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.
Various inhibitors were tested for their inhibitory effects on soybean urease. The Ki values for boric acid, 4-bromophenylboronic acid, butylboronic acid, and phenylboronic acid were 0.20?±?0.05?mM, 0.22?±?0.04?mM, 1.50?±?0.10?mM, and 2.00?±?0.11?mM, respectively. The inhibition was competitive type with boric acid and boronic acids. Heavy metal ions including Ag+, Hg2+, and Cu2+ showed strong inhibition on soybean urease, with the silver ion being a potent inhibitor (IC50 = 2.3?×?10?8 mM). Time-dependent inhibition studies exhibited biphasic kinetics with all heavy metal ions. Furthermore, inhibition studies with sodium salts of mineral acids (NaF, NaCl, NaNO3, and Na2SO4) showed that only F? inhibited soybean urease significantly (IC50 = 2.9?mM). Competitive type of inhibition was observed for this anion with a Ki value of 1.30?mM.  相似文献   

2.
West Nile virus (WNV) is a member of the flavivirus genus belonging to the Flaviviridae family. The viral serine protease NS2B/NS3 has been considered an attractive target for the development of anti-WNV agents. Although several NS2B/NS3 protease inhibitors have been described so far, most of them are reversible inhibitors. Herein, we present a series of α-aminoalkylphosphonate diphenyl esters and their peptidyl derivatives as potent inhibitors of the NS2B/NS3 protease. The most potent inhibitor identified was Cbz-Lys-Arg-(4-GuPhe)P(OPh)2 displaying Ki and k2/Ki values of 0.4 µM and 28 265 M?1s?1, respectively, with no significant inhibition of trypsin, cathepsin G, and HAT protease.  相似文献   

3.
—The blood-brain barrier transport of amino acids has been measured using the carotid injection technique in the rat. The synthetic amino acids, 2-aminobicyclo(2,2,1)heptane-2-carboxylic acid (BCH) and α-(methylamino)isobutyric acid (MeAIB), were model substrates in the Ehrlich cell for the leucine (L) and alanine (A) neutral amino acid transport mechanisms, respectively. The uptake (±)b-[carboxyl-14C]BCH at the same rate for the five brain regions tested suggested a similarity between regions for the L transport mechanism. At injectant concentrations of 0·1 mm (similar to naturally occurring aromatic neutral amino acids), BCH was mainly taken up by a saturable mediated transport mechanism (K1, 0·16 mm and Vmax, 0·03/μmol/g per min). At higher concentrations, uptake by a nonsaturable or diffusional mechanism could be demonstrated. When BCH was added as a second amino acid to l -[3-14C]DOPA, the saturable component of l -DOPA transport was significantly inhibited. MeAIB had no measurable effect on the rate of l -DOPA transport. These results suggested that the mediated transport mechanism for l -DOPA at the cerebral capillaries is similar to the l -neutral amino acid transport system.  相似文献   

4.
The kinetics of the light-driven Cl? uptake pump of Synechococcus R-2 (PCC 7942) were investigated. The kinetics of Cl? uptake were measured in BG-11 medium (pHo, 7·5; [K+]o, 0·35 mol m?3; [Na+]o, 18 mol m?3; [Cl?]o, 0·508 mol m?3) or modified media based on the above. Net36Cl? fluxes (?Cl?o,i) followed Michaelis-Menten kinetics and were stimulated by Na+ [18 mol m?3 Na+ BG-11 ?Cl?max= 3·29±0·60 (49) nmol m?2 s?1 versus Na+-free BG-11 ?Cl?max= 1·02±0·13 (54) nmol m?2 s?1] but the Km was not significantly different in the presence or absence of Na+ at pHo 10; the Km was lower, but not affected by the presence or absence of Na+ [Km = 22·3±3·54 (20) mmol m?3]. Na+ is a non-competitive activator of net ?Cl?o,i. High [K+]o (18 mol m?3) did not stimulate net ?Cl?o,i or change the Km in Na+-free medium. High [K+]o (18 mol m?3) added to Na+ BG-11 medium decreased net ?Cl?o,i [18 mol m?3K+ BG-11; ?Cl?max= 2·50±0·32 (20) nmol m?2 s?1 versus BG-11 medium; ?Cl?max= 3·35±0·56 (20) nmol m?2 s?1] but did not affect the Km 55·8±8·100 (40) mmol m?3]. Na+-stimulation of net ?Cl?o,i followed Michaelis-Menten kinetics up to 2–5 mol m?3 [Na+]o but higher concentrations were inhibitory. The Km for Na+-stimulation of net ?Cl?o,i [K1/2(Na+)] was different at 47 mmol m?3 [Cl?]o (K1/2[Na+] = 123±27 (37) mmol m?3]. Li+ was only about one-third as effective as Na+ in stimulating Cl? uptake but the activation constant was similar [K1/2(Li+) = 88±46 (16) mmol m?3]. Br? was a competitive inhibitor of Cl? uptake. The inhibition constant (Ki) was not significantly different in the presence and absence of Na+. The overall Ki was 297±23 (45) mmol m?3. The discrimination ratio of Cl? over Br? (δCl?/δBr?) was 6·38±0·92 (df = 147). Synechococcus has a single Na+-stimulated Cl? pump because the Km of the Cl? transporter and its discrimination between Cl? and Br? are not significantly different in the presence and absence of Na+. The Cl? pump is probably driven by ATP.  相似文献   

