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.
 The ligand DOTASA was designed and synthesized in the aim of obtaining a kinetically and thermodynamically stable Gd(III) chelate which, through its uncoordinated carboxylate function, will provide an efficient pathway to couple the complex to bio- or macromolecules without affecting the coordination pattern of DOTA. Furthermore, it allows us to study the influence of an extra carboxylate arm on the parameters determining proton relaxivity in comparison to the commercial agent [Gd(DOTA)(H2O)]. A combined variable-temperature 17O NMR, EPR and nuclear magnetic relaxation dispersion study on the Gd(III) chelate resulted in k 298 ex=(6.3±0.2)×106 s–1 for the water exchange rate and τ298 R=125±2 ps for the rotational correlation time. The slight increase in both k 298 ex and τ298 R, as compared to those for [Gd(DOTA)(H2O)], is attributed to the presence of the extra negative charge. The longer rotational correlation time results in a proton relaxivity of 5.03 mM–1 s–1 for [Gd(DOTASA)(H2O)]2–, which is approximately 30% higher than that for [Gd(DOTA)(H2O)]. The increased water exchange rate of [Gd(DOTASA)(H2O)]2– has no consequence for proton relaxivity since this latter is exclusively limited by fast rotation for both complexes. However, for slowly rotating macromolecular agents, which contain a covalently coupled DOTASA unit instead of a coupled DOTA, this increased exchange rate will have a significant positive effect. Received: 31 December 1998 / Accepted: 4 March 1999  相似文献   

2.
This report describes a new method to measure the chloroplastic lumenal proton pool in leaves (tobacco and sunflower). The method is based on measurement of CO2 outbursts from leaves caused by the shift in the CO2 + H2O ↔ HCO3  + H+ equilibrium in the chloroplast stroma as protons return from the lumen after darkening. Protons did not accumulate in the lumen to a significant extent when photosynthesis was light-limited, but a large pool of >100 μmol H+ m−2 accumulated in the lumen as photosynthesis became light-saturated. During thylakoid energization in the light, large amounts of protons are moved from binding sites in the stroma to binding sites in the lumen. The transthylakoidal difference in the chemical potential of free protons (ΔpH) is largely based on the difference in the chemical potential of bound protons in the lumenal and stromal compartments (pK). Over the course of the dark-light induction of photosynthesis protons accumulate in the lumen during reduction of 3-phosphoglycerate. The accumulation of electrons in reduced compounds of the stroma and cytosol is the natural cause for accumulation of a stoichiometric pool of lumenal protons during this transient event.  相似文献   

3.
Depolarization-activated H+-selective currents were studied using whole-cell and excised-patch voltage clamp methods in human monocytic leukemia THP-1 cells, before and after being induced by phorbol ester to differentiate into macrophage-like cells. The H+ conductance, g H, activated slowly during depolarizing pulses, with a sigmoidal time course. Fitted by a single exponential following a delay, the activation time constant, τact was roughly 10 sec at threshold potentials, decreasing at more positive potentials. Tail currents upon repolarization decayed mono-exponentially at all potentials. The tail current time constant, τtail, was voltage dependent, decreasing with hyperpolarization from 2–3 sec at 0 mV to ∼200 msec at −100 mV. Surprisingly, although τact depended strongly on pH o , τtail was completely independent of pH o . H+ currents were inhibited by Zn2+. Increasing pH o or decreasing pH i shifted the voltage-activation relationship to more negative potentials, tending to activate the g H at any given voltage. Studied in excised, inside-out membrane patches, H+ currents were larger and activated much more rapidly at lower bath pH (i.e., pH i ). In THP-1 cells differentiated into macrophages, the H+ current density was reduced by one-half, and τact was slower by about twofold. The properties of H+ channels in THP-1 cells and in other macrophage-related cells are compared. Received: 19 September 1995/Revised: 14 March 1996  相似文献   

