Enzymology: Soluble Epoxide Hydrolase as an Anti-inflammatory Target of the Thrombolytic Stroke Drug SMTP-7 Naoki Matsumoto, Eriko Suzuki, Makoto Ishikawa, Takumi Shirafuji and Keiji Hasumi J. Biol. Chem. 2014, 289:35826-35838. doi: 10.1074/jbc.M114.588087 originally published online October 31, 2014

Find articles, minireviews, Reflections and Classics on similar topics on the JBC Affinity Sites. Alerts: • When this article is cited • When a correction for this article is posted Click here to choose from all of JBC's e-mail alerts This article cites 65 references, 13 of which can be accessed free at http://www.jbc.org/content/289/52/35826.full.html#ref-list-1

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

Access the most updated version of this article at doi: 10.1074/jbc.M114.588087

THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 289, NO. 52, pp. 35826 –35838, December 26, 2014 © 2014 by The American Society for Biochemistry and Molecular Biology, Inc. Published in the U.S.A.

Soluble Epoxide Hydrolase as an Anti-inflammatory Target of the Thrombolytic Stroke Drug SMTP-7* Received for publication, June 16, 2014, and in revised form, September 23, 2014 Published, JBC Papers in Press, October 31, 2014, DOI 10.1074/jbc.M114.588087

Naoki Matsumoto‡1, Eriko Suzuki‡1, Makoto Ishikawa§, Takumi Shirafuji§, and Keiji Hasumi‡¶2 From the ‡Department of Applied Biological Science, Tokyo Noko University, 3-5-8 Saiwaicho, Fuchu, Tokyo 183-8509, Japan, § Pharmaceutical Research Laboratory, Nihon Pharmaceutical, 34 Shin-izumi, Narita, Chiba 286-0825, Japan, and ¶TMS Co., Ltd., 1-32-1-102 Fuchucho, Fuchu, Tokyo 183-0055, Japan Background: The small molecule thrombolytic SMTP-7 excels in the treatment of ischemic stroke through an unidentified anti-inflammatory activity. Results: Affinity purification and inhibition studies reveal that SMTP-7 and its congeners inhibit soluble epoxide hydrolase (sEH), an enzyme responsible for inflammation. Conclusion: sEH is the plausible in vivo anti-inflammatory target of SMTP-7. Significance: Dual-targeting of thrombolysis and sEH is a promising strategy for novel stroke therapy.

Stroke is a major cause of death and disability worldwide (1). Improving the case fatality rates and long term disability after stroke continues to be a challenge (2). During the past decade thrombolytic enzyme tissue plasminogen activator-based treatment has been the standard therapy for acute ischemic stroke (2). However, due to its risk of parenchymal hemorrhage and a limited therapeutic time window (3–5), only a small fraction of patients (1.8 – 8.9%) benefits from tissue plasminogen activator-based therapy (6, 7). The development of an alternative therapeutic agent is urgently needed. In this context, suppressing inflammation within the infarction area to rescue pen-

* This work was supported in part by Japan Science and Technology Agency Grant AS2316911G (to K. H.) a grant from the Ministry of Education, Culture, Sports, Science and Technology (Translational Research Network Program C7; to K. H.). 1 Joint first authors. 2 To whom correspondence should be addressed: Dept. of Applied Biological Science, Tokyo Noko University, 3-5-8 Saiwaicho, Fuchu, Tokyo 183-8509, Japan. Tel.: 81-42-367-5710; Fax: 81-42-367-5708; E-mail: hasumi@cc. tuat.ac.jp.

35826 JOURNAL OF BIOLOGICAL CHEMISTRY

umbral tissue and reduce hemorrhagic transformation is particularly important (8, 9). SMTPs3 (named after Stachybotrys microspora triprenyl phenols) are a family of novel small molecules produced by the fungus Stachybotrys microspora (10 –12). SMTP-7 (Fig. 1A), one of the SMTP family compounds with profound biological activities, enhances plasminogen activation by modulating plasminogen conformation (10, 13–15). SMTP-7 thus promotes plasmin formation and clot clearance in vivo (14, 16), and it has been used to treat thrombotic and embolic strokes in experimental models in rodents and nonhuman primates (17– 20). Unexpectedly, SMTP-7 was demonstrated to reduce hemorrhagic transformation and to have a wide therapeutic time window (17, 18, 20). These excellent activities are partly explained by the finding that, unlike tissue plasminogen activator, SMTP-7 suppresses inflammatory/oxidative responses after thrombolytic reperfusion (16, 18, 19, 21). These beneficial properties prompted the development of SMTP-7 as an alternative stroke drug that can treat patients who do not benefit from tissue plasminogen activator-based therapy (20). The aim of the present study was to explore the anti-inflammatory mechanism of SMTP-7. We first compared the efficacy of two SMTP congeners differing in the potential of thrombolytic activity using three inflammatory disease animal models. The results demonstrate that the anti-inflammatory action of SMTP is independent of its thrombolytic activity. We next searched for an anti-inflammatory target using SMTP affinity beads resulting in the identification of soluble epoxide hydrolase (sEH). sEH is a bifunctional enzyme with an epoxide hydrolase activity at the C-terminal domain (Cterm-EH) and a lipid phosphate phosphatase activity at the N-terminal domain 3

The abbreviations used are: SMTP, S. microspora triprenyl phenol; sEH, soluble epoxide hydrolase; Cterm-EH, C-terminal epoxide hydrolase domain/ activity of sEH; Nterm-phos, N-terminal phosphatase domain/activity of sEH; EET, epoxyeicosatrienoic acid; DHET, dihydroxyeicosatrienoic acid; TNBS, 2,4,6-trinitrobenzene sulfonic acid; AUDA, 12-(3-adamantan-1yl-ureido)dodecanoic acid; PHOME, (3-phenyl-oxiranyl)-acetic acid cyano(6-methyl-naphthalen-2-yl)-methyl ester; PG, prostaglandin; HETE, hydroxyeicosatetraenoic acid; Fmoc, N-(9-fluorenyl)methoxycarbonyl; Bis-Tris, 2-[bis(2-hydroxyethyl)amino]-2-(hydroxymethyl)propane-1,3-diol.

VOLUME 289 • NUMBER 52 • DECEMBER 26, 2014

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

Although ischemic stroke is a major cause of death and disability worldwide, only a small fraction of patients benefit from the current thrombolytic therapy due to a risk of cerebral hemorrhage caused by inflammation. Thus, the development of a new strategy to combat inflammation during thrombolysis is an urgent demand. The small molecule thrombolytic SMTP-7 effectively treats ischemic stroke in several animal models with reducing cerebral hemorrhage. Here we revealed that SMTP-7 targeted soluble epoxide hydrolase (sEH) to suppress inflammation. SMTP-7 inhibited both of the two sEH enzyme activities: epoxide hydrolase (which inactivates anti-inflammatory epoxyfatty acids) and lipid phosphate phosphatase. SMTP-7 suppressed epoxy-fatty acid hydrolysis in HepG2 cells in culture, implicating the sEH inhibition in the anti-inflammatory mechanism. The sEH inhibition by SMTP-7 was independent of its thrombolytic activity. The simultaneous targeting of thrombolysis and sEH by a single molecule is a promising strategy to revolutionize the current stroke therapy.

