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.
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   

2.
Arachidonic acid is converted to prostaglandin G(2) (PGG(2)) by the cyclooxygenase activities of prostaglandin endoperoxide H synthases (PGHSs) 1 and 2. The initial, rate-limiting step is abstraction of the 13-proS hydrogen from arachidonate which, for PGG(2) formation, is followed by insertion of O(2) at C-11, cyclization, and a second O( 2) insertion at C-15. As an accompaniment to ongoing structural studies designed to determine the orientation of arachidonate in the cyclooxygenase site, we analyzed the products formed from arachidonate by (a) solubilized, partially purified ovine (o) PGHS-1; (b) membrane-associated, recombinant oPGHS-1; and (c) a membrane-associated, recombinant active site mutant (V349L oPGHS-1) and determined kinetic values for formation of each product. Native forms of oPGHS-1 produced primarily PGG(2) but also several monohydroxy acids, which, in order of abundance, were 11R-hydroxy-5Z, 8Z,12E,14Z-eicosatetraenoic acid (11R-HETE), 15S-hydroxy-5Z,8Z,11Z, 13E-eicosatetraenoic acid (15S-HETE), and 15R-HETE. V349L oPGHS-1 formed primarily PGG(2), 15S-HETE, and 15R-HETE but only trace amounts of 11R-HETE. With native enzyme, the K(m) values for PGG(2), 11-HETE, and 15-HETE formation were each different (5.5, 12.1, and 19.4 microM, respectively); similarly, the K(m) values for PGG(2) and 15-HETE formation by V349L oPGHS-1 were different (11 and 5 microM, respectively). These results establish that arachidonate can assume at least three catalytically productive arrangements within the cyclooxygenase site of oPGHS-1 leading to PGG(2), 11R-HETE, and 15S-HETE and/or 15R-HETE, respectively. IC(50) values for inhibition of formation of the individual products by the competitive inhibitor, ibuprofen, were determined and found to be the same for a given enzyme form (i.e. 175 microM for oPGHS-1 and 15 microM for V349L oPGHS-1). These latter results are most simply rationalized by a kinetic model in which arachidonate forms various catalytically competent arrangements only after entering the cyclooxygenase active site.  相似文献   

3.
Prostaglandin endoperoxide H synthases-1 and -2 (PGHSs) can oxygenate 18-22 carbon polyunsaturated fatty acids, albeit with varying efficiencies. Here we report the crystal structures of eicosapentaenoic acid (EPA, 20:5 n-3) and linoleic acid (LA, 18:2 n-6) bound in the cyclooxygenase active site of Co(3+) protoporphyrin IX-reconstituted ovine PGHS-1 (Co(3+)-oPGHS-1) and compare the effects of active site substitutions on the rates of oxygenation of EPA, LA, and arachidonic acid (AA). Both EPA and LA bind in the active site with orientations similar to those seen previously with AA and dihomo-gamma-linolenic acid (DHLA). For EPA, the presence of an additional double bond (C-17/C-18) causes this substrate to bind in a "strained" conformation in which C-13 is misaligned with respect to Tyr-385, the residue that abstracts hydrogen from substrate fatty acids. Presumably, this misalignment is responsible for the low rate of EPA oxygenation. For LA, the carboxyl half binds in a more extended configuration than AA, which results in positioning C-11 next to Tyr-385. Val-349 and Ser-530, recently identified as important determinants for efficient oxygenation of DHLA by PGHS-1, play similar roles in the oxygenation of EPA and LA. Approximately 750- and 175-fold reductions in the oxygenation efficiency of EPA and LA were observed with V349A oPGHS-1, compared with a 2-fold change for AA. Val-349 contacts C-2 and C-3 of EPA and C-4 of LA orienting the carboxyl halves of these substrates so that the omega-ends are aligned properly for hydrogen abstraction. An S530T substitution decreases the V(max)/K(m) of EPA and LA by 375- and 140-fold. Ser-530 makes six contacts with EPA and four with LA involving C-8 through C-16; these interactions influence the alignment of the substrate for hydrogen abstraction. Interestingly, replacement of Phe-205 increases the volume of the cyclooxygenase site allowing EPA to be oxygenated more efficiently than with native oPGHS-1.  相似文献   