4.
Several properties of the exchangeable amide protons of the ganglioside GM2 were studied in detail by1H-NMR spectroscopy in fully deuterated dimethylsulfoxide [2H6]DMSO/2% H2O, and compared with data obtained for the simpler constituent glycosphingolipids GA2 and GM3. In addition to chemical shifts,3 J 2,HN coupling constants, and temperature shift coefficients, the kinetics of NH/2H chemical exchange were examined by following the disappearance of the amide resonances in [2H6]DMSO/2%2H2O. The results included observation of an increase in half-life of theN-acetylgalactosamine acetamido HN by more than an order of magnitude in GM2 compared to GA2, attributable to the presence of the additionalN-acetylneuraminic acid residue. Additional one-dimensional dipolar cross relaxation experiments were also performed on nonexchangeable protons of GM2. The results of all of these experiments support a three-dimensional model for the terminal trisaccharide in which a hydrogen bond is formed between theN-acetylgalactosamine acetamido NH and theN-acetylneuraminic acid carboxyl group. The interaction is proposed to be of the -acceptor type, a possibility which has not yet been explored in the literature on carbohydrates. The proposed model is discussed in comparison with that of Sabesanet al. (1984,Can J Chem 62: 1034–45), and the models of GM1 proposed more recently by Acquottiet al. (1990,J Am Chem Soc 112:7772–8) and Scarsdaleet al. (1990,Biochemistry 29:9843–55).  相似文献   

5.
A new off-resonance spin-lock experiment to record relaxation dispersion profiles of amide protons is presented. The sensitivity-enhanced HSQC-type sequence is designed to minimize the interference from cross-relaxation effects and ensure that the dispersion profiles in the absence of μs-ms time-scale dynamics are flat. Toward this end (i) the proton background is eliminated by sample deuteration (Ishima et al., 1998), (ii) 1H spin lock is applied to two-spin modes 2(HxSin ϑ + HzCos ϑ) Nz, and (iii) the tilt angle ϑ ≈ 35° is maintained throughout the series of measurements (Desvaux et al. Mol. Phys., 86 (1995) 1059). The relaxation dispersion profiles recorded in this manner sample a wide range of effective rf field strengths (up to and in excess of 20 kHz) which makes them particularly suitable for studies of motions on the time scale ≤100 μs. The new experiment has been tested on the Ca2+-loaded regulatory domain of cardiac troponin C. Many residues show pronounced dispersions with remarkably similar correlation times of 30 μs. Furthermore, these residues are localized in the regions that have been previously implicated in conformational changes (Spyracopoulos et al. Biochemistry, 36 (1997) 12138) Electronic supplementary material Electronic supplementary material is available for this article at and accessible for authorised users.  相似文献   

6.
The dynamics of the nucleobase and the ribose moieties in a 14-nt RNA cUUCGg hairpin-loop uniformly labeled with 13C and 15N were studied by 13C spin relaxation experiments. R1, R and the 13C-{1H} steady-state NOE of C6 and C1′ in pyrimidine and C8 and C1′ in purine residues were obtained at 298 K. The relaxation data were analyzed by the model-free formalism to yield dynamic information on timescales of pico-, nano- and milli-seconds. An axially symmetric diffusion tensor with an overall rotational correlation time τc of 2.31±0.13 ns and an axial ratio of 1.35±0.02 were determined. Both findings are in agreement with hydrodynamic calculations. For the nucleobase carbons, the validity of different reported 13C chemical shift anisotropy values (Stueber, D. and Grant, D. M., 2002 J. Am. Chem. Soc. 124, 10539–10551; Fiala et al., 2000 J. Biomol. NMR 16, 291–302; Sitkoff, D. and Case, D. A., 1998 Prog. NMR Spectroscopy 32, 165–190) is discussed. The resulting dynamics are in agreement with the structural features of the cUUCGg motif in that all residues are mostly rigid (0.82 < S2 < 0.96) in both the nucleobase and the ribose moiety except for the nucleobase of U7, which is protruding into solution (S2 = 0.76). In general, ribose mobility follows nucleobase dynamics, but is less pronounced. Nucleobase dynamics resulting from the analysis of 13C relaxation rates were found to be in agreement with 15N relaxation data derived dynamic information (Akke et al., 1997 RNA 3, 702–709). Electronic supplementary material Electronic supplementary material is available for this article at and accessible for authorised users.  相似文献   