Anti-inflammatory Target of a Novel Thrombolytic Stroke Drug (Nterm-phos) (22). The Cterm-EH catalyzes the hydrolysis of epoxy fatty acids such as epoxyeicosatrienoic acids (EETs), which are potent endogenous signaling molecules implicated in anti-inflammation, vascular dilation, endothelial cell hyperpolarization, angiogenesis, neuroprotection, and analgesia (antihyperalgesia) (23–27). The Nterm-phos hydrolyzes lipid phosphates such as lysophosphatidic acid and intermediates of the cholesterol biosynthesis (28 –30). SMTP-7 inhibited both of the two activities of sEH: competitively for Cterm-EH and pseudononcompetitively for Nterm-phos. The treatment of HepG2 cells with of SMTP-7 or its congeners suppressed the hydrolysis of EET to dihydroxyeicosatrienoic acid (DHET). Here, we describe the anti-inflammatory action of SMTP-7 and its mechanism.

DECEMBER 26, 2014 • VOLUME 289 • NUMBER 52

JOURNAL OF BIOLOGICAL CHEMISTRY

35827

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

EXPERIMENTAL PROCEDURES Animal Experiments—All of the animal protocols were approved by the institutional animal experiment committees at the Tokyo Noko University and Nihon Pharmaceutical. Male C57BL/6J mice (7 weeks old; Japan SLC, Hamamatsu) and male Lewis rats (6 weeks old; Charles River Laboratories Japan, Yokohama) were used after 1 week of preliminary rearing for the inflammatory disease models. Retired male ICR mice (Japan SLC, Hamamatsu) were used to obtain the livers for the sEH purification. SMTP Congeners—The SMTP congeners used in this study were produced by S. microspora as described previously (11– 13, 31–34). For in vivo studies, SMTP-7 and SMTP-44D were converted to respective sodium salts. Guillain-Barre´ Syndrome Model—On day 0, male Lewis rats (7 weeks old) were given 0.1 ml of a 1:1 mixture of the synthetic P2 peptide (35) (H-TESPFKNTEISFKLGQEFEETTADNR-OH corresponding to Thr-53–Arg-78 of bovine P2 protein; 2.5 mg ml⫺1 in saline) and Freund’s incomplete adjuvant containing Mycobacterium tuberculosis H37 Ra (Difco) by injection into the foot pads of both of the hind limbs. Sodium salts of SMTP-7 and SMTP-44D dissolved in 5% (wt vol⫺1) mannitol was intraperitoneally injected at a dose of 10 mg kg⫺1 through days 7–16. As a standard, sulfonated immunoglobulin formulation (Kaketsuken, Kumamoto, Japan) was given intravenously at a dose of 400 mg kg⫺1 on days 0, 7, 14, 15, and 16. Control animals received no drug treatment. Normal animals received neither P2 peptide nor any drug. The experimental autoimmune neuritis score (score 0, normal; 0.5, reduced tone of the tail; 1, limp tail; 2, moderate paraparesis; 3; severe paraparesis; 4, tetraparesis or death) was determined on days 7, 9, 11, 13, 15, 17, 20, and 24. There were five animals in each group. Ulcerative Colitis Model—Male C57BL/6J mice (8 weeks old) were given dextran sulfate sodium salt (36 –50 kDa) (36) contained in drinking water (2%, wt vol⫺1) daily through days 0 –7. Sodium salt of SMTP-7 or SMTP-44D was intraperitoneally injected at a dose of 10 mg kg⫺1. As standards, 5-aminosalicylic acid (37) (10 mg ml⫺1 in 0.5% carboxymethylcellulose) or sodium prednisolone succinate (2 mg ml⫺1 in phosphate-buffered saline) were given orally at a dose of 100 and 20 mg kg⫺1, respectively. Drugs were given daily through days 0 –7. Control animals received no drug treatment. Normal animals received neither dextran sulfate nor any drug. The disease activity index

score (38, 39), determined based on the change in body weight (score 0, ⬍1%; 1, 1–5%; 2, 5–10%; 3, 10 –20%; 4, ⬎20%), stool inconsistency (score 0, normal; 2, loose; 4, diarrhea), and stool blood (score 0, negative; 2, occult blood; 4, gross bleeding) was measured on days 2, 4, 5, 6, 7, and 8. There were five animals in each group. Crohn Disease Model—Male C57BL/6J mice (8 weeks old) were intrarectally injected with 2,4,6-trinitrobenzene sulfonic acid (TNBS) (40) solution (20 mg ml⫺1 in 50% ethanol) at a dose of 100 mg kg⫺1. SMTP-7, SMTP-44D, 5-aminosalicylic acid, or prednisolone was administered as described for the dextran sulfate model. These treatments were made 30 min before the TNBS injection as well as 24, 48, and 72 h after the TNBS injection. Control animals received no drug treatment. Normal animals received neither TNBS nor any drug. The disease activity index score was measured on days 1, 2, 3, and 4 after the TNBS injection. There were five animal in each group. SMTP-47—SMTP-47 was synthesized by the microbial amine feeding method (31). For the preparation of feeding amine, N␣-tert-butoxycarbonyl-N␦-Fmoc-L-ornithine (60 mg ml⫺1 in tetrahydrofuran) was treated with equal volume of trifluoroacetic acid, affording N␦-Fmoc-L-ornithine. The culture of S. microspora IFO 30018 (100 ml) was fed with 100 mg of N␦-Fmoc-L-ornithine, and the resulting SMTP-47 (89 mg) was purified by reverse-phase HPLC developed with MeOH, 0.1% formic acid (85:15). 1H NMR (acetone-d6, 600 MHz): ␦ 7.83 (2H, d, J ⫽ 7.3 Hz), 7.67 (2H, d, J ⫽ 7.3 Hz), 7.38 (2H, m), 7.29 (2H, t, J ⫽ 7.3 Hz), 6.80 (1H, d, J ⫽ 2.2 Hz), 5.16 (1H, t, J ⫽ 6.6 Hz), 5.06 (1H, m), 4.97 (1H, dd, J ⫽ 4.4, 11.0 Hz), 4.44 (1H, d, J ⫽ 16.1 Hz), ⬃4.31 (3H, m), 4.20 (1H, t, J ⫽ 6.6 Hz), 3.95 (1H, dd, J ⫽ 5.9, 7.3 Hz), 3.23 (2H, m), 3.02 (1H, dd, J ⫽ 5.9, 17.6 Hz), 2.68 (1H, dd, J ⫽ 7.3, 17.6 Hz), 2.20 (2H, m), ⬃2.06 (4H, m), 1.95 (2H, m), 1.71 (2H, m), 1.62 (3H, s), 1.57 (3H, s), ⬃1.56 (2H, m), 1.55 (3H, s), 1.28 (3H, s). 13C NMR (acetone-d6, 150 MHz): ␦ 172.95, 169.71, 157.18, 157.13, 149.88, 145.18, 142.09, 135.70, 132.56, 131.65, 128.43, 127.87, 126.08, 125.19, 125.12, 121.77, 120.73, 112.76, 100.86, 79.93, 67.61, 66.68, 54.30, 48.15, 44.97, 40.90, 40.40, 38.42, ⬃27.67 (2 signals overlapped), ⬃27.64, 27.38, 25.80, 22.23, 18.79, 17.70, 15.99. MALDI-TOF-MS (m/z): [M ⫹ Na]⫹ calculated for C43H50N2NaO8, 745.3465; found, 745.3575. UV (MeOH): ␭max nm (⑀) 208 (86, 390), 263 (29, 330), 289 (7, 370), 300 (9, 100). IR (neat): ␯max cm⫺1 3329, 3064, 2968, 2926, 2866, 1701, 1670, 1620, 1533, 1462, 1356, 1252, 1157, 1076, 847, 744, 542. SMTP-50 —To prepare SMTP-50, SMTP-47 (90 mg) was treated with piperidine/N,N-dimethylformamide (6:1) for 1 h to afford SMTP-50, which was purified by reverse-phase HPLC developed with a linear gradient of MeOH in 0.1% formic acid (60 –100%) for 20 min, yielding 30.3 mg of the purified material. 1 H NMR (CD3OD, 600 MHz): ␦ 6.74 (1H, s), 5.12 (1H, t, J ⫽ 6.9 Hz), 5.06 (1H, t, J ⫽ 6.9 Hz), 4.75 (1H, dd, J ⫽ 4.8, 10.2 Hz), 4.61 (1H, d, J ⫽ 16.8 Hz), 4.26 (1H, d, J ⫽ 17.4 Hz), 3.88 (1H, t, J ⫽ 6.0 Hz), 2.97 (3H, m), 2.64 (1H, dd, J ⫽ 7.2, 17.4 Hz), 2.16 (3H, m), 2.05 (2H, m), 1.96 (2H, m), 1.90 (1H, m), 1.68 (1H, m), 1.64 (3H, s), ⬃1.61 (3H, m), 1.59 (3H, s), 1.57 (3H, s), 1.26 (3H, s). 13C NMR (CD3OD, 150 MHz): ␦ 177.17, 171.83, 157.82, 150.01, 136.39, 132.66, 132.13, 125.39, 125.36, 122.24, 113.21, 100.91, 80.09, 68.22, 57.09, 45.92, 40.81, 40.01, 38.72, 28.61, 27.73 (2

