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.
Bai S  Jain MK  Berg OG 《Biochemistry》2008,47(9):2899-2907
Pig pancreatic IB phospholipase A 2 (PLA2) forms three distinguishable premicellar E i (#) ( i = 1, 2, and 3) complexes at successively higher decylsulfate concentrations. The Hill coefficient for E 1 (#) is n 1 = 1.6, and n 2 and n 3 for E 2 (#) and E 3 (#) are about 8 each. Saturation-transfer difference nuclear magnetic resonance (NMR) and other complementary results with PLA2 show that decylsulfate molecules in E 2 (#) and E 3 (#) are contiguously and cooperatively clustered on the interface-binding surface or i-face that makes contact with the substrate interface. In these complexes, the saturation-transfer difference NMR signatures of (1)H in decylsulfate are different. The decylsulfate epitope for the successive E i (#) complexes increasingly resembles the micellar complex formed by the binding of PLA2 to preformed micelles. Contiguous cooperative amphiphile binding is predominantly driven by the hydrophobic effect with a modest electrostatic shielding of the sulfate head group in contact with PLA2. The formation of the complexes is also associated with structural change in the enzyme. Calcium affinity of E 2 (#) appears to be modestly lower than that of the free enzyme and E 1 (#). Binding of decylsulfate to the i-face does not require the catalytic calcium required for the substrate binding to the active site and for the chemical step. These results show that E i (#) complexes are useful to structurally characterize the cooperative sequential and contiguous binding of amphiphiles on the i-face. We suggest that the allosteric changes associated with the formation of discrete E i (#) complexes are surrogates for the catalytic and allosteric states of the interface activated PLA2.  相似文献   

2.
Yu BZ  Apitz-Castro R  Tsai MD  Jain MK 《Biochemistry》2003,42(20):6293-6301
Pancreatic IB phospholipase A(2) (PLA2) forms aggregates of defined size with monodisperse alkyl sulfates in the premicellar concentration range. As an extension of the interfacial kinetic paradigm, results are interpreted in terms of a model in which several amphiphile molecules bind along their polar headgroup to the interface binding region (i-face) of PLA2. The resulting complex, E(#), has a half-micellar structure, and it acts as an "amphiphile" in the aqueous phase. E(#) not only self-aggregates but also binds hydrophobic probes and interacts with hydrophobic surfaces. As expected, resonance energy transfer from the tryptophan donor in PLA2 to an acceptor probe partitioned in E(#) shows a biphasic dependence as the probe coexisting with PLA2 is diluted at higher alkyl sulfate concentrations. The gel-permeation behavior of PLA2 at premicellar alkyl sulfate concentrations is also biphasic. For example, above 1.2 mM decyl sulfate (CMC = 3.5 mM) PLA2 elutes as a single sharp peak, presumably the self-aggregate of E(#) with apparent molecular mass of 120-150 kDa. At 0.4-1 mM decyl sulfate the retention volume is even larger than that for the 14 kDa PLA2. This anomalous retention is attributed to the interaction of the hydrophobic region of E(#) with the hydrophobic patches on the gel-permeation matrix. Elution behavior of the self-aggregated E(#) form of site-directed mutants in dodecyl sulfate suggests that certain substitutions in the conserved hydrogen-bonding network have a significant effect on the aggregate size. These results suggest a role for the network in the amphiphile binding along the i-face of PLA2, presumably through a change in the anion coordination ligands.  相似文献   

3.
The changes in the microenvironment of the Trp-3 on the i-face of pig pancreatic IB phospholipase A2 (PLA2) provide a measure of the tight contact (Ramirez and Jain, Protein Sci. 9, 229-239, 1991) with the substrate interface during the processive interfacial turnover. Spectral changes from the single Trp-substituent at position 1, 2, 6, 10, 19, 20, 31, 53, 56 or 87 on the surface of W3F PLA2 are used to probe the Trp-environment. Based on our current understanding only the residue 87 is away from i-face, therefore all other mutants are well suited to report modest differences along the i-face. All Trp-mutants bind tightly to anionic vesicles. Only those with Trp at 1, 2 or 3 near the rim of the active site on the i-face cause significant perturbation of the catalytic functions. Most other Trp-mutants showed < 3-fold change in the interfacial processive turnover rate and the competitive inhibition by MJ33. Binding of calcium to the enzyme in the aqueous phase had modest effect on the Trp-emission intensity. However, on the binding of the enzyme to the interface the fluorescence change is large, and the rate of oxidation of the Trp-substituent with N-bromosuccinimide depends on the location of the Trp-substituent. These results show that the solvation environment of the Trp-substituents on the i-face is shielded in the enzyme bound to the interface. Additional changes are noticeable if the active site of the bound enzyme is also occupied, however, the catalytically inert zymogen of PLA2 (proPLA2) does not show such changes. Significance of these results in relation to the changes in the solvent accessibility and desolvation of the i-face of PLA2 at the interface is discussed.  相似文献   

