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 484 毫秒
1.
Since the discovery of the inducible form of prostaglandin (PG) H synthase (PGHS), PGHS-2, considerable effort has been made to design selective inhibitors of this isozyme. N-(2-cyclohexyloxy-4-nitrophenyl) methanesulfonamide (NS-398) and 5-bromo-2-(4-fluorophenyl)-3-(4-methylsulfonyl) thiophene (DuP-697) have been shown to interact reversibly with PGHS-1, while irreversibly inhibiting PGHS-2 in a time-dependent manner. In the present study we have tested the effects of DuP-697 and NS-398 on the activity of PGHS-1 and further explored the interactions between these agents and the inhibition of PGHS-1 by aspirin, indomethacin and ibuprofen. Three independent experimental systems, namely bovine aortic endothelial cells (BAEC), human fibroblasts and ram seminal vesicle microsomes were used to investigate the effects of DuP-697 and NS-398 on PGHS-1. The results show that DuP-697 and NS-398, at concentrations ranges which do not inhibit PGHS-1 activity, significantly attenuated the inhibition of PGHS-1 that was caused by aspirin and indomethacin. The same concentrations of DuP-697 and NS-398 did not affect the inhibition of PGHS-1 that was induced by the competitive reversible inhibitors ibuprofen and naproxen. Similar effects of DuP-697 and NS-393 were obtained with ram seminal vesicle microsomes. These results suggest that PGHS-2 inhibitors DuP-697 and NS-398 possibly interact with PGHS-1 at a site different from the enzyme's catalytic site, thus causing attenuation of PGHS-1 inhibition by aspirin and indomethacin without altering PGHS-1 basal activity or the ibuprofen-induced inhibition.  相似文献   

2.
Prostaglandin H synthase isoforms 1 and -2 (PGHS-1 and -2) react with peroxide to form a radical on Tyr385 that initiates the cyclooxygenase catalysis. The tyrosyl radical EPR signals of PGHS-1 and -2 change over time and are altered by cyclooxygenase inhibitor binding. We characterized the tyrosyl radical dynamics using wild type human PGHS-1 (hPGHS-1) and its Y504F, Y385F, and Y385F/Y504F mutants to determine whether the radical EPR signal changes involve Tyr504 radical formation, Tyr385 radical phenyl ring rotation, or both. Reaction of hPGHS-1 with peroxide produced a wide singlet, whereas its Y504F mutant produced only a wide doublet signal, assigned to the Tyr385 radical. The cyclooxygenase specific activity and KM value for arachidonate of hPGHS-1 were not affected by the Y504F mutation, but the peroxidase specific activity and the KM value for peroxide were increased. The Y385F and Y385F/Y504F mutants retained only a small fraction of the peroxidase activity; the former had a much-reduced yield of peroxide-induced radical and the latter essentially none. After binding of indomethacin, a cyclooxygenase inhibitor, hPGHS-1 produced a narrow singlet but the Y504F mutant did not form a tyrosyl radical. These results indicate that peroxide-induced radicals form on Tyr385 and Tyr504 of hPGHS-1, with radical primarily on Tyr504 in the wild type protein; indomethacin binding prevented radical formation on Tyr385 but allowed radical formation on Tyr504. Thus, hPGHS-1 and -2 have different distributions of peroxide-derived radical between Tyr385 and Tyr504. Y504F mutants in both hPGHS-1 and -2 significantly decreased the cyclooxygenase activation efficiency, indicating that formation of the Tyr504 radical is functionally important for both isoforms.  相似文献   