Anti-inflammatory Target of a Novel Thrombolytic Stroke Drug

35828 JOURNAL OF BIOLOGICAL CHEMISTRY

natant was concentrated to dryness. The resulting materials were dissolved in methanol and subjected to LC-MS analysis for 14,15-EET and 14,15-DHET on a Micromass Quattro Ultima triple quadrupole tandem mass spectrometer equipped with an electrospray ionization interface (Waters, Tokyo, Japan). Samples (10 ␮l) were resolved on a silica-ODS column (100 ⫻ 2 mm; Pegasil ODS SP100 –3, Senshu Scientific, Tokyo, Japan) developed at 0.2 ml min⫺1 with a linear gradient of acetonitrile in 0.1% formic acid (60 –100%) for 15 min). The electrospray ionization was performed in the negative ion mode with a capillary voltage at 3.0 kV. The cone voltages were set at 35 V for 14,15-DHET/14,15-DHET-d11, and 30 V for 14,15EET/14,15-EET-d11. Data were acquired in the multichannel analysis mode and analyzed using the MassLynx software (Version 3.5; Waters). Assay for sEH—The Cterm-EH activity was assayed with 14,15-EET or (3-phenyl-oxiranyl)-acetic acid cyano-(6-methoxy-naphthalen-2-yl)-methyl ester (PHOME) as a substrate. When using PHOME, we preincubated mouse sEH (60 ng) for 10 min in 80 ␮l of 25 mM Bis-Tris-HCl, pH 7.0, containing 0.1 mM MgCl2 and 0.1 mg ml⫺1 bovine serum albumin (buffer C) with or without a compound to be tested. The composition of buffer C was based on the method by Tran et al. (28), unless MgCl2 was added to unify the buffer composition with that for Nterm-phos determination, in which MgCl2 is essential. After adding 20 ␮l of the substrate, fluorescence (excitation, 355 nm; emission, 460 nm) of the reaction product was measured kinetically at 30 °C. The final concentrations of sEH and PHOME in the standard assay conditions were 4.7 nM and 12.5 ␮M, respectively. When using EET, we incubated mouse sEH (0.18 ng) with 14,15-EET in 60 ␮l of buffer C at 30 °C for 20 min. After the addition of 14,15-EET-d11 and 14,15-DHET-d11 (25 pmol and 5 pmol, respectively) as internal standards, 14,15-DHET formed was determined by LC-MS as described above. The Ntermphos activity was assayed with AttoPhos as a substrate. Mouse sEH (30 ng) was preincubated for 10 min in 80 ␮l of buffer C with or without a compound to be tested. After adding 20 ␮l of AttoPhos, fluorescence (excitation, 450 nm; emission, 545 nm) of the reaction product was measured kinetically at 30 °C. The final concentrations of sEH and AttoPhos in the standard assay conditions were 2.3 nM and 5 ␮M, respectively. Assay for 14,15-EET Hydrolysis in Cultured Cells—HepG2 cells (5 ⫻ 104 cells) were seeded on 24-well plates and cultured overnight. Cells were then washed with Hanks’ balanced salt solution containing 20 mM Hepes, pH 7.4, 0.1 mg ml⫺1 bovine serum albumin, and 1 mM MgCl2 (buffer D) and subsequently treated with various concentrations of SMTP-7 (0 –30 ␮M) in 500 ␮l of buffer D for 10 min. After the addition of 14,15-EET (0.3 ␮M) to the culture for 40 min, the reaction was stopped by adding 2-propanol (300 ␮l). After the addition of 14,15-EETd11 and 14,15-DHET-d11 (200 and 10 pmol, respectively), the amounts of 14,15-EET remaining and 14,15-DHET formed were determined by LC-MS as described above. Global Analysis of Arachidonate Metabolites of the Cytochrome P450, Cyclooxygenase, and Lipoxygenase Pathways— Plasma (400 ␮l) obtained from Guillain-Barre´ syndrome model rats 2 h after SMTP-7 or saline treatment (n ⫽ 8) on day 13 was randomly paired within each group, and the mixture (n ⫽ 4 for VOLUME 289 • NUMBER 52 • DECEMBER 26, 2014