7.
The organoruthenium complex [(η6-hmb)Ru(en)(Cl)][PF6] (hmb is hexamethylbenzene, en is ethylenediamine) undergoes facile aquation and then reacts with KSCN in unbuffered solution to give the S-coordinated thiocyanato product [(η6-hmb)Ru(en)(S-SCN)]+ which slowly converts to the thermodynamically favored N-bound complex [(η6-hmb)Ru(en)(N-NCS)]+ (1 +). Complex 1 was synthesized and characterized by X-ray crystallography and mass spectrometry. Despite its lack of hydrolysis over 24 h, complex 1 exhibits moderate cytotoxicity (IC50 24 μM) towards the human ovarian cancer cell line A2780, comparable with that of the chlorido analogue which is thought to be activated (towards potential target DNA) via a rapid aquation (Wang et. al. in Proc Natl Acad Sci USA 102:18269–18274, 2005). Detailed kinetic studies suggest that complex 1 binds to guanosine 5′-monophosphate (GMP) through direct N7 substitution of the N-bound SCN ligand. In the presence of a high concentration of chloride (104 mM), however, complex 1 may bind partly to GMP via Cl substitution.  相似文献   

8.
Superoxide dismutases (SODs) protect cells against oxidative stress by disproportionating O2 to H2O2 and O2. The recent finding of a nickel-containing SOD (Ni-SOD) has widened the diversity of SODs in terms of metal contents and SOD catalytic mechanisms. The coordination and geometrical structure of the metal site and the related electronic structure are the keys to understanding the dismutase mechanism of the enzyme. We performed Q-band 14N,1/2H continuous wave (CW) and pulsed electron–nuclear double resonance (ENDOR) and X-band 14N electron spin echo envelope modulation (ESEEM) on the resting-state Ni-SOD extracted from Streptomyces seoulensis. In-depth analysis of the data obtained from the multifrequency advanced electron paramagnetic resonance techniques detailed the electronic structure of the active site of Ni-SOD. The analysis of the field-dependent Q-band 14N CW ENDOR yielded the nuclear hyperfine and quadrupole coupling tensors of the axial Nδ of the His-1 imidazole ligand. The tensors are coaxial with the g-tensor frame, implying the g-tensor direction is modulated by the imidazole plane. X-band 14N ESEEM characterized the hyperfine coupling of Nε of His-1 imidazole. The nuclear quadrupole coupling constant of the nitrogen suggests that the hydrogen-bonding between Nε–H and OGlu-17 present for the reduced-state Ni-SOD is weakened or broken upon oxidizing the enzyme. Q-band 1H CW ENDOR and pulsed 2H Mims ENDOR showed a strong hyperfine coupling to the protons(s) of the equatorially coordinated His-1 amine and a weak hyperfine coupling to either the proton(s) of a water in the pocket at the side opposite the axial Nδ or the proton of a water hydrogen-bonded to the equatorial thiolate ligand.  相似文献   

9.
The influence of sub-optimal temperatures (T) on the microbial growth rate (μ) has been assessed by means of dimensionless variables: Tdim = [T−Tmin]/[Topt−Tmin] and μdim = μ/μopt. Tmin represents the temperature at which there is no growth, Topt the optimum temperature at which the growth rate, μopt, is maximum. Data sets, growth rate vs temperature, have been taken from the literature for 12 organisms (psychrotrophs, mesophiles and thermophiles). In order to compare these organisms, the power law function has been used: [μdim] = [Tdim]α. The parameters μopt and Topt are determined from direct readings whereas Tmin and αare estimated by means of a non-linear regression. An accurate estimation of Tmin is obtained providing low growth rate data are available. A wide range of optimal temperatures where the growth rate almost equals μopt prevents one from obtaining a narrow confidence interval forα. On the basis of the analysis hereafter developed, thermophiles are characterized by values of the power α less than mesophiles and psychrotrophs. Almost all of these values are significantly different from two, previously determined for Staphylococcus xylosus and widely used for predicting the microbial growth in foods. Received 15 May 1998/ Accepted in revised form 25 September 1998  相似文献   