3.
Malonamidase E2 (MAE2) from Bradyrhizobium japonicum is an enzyme that hydrolyzes malonamate to malonate and has a Ser-cis-Ser-Lys catalytic triad at the active site. The crystal structures of wild type and mutant MAE2 exhibited that the guanido group of Arg-158 could be involved in the binding of malonamate in which the negative charge of the carboxyl group could destabilize a negatively charged transition-state oxyanion in the enzymatic reaction. In an attempt to elucidate the specific roles of Arg-158, site-directed mutants, R158Q, R158E, and R158K, were prepared (see Table 1). The crystal structure of R158Q determined at 2.2 Angstrom resolution showed that the guanido group of Arg-158 was important for the substrate binding with the marginal structural change upon the mutation. The k(cat) value of R158Q significantly decreased by over 1500-fold and the catalytic activity of R158E could not be detected. The k(cat) value of R158K was similar to that of the wild type with the K(m) value drastically increased by 100-fold, suggesting that Lys-158 of R158K can stabilize the negative charge of the carboxylate in the substrate to some extent and contribute to the stabilization of the transition-state oxyanion, but a single amine group of Lys-158 in R158K could not precisely anchor the carboxyl group of malonamate compared with the guanido group of Arg-158. Our kinetic and structural evidences demonstrate that Arg-158 in MAE2 should be critical to both binding the substrate and stabilizing the transition-state oxyanion for the catalytic reaction of MAE2.  相似文献   

4.
Prostaglandin endoperoxide H synthases 1 and 2 (PGHS-1 and -2) are the major targets of nonsteroidal anti-inflammatory drugs. Both isozymes are integral membrane proteins but lack transmembrane domains. X-ray crystallographic studies have led to the hypothesis that PGHS-1 and -2 associate with only one face of the membrane bilayer through a novel, monotopic membrane binding domain (MBD) that is comprised of four short, consecutive, amphipathic alpha-helices (helices A-D) that include residues 74-122 in ovine PGHS-1 (oPGHS-1) and residues 59-108 in human PGHS-2 (hPGHS-2). Previous biochemical studies from our laboratory showed that the MBD of oPGHS-1 lies somewhere between amino acids 25 and 166. In studies reported here, membrane-associated forms of oPGHS-1 and hPGHS-2 were labeled using the hydrophobic, photoactivable reagent 3-trifluoro-3-(m-[(125)I]iodophenyl)diazirine, isolated, and cleaved with AspN and/or GluC, and the photolabeled peptides were sequenced. The results establish that the MBDs of oPGHS-1 and hPGHS-2 reside within residues 74-140 and 59-111, respectively, and thus provide direct provide biochemical support for the hypothesis that PGHS-1 and -2 do associate with membranes through a monotopic MBD. We also prepared HelA, HelB, and HelC mutants of oPGHS-1, in which, for each helix, three or four hydrophobic residues expected to protrude into the membrane were replaced with small, neutral residues. When expressed in COS-1 cells, HelA and HelC mutants exhibited little or no catalytic activity and were present, at least in part, as misfolded aggregates. The HelB mutant retained about 20% of the cyclooxygenase activity of native oPGHS-1 and partitioned in subcellular fractions like native oPGHS-1; however, the HelB mutant exhibited an extra site of N-glycosylation at Asn(104). When this glycosylation site was eliminated (HelB/N104Q mutation), the mutant lacked cyclooxygenase activity. Thus, our mutational analyses indicate that the amphipathic character of each helix is important for the assembly and folding of oPGHS-1 to a cyclooxygenase active form.  相似文献   

5.
Prostaglandin endoperoxide H synthases (PGHSs) catalyze the committed step in the biosynthesis of prostaglandins and thromboxane, the conversion of arachidonic acid, two molecules of O(2), and two electrons to prostaglandin endoperoxide H(2) (PGH(2)). Formation of PGH(2) involves an initial oxygenation of arachidonate to yield PGG(2) catalyzed by the cyclooxygenase activity of the enzyme and then a reduction of the 15-hydroperoxyl group of PGG(2) to form PGH(2) catalyzed by the peroxidase activity. The cyclooxygenase active site is a hydrophobic channel that protrudes from the membrane binding domain into the core of the globular domain of PGHS. In the crystal structure of Co(3+)-heme ovine PGHS-1 complexed with arachidonic acid, 19 cyclooxygenase active site residues are predicted to make a total of 50 contacts with the substrate (Malkowski, M. G, Ginell, S., Smith, W. L., and Garavito, R. M. (2000) Science 289, 1933-1937); two of these are hydrophilic, and 48 involve hydrophobic interactions. We performed mutational analyses to determine the roles of 14 of these residues and 4 other closely neighboring residues in arachidonate binding and oxygenation. Mutants were analyzed for peroxidase and cyclooxygenase activity, and the products formed by various mutants were characterized. Overall, the results indicate that cyclooxygenase active site residues of PGHS-1 fall into five functional categories as follows: (a) residues directly involved in hydrogen abstraction from C-13 of arachidonate (Tyr-385); (b) residues essential for positioning C-13 of arachidonate for hydrogen abstraction (Gly-533 and Tyr-348); (c) residues critical for high affinity arachidonate binding (Arg-120); (d) residues critical for positioning arachidonate in a conformation so that when hydrogen abstraction does occur the molecule is optimally arranged to yield PGG(2) versus monohydroperoxy acid products (Val-349, Trp-387, and Leu-534); and (e) all other active site residues, which individually make less but measurable contributions to optimal catalytic efficiency.  相似文献   