4.
Berg OG  Yu BZ  Apitz-Castro RJ  Jain MK 《Biochemistry》2004,43(7):2080-2090
Phosphatidylinositol-specific phospholipase C (PI-PLC) from Bacillus cereus forms a premicellar complex E(#) with monodisperse diheptanoylphosphatidylcholine (DC(7)PC) that is distinguishable from the E complex formed with micelles. Results are interpreted with the assumption that in both cases amphiphiles bind to the interfacial binding surface (i-face) of PI-PLC but not to the active site. Isothermal calorimetry and fluorescence titration results for the binding of monodisperse DC(7)PC give an apparent dissociation constant of K(2) = 0.2 mM with Hill coefficient of 2. The gel-permeation, spectroscopic, and probe partitioning behaviors of E(#) are distinct from those of the E complex. The aggregation and partitioning behaviors suggest that the acyl chains in E(#) but not in E remain exposed to the aqueous phase. The free (E) and complexed (E(#) and E) forms of PI-PLC, each with distinct spectroscopic signatures, readily equilibrate with changing DC(7)PC concentration. The underlying equilibria are modeled and their significance for the states of the PI-PLC under monomer kinetic conditions is discussed to suggest that the Michaelis-Menten complex formed with monodisperse DC(7)PC is likely to be E(#)S or its aggregate rather than the classical monodisperse ES complex.  相似文献   

5.
During the steady state reaction progress in the scooting mode with highly processive turnover, Bacillus cereus sphingomyelinase (SMase) remains tightly bound to sphingomyelin (SM) vesicles (Yu et al., Biochim. Biophys. Acta 1583, 121-131, 2002). In this paper, we analyze the kinetics of SMase-catalyzed hydrolysis of SM dispersed in diheptanoylphosphatidyl-choline (DC7PC) micelles. Results show that the resulting decrease in the turnover processivity induces the stationary phase in the reaction progress. The exchange of the bound enzyme (E*) between the vesicle during such reaction progress is mediated via the premicellar complexes (E(i)#) of SMase with DC7PC. Biophysical studies indicate that in E(i)# monodisperse DC7PC is bound to the interface binding surface (i-face) of SMase that is also involved in its binding to micelles or vesicles. In the presence of magnesium, required for the catalytic turnover, three different complexes of SMase with monodisperse DC7PC (E(i)# with i=1, 2, 3) are sequentially formed with Hill coefficients of 3, 4 and 8, respectively. As a result, during the stationary phase reaction progress, the initial rate is linear for an extended period and all the substrate in the reaction mixture is hydrolyzed at the end of the reaction progress. At low mole fraction (X) of total added SM, exchange is rapid and the processive turnover is limited by the steps of the interfacial turnover cycle without becoming microscopically limited by local substrate depletion or enzyme exchange. At high X, less DC7PC will be monodisperse, E(i)# does not form and the turnover becomes limited by slow enzyme exchange. Transferred NOESY enhancement results show that monomeric DC7PC in solution is in a rapid exchange with that bound to E(i)# at a rate comparable to that in micelles. Significance of the exchange and equilibrium properties of the E(i)# complexes for the interpretation of the stationary phase reaction progress is discussed.  相似文献   

