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 93 毫秒
1.
A binding component with a high affinity for 5-methyltetrahydrofolate (KD = 0.11μm) is present on the external surface of L1210 cells. The amount of binder (1 pmol/mg protein) corresponds to 8 × 104 sites per cell. The participation of this component in the high-affinity 5-methyltetrahydrofolate/methotrexate transport system is supported by similarities in the KD values for 5-methyltetrahydrofolate and methotrexate binding and the Kt values of these compounds for transport. Relative affinities for other folate substrates (aminopterin, 5-formyltetrahydrofolate, and folate) and various competitive inhibitors (thiamine pyrophosphate, ADP, AMP, arsenate, and phosphate) are also similar for both the binding component and the transport system. The measured binding activity does not represent low-temperature transport of substrate into cells, since it is readily saturable with time and is eliminated by either washing the cells with buffer or by the addition of excess unlabeled substrate.  相似文献   

2.
Myoglobin is a cytoplasmic hemoprotein, expressed solely in cardiac myocytes and oxidative skeletal muscle fibers, that reversibly binds O2 by its heme residue. Myoglobin is an essential oxygen-storage hemoprotein capable of facilitating oxygen transport and modulating nitric oxide homeostasis within cardiac and skeletal myocytes. Functionally, myoglobin is well accepted as an O2- storage protein in muscle, capable of releasing O2 during periods of hypoxia or anoxia. There is no evidence available regarding active sites, ligand binding sites, antigenic determinants and the ASA value of myoglobin in Channa striata. We further document the predicted active sites in the structural model with solvent exposed ASA residues. During this study, the model was built by CPH program and validated through PROCHECK, Verify 3D, ERRAT and ProSA for reliability. The active sites were predicted in the model with further ASA analysis of active site residues. The discussed information thus provides the predicted active sites, ligand binding sites, antigenic determinants and ASA values of myoglobin model in Channa striata.  相似文献   

3.
The SdcF transporter from Bacillus licheniformis (gene BL02343) is a member of the divalent anion sodium symporter (DASS)/SLC13 family that includes Na+/dicarboxylate transporters from bacteria to humans. SdcF was functionally expressed in Escherichia coli (BL21) and assayed in right side out membrane vesicles. ScdF catalyzed the sodium-coupled transport of succinate and α-ketoglutarate. Succinate transport was strongly inhibited by malate, fumarate, tartrate, oxaloacetate and l-aspartate. Similar to the other DASS transporters, succinate transport by SdcF was inhibited by anthranilic acids, N-(p-amylcinnamoyl) anthranilic acid and flufenamate. SdcF transport was cation-dependent, with a K 0.5 for sodium of ~1.5 mM and a K 0.5 for Li+ of ~40 mM. Succinate transport kinetics by SdcF were sigmoidal, suggesting that SdcF may contain two cooperative substrate binding sites. The results support an ordered binding mechanism for SdcF in which sodium binds first and succinate binds last. We conclude that SdcF is a secondary active transporter for four- and five-carbon dicarboxylates that can use Na+ or Li+ as a driving cation.  相似文献   

4.
The phospholipid composition of yeast plasma membrane was manipulated by two different methods: (i) by using two auxotrophic strains KA101 (cho1) and MC13 (Cho+) which required phospholipid bases for growth and (ii) by supplementing Saccharomyces cerevisiae (3059) cells with high concentration of choline or ethanolamine. It was possible to enrich the plasma membrane with phosphatidylcholine (PC) or phosphatidylethanolamine (PE) by both methods. The uptake of amino acids, e.g., glycine, glutamic acid, leucine, lysine methionine, phenylalanine, proline and serine, was significantly reduced in PC- or PE-enriched cells. However, the extent of reduction in transport was variable among different strains. A fluorescent probe, 1-anilino-8-naphthalene sulfonate (ANS), was used to monitor the structural changes induced by altered phospholipid composition. It was observed that the relative fluorescence intensity of bound ANS was decreased as a consequence of PC or PE enrichment. The decrease in fluorescence was probably associated with reduced number of available binding sites (n) and increased apparent dissociation constant (Kd). Furthermore, our results also suggest that a critical level of PE or PC is required for proper functioning of yeast membrane.  相似文献   

