with the protons of hydrogen. In this study, proton-proton elastic recoil coincidence spectrometry has been used to deduce the proton density in the Motoyasu pillar granite. Slices of granite samples were irradiated by a 20 MeV proton beam, and the energies of scattered and recoil protons were measured with a coincidence method. The water concentration in the pillar granite was evaluated to be 0.30 +/- 0.07%wt. This result is consistent with earlier data on adsorptive water (II) and bound water obtained by the Karl Fisher method.  相似文献   

8.
High-resolution proton and carbon-13 NMR of membranes: why sonicate?   总被引:2,自引:0,他引:2  
E Oldfield  J L Bowers  J Forbes 《Biochemistry》1987,26(22):6919-6923
We have obtained high-field (11.7-T) proton and carbon-13 Fourier transform (FT) nuclear magnetic resonance (NMR) spectra of egg lecithin and egg lecithin-cholesterol (1:1) multibilayers, using "magic-angle" sample spinning (MASS) techniques, and sonicated egg lecithin and egg lecithin-cholesterol (1:1) vesicles, using conventional FT NMR methods. Resolution of the proton and carbon-13 MASS NMR spectra of the pure egg lecithin samples is essentially identical with that of sonicated samples, but spectra of the unsonicated lipid, using MASS, can be obtained very much faster than with the more dilute, sonicated systems. With the 1:1 lecithin-cholesterol systems, proton MASS NMR spectra are virtually identical with conventional FT spectra of sonicated samples, while with 13C NMR, we demonstrate that most 13C nuclei in the cholesterol moiety can be monitored, even though these same nuclei are essentially invisible, i.e., are severely broadened, in the corresponding sonicated systems. In addition, 13C MASS NMR, spectra can again be recorded much faster than with sonicated samples, due to concentration effects. Taken together, these results strongly suggest there will seldom be need in the future to resort to ultrasonic disruption of lipid bilayer membranes in order to obtain high-resolution proton or carbon-13 NMR spectra.  相似文献   

9.
Arginine residues are generally considered poor candidates for the role of general bases because they are predominantly protonated at physiological pH. Nonetheless, Arg residues have recently emerged as general bases in several enzymes: IMP dehydrogenase, pectate/pectin lyases, fumarate reductase, and l-aspartate oxidase. The experimental evidence suggesting this mechanistic function is reviewed. Although these enzymes have several different folds and distinct evolutionary origins, a common structural motif is found where the critical Arg residue is solvent accessible and adjacent to carboxylate groups. The chemistry of the guanidine group suggests unique strategies to lower the pK(a) of Arg. Lastly, the presumption that general bases must be predominantly deprotonated is revisited.  相似文献   

10.
The variation of the chemical shift of the protons of 5′-UMP and 5′-AMP is calculated as a function of χCN, ψ and ? torsion angles. The shift of H8 of 5′-AMP and H6 of 5′-UMP is found to be very sensitive to the value of χCN. For the anti conformations the shift of these protons is more sensitive to the value of the rotation about CS′-05′ than about C4′-CS′. For the protons of the ribose the calculations show that for the C2′-endo pucker H3′ and H2′ undergo the largest chemical shift variations when ? and ψ vary. The calculated variations are considered in relation with the role of the conformation of the nucleotides in the chemical shift variation between mono and polynucleotides and between the different helical structures of polynucleotides.  相似文献   

11.
1. The -->H(+)/e(-) quotients for proton release from mitochondria associated with electron flow from succinate and duroquinol to O(2), ferricyanide or ferricytochrome c, and from NNN'N'-tetramethyl-p-phenylenediamine+ascorbate to O(2), were determined from rate measurements of electron flow and proton translocation. 2. Care was taken to avoid, or to take into account, unrelated electron flow and proton translocation, which might take place in addition to the oxido-reductions that were the subject of our analysis. Spectrophotometric techniques were chosen to provide accurate measurement of the rate of consumption of oxidants and reductants. The rate of proton translocation was measured with fast pH meters with a precision of 10(-3) pH unit. 3. The -->H(+)/O quotient for succinate or duroquinol oxidation was, at neutral pH, 4, when computed on the basis of spectrophotometric determinations of the rate of O(2) consumption or duroquinol oxidation. Higher -->H(+)/O quotients for succinate oxidation, obtained from polarographic measurements of O(2) consumption, resulted from underestimation of the respiratory rate. 4. The -->H(+)/2e(-) quotient for electron flow from succinate and duroquinol to ferricyanide or ferricytochrome c ranged from 3.9 to 3.6. 5. Respiration elicited by NNN'N'-tetramethyl-p-phenylenediamine+ascorbate by antimycin-inhibited mitochondria resulted in extra proton release in addition to that produced for oxidation of ascorbate to dehydroascorbate. Accurate spectrophotometric measurement of respiration showed that the -->H(+)/e(-) ratio was only 0.25 and not 0.7-1.0 as obtained with the inadequate polarographic assay of respiration. Proton release was practically suppressed when mitochondria were preincubated aerobically in the absence of antimycin. Furthermore, the rate of scalar proton consumption for water production was lower than that expected from the stoicheiometry. Thus the extra proton release observed during respiration elicited by NNN'N'-tetramethyl-p-phenylenediamine+ascorbate is caused by oxidation of endogenous hydrogenated reductants. 6. It is concluded that (i) the -->H(+)/O quotient for the cytochrome system is, at neutral pH, 4 and not 6 or 8 as reported by others; (ii) all the four protons are released during electron flow from quinol to cytochrome c; (iii) the oxidase transfers electrons from cytochrome c to protons from the matrix aqueous phase and does not pump protons from the matrix to the outer aqueous phase.  相似文献   