6.
We report the structures of the crystallographic dimer of porcine pancreatic IB phospholipase A(2) (PLA2) with either five sulfate or phosphate anions bound. In each structure, one molecule of a tetrahedral mimic MJ33 [1-hexadecyl-3-(trifluoroethyl)-sn-glycero-2-phosphomethanol] and the five anions are shared between the two subunits of the dimer. The sn-2-phosphate of MJ33 is bound in the active site of one subunit (A), and the alkyl chain extends into the active site slot of the second subunit (B) across the subunit-subunit interface. The two subunits are packed together with a large hydrophobic and desolvated surface buried between them along with the five anions that define a plane. The anions bind by direct contact with two cationic residues (R6 and K10) per subunit and through closer-range H-bonding interactions with other polarizable ligands. These features of the "dimer" suggest that the binding of PLA2 to the anionic groups at the anionic interface may be dominated by coordination through H-bonding with only a partial charge compensation needed. Remarkably, the plane defined by the contact surface is similar to the i-face of the enzyme [Ramirez, F., and Jain, M. K. (1991) Proteins: Struct., Funct., Genet. 9, 229-239], which has been proposed to make contact with the substrate interface for the interfacial catalytic turnover. Additionally, these structures not only offer a view of the active PLA2 complexed to an anionic interface but also provide insight into the environment of the tetrahedral intermediate in the rate-limiting chemical step of the turnover cycle. Taken together, our results offer an atomic-resolution structural view of the i-face interactions of the active form of PLA2 associated to an anionic interface.  相似文献   

7.
S-Adenosylmethionine (AdoMet) synthetase catalyzes the only known route of biosynthesis of the primary in vivo alkylating agent. Inhibitors of this enzyme could provide useful modifiers of biological methylation and polyamine biosynthetic processes. The AdoMet synthetase catalyzed reaction converts ATP and L-methionine to AdoMet, PP(i), and P(i), with formation of tripolyphosphate as a tightly bound intermediate. This work describes a nonhydrolyzable analogue of the tripolyphosphate (PPP(i)) reaction intermediate, diimidotriphosphate (O(3)P-NH-PO(2)-NH-PO(3)(5)(-)), as a potent inhibitor. In the presence of AdoMet, PNPNP is a slow-binding inhibitor with an overall inhibition constant (K(i)) of 2 nM and a dissociation rate of 0.6 h(-)(1). In contrast, in the absence of AdoMet PNPNP is a classical competitive inhibitor with a K(i) of 0.5 microM, a slightly higher affinity than PPP(i) itself (K(i) = 3 microM). The imido analogue of the product pyrophosphate, imidodiphosphate (O(3)P-NH-PO(3)(4)(-)) also displays slow onset inhibition only in the presence of AdoMet, with a K(i) of 0.8 microM, compared to K(i) of 250 microM for PP(i). Circular dichroism spectra of the unliganded enzyme and various complexes are indistinguishable indicating that the protein secondary structure is not greatly altered upon complex formation, suggesting local rearrangements at the active site during the slow binding process. A model based on ionization of the bridging -NH- moiety is presented which could account for the potent inhibition by PNP and PNPNP.  相似文献   

8.
Boegeman SC  Deems RA  Dennis EA 《Biochemistry》2004,43(13):3907-3916
Equilibrium dialysis was used to study the binding of two nonhydrolyzable, short chain phospholipid analogues to the secreted group IA phospholipase A(2) (PLA(2)), which has been shown to contain several phospholipid binding sites that dramatically affect activity. This study provides new insight into how these activations occur. One analogue contained a phosphorylethanolamine (DiC(6)SNPE) headgroup, while the other contained a phosphorylcholine (DiC(6)SNPC) headgroup. Using phospholipase D, we incorporated tritium into each analogue. No binding of DiC(6)SNPE to PLA(2) was observed under submicellar conditions. Addition of submicellar amounts of Triton X-100 resulted in a linear nonsaturating response to lipid concentration, suggestive of premicellar aggregation of the DiC(6)SNPE with Triton X-100 and PLA(2). Binding of DiC(6)SNPE when presented as Triton X-100 mixed micelles saturated at 0.93 binding sites per PLA(2) with a K(D) of 38 microM. Addition of sphingomyelin, a potent activator of PLA(2) hydrolysis of phosphorylethanolamine containing compounds, resulted in a 13-fold decrease in the K(D), to 2.8 microM. This suggests that changes in the catalytic site binding affinity contribute to "phosphatidylcholine activation". Binding of DiC(6)SNPC with 2.0 mM Triton X-100 showed positive cooperativity (Hill coefficient of 1.7), which saturated at 2.0 binding sites per PLA(2). No binding of either analogue was observed when the catalytic site was alkylated with p-bromophenacyl bromide. Since p-bromophenacyl bromide does not physically block the phosphatidylcholine activator site, this indicates that the two phosphatidylcholine binding sites interact. The binding studies show that DiC(6)SNPC binds cooperatively to two sites on group IA PLA(2), while DiC(6)SNPE binds to only one site.  相似文献   

