首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Peroxidases typically bind their reducing substrates weakly, with K(d) values in the millimolar range. The binding of benzhydroxamic acid (BHA) to ferric horseradish peroxidase isoenzyme C (HRPC) [K(d) = 2.4 microM; Schonbaum, G. R. (1973) J. Biol. Chem. 248, 502-511] is a notable exception and has provided a useful tool for probing the environment of the peroxidase aromatic-donor-binding site and the distal heme cavity. Knowledge of the underlying thermodynamic driving forces is key to understanding the roles of the various H-bonding and hydrophobic interactions in substrate binding. The isothermal titration calorimetry results of this study on the binding of aromatic hydroxamic acid analogues to ferric HRPC under nonturnover conditions (no H(2)O(2) present) confirm the significance of H-bonding interactions in the distal heme cavity in complex stabilization. For example, the binding of BHA to HRPC is enthalpically driven at pH 7.0, with the H-bond to the distal Arg38 providing the largest contribution (6.74 kcal/mol) to the binding energy. The overall relatively weak binding of the hydroxamic acid analogues to HRPC is due to large entropic barriers (-11.3 to -37.9 eu) around neutral pH, with the distal Arg38 acting as an "entropic gate keeper". Dramatic enthalpy-entropy compensation is observed for BHA and 2-naphthohydroxamic acid binding to HRPC at pH 4.0. The enthalpic loss and entropic gain are likely due to increased flexibility of Arg38 in the complexes at low pH and greater access by water to the active site. Since the Soret absorption band of HRPC is a sensitive probe of the binding of hydroxamic acids and their analogues, it was used to investigate the binding of six donor substrates over the pH range of 4-12. The negligible pH dependence of the K(d) values corrected for substrate ionization suggests that enthalpy-entropy compensation is operative over a wide pH range. Examination of the thermodynamics of binding of ring-substituted hyrazides to HRPC reveals that the binding affinities of aromatic donors are highly sensitive to the position and nature of the ring substituent.  相似文献   