4.
Prostaglandin endoperoxide H synthases (PGHSs)-1 and -2 have a cyclooxygenase (COX) activity involved in forming prostaglandin G2 (PGG2) from arachidonic acid and an associated peroxidase (POX) activity that reduces PGG2 to PGH2. Suicide inactivation processes are observed for both POX and COX reactions. Here we report COX reaction conditions for PGHS-1 under which complete COX inactivation occurs but with > or = 60% retention of POX activity. The rates of POX inactivation were compared for native oPGHS-1 versus Y385F oPGHS-1, a mutant that cannot form the Tyr385 radical of COX Intermediate II; the rates were the same for both native and Y385F oPGHS-1. Our data indicate that a COX Intermediate II/acyl or product complex is the precursor in COX inactivation. However, another species, probably an Intermediate II-like species but with a radical centered on a tyrosine other than Tyr385, is the immediate precursor for POX inactivation.  相似文献   

5.
6.
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   

7.
8.
9.
Arg-120 is located near the mouth of the hydrophobic channel that forms the cyclooxygenase active site of prostaglandin endoperoxide H synthases (PGHSs)-1 and -2. Replacement of Arg-120 of ovine PGHS-1 with a glutamine increases the apparent Km of PGHS-1 for arachidonate by 1,000-fold (Bhattacharyya, D. K., Lecomte, M., Rieke, C. J., Garavito, R. M., and Smith, W. L. (1996) J. Biol. Chem. 271, 2179-2184). This and other evidence indicate that the guanido group of Arg-120 forms an ionic bond with the carboxylate group of arachidonate and that this interaction is an important contributor to the overall strength of arachidonate binding to PGHS-1. In contrast, we report here that R120Q human PGHS-2 (hPGHS-2) and native hPGHS-2 have very similar kinetic properties, but R120L hPGHS-2 catalyzes the oxygenation of arachidonate inefficiently. Our data indicate that the guanido group of Arg-120 of hPGHS-2 interacts with arachidonate through a hydrogen bond rather than an ionic bond and that this interaction is much less important for arachidonate binding to PGHS-2 than to PGHS-1. The Km values of PGHS-1 and -2 for arachidonate are the same, and all but one of the core residues of the active sites of the two isozymes are identical. Thus, the results of our studies of Arg-120 of PGHS-1 and -2 imply that interactions involved in the binding of arachidonate to PGHS-1 and -2 are quite different and that residues within the hydrophobic cyclooxygenase channel must contribute more significantly to arachidonate binding to PGHS-2 than to PGHS-1. As observed previously with R120Q PGHS-1, flurbiprofen was an ineffective inhibitor of R120Q hPGHS-2. PGHS-2-specific inhibitors including NS398, DuP-697, and SC58125 had IC50 values for the R120Q mutant that were up to 1,000-fold less than those observed for native hPGHS-2; thus, the positively charged guanido group of Arg-120 interferes with the binding of these compounds. NS398 did not cause time-dependent inhibition of R120Q hPGHS-2, whereas DuP-697 and SC58125 were time-dependent inhibitors. Thus, Arg-120 is important for the time-dependent inhibition of hPGHS-2 by NS398 but not by DuP-697 or SC58125.  相似文献   

10.
Mouse bone marrow-derived mast cells (BMMC), stimulated with stem cell factor, IL-1beta, and IL-10, secrete IL-6 and demonstrate a delayed phase of PGD(2) generation that is dependent upon the induced expression of PG endoperoxide synthase (PGHS)-2. We have examined the potential for exogenous prostanoids, acting in a paracrine fashion, and endogenous prostanoids, acting in an autocrine fashion, to regulate PGHS-2 induction and IL-6 secretion in mouse BMMC. Exogenous PGE(2), which acts through G protein-coupled receptors, and 15-deoxy-Delta(12,14)-PGJ(2), which is a ligand for peroxisome proliferator-activated receptor (PPAR)gamma, elicited a 2- to 3-fold amplification of PGHS-2 induction, delayed-phase PGD(2) generation, and IL-6 secretion in response to stem cell factor, IL-1beta, and IL-10. The effect of PGE(2) was reproduced by the E prostanoid (EP)1 receptor agonist 17-trinor-PGE(2), and the EP1/EP3 agonist, sulprostone, but not the EP2 receptor agonist, butaprost. Although BMMC express PPARgamma, the effects of 15-deoxy-Delta(12,14)-PGJ(2) were not reproduced by the PPARgamma agonists, troglitazone and ciglitazone. PGHS-2 induction, but not IL-6 secretion, was impaired in cPLA(2)-deficient BMMC. However, there was no impairment of PGHS-2 induction in BMMC deficient in hematopoietic PGD synthase or PGHS-1 in the presence or absence of the PGHS-2 inhibitor, NS-398. Thus, although exogenous prostanoids may contribute to amplification of the inflammatory response by augmenting PGD(2) generation and IL-6 secretion from mast cells, endogenous prostanoids do not play a role.  相似文献   