9.
The role of desolvation in protein binding kinetics is investigated using Brownian dynamics simulations in complexes in which the electrostatic interactions are relatively weak. We find that partial desolvation, modeled by a short-range atomic contact potential, is not only a major contributor to the binding free energy but also substantially increases the diffusion-limited rate for complexes in which long-range electrostatics is weak. This rate enhancement is mostly due to weakly specific pathways leading to a low free-energy attractor, i.e., a precursor state before docking. For alpha-chymotrypsin and human leukocyte elastase, both interacting with turkey ovomucoid third domain, we find that the forward rate constant associated with a collision within a solid angle phi around their corresponding attractor approaches 10(7) and 10(6) M(-1)s(-1), respectively, in the limit phi approximately 2 degrees. Because these estimates agree well with experiments, we conclude that the final bound conformation must be preceded by a small set of well-defined diffusion-accessible precursor states. The inclusion of the otherwise repulsive desolvation interaction also explains the lack of aggregation in proteins by restricting nonspecific association times to approximately 4 ns. Under the same reaction conditions but without short range forces, the association rate would be only approximately 10(3) M(-1)s(-1). Although desolvation increases these rates by three orders of magnitude, desolvation-mediated association is still at least 100-fold slower than the electrostatically assisted binding in complexes such as barnase and barstar.  相似文献   

10.
The changes in the microenvironment of the Trp-3 on the i-face of pig pancreatic IB phospholipase A2 (PLA2) provide a measure of the tight contact (Ramirez and Jain, Protein Sci. 9, 229-239, 1991) with the substrate interface during the processive interfacial turnover. Spectral changes from the single Trp-substituent at position 1, 2, 6, 10, 19, 20, 31, 53, 56 or 87 on the surface of W3F PLA2 are used to probe the Trp-environment. Based on our current understanding only the residue 87 is away from i-face, therefore all other mutants are well suited to report modest differences along the i-face. All Trp-mutants bind tightly to anionic vesicles. Only those with Trp at 1, 2 or 3 near the rim of the active site on the i-face cause significant perturbation of the catalytic functions. Most other Trp-mutants showed < 3-fold change in the interfacial processive turnover rate and the competitive inhibition by MJ33. Binding of calcium to the enzyme in the aqueous phase had modest effect on the Trp-emission intensity. However, on the binding of the enzyme to the interface the fluorescence change is large, and the rate of oxidation of the Trp-substituent with N-bromosuccinimide depends on the location of the Trp-substituent. These results show that the solvation environment of the Trp-substituents on the i-face is shielded in the enzyme bound to the interface. Additional changes are noticeable if the active site of the bound enzyme is also occupied, however, the catalytically inert zymogen of PLA2 (proPLA2) does not show such changes. Significance of these results in relation to the changes in the solvent accessibility and desolvation of the i-face of PLA2 at the interface is discussed.  相似文献   

11.
The interaction of fibrinogen with the mannose-specific lectins concanavalin A (ConA), its acetyl derivative (Ac-ConA) and Lens culinaris agglutinin (LcH) was studied. Both ConA and LcH interact specifically with individual fibrinogen B beta and gamma chains and with denatured fragments D and E. However, analysis of the binding data shows that four moles of Ac-ConA are bound per mole of fibrinogen with two sets of binding sites (Kd1 = 2.4 microM and Kd2 = 16.6 microM; n1 = n2 = 2) while only two moles of LcH are bound per mole of fibrinogen (Kd = 2.6 microM). Ultracentrifugation studies are also in agreement with the presence in the fibrinogen molecule of two and four binding sites for LcH and Ac-ConA, respectively. No aggregates of fibrinogen formed through LcH or Ac-ConA linkages are observed. The use of a crosslinking reagent and ultracentrifugal analysis of the lectin-fibrinogen fragments D1 and E complexes indicated that ConA, as well as Ac-ConA, interact with both fragments D and E while LcH interacts only with fragment D. Furthermore, the binding of ConA to both D and E domains in the intact fibrinogen molecule is clearly demonstrated by using a bifunctional reagent. The bivalent character of ConA tetramers may be misinterpreted as a lack of accessibility of the lectin to two of the four carbohydrate chains of fibrinogen. The differential binding of LcH and ConA to the carbohydrate chains of fibrinogen can be related to a different exposure of the oligosaccharide in D and E fragments and domains and to the different requirements of both lectins for their binding to glycoproteins.  相似文献   