2.
The binding and structural studies of bovine lactoperoxidase with three aromatic ligands, acetylsalicylic acid (ASA), salicylhydoxamic acid (SHA), and benzylhydroxamic acid (BHA) show that all the three compounds bind to lactoperoxidase at the substrate binding site on the distal heme side. The binding of ASA occurs without perturbing the position of conserved heme water molecule W-1, whereas both SHA and BHA displace it by the hydroxyl group of their hydroxamic acid moieties. The acetyl group carbonyl oxygen atom of ASA forms a hydrogen bond with W-1, which in turn makes three other hydrogen-bonds, one each with heme iron, His-109 Nϵ2, and Gln-105 Nϵ2. In contrast, in the complexes of SHA and BHA, the OH group of hydroxamic acid moiety in both complexes interacts with heme iron directly with Fe-OH distances of 3.0 and 3.2Å respectively. The OH is also hydrogen bonded to His-109 Nϵ2 and Gln-105Nϵ2. The plane of benzene ring of ASA is inclined at 70.7° from the plane of heme moiety, whereas the aromatic planes of SHA and BHA are nearly parallel to the heme plane with inclinations of 15.7 and 6.2°, respectively. The mode of ASA binding provides the information about the mechanism of action of aromatic substrates, whereas the binding characteristics of SHA and BHA indicate the mode of inhibitor binding.Lactoperoxidase (LPO)4 (EC 1.11.1.7) is a member of the family of glycosylated mammalian heme-containing peroxidase enzymes which also includes myeloperoxidase (MPO), eosinophil peroxidase (EPO), and thyroid peroxidase. These enzymes also show functional similarities to non-homologous plant and fungal peroxidases because they follow a similar scheme of reaction (13), but their modes of ligand binding differ considerably. Furthermore, the association of the prosthetic heme group in mammalian peroxidases is through covalent bonds (49), whereas covalent linkages are absent in other peroxidases (1014). Among the four mammalian peroxidases, the prosthetic heme group is linked through three covalent bonds in MPO, whereas in LPO, EPO, and thyroid peroxidase only two covalent linkages are formed. So far the detailed crystal structures of only two mammalian peroxidases, MPO and LPO, are known (1520). One of the most striking differences between these two mammalian peroxidases is concerned with the basic structural organization in which MPO exists as a covalently linked dimer, whereas LPO is a monomeric protein. At present the fundamental questions pertaining to mammalian heme peroxidases are (i) what distinguishes between the aromatic ligands that one ligand acts as a substrate, whereas the other ligand works as an inhibitor and (ii) how the substrate and inhibitor specificities differ between two enzymes lactoperoxidase and myeloperoxidase.Lactoperoxidase oxidizes inorganic ions, preferentially thiocyanate (SCN), and to a lesser extent, bromide (Br), whereas in the case of myeloperoxidase the chloride (Cl) ion is a preferred substrate (21, 22). The mammalian peroxidases including LPO are also involved in catalyzing the single electron oxidation of various physiologically important organic aromatic substrates including phenols (23, 24), catecholamines, and catechols (2527) as well as other experimental model compounds such as aromatic amines (28), polychlorinated biphenyls (29), steroid hormones (3032), and polycyclic aromatic hydrocarbons (33). However, the mode of binding of aromatic ligands and associated functional implications are not yet clearly understood. Surprisingly, the structural data on the complexes of mammalian peroxidases with aromatic ligands are conspicuously lacking. The only available structural report is on the complex of MPO with salicylhydroxamic acid (SHA) (34). Even in this case, the coordinates of this structure are not available for a detailed analysis. In the case of non-homologous plant peroxidases, a few crystal structures of their complexes with aromatic compounds are available (3538), but their modes of binding are not very similar to those of mammalian peroxidases because the distal ligand binding sites in mammalian and plant peroxidases differ markedly. In this regard it is pertinent to note that the substrate binding site in peroxidases, in general, is observed at the δ-heme edge of the heme moiety on the distal side; in plant peroxidases an additional ligand binding site has also been observed at γ-heme edge (3941). Unlike those in mammalian peroxidases where the heme moiety is buried deeply in the protein core, in plant peroxidases it is located close to the surface of the protein. Therefore, to characterize the mode of binding of the aromatic substrates and aromatic inhibitors and also for defining the subsites in the substrate binding site, we have determined the crystal structures of three complexes of bovine lactoperoxidase with aromatic ligands, acetylsalicylic acid (ASA), SHA, and benzylhydroxamic acid (BHA). Acetylsalicylic acid can be oxidized by lactoperoxidase to ASA free radical (42), whereas both salicylhydroxamic acid and benzylhydroxamic acid act as potent inhibitors of mammalian peroxidases (4347). The determination of binding characteristics of these compounds having different actions has helped in establishing the relationship between the modes of binding and their potential actions as the substrates and inhibitors. To the best of our knowledge, this is the first structural report on the modes of binding of three aromatic ligands, ASA, SHA, and BHA, to LPO as well as the first structural study of the complexes of any mammalian peroxidase with ASA and BHA. These studies have shown that ASA, SHA, and BHA bind to LPO at the substrate binding site on the distal side. The SHA and BHA directly interact with the heme iron, whereas ASA interacts through the heme water molecule, which in turn is hydrogen-bonded to the heme iron. These studies have provided a greater insight into the mechanisms of substrate and inhibitor binding in the two mammalian peroxidases.  相似文献   