6.
Prostaglandin H synthase-1 and -2 (PGHS-1 and PGHS-2, EC 1.14.99.1) are membrane associated glycoproteins that catalyze the first two steps in prostaglandin synthesis. As the enzymes play an important regulatory role in several physiological and pathophysiological processes, recombinant PGHS isoforms are widely used in biomedical research. In the present study, we expressed human PGHS-2 (hPGHS-2) with and without a six histidine sequence tag (His(6) tag) near the amino- or carboxy-terminus of the protein in the Pichia pastoris (P. pastoris) expression system using native or yeast signal sequences. The recombinant His(6) tagged hPGHS-2 was purified using Ni-affinity and anion exchange chromatography, whereas the purification of the C-terminally His(6) tagged hPGHS-2 was more efficient. K(m), k(cat) and IC(50) values were determined to characterize the protein. The data obtained indicate that both the N- and C-terminally His(6) tagged hPGHS-2 are functional and the catalytic properties of the recombinant protein and the enzyme produced in other expression systems are comparable. As the yeast culture is easy to handle, the P. pastoris system could serve as an alternative to the most commonly used baculovirus-insect cell expression system for the production of the recombinant PGHS-2.  相似文献   

7.
Three arginine residues (Arg-11, Arg-39, Arg-61) are found at the active site of 4-oxalocrotonate tautomerase in the X-ray structure of the affinity-labeled enzyme [Taylor, A. B., Czerwinski, R. M., Johnson, R. M., Jr., Whitman, C. P., and Hackert, M. L. (1998) Biochemistry 37, 14692-14700]. The catalytic roles of these arginines were examined by mutagenesis, kinetic, and heteronuclear NMR studies. With a 1,6-dicarboxylate substrate (2-hydroxymuconate), the R61A mutation showed no kinetic effects, while the R11A mutation decreased k(cat) 88-fold and increased K(m) 8.6-fold, suggesting both binding and catalytic roles for Arg-11. With a 1-monocarboxylate substrate (2-hydroxy-2,4-pentadienoate), no kinetic effects of the R11A mutation were found, indicating that Arg-11 interacts with the 6-carboxylate of the substrate. The stereoselectivity of the R11A-catalyzed protonation at C-5 of the dicarboxylate substrate decreased, while the stereoselectivity of protonation at C-3 of the monocarboxylate substrate increased in comparison with wild-type 4-OT, indicating the importance of Arg-11 in properly orienting the dicarboxylate substrate by interacting with the charged 6-carboxylate group. With 2-hydroxymuconate, the R39A and R39Q mutations decreased k(cat) by 125- and 389-fold and increased K(m) by 1.5- and 2.6-fold, respectively, suggesting a largely catalytic role for Arg-39. The activity of the R11A/R39A double mutant was at least 10(4)-fold lower than that of the wild-type enzyme, indicating approximate additivity of the effects of the two arginine mutants on k(cat). For both R11A and R39Q, 2D (1)H-(15)N HSQC and 3D (1)H-(15)N NOESY-HSQC spectra showed chemical shift changes mainly near the mutated residues, indicating otherwise intact protein structures. The changes in the R39Q mutant were mainly in the beta-hairpin from residues 50 to 57 which covers the active site. HSQC titration of R11A with the substrate analogue cis, cis-muconate yielded a K(d) of 22 mM, 37-fold greater than the K(d) found with wild-type 4-OT (0.6 mM). With the R39Q mutant, cis, cis-muconate showed negative cooperativity in active site binding with two K(d) values, 3.5 and 29 mM. This observation together with the low K(m) of 2-hydroxymuconate (0.47 mM) suggests that only the tight binding sites function catalytically in the R39Q mutant. The (15)Nepsilon resonances of all six Arg residues of 4-OT were assigned, and the assignments of Arg-11, -39, and -61 were confirmed by mutagenesis. The binding of cis,cis-muconate to wild-type 4-OT upshifts Arg-11 Nepsilon (by 0.05 ppm) and downshifts Arg-39 Nepsilon (by 1.19 ppm), indicating differing electronic delocalizations in the guanidinium groups. A mechanism is proposed in which Arg-11 interacts with the 6-carboxylate of the substrate to facilitate both substrate binding and catalysis and Arg-39 interacts with the 1-carboxylate and the 2-keto group of the substrate to promote carbonyl polarization and catalysis, while Pro-1 transfers protons from C-3 to C-5. This mechanism, together with the effects of mutations of catalytic residues on k(cat), provides a quantitative explanation of the 10(7)-fold catalytic power of 4-OT. Despite its presence in the active site in the crystal structure of the affinity-labeled enzyme, Arg-61 does not play a significant role in either substrate binding or catalysis.  相似文献   

