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.
Enzymatic processes are useful for industrially important sugar production, and in vitro two-step isomerization has proven to be an efficient process in utilizing readily available sugar sources. A hypothetical uncharacterized protein encoded by ydaE of Bacillus licheniformis was found to have broad substrate specificities and has shown high catalytic efficiency on d-lyxose, suggesting that the enzyme is d-lyxose isomerase. Escherichia coli BL21 expressing the recombinant protein, of 19.5 kDa, showed higher activity at 40 to 45°C and pH 7.5 to 8.0 in the presence of 1.0 mM Mn2+. The apparent Km values for d-lyxose and d-mannose were 30.4 ± 0.7 mM and 26 ± 0.8 mM, respectively. The catalytic efficiency (kcat/Km) for lyxose (3.2 ± 0.1 mM−1 s−1) was higher than that for d-mannose (1.6 mM−1 s−1). The purified protein was applied to the bioproduction of d-lyxose and d-glucose from d-xylose and d-mannose, respectively, along with the thermostable xylose isomerase of Thermus thermophilus HB08. From an initial concentration of 10 mM d-lyxose and d-mannose, 3.7 mM and 3.8 mM d-lyxose and d-glucose, respectively, were produced by two-step isomerization. This two-step isomerization is an easy method for in vitro catalysis and can be applied to industrial production.  相似文献   

2.
Cytoglobin (Cygb) was investigated for its capacity to function as a NO dioxygenase (NOD) in vitro and in hepatocytes. Ascorbate and cytochrome b5 were found to support a high NOD activity. Cygb-NOD activity shows respective Km values for ascorbate, cytochrome b5, NO, and O2 of 0.25 mm, 0.3 μm, 40 nm, and ∼20 μm and achieves a kcat of 0.5 s−1. Ascorbate and cytochrome b5 reduce the oxidized Cygb-NOD intermediate with apparent second order rate constants of 1000 m−1 s−1 and 3 × 106 m−1 s−1, respectively. In rat hepatocytes engineered to express human Cygb, Cygb-NOD activity shows a similar kcat of 1.2 s−1, a Km(NO) of 40 nm, and a kcat/Km(NO) (kNOD) value of 3 × 107 m−1 s−1, demonstrating the efficiency of catalysis. NO inhibits the activity at [NO]/[O2] ratios >1:500 and limits catalytic turnover. The activity is competitively inhibited by CO, is slowly inactivated by cyanide, and is distinct from the microsomal NOD activity. Cygb-NOD provides protection to the NO-sensitive aconitase. The results define the NOD function of Cygb and demonstrate roles for ascorbate and cytochrome b5 as reductants.  相似文献   

3.
4.
Pah1 is the phosphatidate phosphatase in the yeast Saccharomyces cerevisiae that produces diacylglycerol for triacylglycerol synthesis and concurrently controls the levels of phosphatidate used for phospholipid synthesis. Phosphorylation and dephosphorylation of Pah1 regulate its subcellular location and phosphatidate phosphatase activity. Compared with its phosphorylation by multiple protein kinases, Pah1 is dephosphorylated by a protein phosphatase complex consisting of Nem1 (catalytic subunit) and Spo7 (regulatory subunit). In this work, we characterized the Nem1-Spo7 phosphatase complex for its enzymological, kinetic, and regulatory properties with phosphorylated Pah1. The dephosphorylation of Pah1 by Nem1-Spo7 phosphatase resulted in the stimulation (6-fold) of phosphatidate phosphatase activity. For Pah1 phosphorylated by the Pho85-Pho80 kinase complex, maximum Nem1-Spo7 phosphatase activity required Mg2+ ions (8 mm) and Triton X-100 (0.25 mm) at pH 5.0. The energy of activation for the reaction was 8.4 kcal/mol, and the enzyme was thermally labile at temperatures above 40 °C. The enzyme activity was inhibited by sodium vanadate, sodium fluoride, N-ethylmaleimide, and phenylglyoxal but was not significantly affected by lipids or nucleotides. Nem1-Spo7 phosphatase activity was dependent on the concentrations of Pah1 phosphorylated by Pho85-Pho80, Cdc28-cyclin B, PKA, and PKC with kcat and Km values of 0.29 s−1 and 81 nm, 0.11 s−1 and 127 nm, 0.10 s−1 and 46 nm, and 0.02 s−1 and 38 nm, respectively. Its specificity constant (kcat/Km) for Pah1 phosphorylated by Pho85-Pho80 was 1.6-, 4-, and 6-fold higher, respectively, than that phosphorylated by PKA, Cdc28-cyclin B, and PKC.  相似文献   