3.
The X-ray crystal structure of the complex of salicylhydroxamic acid (SHA) with Arthromyces ramosus peroxidase (ARP) has been determined at 1.9 A resolution. The position of SHA in the active site of ARP is similar to that of the complex of benzhydroxamic acid (BHA) with ARP [Itakura, H., et al. (1997) FEBS Lett. 412, 107-110]. The aromatic ring of SHA binds to a hydrophobic region at the opening of the distal pocket, and the hydroxamic acid moiety forms hydrogen bonds with the His56, Arg52, and Pro154 residues but is not asscoiated with the heme iron. X-ray analyses of ARP-resorcinol and ARP-p-cresol complexes failed to identify the aromatic donor molecules, most likely due to the very low affinities of these aromatic donors for ARP. Therefore, we examined the locations of these and other aromatic donors on ARP by the molecular dynamics method and found that the benzene rings are trapped similarly by hydrophobic interactions with the Ala92, Pro156, Leu192, and Phe230 residues at the entrance of the heme pocket, but the dihedral angles between the benzene rings and the heme plane vary from donor to donor. The distances between the heme iron and protons of SHA and resorcinol are similar to those obtained by NMR relaxation. Although SHA and BHA are usually considered potent inhibitors for peroxidase, they were found to reduce compound I and compound II of ARP and horseradish peroxidase C in the same manner as p-cresol and resorcinol. The aforementioned spatial relationships of these aromatic donors to the heme iron in ARP are discussed with respect to the quantum chemical mechanism of electron transfer in peroxidase reactions.  相似文献   

4.
Bell JH  Pratt RF 《Biochemistry》2002,41(13):4329-4338
The class C beta-lactamase of Enterobacter cloacae P99 is competitively inhibited by low concentrations of 1:1 complexes of vanadate and hydroxamic acids. Structure-activity studies indicated that the hydroxamic acid functional group was essential to this inhibition. Both aryl and alkyl hydroxamic acids form inhibitory ternary complexes with vanadate and the enzyme, although, in certain cases of the latter, the inhibition may not be seen because of the low formation constants of the vanadate-hydroxamic acid complex. After all of the vanadate species present in solution had been taken into account, "real" K(i) values for the vanadate complexes could be determined. The K(i) value of the best of the inhibitors that were investigated, the 1:1 complex of vanadate with 4-nitrobenzohydroxamic acid, was 0.48 microM. Kinetics studies showed that the association and dissociation rate constants of this complex with the enzyme were 1.48 x 10(6) s(-1) M(-1) and 0.73 s(-1), respectively; the magnitude of the latter indicates covalent interaction of the complex with the enzyme. (51)V NMR and UV-vis spectra suggest that the structure of the vanadate complex bound to the enzyme may be very similar to that in solution. A (13)C NMR spectrum of the enzyme complex with 4-nitrobenzo[(13)C]hydroxamic acid and vanadate yields a coordination-induced shift (CIS) of 7.74 ppm. This is significantly larger than that of the vanadate complex in free solution (3.62 ppm), suggesting either, somewhat contrary to the (51)V and UV-vis spectra, greater interaction between vanadium and the hydroxamate carbonyl oxygen in the enzyme complex than in free solution or, more likely, polarization of the hydroxamate by interaction, e.g., hydrogen bonding, with the enzyme. Molecular modeling indicates that a pentacoordinated vanadate complex may well be able to snugly occupy the enzyme active site; Asn 152 is suitably placed to hydrogen bond to the hydroxamic acid oxygen atom. The experimental results are in accord with a model whereby the vanadate-hydroxamate-enzyme complex is a moderately good analogue of the transition state of the reaction of the beta-lactamase with phosphonate inhibitors.  相似文献   

5.
A quantitative structure-activity relationship (QSAR) study has been made on some lipoxygenase inhibitors belonging to the series of omega-phenylalkyl hydroxamic acids, omega-naphthylalkyl hydroxamic acids, eicosatetraenoic acids, and 1H.benzimidazole-4-ols. It was found that the hydrophobic character of the molecules and the size of their substituents selectively govern their lipoxygenase inhibitory activity. The enzyme active site possesses a non-heme ferric ion, a hydrophobic domain, and a carboxylic acid binding site. It was found that while the functional group of inhibitors must interact with the ferric ion, the substituent on one side of it would be involved in hydrophobic interaction and that on the other side in van der Waals interaction with the enzyme so leading to an enhancement in the inhibitory activity of the inhibitors.  相似文献   