8.
The cyclooxygenases (COX-1 and COX-2) are membrane-associated, heme-containing homodimers that generate prostaglandin H2 from arachidonic acid (AA) in the committed step of prostaglandin biogenesis and are the targets for nonsteroidal anti-inflammatory drugs (NSAIDs). N-(2-cyclohexyloxy-4-nitrophenyl) methanesulfonamide (NS-398) was the first in a series of isoform-selective drugs designed to preferentially inhibit COX-2, with the aim of ameliorating many of the toxic gastrointestinal side effects caused by conventional NSAID inhibition. We determined the X-ray crystal structure of murine COX-2 in complex with NS-398 utilizing synchrotron radiation to 3.0 A resolution. NS-398 binds in the cyclooxygenase channel in a conformation that is different than that observed for other COX-2-selective inhibitors, such as celecoxib, with no discernible penetration into the side pocket formed in COX-2 by the isoform-specific substitutions of I434V, H513R, and I523V. Instead, the methanesulfonamide moiety of NS-398 interacts with the side chain of Arg-120 at the opening of the cyclooxygenase channel, similar to that observed for acidic, nonselective NSAIDs such as indomethacin and flurbiprofen. Our structure validates inhibitor studies that identified Arg-120 as a molecular determinant for time-dependent inhibition of COX-2 by NS-398.  相似文献   

9.
Prostaglandin E(2) (PGE(2)) production involves the activity of a multistep biosynthetic pathway. The terminal components of this cascade, two PGE(2) synthases (PGES), have very recently been identified as glutathione-dependent proteins. cPGES is cytoplasmic, apparently identical to the hsp90 chaperone, p23, and associates functionally with prostaglandin-endoperoxide H synthase-1 (PGHS-1), the constitutive cyclooxygenase. A second synthase, designated mPGES, is microsomal and can be regulated. Here we demonstrate that mPGES and PGHS-2 are expressed at very low levels in untreated human orbital fibroblasts. Interleukin (IL)-1beta treatment elicits high levels of PGHS-2 and mPGES expression. The induction of both enzymes occurs at the pretranslational level, is the consequence of enhanced gene promoter activities, and can be blocked by dexamethasone (10 nm). SC58125, a PGHS-2-selective inhibitor, could attenuate the induction of mPGES, suggesting a dependence of this enzyme on PGHS-2 activity. IL-1beta treatment activates p38 and ERK mitogen-activated protein kinases. Induction of both mPGES and PGHS-2 was susceptible to either chemical inhibition or molecular interruption of these pathways with dominant negative constructs. These results indicate that the induction of PGHS-2 and mPGES by IL-1beta underlies robust PGE(2) production in orbital fibroblasts.  相似文献   