10.
We present a systematic study of the effect of the level of exchangeable protons on the observed amide proton linewidth obtained in perdeuterated proteins. Decreasing the amount of D2O employed in the crystallization buffer from 90 to 0%, we observe a fourfold increase in linewidth for both 1H and 15N resonances. At the same time, we find a gradual increase in the signal-to-noise ratio (SNR) for 1H–15N correlations in dipolar coupling based experiments for H2O concentrations of up to 40%. Beyond 40%, a significant reduction in SNR is observed. Scalar-coupling based 1H–15N correlation experiments yield a nearly constant SNR for samples prepared with ≤30% H2O. Samples in which more H2O is employed for crystallization show a significantly reduced NMR intensity. Calculation of the SNR by taking into account the reduction in 1H T 1 in samples containing more protons (SNR per unit time), yields a maximum SNR for samples crystallized using 30 and 40% H2O for scalar and dipolar coupling based experiments, respectively. A sensitivity gain of 3.8 is obtained by increasing the H2O concentration from 10 to 40% in the CP based experiment, whereas the linewidth only becomes 1.5 times broader. In general, we find that CP is more favorable compared to INEPT based transfer when the number of possible 1H,1H interactions increases. At low levels of deuteration (≥60% H2O in the crystallization buffer), resonances from rigid residues are broadened beyond detection. All experiments are carried out at MAS frequency of 24 kHz employing perdeuterated samples of the chicken α-spectrin SH3 domain.  相似文献   

11.
Proton translocation by washed cells of the sulfate-reducing bacterium Desulfovibrio desulfuricans strain Essex 6 was studied by means of pH and sulfide electrodes. Reversible extrusion of protons could be induced either by addition of electron acceptors to cells incubated under hydrogen, or by addition of hydrogen to cells incubated in the presence of an appropriate electron acceptor. Proton translocation was increased in the presence of ionophores that dissipate the membrane potential (thiocyanate, methyl triphenylphosphonium cation, but not valinomycin) and was sensitive to the uncoupler carbonylcyanide m-chlorophenylhydrazone (CCCP). Upon micromolar additions of H2, usually sulfide was formed in stoichiometric amounts, and extrapolated H+/H2 ratios were 1.8±0.5 with sulfate, 2.3±0.3 with sulfite and 0.5±0.1 with thiosulfate. In several experiments hydrogen pulses caused increased proton extrusion not associated with sulfide production. This was a hint that sulfite might be reduced via intermediates. In the absence of H2S formation, extrapolated H+/H2 ratios were 3.1±0.8 with sulfate, 3.4±1.1 with sulfite, 4.4±0.8 with thiosulfate and 6.3±1.2 with oxygen. Micromolar pulses of electron acceptors to cells incubated under H2 caused less proton translocation than H2 pulses in presence of excess of electron acceptor; extrapolated H+/H2 ratios were 1.3±0.4 with sulfite, 3.3±0.9 with nitrite and 4.2±0.5 with oxygen. No proton translocation was observed after micromolar pulses of sulfate, thiosulfate or nitrate to cells incubated under hydrogen in the presence of thiocyanate. Inhibition experiments with CO and CuCl2 revealed that the hydrogenase activity was localized in the intracellular space, and that no periplasmic hydrogenase was present. The results indicate that D. desulfuricans can generate a proton gradient by pumping protons across the cytoplasmic membrane.Abbreviations APS adenosine 5-phosphosulfate - CCCP carbonyl cyanide m-chlorophenylhydrazone - MTTP+ methyl triphenylphosphonium cation  相似文献   