6.
Sharp KH  Moody PC  Brown KA  Raven EL 《Biochemistry》2004,43(27):8644-8651
  相似文献   

7.
Günther U  Mittag T  Schaffhausen B 《Biochemistry》2002,41(39):11658-11669
Few techniques for probing the role of individual amino acids in interactions of a protein with ligands are available. Chemical shift perturbations in NMR spectra provide qualitative information about the response of individual amino acids of a protein to its interactions with ligands. Line shapes derived from (15)N-HSQC spectra recorded for different steps of a ligand titration yield both kinetic constants and insight into mechanisms by which the ligand binds. Here we have analyzed line shapes for 37 signals of amino acids of the N-terminal src homology 2 domain (N-SH2) of the 85 kDa subunit of phosphatidylinositol 3-kinase (PI3-K) upon binding of phosphotyrosine (ptyr)-containing peptides. Kinetic rates at individual amino acids of the SH2 varied throughout the structure. For a subset of SH2 residues, the fine structure of the NMR line shapes indicated slow motions induced by the presence of small amounts of the ligand. These complex line shapes require one or more additional conformational states on the kinetic pathway. Modeling of the observed ligand interactions suggests a quasi-allosteric initial binding step. N-SH2 mutants with altered ligand affinity or specificity were also examined. Analysis of their line shapes revealed three distinct classes of mutants with different kinetic behaviors.  相似文献   

8.
Abstract

A quantitative structure-activity relationship (QSAR) study has been made on some lipoxygenase inhibitors belonging to the series of ω-phenylalkyl hydroxamic acids, ω-naphthylalkyl hydroxamic acids, eicosatetraenoic acids, and 1H.benzimidazole-4-ols. It was found that the hydrophobic character of the molecules and the size of their substituents selectively govern their lipoxygenase inhibitory activity. The enzyme active site possesses a non-heme ferric ion, a hydrophobic domain, and a carboxylic acid binding site. It was found that while the functional group of inhibitors must interact with the ferric ion, the substituent on one side of it would be involved in hydrophobic interaction and that on the other side in van der Waals interaction with the enzyme so leading to an enhancement in the inhibitory activity of the inhibitors.  相似文献   

9.
The cyclooxygenase activity of the bifunctional enzyme prostaglandin H(2) synthase-2 (PGHS-2) is the target of non-steroidal anti-inflammatory drugs. Inhibition of the peroxidase activity of PGHS has been less studied. Using Soret absorption changes, the binding of aromatic hydroxamic acids to the peroxidase site of PGHS-2 was examined to investigate the structural determinants of inhibition. Typical of mammalian peroxidases, the K(d) for benzhydroxamic acid (42mM) is much greater than that for salicylhydroxamic acid (475microM). Binding of the hydroxamic acid tepoxalin (25microM) resulted in only minor Soret changes. However, tepoxalin is an efficient reducing cosubstrate, indicating that it is an alternative electron donor rather than an inhibitor of the peroxidase activity. Aromatic hydrazides are metabolically activated inhibitors of peroxidases. 2-Naphthoichydrazide (2-NZH) caused the time- and concentration-dependent inhibition of both PGHS-2 peroxidase and cyclooxygenase activities. H(2)O(2) was required for the inactivation of both PGHS-2 activities and indomethacin (which binds at the cyclooxygenase site) did not affect the peroxidase inhibitory potency of 2-NZH. A series of aromatic hydrazides were found to be potent inhibitors of PGHS-2 peroxidase activity with IC(50) values in the 6-100microM range for 13 of the 18 hydrazides examined. Selective inhibition of PGHS-2 over myeloperoxidase and horseradish peroxidase isozyme C was increased by certain ring substitutions. In particular, a chloro group para to the hydrazide moiety increased the PGHS-2 selectivity relative to both myeloperoxidase and horseradish peroxidase isozyme C.  相似文献   