5.
1. Phosphomevalonate kinase and 5-pyrophosphomevalonate decarboxylase have been purified from the freeze-dried latex serum of the commercial rubber tree Hevea brasiliensis. 2. The phosphomevalonate kinase was acid- and heat-labile and required the presence of a thiol to maintain activity. 3. The 5-pyrophosphomevalonate decarboxylase was relatively acid-stable and more heat-stable than the phosphokinase. 4. Maximum activity of the phosphokinase was achieved at pH 7.2 with 0.2mm-5-phosphomevalonate (Km 0.042mm), 2.0mm-ATP (Km 0.19mm) and 8mm-Mg2+ at 40°C. The apparent activation energy was 14.8kcal/mol. 5. Maximum activity of 5-pyrophosphomevalonate decarboxylase was achieved at pH5.5–6.5 with 0.1mm-5-pyrophosphomevalonate (Km 0.004mm), 1.5mm-ATP (Km 0.12mm) and 2mm-Mg2+. The apparent activation energy was 13.7kcal/mol. The enzyme was somewhat sensitive to inhibition by its products, isopentenyl pyrophosphate and ADP.  相似文献   

6.
Inactivation of thrombin (T) by the serpins heparin cofactor II (HCII) and antithrombin (AT) is accelerated by a heparin template between the serpin and thrombin exosite II. Unlike AT, HCII also uses an allosteric interaction of its NH2-terminal segment with exosite I. Sucrose octasulfate (SOS) accelerated thrombin inactivation by HCII but not AT by 2000-fold. SOS bound to two sites on thrombin, with dissociation constants (KD) of 10 ± 4 μm and 400 ± 300 μm that were not kinetically resolvable, as evidenced by single hyperbolic SOS concentration dependences of the inactivation rate (kobs). SOS bound HCII with KD 1.45 ± 0.30 mm, and this binding was tightened in the T·SOS·HCII complex, characterized by Kcomplex of ∼0.20 μm. Inactivation data were incompatible with a model solely depending on HCII·SOS but fit an equilibrium linkage model employing T·SOS binding in the pathway to higher order complex formation. Hirudin-(54–65)(SO3) caused a hyperbolic decrease of the inactivation rates, suggesting partial competitive binding of hirudin-(54–65)(SO3) and HCII to exosite I. Meizothrombin(des-fragment 1), binding SOS with KD = 1600 ± 300 μm, and thrombin were inactivated at comparable rates, and an exosite II aptamer had no effect on the inactivation, suggesting limited exosite II involvement. SOS accelerated inactivation of meizothrombin 1000-fold, reflecting the contribution of direct exosite I interaction with HCII. Thrombin generation in plasma was suppressed by SOS, both in HCII-dependent and -independent processes. The ex vivo HCII-dependent process may utilize the proposed model and suggests a potential for oversulfated disaccharides in controlling HCII-regulated thrombin generation.  相似文献   