12.
Summary Proton chemical shifts of a series of disordered linear peptides (H-Gly-Gly-X-Gly-Gly-OH, with X being one of the 20 naturally occurring amino acids) have been obtained using 1D and 2D 1H NMR at pH 5.0 as a function of temperature and solvent composition. The use of 2D methods has allowed some ambiguities in side-chain assignments in previous studies to be resolved. An additional benefit of the temperature data is that they can be used to obtain ‘random coil’ amide proton chemical shifts at any temperature between 278 and 318 K by interpolation. Changes of chemical shift as a function of trifluoroethanol concentration have also been determined at a variety of temperatures for a subset of peptides. Significant changes are found in backbone and side-chain amide proton chemical shifts in these ‘random coil’ peptides with increasing amounts of trifluoroethanol, suggesting that caution is required when interpreting chemical shift changes as a measure of helix formation in peptides in the presence of this solvent. Comparison of the proton chemical shifts obtained here for H-Gly-Gly-X-Gly-Gly-OH with those for H-Gly-Gly-X-Ala-OH [Bundi, A. and Wüthrich, K. (1979) Biopolymers, 18, 285–297] and for Ac-Gly-Gly-X-Ala-Gly-Gly-NH2 [Wishart, D.S., Bigam, C.G., Holm, A., Hodges, R.S. and Sykes, B.D. (1995) J. Biomol. NMR, 5, 67–81] generally shows good agreement for CH protons, but reveals significant variability for NH protons. Amide proton chemical shifts appear to be highly sensitive to local sequence variations and probably also to solution conditions. Caution must therefore be exercised in any structural interpretation based on amide proton chemical shifts.  相似文献   

13.
The interaction of the diastereomeric complexes Λ-[Ru(bpy)2(m-GHK)]Cl2 and Δ-[Ru(bpy)2(m-GHK)]Cl2 (bpy is 2,2′-bipyridine, GHK is glycine–l-histidine–l-lysine) with the deoxynucleotide duplex d(5′-CGCGATCGCG)2 was studied by means of 1H NMR spectroscopy. At a Δ-isomer to DNA ratio of 1:1, significant shifts for the metal complex are observed, whereas there is negligible effect on the oligonucleotide protons and only one intermolecular nuclear Overhauser effect (NOE) is present at the 2D nuclear Overhauser enhancement spectroscopy spectrum. The 1Η NMR spectrum at ratio 2:1 is characterized by a slight shift for the Δ-isomer’s bpy aromatic protons as well as significant shifts for the decanucleotide G4 H1′ and Η2″, A5 H2, G10 H1′, T6 NH and G2 NH protons. Furthermore, at ratio 2:1, 11 intermolecular NOEs are observed. The majority of the NOEs involve the sugar Η2′ and Η2″ protons sited in the major groove of the decanucleotide. Increasing the Δ-isomer to d(CGCGATCGCG)2 ratio to 5:1 results in noteworthy spectral changes. The Δ-isomer’s proton shifts are reduced, whereas significant shifts are observed for the decanucleotide protons, especially the sugar protons, as well as for the exchangeable protons. Interaction is characterized by the presence of only one intermolecular NOE. Furthermore, there is significant broadening of the imino proton signals as the ratio of the Δ-isomer to DΝΑ increases, which is attributed to the opening of the two strands of the duplex. The Λ-isomer, on the other hand, approaches the minor groove of the oligonucleotide and interacts only weakly, possibly by electrostatic interactions. Photocleavage studies were also conducted with the plasmid pUC19 and a 158-bp restriction fragment, showing that both diastereomers cleave DNA with similar efficiency, attacking mainly the guanines of the sequence probably by generating active oxygen species. Electronic supplementary material Supplementary material is available in the online version of this article at and is accessible for authorized users. An erratum to this article can be found at  相似文献   

14.
TheT 1 andT 2 relaxation times of water protons in two cell types in culture derived from Syrian hamster fetuses (normal primary or secondary fetal cells vs BP6T tumor cells derived from the normal cells transformed by carcinogens) were measured at 7.05 Tesla magnetic field (proton frequency =300 MHz). TheT 1/T 2 ratios and the correlation time, τ c , calculated from theT 1/T 2 ratio of cellular water protons, are significantly different in these two fibroblastic cell types of the same biological origin and with similar morphologies and growth rates in culture.  相似文献   