10.
R E Hirsch  M J Lin  C M Park 《Biochemistry》1989,28(4):1851-1855
In erythropoietic protoporphyria and lead poisoning, free protoporphyrin (PPIX) and zinc protoporphyrin (ZPP), respectively, accumulate in erythrocytes. That PPIX and ZPP bind to human hemoglobin A (Hb4) is established, but the site of binding is still a matter of controversy. We investigated the interaction of ZPP with intact, tetrameric oxy Hb4, using batch microcalorimetry, front-face fluorometry, absorption difference spectroscopy, oxygen equilibrium studies, and isoelectric focusing (IEF). In the presence of oxy Hb4 (pH 7.35, 0.05 M phosphate), the fluorescence emission maximum (excitation at 420 nm) of ZPP immediately shifts from 587 nm (ZPP alone) to 594 nm, as expected when binding to protein. The fluorescence intensity increases with time and is correlated with the ZPP:Hb4 mole ratio. A slow, time-dependent reaction is also observed with microcalorimetry: the rate of heat of reaction exhibits both a fast and a slow component. The heats of reaction range from -2.1 to -14.8 mcal depending upon the ZPP:Hb4 ratio of 4:1 (0.4 mM:0.1 mM) to 38:1 (3.8 mM:0.1 mM), respectively, and are typical of weak, noncovalent protein-ligand interactions. The optical difference spectra are a function of the ZPP:Hb4 molar ratio and also exhibit a slow increase in intensity over time. No time-dependent optical difference spectra are observed with ZPP or with Hb4 alone. The oxygen affinity of Hb4 in the presence of ZPP decreases with increasing mole ratio. During IEF, all ZPP separates from Hb4, consistent with a weak, noncovalent interaction at a non-heme pocket site. We conclude that ZPP binds to intact, tetrameric hemoglobin at non-heme pocket sites in a nonspecific, weak, noncovalent interaction.  相似文献   

11.
Collagen IX, located on the surface of collagen fibrils, is crucial for cartilage integrity and stability. The N-terminal NC4 domain of the alpha1(IX) chain is probably important in this because it interacts with various macromolecules such as proteoglycans and cartilage oligomeric matrix protein. At least 17 distinct collagen polypeptides carry an NC4-like unit near their N terminus, but this report, describing the crystal structure of NC4 at 1.8-A resolution, represents the first atomic level structure for these domains. The structure is similar to previously characterized laminin-neurexin-sex hormone binding globulin (LNS) structures, dominated by an antiparallel beta-sheet sandwich. In addition, a zinc ion was found in a position similar to that of the metal binding site of other LNS domains. A partial backbone NMR assignment of NC4 was obtained and utilized in NMR titration studies to investigate the zinc binding in solution state and to quantitate the affinity of metal binding. The K(d) of 11.5 mM suggests a regulatory rather than a structural role for zinc in solution. NMR titration with a heparin tetrasaccharide revealed the presence of a secondary binding site for heparin on NC4, showing structural and functional conservation with thrombospondin-1, but a markedly reduced affinity for the ligand. Also the overall arrangement of the N and C termini of NC4 resembles most closely the N-terminal domain of thrombospondin-1, distinguishing the two from the majority of the published LNS structures.  相似文献   

12.
Dissociation constants (Ks) in the pH range 6.5-9.0 for complexes of luciferin, dimethyloxyluciferin (DMOL), and monomethylluciferin (MMOL) with recombinant wild-type and mutant (His433Tyr) luciferases from the Luciola mingrelica firefly were determined by fluorescent titration. The protonated effectors were bound by the wild-type and mutant luciferases better than the nonprotonated ones. The affinity of DMOL for the mutant luciferase was higher than for the wild-type luciferase at alkaline pH, whereas the affinity of MMOL was higher at all pH values studied. The fluorescence emission and excitation spectra of DMOL and MMOL in buffer solution (pH 7.8) were obtained in the absence and presence of luciferase. The fluorescence maxima of DMOL and MMOL complexes with luciferase were 20 and 100 nm, respectively, shifted to shorter wavelengths as compared to the values in buffer solution. This was explained by nonspecific and specific influence of the protein microenvironment on the fluorescence spectra of DMOL and its specific influence on the MMOL fluorescence spectra.  相似文献   