12.
13.
Light-driven photosynthetic electron transport is coupled to the movement of protons from the chloroplast stroma to the thylakoid lumen. The resulting proton motive force that is generated is used to drive the conformational rotation of the transmembrane thylakoid ATPase enzyme which converts ADP (adenosine diphosphate) and Pi (inorganic phosphate) into ATP (adenosine triphosphate), the energy currency of the plant cell required for carbon fixation and other metabolic processes. According to Mitchell’s chemiosmotic hypothesis, the proton motive force can be parsed into the transmembrane proton gradient (ΔpH) and the electric field gradient (Δψ), which are thermodynamically equivalent. In chloroplasts, the proton motive force has been suggested to be split almost equally between Δψ and ΔpH (Kramer et al., Photosynth Res 60:151–163, 1999). One of the central pieces of evidence for this theory is the existence of a steady-state electrochromic shift (ECS) absorption signal detected ~515 nm in plant leaves during illumination. The interpretation of this signal is complicated, however, by a heavily overlapping absorption change ~535 nm associated with the formation of photoprotective energy dissipation (qE) during illumination. In this study, we present new evidence that dissects the overlapping contributions of the ECS and qE-related absorption changes in wild-type Arabidopsis leaves using specific inhibitors of the ΔpH (nigericin) and Δψ (valinomycin) and separately using leaves of the Arabidopsis lut2npq1 mutant that lacks qE. In both cases, our data show that no steady-state ECS signal persists in the light longer than ~60 s. The consequences of our observations for the suggesting parsing of steady-state thylakoid proton motive force between (ΔpH) and the electric field gradient (Δψ) are discussed.  相似文献   

14.
The four isomers of methyl O-β-d-galactopyranosyl-β-d-galactopyranoside were prepared by condensation of 2,3,4,6-tetra-O-acetyl-α-d-galactopyranosyl bromide with appropriate, partially O-substituted derivatives of methyl β-d-galactopyranoside. Reaction of 3,4,6-tri-O-acetyl-1,2-O-(1-ethoxyethylidene)-α-d-galactopyranose with the same acceptors, in the presence of mercuric bromide, led to the formation of α and β linkages. Thus, it was possible to assign 13C-n.m.r. resonances of α and β anomers of methyl O-d-galactopyranosyl-β-d-galactopyranosides. In terms of application of these shift values and those of related d-galactobioses to the structural analysis of d-galactopyranans by shift comparisons, some generalizations can be made. For β-d-galactopyranans, the resonances of glycosyloxylated carbon atoms of methyl O-β-d-galactopyranosyl-β-d-galactopyranosides are sensitive to structure and appear to have typical values, whereas limited variation was observed with shifts of C-1′ signals. On the other hand, for assigning structures to d-galactopyranans containing α linkages, the C-1′ shifts (at higher field) of methyl O-α-d-galactopyranosyl-β-d-galactopyranosides are sensitive to linkage position, whereas those of glycosyloxylated carbon atoms vary only a little.  相似文献   

15.
16.
The nature of the control of water/proton selectivity in biological channels is a problem of a fundamental importance. Most studies of this issue have proposed that an interference with the orientational requirements of the so-called proton wire is the source of selectivity. The elucidation of the structures of aquaporins, which have evolved to prevent proton transfer (PT), provided a clear benchmark for exploring the selectivity problem. Previous simulations of this system have not examined, however, the actual issue of PT, but only considered the much simpler task of the transfer of water molecules. Here we take aquaporin as a benchmark and quantify the origin of the water/proton selectivity in this and related systems. This is done by evaluating in a consistent way the free energy profile for transferring a proton along the channel and relating this profile to the relevant PT rate constants. It is found that the water/proton selectivity is controlled by the change in solvation free energy upon moving the charged proton from water to the channel. The reason for the focus on the elegant concept of the proton wire and the related Grotthuss-type mechanism is also considered. It is concluded that these mechanisms are clearly important in cases with flat free energy surfaces (e.g., in bulk water, in gas phase water chains, and in infinitely long channels). However, in cases of biological channels, the actual PT mechanism is much less important than the energetics of transferring the proton charge from water to different regions in the channels.  相似文献   