15.
NOEs between the β-protons of cysteine residues across disulfide bonds in proteins provide direct information on the connectivities and conformations of these important cross-links, which are otherwise difficult to investigate. With conventional [U-13C, 15N]-proteins, however, fast spin diffusion processes mediated by strong dipolar interactions between geminal β-protons prohibit the quantitative measurements and thus the analyses of long-range NOEs across disulfide bonds. We describe a robust approach for alleviating such difficulties, by using proteins selectively labeled with an equimolar mixture of (2R, 3S)-[β-13C; α,β-2H2] Cys and (2R, 3R)-[β-13C; α,β-2H2] Cys, but otherwise fully deuterated. Since either one of the prochiral methylene protons, namely β2 (proS) or β3 (proR), is always replaced with a deuteron and no other protons remain in proteins prepared by this labeling scheme, all four of the expected NOEs for the β-protons across disulfide bonds could be measured without any spin diffusion interference, even with long mixing times. Therefore, the NOEs for the β2 and β3 pairs across each of the disulfide bonds could be observed at high sensitivity, even though they are 25% of the theoretical maximum for each pair. With the NOE information, the disulfide bond connectivities can be unambiguously established for proteins with multiple disulfide bonds. In addition, the conformations around disulfide bonds, namely χ2 and χ3, can be determined based on the precise proton distances of the four β-proton pairs, by quantitative measurements of the NOEs across the disulfide bonds. The feasibility of this method is demonstrated for bovine pancreatic trypsin inhibitor, which has three disulfide bonds.  相似文献   

16.
The individual components of the backbone 15N CSA tensor, σ11, σ22, σ33, and the orientation of σ11 relative to the NH bond described by the angle β have been determined for uniformly labeled 15N, 13C ubiquitin from partial alignment in phospholipid bicelles, Pf1 phage, and poly(ethylene glycol) by measuring the residue-specific residual dipolar couplings and chemical shift deviations. No strong correlation between any of the CSA tensor components is observed with any single structural feature. However, the experimentally determined tensor components agree with the previously determined average CSA principal components [Cornilescu and Bax (2000) J. Am. Chem. Soc. 122, 10143–10154]. Significant deviations from the averages coincide with residues in β-strand or extended regions, while α-helical residue tensor components cluster close to the average values.Electronic Supplementary Material Supplementary material is available to authorised users in the online version of this article at .  相似文献   

17.
Transmembrane proton transport is of fundamental importance for life. The list of H+ transporting proteins has been recently expanded with the discovery that some members of the CLC gene family are stoichiometrically coupled Cl/H+ antiporters. Other CLC proteins are instead passive Cl selective anion channels. The gating of these CLC channels is, however, strongly regulated by pH, likely reflecting the evolutionary relationship with CLC Cl/H+ antiporters. The role of protons in the gating of the model Torpedo channel ClC-0 is best understood. ClC-0 is a homodimer with separate pores in each subunit. Each protopore can be opened and closed independently from the other pore by a “fast gate”. A common, slow gate acts on both pores simultaneously. The opening of the fast gate is controlled by a critical glutamate (E166), whose protonation state determines the fast gate’s pH dependence. Extracellular protons likely can arrive directly at E166. In contrast, protonation of E166 from the inside has been proposed to be mediated by the dissociation of an intrapore water molecule. The OH anion resulting from the water dissociation is stabilized in one of the anion binding sites of the channel, competing with intracellular Cl ions. The pH dependence of the slow gate is less well understood. It has been shown that proton translocation drives irreversible gating transitions associated with the slow gate. However, the relationship of the fast gate’s pH dependence on the proton translocation and the molecular basis of the slow gate remain to be discovered.  相似文献   

18.
The efficiency of cell-free protein synthesis combined with combinatorial selective 15N-labelling provides a method for the rapid assignment of 15N-HSQC cross-peaks to the 19 different non-proline amino-acid types from five 15N-HSQC spectra. This strategy was explored with two different constructs of the C-terminal domain V of the τ subunit of the Escherichia coli DNA polymerase III holoenzyme, τC16 and τC14. Since each of the five 15N-HSQC spectra contained only about one third of the cross-peaks present in uniformly labelled samples, spectral overlap was much reduced. All 15N-HSQC cross-peaks of the backbone amides could be assigned to the correct amino-acid type. Availability of the residue-type information greatly assisted the evaluation of the changes in chemical shifts observed for corresponding residues in τC16 vs. those in τC14, and the analysis of the structure and mobility of the C-terminal residues present in τC16 but not in τC14.  相似文献   