12.
For membrane-associated enzymes, which access substrate from either a monolayer or bilayer of the aggregate substrate, the partitioning from the aqueous phase to this phospholipid interface is critical for catalysis. Despite a large and expanding body of knowledge regarding interfacial enzymes, the biophysical steps involved in interfacial recognition and adsorption remain relatively poorly understood. The surface of the enzyme that contacts the phospholipid surface is referred to as its interfacial binding surface, or more simply, its i-face. The interaction of a protein's i-face with the aggregate substrate may simply control access to substrate. However, it can be more complex, and this interaction often serves to allosterically activate the enzyme on this surface. First we briefly review what is currently known about i-face structure and function for a prototypical interfacial enzyme, the secreted Phospholipase A2 (PLA2). Then we develop, characterize, compare, and discuss models of the PLA2 i-face across a subset of five homologous PLA2 family members, groups IA, IB, IIA, V, and X. A homology model of human group-V is included in this comparison, suggesting that a similar approach could be used to explore interfacial function of any of the PLA2 family members. Despite moderate sequence identity, structural homology and sequence similarity are well conserved. We find that the residues predicted to be interfacial, while conserved structurally, are not highly conserved in sequence. Implications for this divergence on interfacial selectivity are discussed.  相似文献   

13.
A new class of surfactant-cobalt(III) complex ions of the type, cis-[Co(X)(2)(C(14)H(29)NH(2))Cl](2+) (where X=ethylenediamine (en), or 2,2'-bipyridyl (bpy), or 1,10-phenanthroline (phen)) and cis-[Co(trien)(C(14)H(29)NH(2))Cl](2+) (trien=triethylenetetramine) were synthesized and characterized by IR, NMR, UV-visible electronic absorption spectra, elemental analysis and metal analysis. The critical micelle concentration (CMC) values of these surfactant-cobalt(III) complexes in aqueous solution were obtained from conductance measurements. Specific conductivity data (at 298, 308, 318 and 328 K) served for the evaluation of the temperature-dependent CMC and the thermodynamics of micellization (DeltaG(0)(m), DeltaH(0)(m) and DeltaS(0)(m)). Interactions between calf thymus DNA and the surfactant-cobalt(III) complexes in aqueous solution have been investigated by electronic absorption spectroscopy, emission spectroscopy and viscosity measurements. The electrostatic interactions, van der Waals interactions and/or partial intercalative binding have been observed in these systems. The surfactant-cobalt(III) complexes were screened for their antibacterial and antifungal activities against various microorganisms. The results were compared with the standard drugs, Ciprofloxacin and Fluconazole respectively.  相似文献   

14.
K5 lyase of coliphage K5A degrades the K5 polysaccharide of encapsulated E. coli strains expressing the K5 antigen thereby contributing to virus binding and infection. We have investigated the affinities of the recombinant enzyme for different GAG ligands by isothermal fluorescence titrations and correlated them with substrate processing and protein structural changes. Chondroitin sulfate (CS) and heparan sulfate (HS) bound to K5 lyase with a Kd of 0.5 microM whereas heparin exhibited a Kd=1.1 microM. The natural substrate K5 polysaccharide displayed a similar apparent affinity as CS and HS but was the only ligand of the enzyme which induced a large structural rearrangement of the protein as detected by far-UV CD spectroscopy. Since significant enzymatic degradation was only found for the K5 polysaccharide peaking at 44 degrees C, but binding was also detected for heparin, we propose that the K5 lyase is able to discriminate between specific (acetylated/non-sulfated) and unspecific (acetylated/sulfated) ligands by its heparin binding motif in the C-terminus. This is proposed to be the origin for the enzyme's residual HS degrading activity.  相似文献   