7.
The crystal structure of the microbial transglutaminase (MTGase) zymogen from Streptomyces mobaraense has been determined at 1.9-Å resolution using the molecular replacement method based on the crystal structure of the mature MTGase. The overall structure of this zymogen is similar to that of the mature form, consisting of a single disk-like domain with a deep active cleft at the edge of the molecule. A major portion of the prosequence (45 additional amino acid residues at the N terminus of the mature transglutaminase) folds into an L-shaped structure, consisting of an extended N-terminal segment linked with a one-turn short helix and a long α-helix. Two key residues in the short helix of the prosequence, Tyr-12 and Tyr-16, are located on top of the catalytic triad (Cys-110, Asp-301, and His-320) to block access of the substrate acyl donors and acceptors. Biochemical characterization of the mature MTGase, using N-α-benzyloxycarbonyl-l-glutaminylglycine as a substrate, revealed apparent Km and kcat/Km values of 52.66 mm and 40.42 mm−1 min−1, respectively. Inhibition studies using the partial prosequence SYAETYR and homologous sequence SQAETYR showed a noncompetitive inhibition mechanism with IC50 values of 0.75 and 0.65 mm, respectively, but no cross-linking product formation. Nevertheless, the prosequence homologous oligopeptide SQAETQR, with Tyr-12 and Tyr-16 each replaced with Gln, exhibited inhibitory activity with the formation of the SQAETQR-monodansylcadaverine fluorophore cross-linking product (SQAETQR-C-DNS). MALDI-TOF tandem MS analysis of SQAETQR-C-DNS revealed molecular masses corresponding to those of NSQAETQC-C-DNS and C-DNS-NQRC sequences, suggesting the incorporation of C-DNS onto the C-terminal Gln residue of the prosequence homologous oligopeptide. These results support the putative functional roles of both Tyr residues in substrate binding and inhibition.  相似文献   

8.
1. all-trans-Retinoic acid at concentrations greater than 10−7m stimulated the incorporation of d-[3H]glucosamine into 8m-urea/5% (w/v) sodium dodecyl sulphate extracts of 1m-CaCl2-separated epidermis from pig ear skin slices cultured for 18h. The incorporation of 35SO42−, l-[14C]fucose and U-14C-labelled l-amino acids was not significantly affected. 2. Electrophoresis of the solubilized epidermis showed increased incorporation of d-[3H]glucosamine into a high-molecular-weight glycosaminoglycan-containing peak when skin slices were cultured in the presence of 10−5m-all-trans-retinoic acid. The labelling of other epidermal components with d-[3H]glucosamine, 35SO42−, l-[14C]fucose and U-14C-labelled l-amino acids was not significantly affected by 10−5m-all-trans-retinoic acid. 3. Trypsinization dispersed the epidermal cells and released 75–85% of the total d-[3H]glucosamine-labelled material in the glycosaminoglycan peak. Thus most of this material was extracellular in both control and 10−5m-all-trans-retinoic acid-treated epidermis. 4. Increased labelling of extracellular epidermal glycosaminoglycans was also observed when human skin slices were treated with all-trans-retinoic acid, indicating a similar mechanism in both tissues. Increased labelling was also found when the epidermis was cultured in the absence of the dermis, suggesting a direct effect of all-trans-retinoic acid on the epidermis. 5. Increased incorporation of d-[3H]-glucosamine into extracellular epidermal glycosaminoglycans in 10−5m-all-trans-retinoic acid-treated skin slices was apparent after 4–8h in culture and continued up to 48h. all-trans-Retinoic acid (10−5m) did not affect the rate of degradation of this material in cultures `chased' with 5mm-unlabelled glucosamine after 4 or 18h. 6. Cellulose acetate electrophoresis at pH7.2 revealed that hyaluronic acid was the major labelled glycosaminoglycan (80–90%) in both control and 10−5m-all-trans-retinoic acid-treated epidermis. 7. The labelling of epidermal plasma membranes isolated from d-[3H]glucosamine-labelled skin slices by sucrose density gradient centrifugation was similar in control and 10−5m-all-trans-retinoic acid-treated tissue. 8. The results indicate that increased synthesis of mainly extracellular glycosaminoglycans (largely hyaluronic acid) may be the first response of the epidermis to excess all-trans-retinoic acid.  相似文献   

9.