13.
Site-directed mutagenesis was employed to examine the role played by specific surface residues in the activity of cytochrome c peroxidase. The double charge, aspartic acid to lysine, point mutations were constructed at positions 37, 79, and 217 on the surface of cytochrome c peroxidase, sites purported to be within or proximal to the recognition site for cytochrome c in an electron-transfer productive complex formed by the two proteins. The resulting mutant peroxidases were examined for catalytic activity by steady-state measurements and binding affinity by two methods, fluorescence binding titration and cytochrome c affinity chromatography. The cloned peroxidases exhibit similar UV-visible spectra to the wild-type yeast protein, indicating that there are no major structural differences between the cloned peroxidases and the wild-type enzyme. The aspartic acid to lysine mutations at positions 79 and 217 exhibited similar turnover numbers and binding affinities to that seen for the "wild type-like" cloned peroxidase. The same change at position 37 caused more than a 10-fold decrease in both turnover of and binding affinity for cytochrome c. This empirical finding localizes a primary recognition region critical to the dynamic complex. Models from the literature proposing structures for the complex between peroxidase and cytochrome c are discussed in light of these findings.  相似文献   

14.
A mass spectrometric protocol for identifying ligands with a wide range of affinities (3-101 microM) and quantitative spectral analysis for non-covalent interactions have been developed using Src SH2 as a target. Dissociation constants of five compounds, three with a phospho moiety, one with a sulphonic acid group and one with carboxylic acid groups only, were determined using one-ligand one-binding-site, two-ligands one-binding site and one-ligand two-binding-sites models. The Kd values determined by ESI-MS of the three compounds containing the phospho moiety (3.2-7.9 microM) were comparable to those obtained from a solution equilibrium fluorescence polarization assay. The compound with a sulphonate group is a much weaker binding ligand (Kd=101 microM by ESI, >300 microM by FP) towards the Src SH2 protein. Two complexes with different stoichiometric ratios 1:1 and 2:1 (ligand-protein) were observed by ESI-MS for the ligand GIXXX630X. Analysis of binding isotherms indicated the presence of two binding sites for the ligand with Kd values of 9.3 and 193 microM. These data confirmed that, for these polar compounds, non-covalent ESI-MS can measure affinity which very closely reflects the affinity measured under true solution equilibrium conditions. ESI-MS has several key advantages over many solution methods: it can identify the existence of and measure the affinity of complexes other than simple 1:1 ligand-enzyme complexes. Moreover, ESI-MS competition experiments can be readily performed to yield data on whether two ligands bind simultaneously or competitively at the same time as measuring the affinity of the ligand.  相似文献   

15.
The present study describes the biophysical characterization of generation-five poly(amidoamine) (PAMAM) dendrimers conjugated with riboflavin (RF) as a cancer-targeting platform. Two new series of dendrimers were designed, each presenting the riboflavin ligand attached at a different site (isoalloxazine at N-3 and d-ribose at N-10) and at varying ligand valency. Isothermal titration calorimetry (ITC) and differential scanning calorimetry (DSC) were used to determine the binding activity for riboflavin binding protein (RfBP) in a cell-free solution. The ITC data shows dendrimer conjugates have K(D) values of ≥ 465 nM on a riboflavin basis, an affinity ~93-fold lower than that of free riboflavin. The N-3 series showed greater binding affinity in comparison with the N-10 series. Notably, the affinity is inversely correlated with ligand valency. These findings are also corroborated by DSC, where greater protein-conjugate stability is achieved with the N-3 series and at lower ligand valency.  相似文献   