15.
Yu BZ  Rogers J  Tsai MD  Pidgeon C  Jain MK 《Biochemistry》1999,38(15):4875-4884
Primary rate and equilibrium parameters for 60 site-directed mutants of bovine pancreatic phospholipase A2 (PLA2) are analyzed so incremental contributions of the substitution of specific residues can be evaluated. The magnitude of the change is evaluated so a functional role in the context of the N- and C-domains of PLA2 can be assigned, and their relationship to the catalytic residues and to the i-face that makes contact with the interface. The effect of substitutions and interfacial charge is characterized by the equilibrium dissociation constant for dissociation of the bound enzyme from the interface (Kd), the dissociation constant for dissociation of a substrate mimic from the active site of the bound enzyme (KL), and the interfacial Michaelis constants, KM and kcat. Activity is lost (>99.9%) on the substitution of H48 and D49, the catalytic residues. A more than 95% decrease in kcat is seen with the substitution of F5, I9, D99, A102, or F106, which form the substrate binding pocket. Certain residues, which are not part of the catalytic site or the substrate binding pocket, also modulate kcat. Interfacial anionic charge lowers Kd, and induces kcat activation through K56, K53, K119, or K120. Significant changes in KL are seen by the substitution of N6, I9, F22, Y52, K53, N71, Y73, A102, or A103. Changes in KM [=(k2+k-1)/k1] are attributed to kcat (=k2) and KL (=k-1/k1). Some substitutions change more than one parameter, implying an allosteric effect of the binding to the interface on KS, and the effect of the interfacial anionic charge on kcat. Interpreted in the context of the overall structure, results provide insights into the role of segments and domains in the microscopic events of catalytic turnover and processivity, and their allosteric regulation. We suggest that the interfacial recognition region (i-face) of PLA2, due to the plasticity of certain segments and domains, exercises an allosteric control on the substrate binding and chemical step.  相似文献   

16.
The murine adipocyte lipid binding protein (ALBP) has been cloned into Escherichia coli, purified from expressing cultures, and its ligand binding and phosphorylation properties studied. In the cloning strategy, the recombinant, pT7-5 rALBP, was transformed into E. coli strain K38 harboring plasmid pGP1-2 which directs the synthesis of T7 RNA polymerase. Upon shifting the temperature from 30 to 42 degrees C to induce T7 RNA polymerase expression, the 14.6-kDa recombinant ALBP (rALBP) was expressed for approximately 2 h and accumulated to about 1% of total E. coli protein. The recombinant ALBP was soluble in E. coli extracts and resistant to bacterial proteolysis. A procedure for purifying rALBP was developed utilizing immuno-chemical detection based upon reactivity with anti-murine ALBP antiserum. A combination of acidic ammonium sulfate fractionation, gel permeation chromatography, and carboxymethyl ion-exchange high performance liquid chromatography separation was used to prepare homogeneous rALBP. Sequence analysis of rALBP indicated that the initiating methionine residue had been removed and the amino-terminal cysteine residue was not blocked. Purified rALBP exhibited stoichiometric, saturable binding of oleic acid (n = 1.0, K0.5 approximately 100 microM) and retinoic acid (n = 1.0, K0.5 approximately 170 microM). Incubation of rALBP with wheat germ agglutinin-purified insulin receptor, ATP, and 100 nM insulin resulted in a 5-fold stimulation of rALBP phosphorylation above the basal state. Kinetic analysis of rALBP phosphorylation by the 3T3-L1 insulin receptor kinase yielded a Michaelis constant (Km) of 50 microM and a maximal velocity of 1 mol of rALBP phosphorylated/min/mol insulin binding sites. Phosphoamino acid analysis indicated that phosphorylation occurred upon tyrosine. These results indicate that murine ALBP has been cloned and expressed in E. coli, purified to homogeneity, and is a substrate for the insulin receptor tyrosyl kinase in vitro.  相似文献   

17.
2-C-Methyl-D-erythritol-4-phosphate synthase (MEP synthase) catalyzes the rearrangement/reduction of 1-D-deoxyxylulose-5-phosphate (DXP) to methylerythritol-4-phosphate (MEP) as the first pathway-specific reaction in the MEP biosynthetic pathway to isoprenoids. Recombinant E. coli MEP was purified by chromatography on DE-52 and phenyl-Sepharose, and its steady-state kinetic constants were determined: k(cat) = 116 +/- 8 s(-1), K(M)(DXP) = 115 +/- 25 microM, and K(M)(NADPH) = 0.5 +/- 0.2 microM. The rearrangement/reduction is reversible; K(eq) = 45 +/- 6 for DXP and MEP at 150 microM NADPH. The mechanism for substrate binding was examined using fosmidomycin and dihydro-NADPH as dead-end inhibitors. Dihydro-NADPH gave a competitive pattern against NADPH and a noncompetitive pattern against DXP. Fosmidomycin was an uncompetitive inhibitor against NADPH and gave a pattern representative of slow, tight-binding competitive inhibition against DXP. These results are consistent with an ordered mechanism where NADPH binds before DXP.  相似文献   