Background and Aims

Habitats occupied by many halophytes are not only saline, but are also prone to flooding. Few studies have evaluated submergence tolerance in halophytes.

Methods

Responses to submergence, at a range of salinity levels, were studied for the halophytic stem-succulent Tecticornia pergranulata subsp. pergranulata (syn. Halosarcia pergranulata subsp. pergranulata). Growth and total sugars in succulent stems were assessed as a function of time after submergence. Underwater net photosynthesis, dark respiration, total sugars, glycinebetaine, Na+, Cl and K+, in succulent stems, were assessed in a NaCl dose-response experiment.

Key Results

Submerged plants ceased to grow, and tissue sugars declined. Photosynthesis by succulent stems was reduced markedly when underwater, as compared with in air. Capacity for underwater net photosynthesis (PN) was not affected by 10–400 mm NaCl, but it was reduced by 30 % at 800 mm. Dark respiration, underwater, increased in succulent stems at 200–800 mm NaCl, as compared with those at 10 mm NaCl. On an ethanol-insoluble dry mass basis, K+ concentration in succulent stems of submerged plants was equal to that in drained controls, across all NaCl treatments. Na+ and Cl concentrations, however, were elevated in stems of submerged plants, but so was glycinebetaine. Submerged stems increased in succulence, so solutes would have been ‘diluted’ on a tissue-water basis.

Conclusions

Tecticornia pergranulata tolerates complete submergence, even in waters of high salinity. A ‘quiescence response’, i.e. no shoot growth, would conserve carbohydrates, but tissue sugars still declined with time. A low K+ : Na+ ratio, typical for tissues of succulent halophytes, was tolerated even during prolonged submergence, as evidenced by maintenance of underwater PN at up to 400 mm NaCl. Underwater PN provides O2 and sugars, and thus should enhance survival of submerged plants.Key words: Flooding, halophyte, Halosarcia pergranulata, inundation, inland salt marsh, respiration, Salicornioideae, salt lake, submergence–salinity interaction, tissue solutes, underwater net photosynthesis  相似文献   

10.
KIF3AB is an N-terminal processive kinesin-2 family member best known for its role in intraflagellar transport. There has been significant interest in KIF3AB in defining the key principles that underlie the processivity of KIF3AB in comparison with homodimeric processive kinesins. To define the ATPase mechanism and coordination of KIF3A and KIF3B stepping, a presteady-state kinetic analysis was pursued. For these studies, a truncated murine KIF3AB was generated. The results presented show that microtubule association was fast at 5.7 μm−1 s−1, followed by rate-limiting ADP release at 12.8 s−1. ATP binding at 7.5 μm−1 s−1 was followed by an ATP-promoted isomerization at 84 s−1 to form the intermediate poised for ATP hydrolysis, which then occurred at 33 s−1. ATP hydrolysis was required for dissociation of the microtubule·KIF3AB complex, which was observed at 22 s−1. The dissociation step showed an apparent affinity for ATP that was very weak (K½,ATP at 133 μm). Moreover, the linear fit of the initial ATP concentration dependence of the dissociation kinetics revealed an apparent second-order rate constant at 0.09 μm−1 s−1, which is inconsistent with fast ATP binding at 7.5 μm−1 s−1 and a Kd,ATP at 6.1 μm. These results suggest that ATP binding per se cannot account for the apparent weak K½,ATP at 133 μm. The steady-state ATPase Km,ATP, as well as the dissociation kinetics, reveal an unusual property of KIF3AB that is not yet well understood and also suggests that the mechanochemistry of KIF3AB is tuned somewhat differently from homodimeric processive kinesins.  相似文献   