19.
The rabbit Na+/glucose cotransporter (SGLT1) exhibits a presteady-state current after step changes in membrane voltage in the absence of sugar. These currents reflect voltage-dependent processes involved in cotransport, and provide insight on the partial reactions of the transport cycle. SGLT1 presteady-state currents were studied as a function of external Na+, membrane voltage V m , phlorizin and temperature. Step changes in membrane voltage—from the holding V h to test values, elicited transient currents that rose rapidly to a peak (at 3–4 msec), before decaying to the steady state, with time constants τ≈4–20 msec, and were blocked by phlorizin (K i ≈30 μm). The total charge Q was equal for the application of the voltage pulse and the subsequent removal, and was a function of V m . The Q-V curves obeyed the Boltzmann relation: the maximal charge Q max was 4–120 nC; V 0.5, the voltage for 50% Q max was −5 to +30 mV; and z, the apparent valence of the moveable charge, was 1. Q max and z were independent of V h (between 0 and −100 mV) and temperature (20–30°C), while increasing temperature shifted V 0.5 towards more negative values. Decreasing [Na+] o decreased Q max, and shifted V 0.5 to more negative voltages 9by −100 mV per 10-fold decrease in [Na+] o ). The time constant τ was voltage dependent: the τ-V relations were bell-shaped, with maximal τmax 8–20 msec. Decreasing [Na+] o decreased τmax, and shifted the τ-V curves towards more negative voltages. Increasing temperature also shifted the τ-V curves, but did not affect τmax. The maximum temperature coefficient Q 10 for τ was 3–4, and corresponds to an activation energy of 25 kcal/mole. Simulations of a 6-state ordered kinetic model for rabbit Na+/glucose cotransport indicate that charge-movements are due to Na+-binding/dissociation and a conformational change of the empty transporter. The model predicts that (i) transient currents rise to a peak before decay to steady-state; (ii) the τ-V relations are bell-shaped, and shift towards more negative voltages as [Na+] o is reduced; (iii) τmax is decreased with decreasing [Na+] o ; and (iv) the Q-V relations are shifted towards negative voltages as [Na+] o is reduced. In general, the kinetic properties of the presteady-state currents are qualitatively predicted by the model. Received: 12 August 1996/Revised: 30 September 1996  相似文献   

20.
There is an assumption in much recent literature that secreted organic anions (OAs) protect the root meristem from Al toxicity by complexation of Al ions. In fact, several possible mechanisms exist by which common OA might afford some degree of protection. Plants can excrete OA which undergo chemical association with protons (hereafter referred to as protonation) in the soil and increase rhizosphere pH. The cost in reduced carbon relative to protons consumed, C:H+, ranges from 2–6. The efficiency of this mechanism can be enhanced in the presence of soil organisms which can oxidise the OA that remain dissociated at soil pH to CO2 and H2O, thereby consuming protons which associate with lower pK functional groups (pK 1.2 to ~ 4). For fully dissociated organic acids the C:H+ ratio decreases to the range 1–3. The C cost to plants is further minimised if MnO2 is the terminal electron acceptor rather than O2, resulting in C:H+<1. OA might also complex or chelate Al. Complexes of Al3+ with oxalate appear to be effective, with some C:H+≤1. However, citrate complexation appears to be more stable in pure solutions and might offer the additional benefit of enhanced P acquisition. Our assessment is that the most efficient strategy for a plant to employ to protect itself from Al toxicity is to increase pH near the root apex by secreting OA into soil where the microbial oxidation of reduced C could be coupled with the reduction of MnO2. This would consume 0.2–0.67 mole of C per H+, which is the order of magnitude better than the C:H+ ratio of 2–6 that would occur if only protonation of OA was to be relied upon. These mechanisms have implications for the effectiveness of programs aimed at selecting cultivars for resistance to acidic soils.  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