5.
The kinetic parameters of the inhibition of pigeon brain acetylchlolinesterase (AChE) by procaine hydrochloride were investigated. Procaine (0·083–1·67 mM) reversibly inhibited AChE activity (15–83 percent) in a concentration dependent manner, the IC50 being about 0·38 mM. The Michaelis-Menten constant (Km) for the hydrolysis of acetylthiocholine iodide was found to be 1·53 × 10?4 M and the Vmax was 1·06 μmol min?1 mg?1 protein. Dixon as well as Lineweaver-Burk plots and their secondary replots indicated that the nature of the inhibition is of the linear mixed type which is considered to be a mixture of partial competitive and pure non-competitive. The values of Ki(slope) and Ki (intercepts) were estimated as 0·14 mM and 0·22 mM respectively by the primary Dixon and by the secondary replots of the Lineweaver-Burk plot. The Ki′/Ki ratio shows that procaine has a greater affinity of binding for the peripheral than for the active site.  相似文献   

6.
Abstract: With a partially purified, membrane-bound (Ca + Mg)-activated ATPase preparation from rat brain, the K0.5 for activation by Ca2+ was 0.8 p μm in the presence of 3 mm -ATP, 6 mm -MgCl2, 100 mM-KCI, and a calcium EGTA buffer system. Optimal ATPase activity under these circumstances was with 6-100 μm -Ca2+, but marked inhibition occurred at higher concentrations. Free Mg2+ increased ATPase activity, with an estimated K0.5, in the presence of 100 μm -CaCl2, of 2.5 mm ; raising the MgCl2 concentration diminished the inhibition due to millimolar concentrations of CaCl2, but antagonized activation by submicromolar concentrations of Ca2+. Dimethylsulfoxide (10%, v/v) had no effect on the K0.5 for activation by Ca2+, but decreased activation by free Mg2+ and increased the inhibition by millimolar CaCl2. The monovalent cations K+, Na+, and TI+ stimulated ATPase activity; for K+ the K0.5 was 8 mm , which was increased to 15 mm in the presence of dimethylsulfoxide. KCI did not affect the apparent affinity for Ca2+ as either activator or inhibitor. The preparation can be phosphorylated at 0°C by [γ-32P]-ATP; on subsequent addition of a large excess of unlabeled ATP the calcium dependent level of phosphorylation declined, with a first-order rate constant of 0.12 s?1. Adding 10 mm -KCI with the unlabeled ATP increased the rate constant to 0.20 s?1, whereas adding 10 mm -NaCl did not affect it measurably. On the other hand, adding dimethyl-sulfoxide slowed the rate of loss, the constant decreasing to 0.06 s?1. Orthovanadate was a potent inhibitor of this enzyme, and inhibition with 1 μm -vanadate was increased by both KCI and dimethylsulfoxide. Properties of the enzyme are thus reminiscent of the plasma membrane (Na + K)-ATPase and the sarcoplasmic reticulum (Ca + Mg)-ATPase, most notably in the K+ stimulation of both dephosphorylation and inhibition by vanadate.  相似文献   