11.
Accumulation of d-leucine, d-allo-isoleucine, and d-valine was observed in the growth medium of a lactic acid bacterium, Lactobacillus otakiensis JCM 15040, and the racemase responsible was purified from the cells and identified. The N-terminal amino acid sequence of the purified enzyme was GKLDKASKLI, which is consistent with that of a putative γ-aminobutyrate aminotransferase from Lactobacillus buchneri. The putative γ-aminobutyrate aminotransferase gene from L. buchneri JCM 1115 was expressed in recombinant Escherichia coli and then purified to homogeneity. The enzyme catalyzed the racemization of a broad spectrum of nonpolar amino acids. In particular, it catalyzed at high rates the epimerization of l-isoleucine to d-allo-isoleucine and d-allo-isoleucine to l-isoleucine. In contrast, the enzyme showed no γ-aminobutyrate aminotransferase activity. The relative molecular masses of the subunit and native enzyme were estimated to be about 49 kDa and 200 kDa, respectively, indicating that the enzyme was composed of four subunits of equal molecular masses. The Km and Vmax values of the enzyme for l-isoleucine were 5.00 mM and 153 μmol·min−1·mg−1, respectively, and those for d-allo-isoleucine were 13.2 mM and 286 μmol·min−1·mg−1, respectively. Hydroxylamine and other inhibitors of pyridoxal 5′-phosphate-dependent enzymes completely blocked the enzyme activity, indicating the enzyme requires pyridoxal 5′-phosphate as a coenzyme. This is the first evidence of an amino acid racemase that specifically catalyzes racemization of nonpolar amino acids at the C-2 position.  相似文献   

12.

Background and Aims

Measuring the Al3+ uptake rate across the plasma membrane of intact root cells is crucial for understanding the mechanisms and time-course of Al toxicity in plants. However, a reliable method with the sufficient spatial and temporal resolution to estimate Al3+ uptake in intact root cells does not exist.

Methods

In the current study, fluorescent lifetime imaging (FLIM) analysis was used to quantify Al3+ uptake in the root-cell cytoplasm in vivo. This was performed via the estimation of the fluorescence lifetime of Al–lumogallion {5-chloro-3[(2,4-dihydroxyphenyl)azo]-2-hydroxybenzenesulfonic acid} complexes and measurements of intracellular pH while exposing arabidopsis seedlings to acidic and Al3+ stresses.

Key Results

The lifetime of Al–lumogallion complexes fluorescence is pH-dependent. The primary sites for Al3+ entry are the meristem and distal elongation zones, while Al3+ uptake via the cortex and epidermis of the mature root zone is limited. The maximum rates of Al uptake into the cytoplasm (2–3 µmol m−3 min−1 for the meristematic root zone and 3–7 µmol m−3 min−1 for the mature zone) were observed after a 30-min exposure to 100 µm AlCl3 (pH 4·2). Intracellular Al concentration increased to 0·4 µm Al within the first 3 h of exposure to 100 µm AlCl3.

Conclusions

FLIM analysis of the fluorescence of Al–lumogallion complexes can be used to reliably quantify Al uptake in the cytoplasm of intact root cells at the initial stages of Al3+ stress.Key words: Acid stress, Al3+, aluminium toxicity, Arabidopsis thaliana, low pH, fluorescent lifetime imaging (FLIM), lumogallion  相似文献   