11.
The inhibitory mechanism of high levels of exogenously added arachidonic acid on activation of washed human platelets was investigated. While low levels of arachidonic acid (5-10 microM) induced aggregation, ATP secretion and increase in cytoplasmic free Ca2+ concentration (first phase of activation), these platelet responses did not occur significantly at high concentrations (30-50 microM). However, much higher concentrations than 80 microM again elicited these responses (second phase). The first phase of platelet activation was inhibited by cyclooxygenase inhibitor, indomethacin, whereas the second one was independent of such treatment. Thromboxane B2 was produced dose-dependently until reaching a plateau at arachidonic acid concentrations higher than 20 microM, irrespective of the lack of aggregation and secretion at high concentrations. After that the amount of free arachidonic acid which remained unmetabolized in platelets gradually increased. High concentrations of arachidonic acid as well as other polyunsaturated fatty acids caused desensitization of platelets in response to U46619, and also depressed the specific [3H]U46619-binding to the receptor as well as other polyunsaturated fatty acids. The amount free arachidonic acid needed in platelets to suppress [3H]U46619 binding corresponded to that needed to inhibit platelet aggregation. Furthermore, arachidonic acid dose-dependently induced fluidization of lipid phase of platelet membranes as detected by 1,6-diphenyl-1,3,5-hexatriene. These results suggest that the inhibition of platelet response by high levels of arachidonic acid can be attributed to interference with endoperoxide/thromboxane A2 binding to the receptor, probably due to perturbation of the membrane lipid phase due to excess amounts of free arachidonic acid remaining in the membranes.  相似文献   

12.
Prostaglandin E(2) (PGE(2)) is the most common prostanoid and has a variety of bioactivities including a crucial role in urogenital function. Multiple enzymes are involved in its biosynthesis. Among 3 PGE(2) terminal synthetic enzymes, membrane-associated PGE(2) synthase-2 (mPGES-2) is the most recently identified, and its role remains uncharacterized. In previous studies, membrane-associated PGE(2) synthase-1 (mPGES-1) and cytosolic PGE(2) synthase (cPGES) were reported to be expressed along the urogenital tracts. Here we report the genomic structure and tissue distribution of mPGES-2 in the urogenital system. Analysis of several bioinformatic databases demonstrated that mouse mPGES-2 spans 7 kb and consists of 7 exons. The mPGES-2 promoter contains multiple Sp1 sites and a GC box without a TATA box motif. Real-time quantitative PCR revealed that constitutive mPGES-2 mRNA was most abundant in the heart, brain, kidney and small intestine. In the urogenital system, mPGES-2 was highly expressed in the renal cortex, followed by the renal medulla and ovary, with lower levels in the ureter, bladder and uterus. Immunohistochemistry studies indicated that mPGES-2 was ubiquitously expressed along the nephron, with much lower levels in the glomeruli. In the ureter and bladder, mPGES-2 was mainly localized to the urothelium. In the reproductive system, mPGES-2 was restricted to the epithelial cells of the testis, epididymis, vas deferens and seminal vesicle in males, and oocytes, stroma cells and corpus luteum of the ovary and epithelial cells of the oviduct and uterus in females. This expression pattern is consistent with an important role for mPGES-2-mediated PGE(2) in urogenital function.  相似文献   