10.
Rat carrageenin-induced pleurisy was used to clarify the role of prostaglandin H synthase (PGHS)-2 in acute inflammation. Intrapleural injection of 0.2 ml of 2% λ-carrageenin induced accumulation of exudate and infiltration of leukocytes into the pleural cavity. When PGHS-1 and -2 proteins in the pleural exudate cells were analyzed by Western blot analysis, PGHS-2 was detectable from 1 hr after carrageenin injection. Its level rose sharply, remained high from 3 to 7 hr after injection, and then fell to near the detection limit. PGHS-1 was also detected, but kept almost the same level throughout the course of the pleurisy. Levels of prostaglandin (PG) E2 and thromboxane (TX) B2 in the exudate increased from hour 3 to hour 7, and then declined. Thus, the changes of the level of PGE2 were closely paralleled those of PGHS-2.The selective PGHS-2 inhibitors NS-398, nimesulide and SC-58125 suppressed the inflammatory reaction and caused a marked decrease in the level of PGE2 but not in those of TXB2 and 6-keto-PGF. These results suggest that the PGHS-2 expressed in the pleural exudate cells may be involved in PGE2 formation at the site of inflammation.  相似文献   

11.
By using the technique of site-directed spin labeling combined with EPR spectroscopy, we have observed that binding of arachidonic acid and nonsteroidal anti-inflammatory drugs induces conformational changes in the human prostaglandin endoperoxide H(2) synthase enzyme (PGHS-2). Line shape broadening resulting from spin-spin coupling of nitroxide pairs introduced into the membrane-binding helices of PGHS-2 was used to calculate the inter-helical distances and changes in these distances that occur in response to binding various ligands. The inter-residue distances determined for the PGHS-2 holoenzyme using EPR were 1-7.9 A shorter than those of the crystal structure of the PGHS-2 holoenzyme. However, inter-helical distances calculated and determined by EPR for PGHS-2 complexed with arachidonic acid, flurbiprofen, and SC-58125 were in close agreement with those obtained from the cognate crystal structures. These results indicate that the structure of the solubilized PGHS-2 holoenzyme measured in solution differs from the crystal structure of PGHS-2 holoenzyme obtained by x-ray analysis. Furthermore, binding of ligands induces a conformational change in the holo-PGHS-2, converting it to a structure similar to those obtained by x-ray analysis. Proteolysis protection assays had previously provided circumstantial evidence that binding of heme and non-steroidal anti-inflammatory drugs alters the conformation of PGHS, but the present experiments are the first to directly measure such changes. The finding that arachidonate can also induce a conformational change in PGHS-2 was unexpected, and the magnitude of changes suggests this structural flexibility may be integral to the cyclooxygenase catalytic mechanism.  相似文献   

12.
Indomethacin (25 microM) inhibited the amount of prostaglandin (PG) E2, PGF2alpha and 6-keto-PGF1alpha synthesized by homogenates of day 5 pregnant rat uterus by 80-92%. In contrast, 25 microM NS-398, a selective inhibitor of prostaglandin H synthase-2 (PGHS-2), inhibited the synthesis of these three prostaglandins by homogenates of the same tissue by only 37-60%. Since it has been reported that idomethacin and NS-398 inhibit PGHS-2 with similar potencies and that indomethacin (unlike NS-398) is a potent inhibitor of PGHS-1, it may be concluded that prostaglandin production by homogenates of the rat uterus on day 5 of pregnancy is due to the activities of both PGHS-1 and PGHS-2. The administration of indomethacin (3 mg/kg) twice daily on days 3 and 4 of pregnancy reduced the implantation rate by 77%, whereas the similar administration of NS-398 (6 mg/kg) had no significant effect on implantation. It may be concluded that either the dose of NS-398 used was too low to affect uterine PGHS-2 activity in vivo sufficiently to prevent implantation, or that inhibiting uterine PGHS-2 activity in vivo has no effect on the implantation mechanisms or is compensated for by increased PGHS-1 activity such that the implantation process is not impaired.  相似文献   