5.
l-[3H]Glutamate binding to synaptic plasma membranes from rat cerebral cortices was carried out at 2–4°C in 50 mM Tris-acetate buffer (pH 7.4) using a microfuge centrifugation method. Binding was increased by repeated freezing-thawing and washing in either crude or partially purified synaptic membranes. Scatchard analysis showed a single binding site (dissociation constant, KD = 697 nM; maximal binding capacity, Bmax = 7.5 pmol/mg protein) in four times distilled water washed crude synaptic membrane. After six times freezing-thawing and washing, a new high affinity site (KD1 = 26 nM, Bmax1 = 1.8 pmol/mg protein) appeared and the number of low affinity site was increased with no apparent change in affinity (KD2 = 662 nM, Bmax2 = 10.5 pmol/mg protein). l-[3H]Glutamate binding was inhibited by acidic amino acid analogues that interact with N-methyl-d-aspartate- and quisqualate-sensitive sites of glutamate receptors. Binding was marginally inhibited by kainate and l-2-amino-4-phosphonobutyrate. These results indicate that repeatedly frozen-thawed and washed synaptic plasma membrane is suitable for studying the subtypes and regulation of glutamate receptors.  相似文献   

6.
7.
The zero-trans influx of 500 μM uridine by CHO, P388, L1210 and L929 cells was inhibited by nitrobenzylthioinosine (NBTI) in a biphasic manner; 60–70% of total uridine influx by CHO cells and about 90% of that in P388, L1210 and L929 cells was inhibited by nmolar concentrations of NBTI (ID50 = 3?10 nM) and is designated NBTI-sensitive transport. The residual transport activity, designated NBTI-resistant transport, was inhibited by NBTI only at concentrations above 1 μM (ID50 = 10?50 μM). S49 cells exhibited only NBTI-sensitive uridine transport, whereas Novikoff cells exhibited only NBTI-resistant uridine transport. In all instances NBTI-sensitive transport correlated with the presence of between 7·104 and 7·105 high-affinity NBTI binding sites/cell (Kd = 0.3?1 nM). Novikoff cells lacked such sites. The two types of nucleoside transport, NBTI-resistant and NBTI-sensitive, were indistinguishable in substrate affinity, temperature dependence, substrate specificity, inhibition by structurally unrelated substances, such as dipyridamole or papaverine, and inhibition by sulfhydryl reagents or hypoxanthine. We suggest, therefore, that a single nucleoside transporter can exist in an NBTI-sensitive and an NBTI-resistant form depending on its disposition in the plasma membrane. The sensitive form expresses a high-affinity NBTI binding site(s) which is probably made up of the substrate binding site plus a hydrophobic region which interacts with the lipophilic nitrobenzyl group of NBTI. The latter site seems to be unavailable in NBTI-resistant transporters. The proportion of NBTI-resistant and sensitive uridine transport was constant during proportion of NBTI-resistant and sensitive uridine transport was constant during progression of P388 cells through the cell cycle and independent of the growth stage of the cells in culture. There were additional differences in uridine transport between cell lines which, however, did not correlate with NBTI sensitivity and might be related to the species origin of the cells. Uridine transport in Novikoff cells was more sensitive to inhibition by dipyridamole and papaverine than that in all other cell lines tested, whereas uridine transport in CHO cells was the most sensitive to inactivation by sulfhydryl reagents.  相似文献   

8.
Corynebacterium glutamicum is equipped with abundant membrane transporters to adapt to a changing environment. Many amino acid transporters have been identified in C. glutamicum, but histidine uptake has not been investigated in detail. Here, we identified the aromatic amino acid transporter encoded by aroP as a histidine transporter in C. glutamicum by a combination of the growth and histidine uptake features. Characterization of histidine uptake showed that AroP has a moderate affinity for histidine, with a Km value of 11.40 ± 2.03 μM, and histidine uptake by AroP is competitively inhibited by the aromatic amino acids. Among the four substrates, AroP exhibits a stronger preference for tryptophan than for tyrosine, phenylalanine, and histidine. Homology structure modeling and molecular docking were performed to predict the substrate binding modes and conformational changes during substrate transport. These results suggested that tryptophan is best accommodated in the binding pocket due to shape compatibility, strong hydrophobic interactions, and the lowest binding energy, which is consistent with the observed substrate preference of AroP. Furthermore, the missense mutations of the putative substrate binding sites verified that Ser24, Ala28, and Gly29 play crucial roles in substrate binding and are highly conserved in the Gram-positive bacteria. Finally, the expression of aroP is not significantly affected by extracellular histidine or aromatic amino acids, indicating that the physiological role of AroP may be correlated with the increased fitness of C. glutamicum to assimilate extracellular amino acid for avoiding the high energy cost of amino acid biosynthesis.  相似文献   