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

signals overlapped), 25.89, 25.81, 22.53, 18.95, 17.76, 16.04. Electrospray ionization-TOF-MS (m/z): [M ⫹ H]⫹ calculated for C28H41N2O6, 501.2965; found, 501.3003. UV (MeOH): ␭max nm (⑀): 216 (41, 700), 260 (9, 600), 300 (2, 800). IR (neat): ␯max cm⫺1 3386, 3208, 2967, 2926, 1657, 1611, 1466, 1381, 1245, 1213, 1163, 1080, 1047. [␣]D ⫽ ⫺12.1° (c. 1.0, MeOH). SMTP-50-coupled Affinity Matrix—One column volume (1 ml for target identification or 5 ml for target purification) of 1.2 mM SMTP-50, dissolved in 10 mM sodium phosphate, pH 8.3, was applied to a HiTrap NHS-activated HP column (GE Healthcare) at room temperature for 30 min. The column was then treated with monoethanolamine to block residual N-hydroxysuccinimidyl group of the matrix. Identification of SMTP-binding Protein—The livers from male ICR mice were perfused with ice-cold saline and homogenized in 4 volumes of 25 mM Tris-HCl, pH 7.4, containing 150 mM NaCl and 0.2% (wt vol⫺1) sodium deoxycholate (buffer A). A supernatant fraction was obtained after centrifugation at 1000 ⫻ g for 15 min followed by 20,000 ⫻ g for 15 min at 4 °C. After filtration, 15 ml of the supernatant was applied to a 1-ml SMTP-affinity column pre-equilibrated with buffer A at 20 °C. The column was washed with 10 ml of buffer A and developed with 10 ml of buffer A containing additional NaCl (500 mM, finally). Aliquots of the eluate were resolved by reduced SDSpolyacrylamide gel electrophoresis. Coomassie Brilliant Blue R250-stained protein bands were excised from the gel and digested with trypsin. The resulting peptides were subjected to chemically assisted fragmentation post-source decay analysis by matrix-assisted laser desorption/ionization (MALDI) timeof-flight (TOF) mass spectrometry on a Ettan MALDI-TOF Pro (Amersham Biosciences) or LC-MALDI-TOF/TOF analysis on a 4700 Proteomics Analyzer (Applied Biosystems, Framingham, MA). Detected masses were subjected to amino acid sequence analysis and to comparison with theoretical peptide masses on MASCOT search engine (Matrixscience, Boston, MA) to identify the protein species. Purification of sEH by SMTP Affinity Chromatography—All of the following operations were carried out at 0 – 4 °C. The livers from male ICR mice were homogenized in 11.5 volumes of 76 mM sodium phosphate, pH 7.4. A supernatant fraction was obtained after centrifugation at 1000 ⫻ g for 10 min followed by 10,000 ⫻ g for 25 min and then 100,000 ⫻ g for 60 min. The resulting cytosol fraction (250 ml) was applied to a 5-ml SMTPcolumn pre-equilibrated with 76 mM sodium phosphate, pH 7.4, containing 0.1 mM EDTA (buffer B). After washing with buffer B (200 ml), the column was developed with 60 ml of buffer B containing 10 ␮M 12-(3-adamantan-1-yl-ureido)dodecanoic acid (AUDA). The eluate was dialyzed against buffer B to remove AUDA and ultrafiltered to concentrate and exchange buffer to 100 mM sodium phosphate, pH 7.4, containing 3 mM dithiothreitol. From 3 batches of affinity chromatography, 2.2 mg of homogeneous sEH was purified. The purified sEH had specific activities of 511 nmol min⫺1 mg⫺1 for the Cterm-EH and 2850 nmol min⫺1 mg⫺1 for the Nterm-phos when we used trans-stilbene oxide and p-nitrophenyl phosphate as respective substrates. LC-MS Analysis of EET and DHET—Samples to be analyzed were extracted with ethyl acetate. After centrifugation, super-

Anti-inflammatory Target of a Novel Thrombolytic Stroke Drug

each group) was centrifuged at 5000 ⫻ g for 10 min. The resulting supernatant was mixed with 80 ␮l of formic acid, and the mixture was applied to Sep-Pak C18 Plus Short Cartridges (Waters, Tokyo, Japan). The column was washed with EtOHwater-formic acid (10:100:1, vol⫺1 vol⫺1), and metabolites of interest were eluted with 5 ml of EtOH. The eluate was evaporated and dissolved with 40 ␮l of 50% aqueous MeOH. Aliquots (10 ␮l) were subjected to LC-MS/MS analysis on an L-column2 ODS (2 ␮m, 1 ⫻ 150 mm, CERI, Tokyo, Japan) developed at a rate of 0.1 ml⫺1 with a liner gradient (10–85%) of acetonitrile in 5 mM ammonium formate-formic acid (1000:1, vol⫺1 vol⫺1) for 26 min. Eluates were ionized with electrospray ionization, and negative ions of oxylipins were monitored on API 3200 QTRAP (AB SCIEX, Tokyo, Japan). Metabolites to be analyzed were the following 48: (⫾)-5,6-DHET, (⫾)-8,9-DHET, (⫾)-11,12DHET, (⫾)-14,15-DHET, (⫾)-5,6-EET, (⫾)-8,9-EET, (⫾)11,12-EET, (⫾)-14,15-EET, prostaglandin (PG) A2, PGB2, PGD2, 6-keto-PGE1, PGE2, 6-keto-PGF1␣, PGF2␣, 15-ketoPGF2␣, 2,3-dinor-8-iso-PGF2␣, PGJ2, ␦-12-PGJ2, 15-deoxy-␦12,14-PGJ2, thromboxane (TX)B2, 11-dehydro-TXB2, leukotriene (LT) B4, 12-keto-LTB4, 20-COOH-LTB4, LTC4, LTD4, LTE4, LTF4, 20-OH-LTB4, 5(S)-hydroxyeicosatetraenoic acid DECEMBER 26, 2014 • VOLUME 289 • NUMBER 52

(HETE), 8(R)-HETE, 9(S)-HETE, 11(R)-HETE, 12(R)-HETE, 15(S)-HETE, 16(R)-HETE, 19(S)-HETE, 20-HETE, 5(S)-hydroperoxyeicosatetraenoic acid (HPETE), 12(S)-HPETE, 15(S)HPETE, 5-oxo-eicosatetraenoic acid (ETE), 12-oxoETE, 15oxoETE, 5(S),6(S)-lipoxin (LX) A4, 5(S),14(R)-LXB4, and hepoxilin A3.

RESULTS Anti-inflammatory Action Independent of Plasminogen Modulation—Although SMTP-7 reduces inflammatory responses in thromboembolic stroke models, it was unclear whether or not this outcome is a consequence of the recanalization by thrombolytic enhancement. We thus employed inflammatory disease models (Guillain-Barre´ syndrome, ulcerative colitis, and Crohn disease models) that were irrelevant to thromboembolic complications to directly assess the anti-inflammatory action of SMTP-7. We compared the efficacy of SMTP-7 with that of its congener, SMTP-44D (11) (Fig. 1A), which is essentially inactive in the plasminogen modulation activity (Fig. 1B). In the Guillain-Barre´ syndrome model in rats, SMTP-7 (10 mg kg⫺1) and SMTP-44D (10 mg kg⫺1) both ameliorated neuritis symptoms as did the clinically used immunoglobulin formulaJOURNAL OF BIOLOGICAL CHEMISTRY

35829

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

FIGURE 1. Both SMTP-7 and SMTP-44D, a thrombolytically inactive congener, are effective in treating inflammatory disease in rodent models. A, the structures of SMTP-7 and SMTP-44D. B, plasminogen modulation activity. The figure was reproduced based on published data (11). C, efficacy in a rat Guillain-Barre´ syndrome model. EAN score, experimental autoimmune neuritis score. D–G, efficacy in a mouse ulcerative colitis model. Disease activity index (DAI) score, disease activity index score. H–K, efficacy in a mouse Crohn disease model. SMTP-7, 10 mg kg⫺1; SMTP-44D, 10 mg kg⫺1; 5-aminosalicylic acid (5-ASA), 100 mg kg⫺1; prednisolone, 20 mg kg⫺1; sulfonated immunoglobulin formulation, 400 mg kg⫺1. *, p ⬍ 0.05; **, p ⬍ 0.01 by Dunnett’s test in comparison with control. #, p ⬍ 0.05 by Mann-Whitney’s U test in comparison with control. Each value represents the mean ⫾ S.D. obtained from five animals.