16.
4-Nitrocatechol is examined as an active site probe for non-heme iron dioxygenases and found to be of value, particularly with those containing iron in the Fe(II) oxidation state. 4-Nitrocatechol is astrong competitive inhibitor of substrate oxygenation by protocatechuate 3,4-dioxygenase, forming a reversible complex with this enzyme, and by pyrocatechase. The number of binding sites per enzyme molecule titrated spectrophotometrically with 4-nitrocatechol agrees with results from previous studies with either the principal substrate or other analogues, as expected of an effective probe. Despite these facts and the observation that both enzymes cleave the same substrates at the same carbon-carbon bond, the optical and electron paramagnetic resonance (EPR) spectra of their 4-nitrocatechol complexes are remarkably different. The 4-nitocatechol-protocatechuate 3,4-dioxygenase optical spectra resemble that of the 4-nitrocatecholate ion shifted 20 to 30 nm to longer wavelength. Concomitant with this change the EPR signal centered at g equal 4.28 shows increased rhombicity (g values at 4.74, 4.28, and 3.74).In contrast, the spectrum of the 4-nitrocatechol-pyrocatechase complex has a maximum at the same wavelength as that of a 1:1 solution of free Fe(II) and 4-nitrocatechol in the absence of enzyme after titration of the catecholic protons with base and the g equal 4.28 EPR signal is not resolved at liquid N-2 temperature. These changes are interpreted as resulting in part from a pronounced change in the ligand fields about the irons at the active sites which in the case of protocatechuate 3,4-dioxygenase leads to enzyme inactivation. The results also are the first indication that substrate analogues change their ionization form upon complexation with Fe (III) dioxygenases. The interaction of the probe with metapyrocatechase, an Fe(III) containing dioxygenase, and with several additional oxygenases and hydroperoxidases is also briefly examined. The probe is not specific for any particular class of non-heme iron dioxygenases.  相似文献   

17.
Jezewska MJ  Bujalowski W 《Biochemistry》2000,39(34):10454-10467
Quantitative analyses of the interactions of the Escherichia coli replicative helicase PriA protein with a single-stranded DNA have been performed, using the thermodynamically rigorous fluorescence titration technique. The analysis of the PriA helicase interactions with nonfluorescent, unmodified nucleic acids has been performed, using the macromolecular competition titration (MCT) method. Thermodynamic studies of the PriA helicase binding to ssDNA oligomers, as well as competition studies, show that independently of the type of nucleic acid base, as well as the salt concentration, the type of salt in solution, and nucleotide cofactors, the PriA helicase binds the ssDNA as a monomer. The enzyme binds the ssDNA with significant affinity in the absence of any nucleotide cofactors. Moreover, the presence of AMP-PNP diminishes the intrinsic affinity of the PriA protein for the ssDNA by a factor approximately 4, while ADP has no detectable effect. Analyses of the PriA interactions with different ssDNA oligomers, over a large range of nucleic acid concentrations, indicates that the enzyme has a single, strong ssDNA-binding site. The intrinsic affinities are salt-dependent. The formation of the helicase-ssDNA complexes is accompanied by a net release of 3-4 ions. The experiments have been performed with ssDNA oligomers encompassing the total site size of the helicase-ssDNA complex and with oligomers long enough to encompass only the ssDNA-binding site of the enzyme. The obtained results indicate that salt dependence of the intrinsic affinity results predominantly, if not exclusively, from the interactions of the ssDNA-binding site of the helicase with the nucleic acid. There is an anion effect on the studied interactions, which suggests that released ions originate from both the protein and the nucleic acid. Contrary to the intrinsic affinities, cooperative interactions between bound PriA molecules are accompanied by a net uptake of approximately 3 ions. The PriA protein shows preferential intrinsic affinity for pyrimidine ssDNA oligomers. In our standard conditions (pH 7.0, 10 degrees C, 100 mM NaCl), the intrinsic binding constant for the pyrimidine oligomers is approximately 1 order of magnitude higher than the intrinsic binding constant for the purine oligomers. The significance of these results for the mechanism of action of the PriA helicase is discussed.  相似文献   

18.
Using circular dichroism (CD) and electronic absorption spectroscopy techniques, interaction of the natural dietary cis-carotenoid bixin with an important human plasma protein in vitro was demonstrated for the first time. The induced CD spectra of bixin obtained under physiological conditions (pH 7.4, 37 degrees C) revealed its binding to the serum acute-phase reactant alpha(1)-acid glycoprotein (AGP), a member of the lipocalin protein family. Spectral features of the extrinsic Cotton effects of bixin suggested the inclusion of a single, chirally distorted ligand molecule into the asymmetric protein environment. Compared with the absorption spectra obtained in ethanol and benzene, the strong red shift of the main absorption peak of AGP-bound bixin indicated that the proposed binding site was rich in aromatic residues, and also suggested that hydrophobic interactions were involved in the binding. Using the data obtained from the CD titration experiments, the association constant (Ka=4.5x10(5)M-1) and stoichiometry of the binding (0.15) were calculated. The low value of the stoichiometry was attributed to the structural polymorphism of AGP. To the authors' knowledge, the current study represents the first human lipocalin protein for which carotenoid binding affinity has been explored in vitro with these techniques.  相似文献   