13.
Plasmalemma electrical properties were used to investigate K+ transport and its control in internodal cells of Chara corallina Klein ex Willd., em R.D.W. Cell exposure to solutions containing 10 mm KCl caused the potential, normally −250 millivolts (average), to depolarize in two steps. The first step was a 21 millivolt depolarization that lasted from 1 to 40 minutes. The second step started with an action potential and left the membrane potential at −91 millivolts, with a 10-fold reduction in resistance. We suggest that the second step was caused by the opening of K+ -channels in the membrane. This lowered the resistance and provided a current pathway that partially short-circuited the electrogenic pump. Although largely short-circuited, the electrogenic pump was still operating as indicated by: (a) the depolarized potential of −91 millivolts was more negative than Ek (=−42 millivolts in 10 mm K+); (b) a large net K+ uptake occurred while the cell was depolarized; (c) both the electrogenic pump inhibitor, diethylstilbestrol, and the sulfhydryl-reagent N-ethylmaleimide (which increased the passive membrane permeability) further depolarized the potential in 10 mm KCl.A two-phase recovery back to normal cell potentials occurred upon lowering the K+ concentration from 10 to 0.2 mm. The first phase was an apparent Nernst potential response to the change in external K+ concentration. The second phase was a sudden hyperpolarization accompanied by a large increase in membrane resistance. We attribute the second phase to the closing of K+ -channels and the removal of the associated short-circuiting effect on the electrogenic pump, thereby allowing the membrane to hyperpolarize. Further experiments indicated that the K+ -channel required Ca2+ for normal closure, but other ions could substitute, including: Na+, tetraethylammonium, and 2,4,6-triaminopyrimidine. Apparently, K+ -channel conductance is determined by competition between Ca2+ and K+ for a control (gating?) binding site.  相似文献   

14.
C-reactive protein (CRP) is an acute phase protein of the pentraxin family that binds ligands in a Ca2+-dependent manner, and activates complement. Knowledge of its oligomeric state in solution and at surfaces is essential for functional studies. Analytical ultracentrifugation showed that CRP in 2 mm Ca2+ exhibits a rapid pentamer-decamer equilibrium. The proportion of decamer decreased with an increase in NaCl concentration. The sedimentation coefficients s20,w0 of pentameric and decameric CRP were 6.4 S and in excess of 7.6 S, respectively. In the absence of Ca2+, CRP partially dissociates into its protomers and the NaCl concentration dependence of the pentamer-decamer equilibrium is much reduced. By x-ray scattering, the radius of gyration RG values ranged from 3.7 nm for the pentamer to above 4.0 nm for the decamer. An averaged KD value of 21 μm in solution (140 mm NaCl, 2 mm Ca2+) was determined by x-ray scattering and modeling based on crystal structures for the pentamer and decamer. Surface plasmon resonance showed that CRP self-associates on a surface with immobilized CRP with a similar KD value of 23 μm (140 mm NaCl, 2 mm Ca2+), whereas CRP aggregates in low salt. It is concluded that CRP is reproducibly observed in a pentamer-decamer equilibrium in physiologically relevant concentrations both in solution and on surfaces. Both 2 mm Ca2+ and 140 mm NaCl are essential for the integrity of CRP in functional studies and understanding the role of CRP in the acute phase response.  相似文献   

15.
DNA polymerase ν (pol ν) is a low fidelity A-family polymerase with a putative role in interstrand cross-link repair and homologous recombination. We carried out pre-steady-state kinetic analysis to elucidate the kinetic mechanism of this enzyme. We found that the mechanism consists of seven steps, similar that of other A-family polymerases. pol ν binds to DNA with a Kd for DNA of 9.2 nm, with an off-rate constant of 0.013 s−1and an on-rate constant of 14 μm−1 s−1. dNTP binding is rapid with Kd values of 20 and 476 μm for the correct and incorrect dNTP, respectively. Pyrophosphorylation occurs with a Kd value for PPi of 3.7 mm and a maximal rate constant of 11 s−1. Pre-steady-state kinetics, examination of the elemental effect using dNTPαS, and pulse-chase experiments indicate that a rapid phosphodiester bond formation step is flanked by slow conformational changes for both correct and incorrect base pair formation. These experiments in combination with computer simulations indicate that the first conformational change occurs with rate constants of 75 and 20 s−1; rapid phosphodiester bond formation occurs with a Keq of 2.2 and 1.7, and the second conformational change occurs with rate constants of 2.1 and 0.5 s−1, for correct and incorrect base pair formation, respectively. The presence of a mispair does not induce the polymerase to adopt a low catalytic conformation. pol ν catalyzes both correct and mispair formation with high catalytic efficiency.  相似文献   