9.
Coliphage T4 endonuclease II (EndoII), encoded by gene denA, is a small (16 kDa, 136 aa) enzyme belonging to the GIY-YIG family of endonucleases, which lacks a C-terminal domain corresponding to that providing most of the binding energy in the structurally characterized GIY-YIG endonucleases, I-TevI and UvrC. In vivo, it is involved in degradation of host DNA, permitting scavenging of host-derived nucleotides for phage DNA synthesis. EndoII primarily catalyzes single-stranded nicking of DNA; 5- to 10-fold less frequently double-stranded breaks are produced. The Glu118Ala mutant of EndoII was crystallized in space group P21 with four monomers in the asymmetric unit. The fold of the EndoII monomer is similar to that of the catalytic domains of UvrC and I-TevI. In contrast to these enzymes, EndoII forms a striking X-shaped tetrameric structure composed as a dimer of dimers, with a protruding hairpin domain not present in UvrC or I-TevI providing most of the dimerization and tetramerization interfaces. A bound phosphate ion in one of the four active sites of EndoII likely mimics the scissile phosphate in a true substrate complex. In silico docking experiments showed that a protruding loop containing a nuclease-associated modular domain 3 element is likely to be involved in substrate binding, as well as residues forming a separate nucleic acid binding surface adjacent to the active site. The positioning of these sites within the EndoII primary dimer suggests that the substrate would bind to a primary EndoII dimer diagonally over the active sites, requiring significant distortion of the enzyme or the substrate DNA, or both, for simultaneous nicking of both DNA strands. The scarcity of potential nucleic acid binding residues between the active sites indicates that EndoII may bind its substrate inefficiently across the two sites in the dimer, offering a plausible explanation for the catalytic preponderance of single-strand nicks. Mutations analyzed in earlier functional studies are discussed in their structural context.  相似文献   

10.
2-Nitro-4-azidocarbonylcyanide phenylhydrazone (N3CCP), a potent water-soluble uncoupler at pH 6–8, was used to determine the nature of binding of the uncoupler to the mitochondrial membrane. Equilibrium binding studies with N3CCP showed that isolated pigeon heart mitochondria contain 1.6 ± 0.3 high-affinity binding sites per cytochrome a. Several different types of chemical uncouplers were also found to bind to the same high-affinity site as evidenced by their observed competition with N3CCP. The potassium ionophore valinomycin and the respiratory inhibitor antimycin A did not affect uncoupler binding to the high-affinity sites nor did active respiration of the mitochondria. The number of high-affinity binding sites was essentially unchanged by extraction of 80% of the mitochondrial phospholipids. The ability of the uncouplers to bind to the high-affinity binding sites is proportional to the uncoupler activities. These data support the idea that the high-affinity binding sites of mitochondria are protein(s) which are involved in the coupling reactions of oxidative phosphorylation and that uncoupler bound at these sites is responsible for the uncoupling activity.  相似文献   

11.
Summary Time courses of phlorizin binding to the outside of membrane vesicles from porcine renal outer cortex and outer medulla were measured and the obtained families of binding curves were fitted to different binding models. To fit the experimental data a model with two binding sites was required. Optimal fits were obtained if a ratio of low and high affinity phlorizin binding sites of 1:1 was assumed. Na+ increased the affinity of both binding sites. By an inside-negative membrane potential the affinity of the high affinity binding site (measured in the presence of 3 mM Na+) and of the low affinity binding site (measured in the presence of 3 or 90 mM Na+) was increased. Optimal fits were obtained when the rate constants of dissociation were not changed by the membrane potential. In the presence of 90 mM Na+ on both membrane sides and with a clamped membrane potential,K D values of 0.4 and 7.9 M were calculated for the low and high affinity phlorizin binding sites which were observed in outer cortex and in outer medulla. Apparent low and high affinity transport sites were detected by measuring the substrate dependence ofd-glucose uptake in membrane vesicles from outer cortex and outer medulla which is stimulated by an initial gradient of 90 mM Na+(out>in). Low and high affinity transport could be fitted with identicalK m values in outer cortex and outer medulla. An inside-negative membrane potential decreased the apparentK m ofhigh affinity transport whereas the apparentK m of low affinity transport was not changed. The data show that in outer cortex and outer medulla of pighigh and low affinity Na+-d-glucose cotransporters are present which containlow and high affinity phlorizin binding sites, respectively. It has to be elucidated from future experiments whether equal amounts of low and high affinity transporters are expressed in both kidney regions or whether the low and high affinity transporter are parts of the same glucose transport moleculc.  相似文献   