18.
The membrane/water partition coefficients, K, of 15 electrically neutral (non-charged or zwitterionic) detergents were measured with phospholipid vesicles by using isothermal titration calorimetry, and were compared to the corresponding critical micellar concentrations, cmc. The detergents measured were oligo(ethylene oxide) alkyl ethers (C(m)EO(n) with m = 10/n = 3, 7 and m = 12/n = 3.8); alkylglucosides (octyl, decyl); alkylmaltosides (octyl, decyl, dodecyl); diheptanoylphosphatidylcholine; Tritons (X-100, X-114) and CHAPS. A linear relation between the free energies of partitioning into the membrane and micelle formation was found such that K. CMC approximately 1. The identity K. CMC = 1 was used to classify detergents with respect to their membrane disruption potency. "Strong" detergents are characterized by K. CMC < 1 and solubilize lipid membranes at detergent-to-lipid ratios X(b) < 1 (alkylmaltosides, tritons, heptaethylene glycol alkyl ethers). "Weak" detergents are characterized by K. CMC > 1 and accumulate in the membrane- to detergent-to-lipid ratios X(b) > 1 before the bilayer disintegrates (alkylglucosides, pentaethylene glycol dodecyl ether).  相似文献   

19.
The binding of NADH and NAD+ to the human liver cytoplasmic, E1, and mitochondrial, E2, isozymes at pH 7.0 and 25 degrees C was studied by the NADH fluorescence enhancement technique, the sedimentation technique, and steady-state kinetics. The binding of radiolabeled [14C]NADH and [14C]NAD+ to the E1 isozyme when measured by the sedimentation technique yielded linear Scatchard plots with a dissociation constant of 17.6 microM for NADH and 21.4 microM for NAD+ and a stoichiometry of ca. two coenzyme molecules bound per enzyme tetramer. The dissociation constant, 19.2 microM, for NADH as competitive inhibitor was found from steady-state kinetics. With the mitochondrial E2 isozyme, the NADH fluorescence enhancement technique showed only one, high-affinity binding site (KD = 0.5 microM). When the sedimentation technique and radiolabeled coenzymes were used, the binding studies showed nonlinear Scatchard plots. A minimum of two binding sites with lower affinity was indicated for NADH (KD = 3-6 microM and KD = 25-30 microM) and also for NAD+ (KD = 5-7 microM and KD = 15-30 microM). A fourth binding site with the lowest affinity (KD = 184 microM for NADH and KD = 102 microM for NAD+) was observed from the steady-state kinetics. The dissociation constant for NAD+, determined by the competition with NADH via fluorescence titration, was found to be 116 microM. The number of binding sites found by the fluorescence titration (n = 1 for NADH) differs from that found by the sedimentation technique (n = 1.8-2.2 for NADH and n = 1.2-1.6 for NAD+).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Nitrophorins 1-4 (NP1-4) are ferriheme proteins from the blood-sucking insect Rhodnius prolixus that transport nitric oxide (NO) to the victim, sequester histamine, and inhibit blood coagulation. Here, we report kinetic and thermodynamic analyses for ligand binding by all four proteins and their reduction potentials. All four undergo biphasic association and dissociation reactions with NO. The initial association is fast (1.5-33 microM(-)(1) s(-)(1)) and similar to that of elephant metmyoglobin. However, unlike in metmyoglobin, a slower second phase follows ( approximately 50 s(-)(1)), and the stabilized final complexes are resistant to autoreduction (E degrees = +3 to +154 mV vs normal hydrogen electrode). NO dissociation begins with a slow, pH-dependent step (0.02-1.4 s(-)(1)), followed by a faster phase that is again similar to that of metmyoglobin (3-52 s(-)(1)). The equilibrium dissociation constants are quite small (1-850 nM). NP1 and NP4 display larger release rate constants and smaller association rate constants than NP2 and NP3, leading to values for K(d) that are about 10-fold greater. The results are discussed in light of the recent crystal structures of NP1, NP2, and NP4, which display open, polar distal pockets, and of NP4-NO, which displays an NO-induced conformational change that leads to expulsion of solvent and complete burial of the NO ligand in a now nonpolar distal pocket. Taken together, the results suggest that tighter NO binding in the nitrophorins is due to the trapping of the molecule in a nonpolar distal pocket rather than through formation of particularly strong Fe-NO or hydrogen bonds.  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