13.
We present a plausible productive conformation obtained by docking calculations for the binding of prostaglandin G2 (PGG2) to the peroxidase site of prostaglandin endoperoxide H synthase-1 (PGHS-1, COX-1). The enzyme-substrate complex stability was verified by molecular dynamics. Structural analysis reveals the requirements for enzyme-substrate recognition and binding: the PGG2 15-hydroperoxide group is in the proximity of the heme iron and participates in a hydrogen bond network with the conserved His207 and Gln203 and a water molecule, whereas the carboxylate group forms salt bridges with the remote Lys215 and Lys222. Site-directed mutagenesis showed that a single mutation of Lys215 or Lys222 does not affect enzyme activity, whereas dual mutation of these residues, to either alanine or glutamate, significantly decreases turnover. This indicates that the conserved cationic pocket is involved in enzyme-substrate binding.  相似文献   

14.
Prostaglandin endoperoxide H synthase-1 (PGHS-1) is an abundant enzyme in platelets, where it plays a key role in the cascade of prostanoid formation. In platelets, the primary site of PGHS-1 synthesis is in precursor megakaryocytic cells. We have previously shown that in megakaryocytic MEG-01 cells, TPA induces an increase of PGHS-1 mRNA within a few hours, whereas protein increase occurs after several days of treatment. We now report that the delayed increase in PGHS-1 protein is caused by translational regulation. De novo PGHS-1 synthesis, measured using [(35)S]methionine pulse labeling followed by immunoprecipitation, was detected at day 4 after TPA treatment but not at day 1. To identify a potential element of PGHS-1 mRNA controlling translation, we compared the 3'-untranslated region from different species and identified a 20-nt segment perfectly conserved. The 20-nt segment was used as a probe in RNA gel mobility-shift assays using MEG-01 extracts from control cells or from TPA-treated cells. Four complexes were formed with extracts from control cells or cells treated with TPA for 1 day but were not observed with extracts from cells treated for 4 days. Of the 4 complexes, one was sequence-specific and binding involved uridylate residues and interactions with a 45-kDa protein and a protein doublet of 116 kDa. Binding of this 45/116-kDa complex to the 20-nt conserved cis element most likely regulates negatively PGHS-1 protein accumulation. We have provided evidence that the PGHS-1 gene is regulated at the translational level.  相似文献   

15.
We investigated the utilization of exogenous 14C-labelled arachidonic acid by the cyclooxygenase system of the gastric mucosa and its alteration by cytosolic factors, protein binding, glutathione peroxidase (GSH-Px), and hydrogen peroxides. Total prostaglandin (PG) synthesis from gastric microsomes was reduced in a dose- dependent manner to 12% and 68% of controls by increasing amounts of the 105,000g supernatant or albumin (8mg protein/ml), respectively (p less than 0.01). The inhibitory cytosolic factor was heat labile, protease sensitive, and was retained by a 300,000 Dalton ultrafiltration membrane. Thus, it was likely a protein. Other possible inhibitory mechanisms like protease- or heme-induced destabilization of the cyclooxygenase, haptoglobin-mediated inhibition, or self-inactivation by endogenous substrate were excluded. N-ethylmaleimide (NEM), an agent that alkylates sulfhydryl-groups thereby inhibiting GSH-Px, abolished the inhibitory effect of cytosol in a dose-dependent fashion. In contrast to their inhibition of prostaglandin synthesis, the binding of arachidonic acid by albumin or cytosolic proteins accounted to 75% and 19% under comparable conditions, respectively, however, cytosolic fatty acid binding was unaffected by NEM. Thus, it was concluded that the inhibitory effect of cytosol, in contrast to albumin, was mediated by a sulfhydryl-depending process, probably a GSH-Px. This conclusion was supported by a qualitatively comparable inhibition by a purified GSH-Px from bovine erythrocytes. The inhibitory action of cytosolic proteins was reduced significantly by increasing concentrations or repeated application of arachidonic acid; therefore, cytosolic GSH-Px was likely to affect substrate utilization by the microsomal PGH synthase through reduction of activating substrate peroxides. Similarly, the in vitro formation of cyclooxygenase products by mucosal homogenate or gastric microsomes in the absence of cytosol was limited at substrate concentrations below 80 microM, despite sufficient nonesterified arachidonic acid remaining in the incubate. This limitation was mediated only partially by self-inactivation of the prostaglandin cyclooxygenase. Neither N-ethylmaleimide nor repeated application of hydrogen peroxides increased substrate utilization by isolated microsomes, excluding contamination by GSH-Px or simply a lack of hydrogen peroxides as possible mechanisms for the limited utilization. From these results, a special role of substrate-linked lipid peroxides in the activation of mucosal prostaglandin synthesis is proposed. The reduction of these peroxides by glutathione dependent or independent peroxidases, e.g. the PGH synthase-linked hydroperoxidase activity itself, could explain the reduced utilization of nonesterified arachidonic acid by the gastric mucosa.  相似文献   