16.
Pyranose 2-oxidase (P2O) catalyzes the oxidation by O2 of d-glucose and several aldopyranoses to yield the 2-ketoaldoses and H2O2. Based on crystal structures, in one rotamer conformation, the threonine hydroxyl of Thr169 forms H-bonds to the flavin-N5/O4 locus, whereas, in a different rotamer, it may interact with either sugar or other parts of the P2O·sugar complex. Transient kinetics of wild-type (WT) and Thr169 → S/N/G/A replacement variants show that d-Glc binds to T169S, T169N, and WT with the same Kd (45–47 mm), and the hydride transfer rate constants (kred) are similar (15.3–9.7 s−1 at 4 °C). kred of T169G with d-glucose (0.7 s−1, 4 °C) is significantly less than that of WT but not as severely affected as in T169A (kred of 0.03 s−1 at 25 °C). Transient kinetics of WT and mutants using d-galactose show that P2O binds d-galactose with a one-step binding process, different from binding of d-glucose. In T169S, T169N, and T169G, the overall turnover with d-Gal is faster than that of WT due to an increase of kred. In the crystal structure of T169S, Ser169 Oγ assumes a position identical to that of Oγ1 in Thr169; in T169G, solvent molecules may be able to rescue H-bonding. Our data suggest that a competent reductive half-reaction requires a side chain at position 169 that is able to form an H-bond within the ES complex. During the oxidative half-reaction, all mutants failed to stabilize a C4a-hydroperoxyflavin intermediate, thus suggesting that the precise position and geometry of the Thr169 side chain are required for intermediate stabilization.  相似文献   

17.
Calcium compartments and fluxes were measured by kinetic analyses in kidney cell suspensions in a three-compartment closed system. The fast phase influx and compartment size increase linearly with the medium calcium and the half-time of exchange is only 1.3 min which suggests that the fast component is extracellular. The slow phase compartment rises linearly from 0.1 to 0.5 mmole calcium/kg cell water when the medium calcium is raised from 0.02 to 2.5 mM. The slow phase calcium influx exhibits the pattern of saturation kinetics with a V max of 0.065 µµmole cm-2 sec-1 and a Km of 0.3 mM indicating that it is a carrier-mediated transport process. PTH has no effect on the fast phase of calcium influx, but increases both calcium influx and the calcium pool size of the slow component. The maximum effect is obtained at medium calcium concentration of 1.3 mM. Below 0.3 mM extracellular calcium, the effects of the hormone cannot be demonstrated. PTH increases the V max of calcium influx from 0.065 to 0.128 µµmole cm-2 sec-1 while the Km rises from 0.3 to 1.15 mM. These findings suggest that PTH increases the translocation of the calcium-carrier complex across the membrane and not the carrier concentration or its binding affinity for calcium.  相似文献   

18.

Background and Aims

Soil salinity is often heterogeneous, yet the physiology of halophytes has typically been studied with uniform salinity treatments. An evaluation was made of the growth, net photosynthesis, water use, water relations and tissue ions in the halophytic shrub Atriplex nummularia in response to non-uniform NaCl concentrations in a split-root system.

Methods

Atriplex nummularia was grown in a split-root system for 21 d, with either the same or two different NaCl concentrations (ranging from 10 to 670 mm), in aerated nutrient solution bathing each root half.

Key Results

Non-uniform salinity, with high NaCl in one root half (up to 670 mm) and 10 mm in the other half, had no effect on shoot ethanol-insoluble dry mass, net photosynthesis or shoot pre-dawn water potential. In contrast, a modest effect occurred for leaf osmotic potential (up to 30 % more solutes compared with uniform 10 mm NaCl treatment). With non-uniform NaCl concentrations (10/670 mm), 90 % of water was absorbed from the low salinity side, and the reduction in water use from the high salinity side caused whole-plant water use to decrease by about 30 %; there was no compensatory water uptake from the low salinity side. Leaf Na+ and Cl concentrations were 1·9- to 2·3-fold higher in the uniform 670 mm treatment than in the 10/670 mm treatment, whereas leaf K+ concentrations were 1·2- to 2·0-fold higher in the non-uniform treatment.