19.
In the present report, a method based on chip-based nanoelectrospray mass spectrometry (nanoESI-MS) is described to detect noncovalent ligand binding to the human estrogen receptor alpha ligand-binding domain (hERalpha LBD). This system represents an important environmental interest, because a wide variety of molecules, known as endocrine disruptors, can bind to the estrogen receptor (ER) and induce adverse health effects in wildlife and humans. Using proper experimental conditions, the nanoESI-MS approach allowed for the detection of specific ligand interactions with hERalpha LBD. The relative gas-phase stability of selected hERalpha LBD-ligand complexes did not mirror the binding affinity in solution, a result that demonstrates the prominent role of hydrophobic contacts for stabilizing ER-ligand complexes in solution. The best approach to evaluate relative solution-binding affinity by nanoESI-MS was to perform competitive binding experiments with 17beta-estradiol (E2) used as a reference ligand. Among the ligands tested, the relative binding affinity for hERalpha LBD measured by nanoESI-MS was 4-hydroxtamoxifen approximately diethylstilbestrol > E2 > genistein > bisphenol A, consistent with the order of the binding affinities in solution. The limited reproducibility of the bound to free protein ratio measured by nanoESI-MS for this system only allowed the binding constants (K(d)) to be estimated (low nanomolar range for E2). The specificity of nanoESI-MS combined with its speed (1 min/ligand), low sample consumption (90 pmol protein/ligand), and its sensitivity for ligand (30 ng/mL) demonstrates that this technique is a promising method for screening suspected endocrine disrupting compounds and to qualitatively evaluate their binding affinity.  相似文献   

20.
Modification of heme·heme oxygenase by iron(III) and cobalt(II) tetrasulfonated phthalocyanines has been performed. New compounds have been isolated and their properties have been investigated by difference spectroscopy, electrophoresis, molecular weight estimation, electron paramagnetic resonance (EPR) and carboxymethylation at histidyl groups. Spectrophotometric titration data indicate the ratio of the reagents in this process to be 1:1. The visible absorption spectra show the main peak at 650 nm for the iron compound and 682 nm for the cobalt one. Electrophoresis and molecular weight estimation show both complexes to be monomers. Cobalt(II) tetrasulfonated phthalocyanine, under aerobic conditions with heme oxygenase protein, undergoes autooxidation to the cobalt(III) complex, as has been proved by EPR and spectroscopic data. Iron and cobalt phthalocyanine modified heme·heme oxygenase with excess dithionite is reduced at the phthalocyanine ligand. In the presence of oxygen, the reduction product transforms into oxygenated Fe(III)Lheme oxygenase or Co(III)heme oxygenase, respectively. Reduction of the iron(III) model complex with ascorbic acid under anaerobic conditions leads to degradation of the phthalocyanine moiety, while Co(III)heme oxygenase with ascorbic acid is reduced to Co(II)Lheme oxygenase. As has been shown by carboxymethylation of the heme oxygenase protein at the histidine residues, the predominant binding site of both phthalocyanine complexes is the heme-binding histidyl residue. There is evidence that there is a second binding site with lower affinity towards Co(II)L on the heme oxygenase protein. Iron and cobalt tetrasulfonated phthalocyanines are not able to displace heme from the heme·heme oxygenase complex. In this reaction the iron complex undergoes degradation and the cobalt one gives a hybrid compound with heme·heme oxygenaseHeme oxygenase protein complexes with iron and cobalt tetrasulfonated phthalocyanines do not exhibit activity in their oxidative degradation.  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