7.
Abstract— A method for a partial purification of ATP:citrate oxaloacetate lyase from rat brain is described. The Lineweaver–Burk plots of velocity vs citrate concentration are biphasic in the presence of fixed concentrations of MgCl2. Therefore two values of Km, corresponding to low and high concentrations of citrate, can be determined. When MgCl2 is added in equimolar concentrations with citrate, a monophasic plot with one Km of 0.13 mm is obtained. The Km value for MgATP2- was independent of citrate concentration, being equal to 0.40–0.43 mm. The Km for CoA was 0.0007 mm. ADP and Pi are competitive inhibitors with respect to ATP. Ki for MgADP is equal to 0.13 mm. dl -isocitrate and cis-aconitate are partially competitive inhibitors with respect to citrate with Ki values of 5.8 and 4.8 mm, respectively. α-Ketoglutarate and pyruvate are noncompetitive inhibitors with respect to ATP and citrate, with Ki values equal to 9 and 45 mm, respectively. The physiological significance of these effectors for the regulation of citrate lyase activity in brain is discussed.  相似文献   

8.
The rates of formation and dissociation of concanavalin A with some 4-methylumbelliferyl and p-nitrophenyl derivatives of α- and β-D-mannopyranosides and glucopyranosides were measured by fluorescence and spectral stopped-flow methods. All process examined were uniphasic. The second-order formation rate constants varied only from 6.8 · 104 to 12.8 · 104 M?. s?1, whereas the first-order dissociation rate constants ranged from 4.1. to 220 s?1, all at ph 5.0, I = 0.3 M, and 25°C. Dissociation rates thus controlled the value of binding constant. The effect of temperature on these reactions was examined, from which enthalpies and entropies of activation and of reaction could be calculated. The effects of pH at 25°C on the reaction rates of 4-methylumbelliferyl α-D-mannopyranoside and 4-methylumbelliferyl α-D-glucopyranoside with concanavalin A were examined. The value of the binding constant Kap (derived from the kinetics) at any pH could be related to the intrinsic binding constant K by the expression Kap = KaK(Ka + [H+])?1. The values of Ka, the ionization constant of the protein segment responsive to sugar binding, were 3 · 10?4 M and 1 · 10?4 M for 4-methylumbelliferyl α-D-mannopyranoside and 4-methylumbelliferyl α-D-glucopyranoside, respectively. The binding constant of p-nitrophenyl α-D-mannopyranoside is surprisingly much less sensitive to a pH change from 5.0 to 2.7. Ionic strength had little effect on the binding characteristics of 4-methylumbelliferyl α-D-mannopyranoside to concanavalin A at pH 5.2 and 25°C.  相似文献   

9.
—The hydrolysis of ThTP by rat brain membrane-bound ThTPase is inhibited by nucleoside diphosphates and triphosphates. ATP and ADP are most effective, reducing hydrolysis by 50% at concentrations of 2 × 10?5m and 7·5 × 10?5m respectively. Nucleoside monophosphates and free nuclcosides as well as Pi have no effect on enzyme activity. ThMP and ThDP also fail to inhibit hydrolysis in concentrations up to 5 × 10?3m . Non-hydrolysable methylene phosphate analogs of ATP and ADP were used in further kinetic studies with the ThTPase. The mechanism of inhibition by these analogs is shown to be of mixed non-competitive nature for both compounds. An observed Ki, of 4 × 10?5m for the ATP analog adenosine-PPCP and 9 × 10?5m for the ADP analog adenosine-PCP is calculated at pH 6·5. Formation of the true enzyme substrate, the [Mg2+. ThTP] complex, is not significantly affected by concentrations of analogs producing maximal (>95%) inhibition of enzyme activity. Likewise the relationships between pH and observed Km and pH and Vmax are not shifted by the presence of similar concentrations of inhibitor.  相似文献   