12.
A theoretical analysis of non-stationary states in membrane pores is given in this paper, which is based on the rate-theory treatment of transport processes. The principal aim of this study is to give a basis for the interpretation of relaxation experiments in which an external parameter, such as the voltage across the membrane, is suddenly displaced. From the time course of the membrane current information about the microscopic properties of the pore may be obtained. The pore is considered as a sequence of binding sites, separated by energy barriers over which the ion has to jump. It is found that under certain conditions damped oscillations occur after the initial perturbations of the membrane. In all other cases the approach towards the steady state may be described by a discrete spectrum of n relaxation times, where n is the number of binding sites within the pore. In the case of a pore with regular energy profile (internal barriers of identical height) the relaxation times may be obtained as the roots of Tchebycheff polynomials for arbitrary n. It is shown that the present treatment becomes identical with the continuum analysis of transport processes in the limit of large n.  相似文献   

13.
The mitochondrial glutamate-aspartate exchange carrier catalyzes the electrogenic exchange of intramitochondrial aspartate for extramitochondrial glutamate. Protons are cotransported with glutamate in a 1:1 ratio. In the present study, the effects of pH and glutamate concentration on glutamate entry into intact mitochondria were determined. Hydrogen ions were found to decrease the Km for glutamate entry. In addition, using glutamate-loaded submitochondrial particles, aspartate transport into the particles was measured as a function of internal and external glutamate concentrations, pH, and electrical potential across the membrane. Glutamate, was a competitive inhibitor of aspartate transport when both amino acids were present on the same side of the membrane, while H+ was a noncompetitive inhibitor of aspartate entry into the particles. A decrease in glutamate concentration on the inside of the particles brought about a parallel decrease in V and Km for aspartate outside of the particles, thus suggesting a ping-pong mechanism for the carrier. The uncoupling agent, carbonyl cyanide p-trifluoromethoxy-phenylhydrazone (FCCP), lowered both the Km and V of aspartate transport, while the effect on V was somewhat larger. Data obtained in the presence of KSCN was similar to that obtained with FCCP, and therefore it is concluded that both Km and V changes are dependent on a change of electrical potential across the membrane. A model for the carrier is proposed, which is consistent with the data presented. The model includes a single binding site specific for either glutamate or aspartate, and a separate binding site for the cotransported proton. The affinity of the binding site for protons is increased by simultaneous glutamate binding, but decreased by aspartate binding. The data suggest that an increase in the membrane potential increases the mobility of the charged carrier-aspartate complex, but also facilitates some additional step in the exchange cycle involving subsequent return of the carrier to the matrix side of the membrane. The additional membrane-potential-dependent step could be proton binding on the cytosolic side of the carrier.  相似文献   

14.
Secondary active transporters couple the free energy of the electrochemical potential of one solute to the transmembrane movement of another. As a basic mechanistic explanation for their transport function the model of alternating access was put forward more than 40 years ago, and has been supported by numerous kinetic, biochemical and biophysical studies. According to this model, the transporter exposes its substrate binding site(s) to one side of the membrane or the other during transport catalysis, requiring a substantial conformational change of the carrier protein. In the light of recent structural data for a number of secondary transport proteins, we analyze the model of alternating access in more detail, and correlate it with specific structural and chemical properties of the transporters, such as their assignment to different functional states in the catalytic cycle of the respective transporter, the definition of substrate binding sites, the type of movement of the central part of the carrier harboring the substrate binding site, as well as the impact of symmetry on fold-specific conformational changes. Besides mediating the transmembrane movement of solutes, the mechanism of secondary carriers inherently involves a mechanistic coupling of substrate flux to the electrochemical potential of co-substrate ions or solutes. Mainly because of limitations in resolution of available transporter structures, this important aspect of secondary transport cannot yet be substantiated by structural data to the same extent as the conformational change aspect. We summarize the concepts of coupling in secondary transport and discuss them in the context of the available evidence for ion binding to specific sites and the impact of the ions on the conformational state of the carrier protein, which together lead to mechanistic models for coupling.  相似文献   