Anti-inflammatory Target of a Novel Thrombolytic Stroke Drug

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

FIGURE 2. Identification of sEH as an intracellular target of SMTP. A, the synthesis of SMTP-50-coupled affinity matrix. We first synthesized SMTP-47 by the precursor amine-fed culture of S. microspora IFO 30018 using N␦-Fmoc-L-ornithine as a feeding amine. The Fmoc group was removed to afford SMTP-50, which was then coupled with N-hydroxysuccinimidyl-activated cross-linked agarose, affording SMTP-50-coupled affinity matrix. B, SDS-polyacrylamide gel electrophoresis of the eluates from the affinity chromatography of mouse liver homogenates. Arrowheads denote the protein bands specifically bound to the affinity matrix, and asterisks denote nonspecifically bound protein bands as judged by comparison with the results obtained with control matrix. C, summary of the peptide mass fingerprinting analysis of protein bands 1– 4. D, results of the peptide mass fingerprinting analysis of protein band 1. The bars are for the found peptides cover the peptide sequences in Table 1. E, SDS-polyacrylamide gel electrophoresis of the preparation specifically eluted with AUDA by SMTP-affinity chromatography.

tion (400 mg kg⫺1) (Fig. 1C). Both SMTP-7 (10 mg kg⫺1) and SMTP-44D (10 mg kg⫺1) alleviated the disease-associated body weight loss, stool inconsistency, and intestinal bleeding in the models of ulcerative colitis and Crohn disease in mice (Fig. 1, D–G and H–K). These effects were comparable with more prominent than the effects of the standard drug 5-aminosalicylic acid (100 mg kg⫺1) and those of prednisolone (20 mg

35830 JOURNAL OF BIOLOGICAL CHEMISTRY

kg⫺1). Thus, the anti-inflammatory action of SMTP is independent of the plasminogen modulation activity. Identification of sEH as a Target—To identify the molecule that is involved in the anti-inflammatory mechanism of SMTP, we designed an affinity matrix that contained an essential part of SMTP, tricyclic ␥-lactam with a geranylmethyl side chain (Fig. 2A). To prepare the SMTP congener with a primary amine VOLUME 289 • NUMBER 52 • DECEMBER 26, 2014

Anti-inflammatory Target of a Novel Thrombolytic Stroke Drug TABLE 1 Results of the peptide mass fingerprinting analysis of protein band 1 in Fig. 2B The propionamide modification of cysteine residue may be due to the reaction with acrylamide monomer remaining in polyacrylamide.

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

DECEMBER 26, 2014 • VOLUME 289 • NUMBER 52

JOURNAL OF BIOLOGICAL CHEMISTRY

35831

Anti-inflammatory Target of a Novel Thrombolytic Stroke Drug

at the N-linked side chain (SMTP-50), we first synthesized a precursor with the primary amine of SMTP-50 protected by a 9-fluorenylmethyloxycarbonyl group (SMTP-47) (Fig. 2A) by the precursor amine-fed culture method (31). After eliminating the protective group, the resulting SMTP-50 was coupled to cross-linked agarose beads. Detergent-solubilized homogenates of the mouse liver were subjected to affinity chromatography on the SMTP-coupled beads. The resulting eluate gave several protein bands on SDSpolyacrylamide gel electrophoresis. We subjected four specifically eluted protein species to peptide-mass fingerprinting analysis (Fig. 2B), and every protein species was identified as a full-length form or a fragment of sEH (Fig. 2, C and D, and Table 1). The purification of sEH to homogeneity was achieved when SMTP-affinity chromatography was performed using a cytosol fraction of detergent-free homogenates and an elution buffer containing AUDA, a competitive inhibitor of the Cterm-EH of sEH (41) (Fig. 2E). Inhibition of sEH—SMTP-7 inhibited both the Cterm-EH and the Nterm-phos of sEH (their respective IC50 values were 23 ⫾ 1 and 6 ⫾ 1 ␮M) when we used the synthetic substrates PHOME and AttoPhos, respectively (Fig. 3, A and B). Similarly, SMTP-44D inhibited Cterm-EH and Nterm-phos (IC50 27 ⫾ 2 and 24 ⫾ 3 ␮M, respectively) (Fig. 3, A and B). In addition, the structurally simplest congener SMTP-0 (which lacks the N-linked side chain) was inhibitory to both activities (IC50 6 ⫾ 1 and 14 ⫾ 1 ␮M, respectively) (Fig. 3, A and B). Thus, the structural requirement for sEH inhibition is clearly distinguishable from that for plasminogen modulation activity in which the N-linked side chain plays a crucial role (11, 12, 32). We performed detailed kinetic analyses of sEH inhibition using SMTP-0 and the natural substrate 14,15-EET. The use of SMTP-0 was to avoid complexity of data analysis: SMTP-0 consists of a single core unit of SMTP (Fig. 3), whereas SMTP-7 has two core units that are asymmetrically configured (Fig. 1A), and each of these may differently interact with the enzyme. The kinetic results revealed a positive cooperativity for the hydrol-

35832 JOURNAL OF BIOLOGICAL CHEMISTRY

ysis of 14,15-EET (Fig. 4A) (Hill coefficient of 1.9 for the substrate binding; Fig. 4B), suggesting an allosteric interaction between the two monomers of sEH. The data were, therefore, analyzed based on a nonlinear mathematical model that allowed allostericity between the two equivalent catalytic sites of Cterm-EH (Fig. 4C). The pattern of the Cterm-EH inhibition by SMTP-0 fitted well to a competitive model (Fig. 4, A and C). Moreover, Cterm-EH inhibition by SMTP-0 was competed for by AUDA (this class of inhibitor binds to the catalytic site in Cterm-EH (42)) (Fig. 4, D and E). These results are consistent with the observation that sEH is specifically eluted with AUDA in SMTP affinity chromatography. Regarding the Nterm-phos, the kinetic data fitted to a linear mathematical model (Fig. 5, A–C), suggesting no cooperativity between the two Ntermphos domains (Hill coefficient of 0.91; Fig. 5B). The inhibition of Nterm-phos by SMTP-0 was pseudo-noncompetitive with respect to the substrate AttoPhos (Fig. 5, A and B). The pseudononcompetitive mechanism suggests that the Nterm-phos inhibition is mediated by the SMTP-0 binding to an allosteric site other than the substrate binding site in the Nterm-phos. Because there remained a possibility that the binding of SMTP-0 to the substrate binding site in the Cterm-EH might affect the activity of the Nterm-phos, we assessed the inhibitory activity of SMTP-0 toward the Nterm-phos in the presence of AUDA, which competed with SMTP-0 for binding to the catalytic site in the Cterm-EH (Fig. 4D). As a result, the presence of AUDA did not affect the SMTP-0 inhibition of the Nterm-phos (Fig. 5, D and E). Thus, SMTP-0 should bind to two distinct sites in sEH; one is the catalytic site in the Cterm-EH, and the other is an allosteric site that affects the Nterm-phos. The structure-activity relationships of SMTP congeners differing in the N-linked side chain are summarized in Fig. 6. Although the minimum structural requirement for the sEH inhibition was the tricyclic ␥-lactam with a geranylmethyl side chain (represented by SMTP-0), the difference in the N-linked side chain variably affected the potency of the inhibition of the Cterm-EH and Nterm-phos. The congener with a naphthalene VOLUME 289 • NUMBER 52 • DECEMBER 26, 2014

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

FIGURE 3. Inhibition of sEH by SMTPs. Cterm-EH (A) and Nterm-phos (B) were determined in the presence of the indicated concentrations of SMTP-0, SMTP-7, and SMTP-44D using the following concentrations of enzyme and substrate: 4.7 nM sEH and 12.5 ␮M PHOME for Cterm-EH and 2.3 nM sEH and 5 ␮M AttoPhos for Nterm-phos. Each value represents the mean ⫾ S.D. from triplicate determinations. The structure of SMTP-0 is shown.