10.
Mitochondria were isolated from the hepatopancreas of the Florida spiny lobster Panulirus argus using a high osmolarity medium containing 600 mm mannitol, 83 mm sucrose, 5 mm 4-morpholinepropanesulfonic acid, pH 7.6, 0.5% bovine serum albumin (BSA), and 1 mm EDTA. O2 uptake and Ca2+ transport were measured by electrode methods in similar media (plus 4 mm KPi, 3.3 mm MgCl2, and 0.67 mg/ml BSA, with 80 mm KCl replacing a portion of the osmotic support). Substrate-supported respiration was observed to be coupled to phosphorylation of ADP or uptake of Ca2+ ions. State 3 rates (nanogram atoms O × minute?1 × milligram protein?1 ± SEM (N)) were: 49.2 ± 3.9 (19), succinate; 30.9 ± 3.9 (6), dl-palmitoyl carnitine; 29.0 ± 2.7 (9), l-malate; 40.0 ± 2.3 (3), l-glutamate; 27.7 ± 2.2 (5), d-3-hydroxybutyrate; and 26.4 ± 2.4 (18), l-proline ± pyruvate. α-Glycerol phosphate was not oxidized. Ca2+ uptake driven by succinate oxidation proceeded with Ca:O ratios of 4.0 ± 0.2 (SEM). Hepatopancreas mitochondria were not uncoupled by Ca2+ uptake in excess of 1100 ng atoms × mg protein?1. Ca2+ efflux could be induced by ruthenium red, indicating the presence of an active Ca2+ cycle. These mitochondria may provide a favorable model system in which to study regulation of the Ca2+ cycle.  相似文献   

11.
ATP sulfurylase from Penicillium chrysogenum was purified to homogeneity. The enzyme binds 8 mol of free ATP (Ks = 0.53 mM) or AMP (Ks = 0.50 mM) per 440,000 g. The results are consistent with our earlier report that the enzyme is composed of eight identical subunits of Mr 55,000 (J. W. Tweedie and I. H. Segel, 1971, Prep. Biochem. 1, 91–117; J. Biol. Chem. 246, 2438–2446). In the absence of cosubstrates, the purified enzyme catalyzes the hydrolysis of MgATP (to AMP and MgPPi) and adenosine 5′-phosphosulfate (APS) (to AMP and SO42?). MgATP hydrolysis is inhibited by nonreactive sulfate analogs such as nitrate, chlorate, and formate (uncompetitive with MgATP). In spite of the hydrolytic reactions it is possible to observe the binding of MgATP and APS to the enzyme in a qualitative (nonequilibrium) manner. Neither inorganic sulfate (the cosubstrate of the forward reaction) nor formate or inorganic phosphate (inhibitors competitive with sulfate) will bind to the free enzyme in detectable amounts in the absence or in the presence of Mg2+, Ca2+, free ATP, or a nonreactive analog of MgATP such as Mg-α,β-methylene-ATP. Similarly, inorganic pyrophosphate (the cosubstrate of the reverse reaction) will not bind in the absence or in the presence of Mg2+ or Ca2+. The induced binding of 32Pi (presumably to the sulfate site) can be observed in the presence of MgATP. The results are consistent with the obligately ordered binding sequence deduced from the steady-state kinetics (J. Farley et al., 1976, J. Biol. Chem. 251, 4389–4397) and suggest that the subsites for SO2?4 or MgPPi appear only after nucleotide cleavage to form E~AMP · MgPPi or E~AMP · SO4 complexes. The suggestion is supported by the relative values of Kia (ca. 1 mm for MgATP) and Kiq (ca. 1 αm for APS) and by the inconsistent value of k?1 calculated from VfKiaKmA (The value is considerably less than Vr) Purified ATP sulfurylase will also catalyze a Mg32PPi-MgATP exchange in the absence of SO42?. A 35SO42?-APS exchange could not be demonstrated in the absence or presence of MgPPi. This result was not unexpected: The rate of APS hydrolysis (or conversion to MgATP) is extremely rapid compared to the expected exchange rate. Also, the pool of APS at equilibrium is extremely small compared to the sulfate pool. The V values for molybdolysis, APS hydrolysis (in the absence of PPi), ATP synthesis (from APS + MgPPi), and Mg32PPi-MgATP exchange at saturating sulfate are all about equal (12–19 μmol × min?1 × mg of enzyme?1). The rates of Mg32PPi-MgATP exchange in the absence of sulfate, APS synthesis (from MgATP + sulfate), and MgATP hydrolysis (in the absence of sulfate) are considerably slower (0.10 – 0.35 μmol × min?1 × mg of enzyme?1). These results and the fact that k4 calculated from VrKiqKmQ is considerably larger than Vf suggest that the rate-limiting step in the overall forward reaction is the isomerization reaction E~AMP-SO2?4 → EAPS. In the reverse direction the rate-limiting step may be SO2?4 release or isomerization of the E~AMP · MgPPi · SO42? complex. (The reaction appears to be rapid equilibrium ordered.) Reactions involving the synthesis or cleavage of APS are specific for Mg2+. Reactions involving the synthesis or cleavage of ATP will proceed with Mg2+, with Mn2+, and, at a lower rate, with Co2+. The results suggest that the enzyme possesses a Mg2+-preferring divalent cation (activator) binding site that is involved in APS synthesis and cleavage and is distinct from the MeATP or MePPi site. The equilibrium binding of about one atom of 45Ca2+ per subunit (possibly to the activator site) could be demonstrated (Ks = 1.4 mM).  相似文献   