15.
Cytochrome c2 was removed by washing from heavy chromatophores prepared from Rhodopseudomonas capsulata cells. The easy removal of the cytochrome could indicate that it was attached on the outside of the membrane. Therefore, the membrane was probably oriented inside out in relation to the membrane of regular chromatophores, from which cytochrome c2 could not be removed. Washing of the heavy chromatophores caused loss of photphosphorylation activity. The activity was restored to the resolved heavy chromatophores by the supernatant obtained during the washing or by the native cytochrome c2, which was found to be the active component in this supernatant. The activity could not be restored by other c-type cytochromes. Ascorbate, which enhanced photophosphorylation activity in the heavy chromatophores at the optimal concentration of 8 mm, restored this activity to the washed heavy chromatophores, but at an optimum concentration of 50 mm. Cytochrome c2 and dichlorophenol indophenol reduced the optimum of the ascorbate concentration to 7 mm. This might indicate that the effect of ascorbate is mediated through cytochrome c2. Washing the heavy chromatophores caused 70% loss of the light-induced electron transport from ascorbate and from ascorbate-reduced dichlorophenol indophenol to O2. However, this effect was only observed with the lower concentrations of ascorbate and the dye. The activity was restored either by the supernatant obtained from the washing or by various c-type cytochromes, reduced by ascorbate. Washing the heavy chromatophores did not affect succinate oxidation in the dark. It is suggested that cytochrome c2 is one of the cytochromes catalyzing the photosynthetic cyclic electron transport, as has been seen from its high specificity in the reconstitution experiments. Light can induce oxidation of various c-type cytochromes and other redox reagents. However, reduction was specific for cytochrome c2 from Rps. capuslata, since it was the only one which could be both reduced and oxidized as required from a component which is part of a cyclic electron transport chain. It is also suggested that cytochrome c2 was not part of the succinate oxidase system.  相似文献   

16.
The effects of phospholipase A2 treatment on the tetrodotoxin receptors in Electrophorus electricus was studied. (1) The binding of [3H]tetrodotoxin to electroplaque membranes was substantially reduced by treatment of the membranes with low concentrations of phospholipase A2 from a number of sources, including bee venom, Vipera russelli and Crotalus adamanteus and by β-bungarotoxin. (2) Phospholipase A2 from bee venom and from C. adamanteus both caused extensive hydrolysis of electroplaque membrane phospholipids although the substrate specificity differed. Analysis of the phospholipid classes hydrolyzed revealed a striking correlation between loss of toxin binding and hydrolysis of phosphatidylethanolamine but not of phosphatidylserine. (3) The loss of toxin binding could be partially reversed by treatment of the membranes with bovine serum albumin, conditions which are known to remove hydrolysis products from the membrane. (4) Equilibrium binding studies on the effects of phospholipase A2 treatment on [3H]tetrodotoxin binding showed that the reduction reflected loss of binding sites and not a change in affinity. (5) These results are interpreted in terms of multiple equilibrium states of the tetrodotoxin-receptors with conformations determined by the phospholipid environment.  相似文献   

17.
Bush DR 《Plant physiology》1990,93(4):1590-1596
The electrogenicity, pH-dependence, and stoichiometry of the proton-sucrose symport were examined in plasma membrane vesicles isolated from sugar beet (Beta vulgaris L. cv Great Western) leaves. Symport mediated sucrose transport was electrogenic as demonstrated by the effect of membrane potential on ΔpH-dependent flux. In the absence of significant charge compensation, a low rate of sucrose transport was observed. When membrane potential was clamped at zero with symmetric potassium concentrations and valinomycin, the rate of sucrose flux was stimulated fourfold. In the presence of a negative membrane potential, transport increased six-fold. These results are consistent with electrogenic sucrose transport which results in a net flux of positive charge into the vesicles. The effect of membrane potential on the kinetics of sucrose transport was on Vmax only with no apparent change in Km. Sucrose transport rates driven by membrane potential only, i.e. in the absence of ΔpH, were comparable to ΔpH-driven flux. Both membrane potential and ΔpH-driven sucrose transport were used to examine proton binding to the symport and the apparent Km for H+ was 0.7 micromolar. The kinetics of sucrose transport as a function of proton concentration exhibited a simple hyperbolic relationship. This observation is consistent with kinetic models of ion-cotransport systems when the stoichiometry of the system, ion:substrate, is 1:1. Quantitative measurements of proton and sucrose fluxes through the symport support a 1:1 stoichiometry. The biochemical details of protoncoupled sucrose transport reported here provide further evidence in support of the chemiosmotic hypothesis of nutrient transport across the plant cell plasma membrane.  相似文献   