Anti-inflammatory Target of a Novel Thrombolytic Stroke Drug

(SMTP-16) (32) or a glucose moiety as the N-linked side chain (SMTP-33) (32) was essentially inactive in inhibiting the Cterm-EH and Nterm-phos. SMTP-54 (with a glutamine moiety) (33) and SMTP-55 (with a glutamic acid moiety) (33) were relatively specific for the inhibition of the Cterm-EH. On the other hand, SMTP-5D (with a D-leucine moiety) (34) was relatively specific for the inhibition of the Nterm-phos. The inhibitory activity of the congeners with a phenylamine derivative as the N-linked side chain (such as SMTP-26, SMTP-27, and SMTP-28) (12) were potent with respect to both the Cterm-EH and Nterm-phos inhibitions. The variability of the inhibition selectivity (IC50 for the Cterm-EH versus that for the Ntermphos) supports the idea that SMTPs bind to two distinct sites in sEH. DECEMBER 26, 2014 • VOLUME 289 • NUMBER 52

Inhibition of EET Metabolism—HepG2 cells had an ability to hydrolyze 14,15-EET to 14,15-DHET, an inactive diol. SMTP-7 inhibited the formation of 14,15-DHET from 14,15-EET added to the culture medium (IC50 6.5 ␮M) (Fig. 7A). Along with the inhibition of the 14,15-EET hydrolysis, the level of 14,15-EET was elevated in the presence of SMTP-7 (Fig. 7A). SMTP-0 and SMTP-44D were also active in inhibiting 14,15-DHET formation from 14,15-EET in HepG2 cells (IC50 1.2 and 9.2 ␮M, respectively) (Fig. 7B). Effects on the Level of Arachidonate-derived Lipid Mediators in Plasma from Guillain-Barre´ Syndrome Model Rats—To confirm the impact of sEH inhibition by SMTP-7 and its selectivity on sEH, we measured the levels of metabolites in the cyclooxygenase, lipoxygenase, and cytochrome P450 pathways using JOURNAL OF BIOLOGICAL CHEMISTRY

35833

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

FIGURE 4. Kinetic analysis of Cterm-EH inhibition by SMTP-0. A, kinetic analysis of the Cterm-EH inhibition by SMTP-0 determined using 14,15-EET. Dashed lines represent the theoretical lines derived from the mathematical model in panel (C). B, Hill plots of the control data in panel A. C, a diagram of the equilibrium in which allostericity for the substrate binding between the two equivalent catalytic sites of the Cterm-EH is allowed (as sigmoidal velocity curves with a Hill coefficient of 1.9 that suggested an allosteric kinetic model for the substrate binding). As sEH consists of two identical monomers (E1 and E2), we hypothesized that the equilibrium substrate dissociation constants (KS1, the primary substrate dissociation constant, and KS2, the secondary substrate dissociation constant) for each subunit were identical (this means that E1 and E2 are indistinguishable from each other in this model). Allostericity for the inhibitor binding to each monomer is not considered in this model. In addition, no change in the dissociation constant for the substrate binding to the remaining vacant site is taken into account. S, substrate (14,15-EET); I, inhibitor (SMTP-0); P, product (14,15-DHET); KI, inhibitor dissociation constant. The equation is derived from the model. v, reaction velocity; Vmax, maximum reaction velocity. To determine each constant, we tried various combinations of kinetic parameters to fit the velocity results to the equation. Among the parameters tested, the most probable values obtained were: Vmax 11.5 ␮mol min⫺1 mg⫺1; KS1 59 ␮M; KS2 0.21 ␮M; KI 21.2 ␮M; r2 0.99. D, competition of SMTP-0 with AUDA in the Cterm-EH inhibition. The inhibition of Cterm-EH by SMTP-0 was determined with 4.7 nM sEH and 12.5 ␮M PHOME in the presence of the indicated concentrations of AUDA. Inset shows the secondary plots. Dashed lines represent the theoretical lines derived from the mathematical model in panel E fitted to the experimental data. E, a diagram of the equilibrium in which two inhibitors (SMTP-0, I1, and AUDA, I2) competes with each other for binding to the catalytic site in an enzyme (Cterm-EH, E). S, substrate (PHOME); P, product; KI1, inhibitor dissociation constant for I1; KI2, inhibitor dissociation constant for I2. The equation is derived from the model. v, reaction velocity; Vmax, maximum reaction velocity. To determine each constant, we fitted the velocity results obtained in the presence of 5 ␮M of S (PHOME) with various concentrations of I1 and I2 to the equation using the value of Vmax and KS obtained in the absence of the inhibitor (Vmax 1.8 ⫻ 104 ⌬fluorescence min⫺1, KS 59 ␮M). The values obtained were: KI1, 6.8 ␮M; KI2, 6.3 nM; r2, 0.97. Each value represents the mean ⫾ S.D. from triplicate determinations.

Anti-inflammatory Target of a Novel Thrombolytic Stroke Drug

plasma obtained from Guillain-Barre´ syndrome model rats. A global analysis revealed that SMTP-7 did not significantly affect the levels of 47 out of 48 metabolites examined (see “Experimental Procedures” for metabolites analyzed), resulting in a small change in the distribution of the three classes of arachidonate metabolites (Fig. 8A). The only one metabolite that was significantly affected by the treatment with SMTP-7 was 11,12DHET. The level (% distribution among the 48 metabolites) of 11,12-DHET in SMTP-7-treated rats was significantly lower than that in saline-treated rats (13.5 ⫾ 0.17% compared with 20.2 ⫾ 0.04% in saline group, p ⬍ 0.05; Fig. 8B). The levels of 5,6-, 8,9-, and 14,15-DHETs, however, were not significantly changed by the treatment with SMTP-7 (Fig. 8B). The levels of all regioisomers of EETs were too low to be detected by the method employed. The result that SMTP-7 selectively decreased 11,12-DHET can partly be explained by inhibition of sEH. Details of this interpretation are described under “Discussion.”

35834 JOURNAL OF BIOLOGICAL CHEMISTRY

DISCUSSION SMTP-7 has a plasminogen modulation activity that leads to a thrombolytic enhancement as observed in several animal models (14, 16, 20). SMTP-7 effectively treats thrombotic and embolic stroke models (17, 18, 20), and the involvement of an additional mechanism that leads to anti-inflammation has been suggested (16, 18, 19, 21). In the present study we observed the SMTP anti-inflammatory action that is independent of the plasminogen modulation activity. Although the in vivo models used in this study (Guillain-Barre´ syndrome, ulcerative colitis, and Crohn disease) are apparently irrelevant to thromboembolic complication, there remained a possibility that a local generation of plasmin might affect disease progress. Therefore, we compared the effect of SMTP-7 with that of the congener SMTP-44D, which is essentially inactive in plasminogen modulation activity (11). The results clearly demonstrate that both compounds are effective in treating these inflammatory disease VOLUME 289 • NUMBER 52 • DECEMBER 26, 2014