13.
Dihydrofolate reductases from different species contain several highly conserved arginines, some of which have been shown by x-ray crystallography to have their guanido groups near the p-aminobenzoyl glutamate moiety of enzyme-bound methotrexate. The orientation of one of these (Arg-52) appears to be completely reversed in comparing the crystal structures of Escherichia coli with Lactobacillus casei enzyme (Bolin, J. T., Filman, D. J., Matthews, D. A., Hamlin, R. C., and Kraut, J. (1982). J. Biol. Chem. 257, 13650-13662). We synthesized a novel antifolate containing a glyoxal group designed to react specifically with active-site guanido groups which are able to approach the p-aminobenzoyl carbonyl of methotrexate. The binding of this compound to the enzyme was competitive with dihydrofolate (DHF) in ordinary buffers. In borate buffer at pH 8.0 it inactivated dihydrofolate reductases from both E. coli and L. casei at similar maximum rates, while the chicken liver enzyme was more slowly inactivated. The inactivation was stoichiometric, paralleled the loss of the glyoxal chromophore, and showed saturation kinetics. Inhibitor binding and thus inactivation was enhanced by NADPH, while DHF protected the enzyme. This allowed calculation of the Kd for DHF which was found to be identical with its Km. The stoichiometrically inactivated enzyme displayed the 340-nm chromophore characteristic of 4-aminopteridines bound to dihydrofolate reductase confirming active-site labeling with normal orientation of the ligand. The ligand remained covalently bound to inactivated enzyme upon denaturation at low pH but dissociated at neutral pH. Computer graphic modeling of the crystal structures predicted reaction of Arg-31 but not Arg-52 in L. casei dihydrofolate reductase and of only Arg-52 in the E. coli enzyme. Purification of the CNBr fragments from the inactivated enzymes gave a single labeled peptide for each species. The particular peptide tagged in each case was unaffected by the presence of NADPH and was in excellent agreement with the crystallographic predictions.  相似文献   

14.
KAT-50, an established human thyrocyte cell line, expresses constitutively high levels of prostaglandin endoperoxide H synthase-2 (PGHS-2), the inflammatory cyclooxygenase. Here, we examine primary human thyrocytes. We find that they, too, express PGHS-2 mRNA and protein under control culture conditions. A substantial fraction of the basal prostaglandin E(2) (PGE(2)) produced by these cells can be inhibited by SC-58125 (5 microM), a PGHS-2-selective inhibitor. Interleukin (IL)-1beta (10 ng/ml) induces PGHS-2 expression and PGE(2) production in primary thyrocytes. The induction of PGHS-2 and PGE(2) synthesis by IL-1beta could be blocked by glucocorticoid treatment. Unlike KAT-50, most of the culture strains also express PGHS-1 protein. Our observations suggest that both cyclooxygenase isoforms may have functional roles in primary human thyroid epithelial cells, and PGHS-2 might predominate under basal and cytokine-activated culture conditions.  相似文献   

15.
The cyclooxygenases (COX-1 and COX-2) generate prostaglandin H(2) from arachidonic acid (AA). In its catalytically productive conformation, AA binds within the cyclooxygenase channel with its carboxylate near Arg-120 and Tyr-355 and ω-end located within a hydrophobic groove above Ser-530. Although AA is the preferred substrate for both isoforms, COX-2 can oxygenate a broad spectrum of substrates. Mutational analyses have established that an interaction of the carboxylate of AA with Arg-120 is required for high affinity binding by COX-1 but not COX-2, suggesting that hydrophobic interactions between the ω-end of substrates and cyclooxygenase channel residues play a significant role in COX-2-mediated oxygenation. We used structure-function analyses to investigate the role that Arg-120 and residues lining the hydrophobic groove play in the binding and oxygenation of substrates by murine (mu) COX-2. Mutations to individual amino acids within the hydrophobic groove exhibited decreased rates of oxygenation toward AA with little effect on binding. R120A muCOX-2 oxygenated 18-carbon ω-6 and ω-3 substrates albeit at reduced rates, indicating that an interaction with Arg-120 is not required for catalysis. Structural determinations of Co(3+)-protoporphyrin IX-reconstituted muCOX-2 with α-linolenic acid and G533V muCOX-2 with AA indicate that proper bisallylic carbon alignment is the major determinant for efficient substrate oxygenation by COX-2. Overall, these findings implicate Arg-120 and hydrophobic groove residues as determinants that govern proper alignment of the bisallylic carbon below Tyr-385 for catalysis in COX-2 and confirm nuances between COX isoforms that explain substrate promiscuity.  相似文献   