18.
Prokaryotic and eukaryotic Na+-driven transporters couple the movement of one or more Na+ ions down their electrochemical gradient to the active transport of a variety of solutes. When more than one Na+ is involved, Na+-binding data are usually analyzed using the Hill equation with a non-integer exponent n. The results of this analysis are an overall Kd-like constant equal to the concentration of ligand that produces half saturation and n, a measure of cooperativity. This information is usually insufficient to provide the basis for mechanistic models. In the case of transport using two Na+ ions, an n < 2 indicates that molecules with only one of the two sites occupied are present at low saturation. Here, we propose a new way of analyzing Na+-binding data for the case of two Na+ ions that, by taking into account binding to individual sites, provides far more information than can be obtained by using the Hill equation with a non-integer coefficient: it yields pairs of possible values for the Na+ affinities of the individual sites that can only vary within narrowly bounded ranges. To illustrate the advantages of the method, we present experimental scintillation proximity assay (SPA) data on binding of Na+ to the Na+/I symporter (NIS). SPA is a method widely used to study the binding of Na+ to Na+-driven transporters. NIS is the key plasma membrane protein that mediates active I transport in the thyroid gland, the first step in the biosynthesis of the thyroid hormones, of which iodine is an essential constituent. NIS activity is electrogenic, with a 2:1 Na+/I transport stoichiometry. The formalism proposed here is general and can be used to analyze data on other proteins with two binding sites for the same substrate.  相似文献   

19.
ω-Transaminase (ω-TA) is the only naturally occurring enzyme allowing asymmetric amination of ketones for production of chiral amines. The active site of the enzyme was proposed to consist of two differently sized substrate binding pockets and the stringent steric constraint in the small pocket has presented a significant challenge to production of structurally diverse chiral amines. To provide a mechanistic understanding of how the (S)-specific ω-TA from Paracoccus denitrificans achieves the steric constraint in the small pocket, we developed a free energy analysis enabling quantification of individual contributions of binding and catalytic steps to changes in the total activation energy caused by structural differences in the substrate moiety that is to be accommodated by the small pocket. The analysis exploited kinetic and thermodynamic investigations using structurally similar substrates and the structural differences among substrates were regarded as probes to assess how much relative destabilizations of the reaction intermediates, i.e. the Michaelis complex and the transition state, were induced by the slight change of the substrate moiety inside the small pocket. We found that ≈80% of changes in the total activation energy resulted from changes in the enzyme-substrate binding energy, indicating that substrate selectivity in the small pocket is controlled predominantly by the binding step (KM) rather than the catalytic step (kcat). In addition, we examined the pH dependence of the kinetic parameters and the pH profiles of the KM and kcat values suggested that key active site residues involved in the binding and catalytic steps are decoupled. Taken together, these findings suggest that the active site residues forming the small pocket are mainly engaged in the binding step but not significantly involved in the catalytic step, which may provide insights into how to design a rational strategy for engineering of the small pocket to relieve the steric constraint toward bulky substituents.  相似文献   

20.
Rat intestinal alkaline phosphatase is a dimeric enzyme with identical subunits and thus possesses two presumably identical active sites. Binding studies with Pi and l-phenylalanine and pre-steady-state “burst” titrations confirm the existence of two active sites per molecule of enzyme. The sites appear to be nonequivalent with respect to Pi binding, both at low pH, where an enzyme (E)-Pi covalent complex is formed, and at high Pi, where an E-Pi noncovalent complex predominates. The binding affinity of the first site is 100-fold greater than that of the second, i.e., there is negative cooperativity. The Ki value for competitive inhibition of substrate hydrolysis by Pi corresponds to the higher affinity site. The negative cooperativity appears not to be an artifact resulting from contaminating Pi in the purified enzyme preparation. l-Phenylalanine does not bind to the enzyme unless Pi is present, as expected from the previously proposed mechanism of uncompetitive inhibition by the amino acid. No negative cooperativity is seen in l-phenylalanine binding, but the number of moles of amino acid bound at saturation depends on the degree of saturation by Pi The enzyme is also inhibited uncompetitively by NADH, which can compete with l-phenylalanine for the same site on alkaline phosphatase.  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