16.
Despite the high degree of sequence homology between two mouse proteins AIC2A and AIC2B (91% at the amino acid level), only the AIC2A protein binds interleukin 3 (IL-3). Soluble AIC2A protein bound IL-3 with affinity similar to the membrane-bound AIC2A protein, indicating that binding of IL-3 to AIC2A was mediated by the external domain alone. The extracellular domain of the AIC2A protein has two repeats of the common motif shared by members of the cytokine receptor family. Neither one of these repeats alone bound IL-3. Hybrids of AIC2A and AIC2B revealed that the first domain of the cytokine receptor motif could be replaced with the AIC2B sequence without an affinity change, suggesting the importance of the second domain. By changing individual amino acid residues of AIC2A in the second domain which differ from those of AIC2B, we identified several amino acid residues critical for IL-3 binding. All these residues are located at the putative hinge region within the second domain.  相似文献   

17.
Arachidonic acid (AA) incorporation into phospholipids and cyclooxygenase and lipoxygenase mediated metabolism of arachidonic acid were studied in homogenized and intact Neuro-2A cells. When 3H8-AA was added to homogenized cells and incubated 20 minutes, 39% of the label was converted to prostaglandins (PGs), 10% to hydroxy-eicosatetraenoic acid (HETE) and 26% was incorporated into phospholipids. PGE2 and PGF2a were the major PGs produced. Synthesis of PGs was blocked by 10 microM indomethacin and synthesis of PGs and HETE was blocked by 10 microM eicosatetraynoic acid (ETYA). The cell homogenate produced the 13,14-dihydro-15-keto metabolites of PGE2 and PGF2a from 3H8-AA and also converted exogenous 3H7-PGE2 and 3H8-PGF2a to metabolites. When intact cells were labeled for 24 hours with 14C1-AA and the cells and media then analyzed, 75% of the radioactivity was incorporated into cellular phospholipids, 0.8% was converted to PGs and metabolites and 0.7% converted to HETE. Cells prelabeled for 24 hours were washed and incubated for 30 minutes in fatty acid free media. There was a 23% release of AA from phospholipids. One-fifth of the released AA was converted to HETE. PG synthesis in the intact resting cells was low. In summary, the Neuro-2A cell provides a good model system for studying arachidonic acid metabolism and incorporation into phospholipids in cells of neuronal origin.  相似文献   

18.
Apolipoprotein (apo-) E3, when combined with the phospholipid dimyristoylphosphatidylcholine (DMPC), binds avidly to apo-B,E (low density lipoprotein) receptors on human fibroblasts. Apolipoprotein E2 isolated from type III hyperlipoproteinemic subjects, which differs from apo-E3 by the presence of cysteine instead of arginine at residue 158, possesses only about 1% of the receptor binding activity of apo-E3. Modification of apo-E2 with cysteamine, which converts the cysteine at position 158 to a positively charged lysine analogue, activates receptor binding approximately 13-fold. In the present experiments, thrombin was used to cleave apo-E2 into two fragments (Mr = 22,000 and Mr = 10,000). The larger fragment, which has been shown to possess the receptor binding domain, displayed binding activity up to 12-fold greater than intact apo-E2 or equivalent to apo-E2 treated with cysteamine. When the Mr = 22,000 fragment was modified with cysteamine and combined with DMPC, receptor binding was further enhanced, attaining the level of activity of normal apo-E3 X DMPC, a 100-fold increase over apo-E2 X DMPC binding. When the cysteamine modification was reversed by incubation with beta-mercaptoethanol, the Mr = 22,000 fragment retained most of its binding activity. However, when the same sample was tested 24 h later, the level of binding activity dropped significantly. The receptor binding of apo-E2-containing beta-very low density lipoproteins could also be activated by cysteamine treatment, with the same retention of enhanced binding activity occurring after the reversal of the modification. These results indicate that apo-E2 can attain full binding activity by the removal of the carboxyl-terminal one-third of the molecule and the addition of a positive charge at residue 158 of the molecule. The retention of enhanced binding after the reversal of the cysteamine modification indicates that the enhanced binding is probably due to conformational changes induced in the binding domain (and maintained by the phospholipid) and not merely to the presence of the positive charge at residue 158.  相似文献   