17.
From pollen grains of Typha davidiana, T. latifolia, T. angustata the same eight flavonoids have been isolated. They are identified as naringenin I, isorhamnetin II, quercetin III, isorhamnetin-3-O-(2G-α-L-rhamnopyranosyl)-rutioside IV, quercetin-3-O-(2G-α-L-rhamnopyranosyl)-rutinosida, V, isorhamnetio-3-O-rutinoside VI, isorhamnetino-3-O-neohesperidoside VII, kampferol-3-O-neohesperidoside VIII. Flavonoids of pollen grains of five species of Typha, including the above three species, were analysed by TLC with the result showing that the constituents in the pollen grains of the five species are very similar. The chemical comparison among Typha and Sparganium and 16 possibly related families shows that Typha is different from Pandanaceae or Pandanales and is similar to Restionaceae, Flagellariaceae, Juncaceae and Cyperaceae in some respects. Typha and Sparganium are very similar in many respects, and they could be treated in the same family, Typhaceae, which merit the rank of order, Typhales.  相似文献   

18.
Most bacterial genomes have five to nine distinct genes predicted to encode transporters that exchange cytoplasmic Na+ and/or K+ for H+ from outside the cell, i.e. monovalent cation/proton antiporters. By contrast, pathogens that live primarily inside host cells usually possess zero to one such antiporter while other stress-exposed bacteria exhibit even higher numbers. The monovalent cation/proton antiporters encoded by these diverse genes fall into at least eight different transporter protein families based on sequence similarity. They enable bacteria to meet challenges of high or fluctuating pH, salt, temperature or osmolarity, but we lack explanations for why so many antiporters are needed and for the value added by specific antiporter types in specific settings. In this issue of Molecular Microbiology, analyses of the pH dependence of cytoplasmic [Na+], [K+], pH and transmembrane electrical potential in the 'poly extremophile' Natranaerobius thermophilus are the context for assessment of the catalytic properties of 12 predicted monovalent cation/proton antiporters in the genome of this thermophilic haloalkaliphile. The results provide a profile of adaptations of the poly extremophilic anaerobe, including a proposed role of cytoplasmic buffering capacity. They also provide new perspectives on two large monovalent cation/proton antiporter families, the NhaC and the cation/proton antiporter-3 antiporter families.  相似文献   

19.
Nutrient deficiencies are an ongoing problem in many populations and ascorbic acid is a key vitamin whose mild or acute absence leads to a number of conditions including the famously debilitating scurvy. As such, the biochemical effects of ascorbate deficiency merit ongoing scrutiny, and the Gulo knockout mouse provides a useful model for the metabolomic examination of vitamin C deficiency. Like humans, these animals are incapable of synthesizing ascorbic acid but with dietary supplements are otherwise healthy and grow normally. In this study, all vitamin C sources were removed after weaning from the diet of Gulo-/- mice (n?=?7) and wild type controls (n?=?7) for 12?weeks before collection of serum. A replicate study was performed with similar parameters but animals were harvested pre-symptomatically after 2-3?weeks. The serum concentration of 50 metabolites was determined by quantitative profiling of 1D proton NMR spectra. Multivariate statistical models were used to describe metabolic changes as compared to control animals; replicate study animals were used for external validation of the resulting models. The results of the study highlight the metabolites and pathways known to require ascorbate for proper flux.  相似文献   

20.
Sodium–proton antiporters rapidly exchange protons and sodium ions across the membrane to regulate intracellular pH, cell volume, and sodium concentration. How ion binding and release is coupled to the conformational changes associated with transport is not clear. Here, we report a crystal form of the prototypical sodium–proton antiporter NhaA from Escherichia coli in which the protein is seen as a dimer. In this new structure, we observe a salt bridge between an essential aspartic acid (Asp163) and a conserved lysine (Lys300). An equivalent salt bridge is present in the homologous transporter NapA, but not in the only other known crystal structure of NhaA, which provides the foundation of most existing structural models of electrogenic sodium–proton antiport. Molecular dynamics simulations show that the stability of the salt bridge is weakened by sodium ions binding to Asp164 and the neighboring Asp163. This suggests that the transport mechanism involves Asp163 switching between forming a salt bridge with Lys300 and interacting with the sodium ion. pKa calculations suggest that Asp163 is highly unlikely to be protonated when involved in the salt bridge. As it has been previously suggested that Asp163 is one of the two residues through which proton transport occurs, these results have clear implications to the current mechanistic models of sodium–proton antiport in NhaA.  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