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

FIGURE 5. Kinetic analysis of Nterm-phos inhibition by SMTP-0. A, kinetic analysis of the Nterm-phos inhibition by SMTP-0 determined using AttoPhos. Dashed lines represent the theoretical lines derived from the mathematical model in panel C. B, Hill plots of the control data in panel A. C, a diagram of the equilibrium in which no allostericity for the substrate binding between the two equivalent catalytic sites of the Nterm-phos is accounted (this was because hyperbolic velocity curves with a Hill coefficient of 0.91 was obtained). Also, allostericity for the inhibitor binding to each monomer is not considered. E, Ntem-phos; S, substrate (AttoPhos); I, inhibitor (SMTP-0); P, product; Ks, substrate dissociation constant; KI-S, dissociation constant between EI complex and S; KI, inhibitor dissociation constant. The equation is derived from the model. v, reaction velocity; Vmax, maximum reaction velocity. To determine each constant, the velocity results obtained with various concentrations of I and S were fitted to the equation. The values obtained were: Vmax 230 ⌬fluorescence min⫺1; KS 2.1 ␮M; KI-S, 2.7 ␮M; Ki 12 ␮M; r2 0.98. D, absence of competition of SMTP-0 with AUDA in the Nterm-phos inhibition. The inhibition of Nterm-phos by SMTP-0 was determined with 2.3 nM sEH and 5 ␮M AttoPhos in the presence of the indicated concentrations of AUDA. The green dashed line represents the theoretical line that considers the competition of SMTP-0 with AUDA derived from a model in which the inhibition of Nterm-phos by SMTP-0 is competitively affected by AUDA. Experimental data demonstrated that AUDA did not affect the activity of the Nterm-phos. Each value represents the mean ⫾ S.D. from triplicate determinations.

Anti-inflammatory Target of a Novel Thrombolytic Stroke Drug

FIGURE 7. Inhibition of 14,15-EET hydrolysis in HepG2 cells in culture. A and B, inhibition of 14,15-EET hydrolysis in HepG2 cells by SMTP-7 (A), SMTP-0 and SMTP-44D (B). Each value represents the mean ⫾ S.D. from triplicate determinations.

models (Fig. 1). Thus we conclude that SMTPs have an antiinflammatory activity independent of their plasminogen modulation activity. A target molecule for the anti-inflammatory action should, therefore, exist. In addition, the structural requirement for the anti-inflammatory action could be different from that for the plasminogen modulation activity. To identify the anti-inflammatory target, we designed an affinity matrix that can bind a target protein. In our preliminary experiments, we observed anti-inflammatory effects with SMTP congeners with varying N-linked side-chain structures. We, therefore, employed the strategy of coupling the core DECEMBER 26, 2014 • VOLUME 289 • NUMBER 52

SMTP structural moiety (tricyclic ␥-lactam with a geranylmethyl side-chain) to a bead via an N-linked side chain. The major protein that bound to the resulting affinity matrix was sEH (Fig. 2). We purified a homogeneous sEH preparation by performing a single-step chromatography on this matrix using AUDA, which binds to the catalytic site of the Cterm-EH, as an eluent. This result is consistent with the fact that SMTP-0 competitively inhibits the Cterm-EH (Fig. 4). In addition to its inhibition of Cterm-EH, SMTP-0 inhibits the Nterm-phos pseudononcompetitively (Fig. 5). The result that the inhibition of the Cterm-EH, but not the Nterm-phos, by SMTP-0 is competed JOURNAL OF BIOLOGICAL CHEMISTRY

35835

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

FIGURE 6. Structure-activity relationships of SMTP congeners in the inhibition of Cterm-EH and Nterm-phos. The IC50 values of the Cterm-EH and Nterm-phos inhibitions, determined with PHOME and AttoPhos, respectively, are plotted for each congener, shown as the SMTP number. The following concentrations of enzyme and substrate were used: 4.7 nM sEH and 12.5 ␮M PHOME for Cterm-EH and 2.3 nM sEH and 5 ␮M AttoPhos for Nterm-phos. The inhibition selectivity (the ratio of the IC50 of the Cterm-EH over that of the Nterm-phos) of each SMTP congener is also shown as a bar graph. The N-linked side-chain structure for each congener is shown. IC50 values (␮M) of each SMTP congener for Cterm-EH and Nterm-phos, respectively, were the following: SMTP-0, 6 ⫾ 1 and 14 ⫾ 1; SMTP-5D, 70 ⫾ 7 and 8 ⫾ 0; SMTP-7, 23 ⫾ 1 and 6 ⫾ 1; SMTP-16, ⬎ 100 and ⬎ 100; SMTP-26, 12 ⫾ 1 and 12 ⫾ 2; SMTP-27, 8 ⫾ 1 and 14 ⫾ 1; SMTP-28, 5 ⫾ 2 and 9 ⫾ 1; SMTP-33, ⬎100 and ⬎100; SMTP-44D, 27 ⫾ 2 and 24 ⫾ 3; SMTP-54, 17 ⫾ 1 and 67 ⫾ 5; SMTP-55, 19 ⫾ 1 and 60 ⫾ 6. Each value represents the mean ⫾ 95% confidential interval from triplicate determinations.

Anti-inflammatory Target of a Novel Thrombolytic Stroke Drug

FIGURE 8. Global analysis of arachidonate-derived lipid mediators in plasma derived from Guillain-Barre´ syndrome model rats. A, distribution of the arachidonate-derived lipid mediators as categorized by the cyclooxygenase (COX), lipoxygenase (LOX), and cytochrome P450 (CYP) pathways. The total of the levels of 48 metabolites tested (see “Experimental Procedures”) in each group (saline-treated or SMTP-7 treated) was taken as 100%. B, effects of SMTP-7 treatment on the levels of 5,6-, 8,9-, 11,12, and 14,15-DHETs. *, p ⬍ 0.05 by unpaired Student’s t test. Each value represents the mean ⫾ S.D. obtained from four determinations.

35836 JOURNAL OF BIOLOGICAL CHEMISTRY

metabolites can be explained by the low catalytic activity of sEH toward these regioisomers (43, 60), letting sEH to contribute lesser to these levels. We speculate the lack of effect on the 14,15-DHET level is due to complex mechanisms of the formation/catabolism (degradation and excretion) of EETs and DHETs under physiological conditions. According to previous literatures (43, 60, 61) the specific activity of human sEH for the hydrolysis of 14,15-EET is approximately ⬎2 times higher than that for 11,12-EET. In our animal model, the circulating level of 14,15-DHET was only 1.1 times higher than that of 11,12DHET in control animals. Thus, it is likely that the rates of degradation and/or excretion of 14,15-DHET in this model is higher than those of 11,12-DHET. This may make it difficult to reflect the impact of sEH inhibition by SMTP-7. The selective change in the 11,12-DHET level by SMTP-7 suggest a specificity of SMTP-7 in the arachidonate metabolisms. Along with the in vitro data, this in vivo result supports the idea that sEH is an anti-inflammatory target of SMTP-7. The deficiency of sEH and the inhibition of the Cterm-EH have been reported to be protective against disease progression in animal models of inflammatory bowel diseases (62). In thrombotic and embolic stroke models in rodents and primates, SMTP-7 exhibited excellent activities, with a wide therapeutic time window and a reduced cerebral hemorrhage (17, 18, 20), that were not achieved by the conventional thrombolytic therapy. The finding that SMTP-7 inhibits sEH suggests that this activity, aside from the thrombolytic enhancement by plasminogen modulation, may account for these additional pharmacological potentials of SMTP-7. The observations that the sEH deficiency or the Cterm-EH inhibition is protective against experimental stroke (63– 65) support this hypothesis. The pharmacological significance of the SMTP-7 ability to inhibit the Nterm-phos remains to be investigated. SMTP congeners that selectivity inhibit the Nterm-phos can be useful tools to investigate the physiological role of the Nterm-phos. In conclusion, our results demonstrated SMTP-7’s anti-inflammatory action that is independent of its plasminogen modulation activity. The finding that SMTP-7 inhibits sEH and the resulting hydrolysis of the anti-inflammatory signaling molecule EET suggests that sEH inhibition is involved in the antiinflammatory action of SMTP-7. The combination of thromVOLUME 289 • NUMBER 52 • DECEMBER 26, 2014