19.
We investigated the utilization of exogenous 14C-labelled arachidonic acid by the cyclooxygenase system of the gastric mucosa and its alteration by cytosolic factors, protein binding, glutathione peroxidase (GSH-Px), and hydrogen peroxides.Total prostaglandin (PG) synthesis from gastric microsomes was reduced in a dose- dependent manner to 12% and 68% of controls by increasing amounts of the 105,000g supernatant or albumin (8mg protein/ml), respectively (p<0.01). The inhibitory cytosolic factor was heat labile, protease sensitive, and was retained by a 300,000 Dalton ultrafiltration membrane. Thus, it was likely a protein. Other possible inhibitory mechanisms like protease- or heme-induced destabilization of the cyclooxygenase, haptoglobin-mediated inhibition, or self-inactivation by endogenous substrate were excluded.N-ethylmaleimide (NEM), an agent that alkylates sulfhydryl-groups thereby inhibiting GSH-Px, abolished the inhibitory effect of cytosol in a dose-dependent fashion. In contrast to their inhibition of prostaglandin synthesis, the binding of arachidonic acid by albumin or cytosolic proteins accounted to 75% and 19% under comparable conditions, respectively, however, cytosolic fatty acid binding was unaffected by NEM. Thus, it was concluded that the inhibitory effect of cytosol, in contrast to albumin, was mediated by a sulfhydryl-depending process, probably a GSH-Px. This conclusion was supported by a qualitatively comparable inhibition by a purified GSH-Px from bovine erythrocytes.The inhibitory action of cytosolic proteins was reduced significantly by increasing concentrations or repeated application of arachidonic acid; therefore, cytosolic GSH-Px was likely to affect substrate utilization by the microsomal PGH synthase through reduction of activating substrate peroxides.Similarly, the in vitro formation of cyclooxygenase products by mucosal homogenate or gastric microsomes in the absence of cytosol was limited at substrate concentrations below 80μM, despite sufficient nonesterified arachidonic acid remaining in the incubate. This limitation was mediated only partially by self-inactivation of the prostaglandin cyclooxygenase. Neither N-ethylmaleimide nor repeated application of hydrogen peroxides increased substrate utilization by isolated microsomes, excluding contamination by GSH-Px or simply a lack of hydrogen peroxides as possible mechanisms for the limited utilization. From these results, a special role of substrate-linked lipid peroxides in the activation of mucosal prostaglandin synthesis is proposed. The reduction of these peroxides by glutathione dependent or independent peroxidases, e.g. the PGH synthase-linked hydroperoxidase activity itself, could explain the reduced utilization of nonesterified arachidonic acid by the gastric mucosa.  相似文献   

20.
Two mutant forms of fumarase C from E. coli have been made using PCR and recombinant DNA. The recombinant form of the protein included a histidine arm on the C-terminal facilitating purification. Based on earlier studies, two different carboxylic acid binding sites, labeled A- and B-, were observed in crystal structures of the wild type and inhibited forms of the enzyme. A histidine at each of the sites was mutated to an asparagine. H188N at the A-site resulted in a large decrease in specific activity, while the H129N mutation at the B-site had essentially no effect. From the results, we conclude that the A-site is indeed the active site, and a dual role for H188 as a potential catalytic base is proposed. Crystal structures of the two mutant proteins produced some unexpected results. Both mutations reduced the affinity for the carboxylic acids at their respective sites. The H129N mutant should be particularly useful in future kinetic studies because it sterically blocks the B-site with the carboxyamide of asparagine assuming the position of the ligand's carboxylate. In the H188N mutation at the active site, the new asparagine side chain still interacts with an active site water that appears to have moved slightly as a result of the mutation.  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