Conclusions

Atriplex nummularia with one root half in 10 mm NaCl maintained net photosynthesis, shoot growth and shoot water potential even when the other root half was exposed to 670 mm NaCl, a concentration that inhibits growth by 65 % when uniform in the root zone. Given the likelihood of non-uniform salinity in many field situations, this situation would presumably benefit halophyte growth and physiology in saline environments.Key words: Split-root system, salinity heterogeneity, root zone heterogeneity, water potential, water use, stomatal conductance, saltbush, leaf ions, photosynthesis, NaCl  相似文献   

19.
The nicotinamide adenine dinucleotide-specific glutamate dehydrogenase (l-glutamate:NAD+ oxidoreductase, EC 1.4.1.2) of Chlorella sorokiniana was purified 1,000-fold to electrophoretic homogeneity. The native enzyme was shown to have a molecular weight of 180,000 and to be composed of four identical subunits with a molecular weight of 45,000. The N-terminal amino acid was determined to be lysine. The pH optima for the aminating and deaminating reactions were approximately 8 and 9, respectively. The Km values for α-ketoglutarate, NADH, NH4+, NAD+, and l-glutamate were 2 mm, 0.15 mm, 40 mm, 0.15 mm, and 60 mm, respectively. Whereas the Km for α-ketoglutarate and l-glutamate increased 10-fold, 1 pH unit above or below the pH optima for the aminating or deaminating reactions, respectively, the Km values for NADH and NAD+ were independent of change in pH from 7 to 9.6. By initial velocity, product inhibition, and equilibrium substrate exchange studies, the kinetic mechanism of enzyme was shown to be consistent with a bi uni uni uni ping-pong addition sequence. Although this kinetic mechanism differs from that reported for any other glutamate dehydrogenase, the chemical mechanism still appears to involve the formation of a Schiff base between α-ketoglutarate and an ε-amino group of a lysine residue in the enzyme. The physical, chemical, and kinetic properties of this enzyme differ greatly from those reported for the NH4+-inducible glutamate dehydrogenase in this organism.  相似文献   

20.
Both cyclic AMP phosphodiesterase and cyclic GMP phosphodiesterase were recovered mainly from the supernatant fractions of guinea-pig pancreas, but a higher proportion of the activity of the former was associated with the pellet fractions. The activities in the supernatant were not separated by gel filtration, but were clearly separated by subsequent chromatography on an anion-exchange resin. The activities of cyclic AMP phosphodiesterase and cyclic GMP phosphodiesterase had high-affinity (Km 6.5±1.1μm and 31.9±3.9μm respectively) and low-affinity (Km 0.56±0.05mm and 0.32±0.03mm respectively) components. The activity of neither enzyme was affected by the pancreatic secretogens, cholecystokinin-pancreozymin, secretin and carbachol. Removal of ions by gel filtration resulted in a marked reduction in cyclic nucleotide phosphodiesterase activity, which could be restored by addition of Mg2+. Mn2+ (3mm) was as effective as Mg2+ (3mm) in the case of cyclic AMP phosphodiesterase, but was less than half as effective in the case of cyclic GMP phosphodiesterase. The metal-ion chelators, EDTA and EGTA, also decreased activity. Ca2+ (1mm) did not affect the activity of cyclic nucleotide phosphodiesterase when the concentration of Mg2+ was 3mm. At concentrations of Mg2+ between 0.1 and 1mm, 1mm-Ca2+ was activatory, and at concentrations of Mg2+ below 0.1mm, 1mm-Ca2+ was inhibitory. These results are discussed in terms of the possible significance of cyclic nucleotide phosphodiesterase in the physiological control of cyclic nucleotide concentrations during stimulus–secretion coupling.  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