12.
Mimicry of structural motifs is a common feature in proteins. The 10‐membered hydrogen‐bonded ring involving the main‐chain C?O in a β‐turn can be formed using a side‐chain carbonyl group leading to Asx‐turn. We show that the N? H component of hydrogen bond can be replaced by a Cγ‐H group in the side chain, culminating in a nonconventional C? H···O interaction. Because of its shape this β‐turn mimic is designated as ω‐turn, which is found to occur ~three times per 100 residues. Three residues (i to i + 2) constitute the turn with the C? H···O interaction occurring between the terminal residues, constraining the torsion angles ?i + 1, ψi + 1, ?i + 2 and χ1(i + 2) (using the interacting Cγ atom). Based on these angles there are two types of ω‐turns, each of which can be further divided into two groups. Cβ‐branched side‐chains, and Met and Gln have high propensities to occur at i + 2; for the last two residues the carbonyl oxygen may participate in an additional interaction involving the S and amino group, respectively. With Cys occupying the i + 1 position, such turns are found in the metal‐binding sites. N‐linked glycosylation occurs at the consensus pattern Asn‐Xaa‐Ser/Thr; with Thr at i + 2, the sequence can adopt the secondary structure of a ω‐turn, which may be the recognition site for protein modification. Location between two β‐strands is the most common occurrence in protein tertiary structure, and being generally exposed ω‐turn may constitute the antigenic determinant site. It is a stable scaffold and may be used in protein engineering and peptide design. Proteins 2015; 83:203–214. © 2014 Wiley Periodicals, Inc.  相似文献   

13.
The activities of novel Cbz-N-protected α-aminophosphonic phenyl esters, analogs of leucine (1–15) and phenylalanine (17–29), which are substituted at the phenyl ester rings, as well as of their peptidic derivatives (31–43), were investigated for their inhibitory effects on chymotrypsin and subtilisin. The chemical nature and position of the examined substituents clearly demonstrated a strong structure–activity relationship. Among all synthesized compounds the most potent phosphonic-type inhibitors of subtilisin and chymotrypsin were identified, with k2/Ki values 114,380?M?1s?1 and 307,380?M?1s?1, respectively.  相似文献   

14.
Human serum butyrylcholinesterase (BChE) has been converted into a stable but less active desensitized form when heated at 45°C for 24 h. The desensitized BChE follows Michaelis-Menten kinetics, whereas native enzyme exhibits slightly negative cooperativity with respect to butyrylthiocholine binding. In this study, we investigated the effects of Ni2+, Co2+, and Mn2+ on the desensitized BChE. It is found that all three ions were noncompetitive inhibitors of the desensitized BChE, and K i values have been determined as 7.816±1.060 mM, 48.722±4.635 mM, and 84.795±5.249 mM for Ni2+, Co2+, and Mn2+, respectively. In our previous study, these ions were linear mixed-type inhibitors of the native BChE. This finding confirms that desensitized BChE changes to a different conformation than native BChE. From the comparison of K i values of the trace elements, it can be said that Ni2+ is a more effective inhibitor of the desensitized BChE than Co2+ and Mn2+.  相似文献   

15.
Uptake of d-glucosamine by rat brain synaptosomes occurs via a saturable transport process (Km 2.1 mM, V 3.0 nmol/mg per min) which was clearly distinguishable from simple diffusion. This transport process is highly sensitive to cytochalasin (Ki = 7 · 10?5 mM. d-Glucose competitively inhibits d-glucosamine uptake with a Ki value of 8 · 10?1 mM.  相似文献   