16.
Because the induction of new lipid body formation in leukocytes correlates with and likely contributes to their enhanced 'primed' prostaglandin and leukotriene formation, we evaluated two selective cyclooxygenase (COX)-2 inhibitors. Three types of stimuli, cis -unsaturated fatty acids, platelet activating factor and protein kinase C activators, stimulate lipid body formation. NS-398 (0.1-10 microM), but not another COX-2 inhibitor, SC58125 (0.1- 10 microM), blocked leukocyte lipid body formation elicited by all three types of stimuli and also blocked priming for enhanced LTB(4) production and PGE(2) production. The effect of NS-398 on lipid body formation was independent of its inhibitory effects on COX-2 since arachidonate-induced lipid body formation in COX-2-deficient mouse leukocytes was also inhibited by NS-398. By means of its ability to inhibit leukocyte lipid body formation, NS-398 may exert actions independent of its COX-2 inhibition and more broadly contribute to the suppression of formation of COX-1 and lipoxygenase-derived eicosanoids.  相似文献   

17.
Prostaglandin endoperoxide H synthase (PGHS) catalyzes the oxidation of arachidonate to prostaglandin H(2). We have previously synthesized and chemically characterized nitroarachidonic acid (AANO(2)), a novel anti-inflammatory signaling mediator. Herein, the interaction of AANO(2) with PGHS was analyzed. AANO(2) inhibited oxygenase activity of PGHS-1 but not PGHS-2. AANO(2) exhibited time- and concentration-dependent inhibition of peroxidase activity in both PGHS-1 and -2. The plot of k(obs) versus AANO(2) concentrations showed a hyperbolic function with k(inact) = 0.045 s(-1) and K(i)(*app) = 0.019 μM for PGHS-1 and k(inact) = 0.057 s(-1) and K(i)(*app) = 0.020 μM for PGHS-2. Kinetic analysis suggests that inactivation of PGHS by AANO(2) involves two sequential steps: an initial reversible binding event (described by K(i)) followed by a practically irreversible event (K(i)(*app)) leading to an inactivated enzyme. Inactivation was associated with irreversible disruption of heme binding to the protein. The inhibitory effects of AANO(2) were selective because other nitro-fatty acids tested, such as nitrooleic acid and nitrolinoleic acid, were unable to inhibit enzyme activity. In activated human platelets, AANO(2) significantly decreased PGHS-1-dependent thromboxane B(2) formation in parallel with a decrease in platelet aggregation, thus confirming the biological relevance of this novel inhibitory pathway.  相似文献   

18.
The antiinflammatory agent darbufelone, ((Z)-5-[[3,5-bis(1,1-dimethylethyl)-4-hydroxyphenyl] methylene]-2-imino-4-thiazolidinone, methanesulfonate salt), was discovered as a dual inhibitor of cellular prostaglandin and leukotriene production. To study the mechanism of action of this drug, we expressed human prostaglandin endoperoxide synthase-1 (PGHS-1) and PGHS-2 and purified the recombinant enzymes using buffers that contain octylglucoside. In cyclooxygenase assays following a 15-min incubation of enzyme with inhibitor, darbufelone potently inhibits PGHS-2 (IC(50) = 0.19 microM) but is much less potent with PGHS-1 (IC(50) = 20 microM). Interestingly, when the assay buffer contains traces of Tween 20 (0.0001%), darbufelone appears inactive with PGHS-2 due to a detergent interaction that is detectable by absorption spectroscopy. We therefore used octylglucoside, which does not affect darbufelone in this way, in place of Tween 20 in our PGHS buffers. Inhibition of PGHS-2 with darbufelone is time dependent: with no preincubation, darbufelone is a weak inhibitor (IC(50) = 14 microM), but after a 30-min incubation it is 20-fold more potent. Plots of PGHS-2 activity vs preincubation time at various darbufelone concentrations reach a plateau. This finding is inconsistent with irreversible or one-step slow-binding inhibition. A two-step slow-binding inhibition model is proposed in which the E.I complex (K(i) = 6.2 +/- 1.9 to 14 +/- 1 microM) slowly transforms (k(5) = 0.015-0.030 s(-)(1)) to a tightly bound E.I form with K(i) = 0.63 +/- 0.07 microM and k(6) = 0.0034 s(-)(1). In steady-state kinetics inhibition experiments performed with no preincubation, we find that darbufelone is a noncompetitive inhibitor of PGHS-2 (K(i) = 10 +/- 5 microM). Darbufelone quenches the fluorescence of PGHS-2 at 325 nm (lambda(ex) = 280 nm) with K(d) = 0.98 +/- 0.03 microM. The PGHS substrate, arachidonate, and various cyclooxygenase inhibitors do not alter this binding affinity of darbufelone but a structural analogue of darbufelone competes directly for binding to PGHS-2. Di-tert-butyl phenols such as darbufelone may inhibit PGHS-2 by exploiting a previously unrecognized binding site on the enzyme.  相似文献   