16.
Synechococcus R-2 (PCC 7942) actively accumulated Cl? in the light and dark, under control conditions (BG-11 media: pHo, 7·5; [Na+]o, 18 mol m?3; [Cl?]o, 0·508 molm?3). In BG-11 medium [Cl?], was 17·2±0·848 mol m?3 (light), electrochemical potential of Cl? (ΔμCl?i,o) =+211±2mV; [Cl?]i= 1·24±0·11 mol m?3(dark), ΔμCl?i,o=+133±4mV. Cl? fluxes, but not permeabilities, were much higher in the light: ?Cl?i,o= 4·01±5·4 nmol m?2 s?1, PCl?i,o= 47±5pm s?1 (light); ?Cl?i,o= 0·395±0·071 nmol m?2 s?1, PCl?i,o= 69±14 pm s?1 (dark). Chloride fluxes are inhibited by acid pHo (pHo 5; ?Cl?i,o= 0·14±0·04 nmol m?2 s?1); optimal at pHo 7·5 and not strongly inhibited by alkaline pHo (pHo 10; ?Cl?1i,o= 1·7±0·14 nmol m?2 s?1). A Cl?in/2H+in coporter could not account for the accumulation of Cl? alkaline pHo. Permeability of Cl? is very low, below 100pm s?1 under all conditions used, and appears to be maximal at pHo 7·5 (50–70 pm s?1) and minimal in acid pHo (20pm s?1). DCCD (dicyclohexyl-carbodiimide) inhibited ?Cl?i,o in the light about 75% and [Cl?]i fell to 2·2±0·26 (4) mol m?3. Valinomycin had no effect but monensin severely inhibited Cl? uptake ([Cl?]i= 1·02±0·32 mol m?3; ?Cl?i,o= 0·20±0·1 nmol m?2 s?1). Vanadate (200 mmol m?3) accelerated the Cl? flux (?Cl?i,o= 5·28±0·64 nmol m?2 s?1) but slightly decreased accumulation of Cl? ([Cl?], = 13·9±1·3 mol m?3) in BG-11 medium but had no significant effect in Na+-free media. DCMU (dichlorophenyldimethylurea) did not reduce [Cl?], or ?Cl?i,o to that found in the dark ([Cl?]i= 8·41±0·76 mol m?3; ?Cl?i,o= 2·06±0·36 nmol m?2 s?1). Synechococcus also actively accumulated Cl? in Na+-free media, [Cl?]i was lower but ΔΨi,o hyperpolarized in Na+-free media and so the ΔμCl?i,o was little changed ([Cl?]i= 7·98±0·698 mol m?3; ΔμCl?i,o=+203±3 mV). Net Cl? uptake was stimulated by Na+; Li+ acted as a partial analogue for Na+. Synechococcus has a Na+ activated Cl? transporter which is probably a primary 2Cl?/ATP pump. The Cl? pump is voltage sensitive. ΔμCl?i,o is directly proportional to ΔΨi,o(P»0·01%): ΔμCl?i,o= -1·487 (±0·102) ×ΔΨi,o, r= -0·983, n= 31. The ΔμCl?i,o increased (more positive) as the Δμi,o became more negative. The ΔμCl?i,o has no known function, but might provide a driving force for the uptake of micronutrients.  相似文献   

17.
Branches of 22-year-old loblolly pine (Pinus taeda, L.) trees growing in a plantation were exposed to ambient CO2, ambient + 165 μmol mol?1 CO2 or ambient + 330 μmol mol?1 CO2 concentrations in combination with ambient or ambient + 2°C air temperatures for 3 years. Field measurements in the third year indicated that net carbon assimilation was enhanced in the elevated CO2 treatments in all seasons. On the basis of A/Ci, curves, there was no indication of photosynthetic down-regulation. Branch growth and leaf area also increased significantly in the elevated CO2 treatments. The imposed 2°C increase in air temperature only had slight effects on net assimilation and growth. Compared with the ambient CO2 treatment, rates of net assimilation were ~1·6 times greater in the ambient + 165 μmol mol?1 CO2 treatment and 2·2 times greater in the ambient + 330 μmol mol?1 CO2 treatment. These ratios did not change appreciably in measurements made in all four seasons even though mean ambient air temperatures during the measurement periods ranged from 12·6 to 28·2°C. This indicated that the effect of elevated CO2 concentrations on net assimilation under field conditions was primarily additive. The results also indicated that the effect of elevated CO2 (+ 165 or + 330 μmol mol?1) was much greater than the effect of a 2°C increase in air temperature on net assimilation and growth in this species.  相似文献   