19.
Cyclooxygenase (COX) inhibitors were regarded as anticarcinogenic agents for lung cancer at least partly via PGE2; but these were based on cytokin stimulation experiment on A549 cell. In order to clarify whether COX inhibitors directly inhibit A549 cell, three COX inhibitors, NS398 (selective COX-2 inhibitor), SC560 (selective COX-1 inhibitor), and acetyl salicylic acid (ASA, non-selective COX inhibitor), were studied. NS398, and ASA, can inhibit PGE2 generation via COX-2 inhibition. The viability of A549 cell was assayed by MTT. However, without cytokin stimulation, all the three inhibitors (NS398 0.2-20 microM; SC560 1.0-100 nM; ASA 0.01-1.0 mM) were not able to inhibit A549 cell proliferation, in the other way round, NS398 promoted cell growth. And arachidonic acid (AA) and lipopolysaccharide (LPS) did not disturb the property of its growth. These data suggested that without cytokin stimulation, COX and PGE2 may not be the kernel molecules involved in A549 cell proliferation, and COX inhibitors could not inhibit A549 cell growth directly.  相似文献   

20.
Phosphoenolpyruvate carboxylases (PEPC, EC 4.1.1.31) from higher plants are regulated by both allosteric effects and reversible phosphorylation. Previous x-ray crystallographic analysis of Zea mays PEPC has revealed a binding site for sulfate ion, speculated to be the site for an allosteric activator, glucose 6-phosphate (Glc-6-P) (Matsumura, H., Xie, Y., Shirakata, S., Inoue, T., Yoshinaga, T., Ueno, Y., Izui, K., and Kai, Y. (2002) Structure (Lond.) 10, 1721-1730). Because kinetic experiments have also supported this notion, each of the four basic residues (Arg-183, -184, -231, and -372' on the adjacent subunit) located at or near the binding site was replaced by Gln, and the kinetic properties of recombinant mutant enzymes were investigated. Complete desensitization to Glc-6-P was observed for R183Q, R184Q, R183Q/R184Q (double mutant), and R372Q, as was a marked decrease in the sensitivity for R231Q. The heterotropic effect of Glc-6-P on an allosteric inhibitor, l-malate, was also abolished, but sensitivity to Gly, another allosteric activator of monocot PEPC, was essentially not affected, suggesting the distinctness of their binding sites. Considering the kinetic and structural data, Arg-183 and Arg-231 were suggested to be involved directly in the binding with phosphate group of Glc-6-P, and the residues Arg-184 and Arg-372 were thought to be involved in making up the site for Glc-6-P and/or in the transmission of an allosteric regulatory signal. Most unexpectedly, the mutant enzymes had almost lost responsiveness to regulatory phosphorylation at Ser-15. An apparent lack of kinetic competition between the phosphate groups of Glc-6-P and of phospho-Ser at 15 suggested the distinctness of their binding sites. The possible roles of these Arg residues are discussed.  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