18.
Transport of l-proline into Saccharomyces cerevisiae K is mediated by two systems, one with a KT of 31 μM and Jmax of 40 nmol · s?1 · (g dry wt.)?1, the other with KT > 2.5 mM and Jmax of 150–165 nmol · s?1 · (g dry wt.)?1, The kinetic properties of the high-affinity system were studied in detail. It proved to be highly specific, the only potent competitive inhibitors being (i) l-proline and its analogs l-azetidine-2-carboxylic acid, sarcosine, d-proline and 3,4-dehydro-dl-proline, and (ii) l-alanine. The other amino acids tested behaved as noncompetitive inhibitors. The high-affinity system is active, has a sharp pH optimum at 5.8–5.9 and, in an Arrhenius plot, exhibits two inflection points at 15°C and 20–21°C. It is trans-inhibited by most amino acids (but probably only the natural substrates act in a trans-noncompetitive manner) and its activity depends to a considerable extent on growth conditions. In cells grown in a rich medium with yeast extract maximum activity is attained during the stationary phase, on a poor medium it is maximal during the early exponential phase. Some 50–60% of accumulated l-proline can leave cells in 90 min (and more if washing is done repeatedly), the efflux being insensitive to 0.5 mM 2,4-dinitrophenol and uranyl ions, to pH between 3 and 7.3, as well as to the presence of 10–100 mM unlabeled l-proline in the outside medium. Its rate and extent are increased by 1% d-glucose and by 10 μg nystatin per ml.  相似文献   

19.
Particulate fractions (10,000g) from pupae of Stomoxys calcitrans transfer [14C]-mannose from GDP-[14C]-mannose to dolichol monophosphate and proteins. Production of the mannosyl lipid was inhibited by Mn2+, UDP, GMP, GDP, and EDTA. The insect growth regulator diflubenzuron had no effect on mannosyl transferase activity. Dolichol monophosphate and Mg2+ stimulated mannosyl transferase activity. The mannosyl lipid product was identified as mannosyl-phosphoryl-dolichol (Man-P-Dol). The apparent Km and Vmax values for the formation of Man-P-Dol using GDP-[14C]-Man while holding dolichol phosphate constant were 2.4 ± 0.9 μM and 9.4 ± 2.3 pmol Man-P-Dol·min?1·mg?1 protein, respectively. The apparent Km and Vmax values using dólichol phosphate while holding GDP-Man constant were 2.2 ± 1.2 μM and 18.5 ± 1.7 pmol Man-P-Dol·min?1·mg?1 protein.  相似文献   

20.
Approximately 80 per cent of tyrosine hydroxylase activity in bovine mandibular nerve and rabbit sciatic nerve was soluble, and the rest of the activity was particle-bound. The soluble enzyme in bovine mandibular nerve was isolated by ammonium sulphate fractionation (25–35 per cent saturation). The enzyme had a pH optimum at 5·9 in Tris-acetate buffer, and at 6·5 in Tris-HCl or phosphate buffer. The enzyme required a tetrahydropteridine cofactor. Km values toward various tetrahydropteridines such as l -erythro-tetrahydrobiopterin (a probable natural cofactor), 2-amino-4-hydroxy-6-methyltetrahydropteridine, and 2-amino-4-hydroxy-6,7-dimethyltetrahydropteridine were 2 × 10−5m , 5 × 10−5m and 4 × 10−4m , respectively. The Km value for tyrosine at 1 × 10−3m -2-amino-4-hydroxy-6-methyltetrahydropteridine as a cofactor was 5 × 10−5m . The enzyme activity was markedly stimulated with Fe2+ or catalase, but Fe2+ gave higher activity. The activity was inhibited with α, α′-dipyridyl, l -α-methyl-p-tyrosine, and various catecholamines. Among catecholamines, dopamine was the most potent inhibitor. l -5-Hydroxytryptophan was an inhibitor as potent as dopamine. Neither d -5-hydroxytryptophan nor 5-hydroxytryptamine inhibited the enzyme. The inhibition by l -5-hydroxytryptophan was partially competitive with tetrahydrobiopterin at concentrations higher than 9 × 10−5m , and partially uncompetitive at concentrations lower than 9 × 10−5m . The addition of heparin or lysolecithin did not affect enzyme activity with tetrahydrobiopterin as cofactor.  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