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

for by AUDA suggests a mechanism in which SMTP-0 binds to two distinct sites in sEH; one is the substrate binding site in the Cterm-EH and the other is an allosteric site that affects the Nterm-phos. We observed a positive cooperativity in CtermEH, whereas no cooperativity (43) or a negative cooperativity in Cterm-EH (44) has been reported in previous studies. These variable findings may be due to the use of enzyme from different sources (native or recombinant) and/or buffer compositions. As expected from the crystal structure, sEH has higher order intradomain and interdomain interactions (45), and environmental conditions would affect the conformational status of sEH to exhibit cooperativity. Nevertheless, kinetic parameters obtained with these three investigations are relatively close (Vmax 11.5 ␮mol min⫺1 mg⫺1 and apparent Km ((KS1 ⫻ KS2)1/2] 3.52 ␮M in this study (Fig. 4); Vmax 9.0 ␮mol min⫺1 mg⫺1 and Km 4 ␮M for 14(R),15(R)-EET, and Vmax 1.36 ␮mol min⫺1 mg⫺1 and Km 5 ␮M for 14(S),15(S)-EET by Zeldin et al. (43); Vmax 20 ␮mol min⫺1 mg⫺1 and apparent Km ((KS1 ⫻ KS2)1/2) 6.3 ␮M by Marowsky et al. (44)). The structure-activity relationship results, which reveal variable inhibition selectivity (the ratio of the IC50 of the Cterm-EH over that of the Nterm-phos) among congeners with different N-linked side chains (Fig. 6), support this idea. Although the physiological function of the Cterm-EH has been extensively characterized by the use of the combination of sEH-deficient animals and specific inhibitors of the Cterm-EH (46–56), the role of the Nterm-phos remains elusive because of the lack of information about the relevant substrate and a potent specific inhibitor. The existence of an allosteric site that affects the Nterm-phos has not been reported. Our structure-activity relationship data suggest the possibility that a selective Nterm-phos inhibitor can be developed based on the allosteric mechanism. EETs are signaling molecules implicated in anti-inflammation (57, 58). SMTP-7 and its congeners inhibited the hydrolysis of 14,15-EET to the inactive 14,15-DHET in HepG2 cells (Fig. 7). In addition, the plasma level of 11,12-DHET in GuillainBarre´ syndrome model rats was significantly decreased by SMTP-7 treatment (Fig. 8), whereas the levels of 5,6-, 8,9-, and 14,15-DHETs were unaffected. The levels of 5,6- and 8,9DHETs were low, which is consistent with the finding by Li et al. (59). The lack of effect of SMTP-7 on the levels of these

Anti-inflammatory Target of a Novel Thrombolytic Stroke Drug bolysis and sEH inhibition explains the excellent activity of SMTP-7 in treating thrombotic and embolic strokes. SMTP-7 is under development as an alternative drug that could be effective in stroke patients who do not benefit from the standard thrombolytic therapy.

16.

17.

Acknowledgments—We thank Haruki Koide (Tokyo Noko University) for synthesizing SMTP-47 and Naoko Nishimura and Keiko Hasegawa (TMS Co., Ltd.) for SMTP-7 and congeners.

18.

REFERENCES

DECEMBER 26, 2014 • VOLUME 289 • NUMBER 52

19.

20.

21.

22.

23.

24.

25. 26.

27.

28.

29.

30.

31.

32.

33.

34.

JOURNAL OF BIOLOGICAL CHEMISTRY

35837

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

1. Donnan, G. A., Fisher, M., Macleod, M., and Davis, S. M. (2008) Stroke. Lancet 371, 1612–1623 2. Donnan, G. A., Davis, S. M., Parsons, M. W., Ma, H., Dewey, H. M., and Howells, D. W. (2011) How to make better use of thrombolytic therapy in acute ischemic stroke. Nat. Rev. Neurol. 7, 400 – 409 3. Lees, K. R., Bluhmki, E., von Kummer, R., Brott, T. G., Toni, D., Grotta, J. C., Albers, G. W., Kaste, M., Marler, J. R., Hamilton, S. A., Tilley, B. C., Davis, S. M., Donnan, G. A., Hacke, W., ECASS, ATLANTIS, NINDS and EPITHET rt-PA Study Group, Allen, K., Mau, J., Meier, D., del Zoppo, G., De Silva, D. A., Butcher, K. S., Parsons, M. W., Barber, P. A., Levi, C., Bladin, C., and Byrnes, G. (2010) Time to treatment with intravenous alteplase and outcome in stroke: an updated pooled analysis of ECASS, ATLANTIS, NINDS, and EPITHET trials. Lancet 375, 1695–1703 4. IST-3 collaborative group, Sandercock, P., Wardlaw, J. M., Lindley, R. I., Dennis, M., Cohen, G., Murray, G., Innes, K., Venables, G., Czlonkowska, A., Kobayashi, A., Ricci, S., Murray, V., Berge, E., Slot, K. B., Hankey, G. J., Correia, M., Peeters, A., Matz, K., Lyrer, P., Gubitz, G., Phillips, S. J., and Arauz, A. (2012) The benefits and harms of intravenous thrombolysis with recombinant tissue plasminogen activator within 6 h of acute ischaemic stroke (the third international stroke trial [IST-3]): a randomised controlled trial. Lancet 379, 2352–2363 5. Wardlaw, J. M., Murray, V., Berge, E., del Zoppo, G., Sandercock, P., Lindley, R. L., and Cohen, G. (2012) Recombinant tissue plasminogen activator for acute ischaemic stroke: an updated systematic review and meta-analysis. Lancet 379, 2364 –2372 6. Kleindorfer, D., Lindsell, C. J., Brass, L., Koroshetz, W., and Broderick, J. P. (2008) National US estimates of recombinant tissue plasminogen activator use: ICD-9 codes substantially underestimate. Stroke 39, 924 –928 7. Nadeau, J. O., Shi, S., Fang, J., Kapral, M. K., Richards, J. A., Silver, F. L., Hill, M. D., and Investigators for the Registry of the Canadian Stroke Network. (2005) TPA use for stroke in the Registry of the Canadian Stroke Network. Can. J. Neurol. Sci. 32, 433– 439 8. Kleinig, T. J., and Vink, R. (2009) Suppression of inflammation in ischemic and hemorrhagic stroke: therapeutic options. Curr. Opin. Neurol. 22, 294 –301 9. Yepes, M., Roussel, B. D., Ali, C., and Vivien, D. (2009) Tissue-type plasminogen activator in the ischemic brain: more than a thrombolytic. Trends Neurosci. 32, 48 –55 10. Hasumi, K., Yamamichi, S., and Harada, T. (2010) Small-molecule modulators of the zymogen activation in the fibrinolytic and coagulation systems. FEBS J. 277, 3675–3687 11. Koide, H., Narasaki, R., Hasegawa, K., Nishimura, N., and Hasumi, K. (2012) A new series of the SMTP plasminogen modulator with a phenylglycine-based side-chain. J. Antibiot. 65, 91–93 12. Koide, H., Hasegawa, K., Nishimura, N., Narasaki, R., and Hasumi, K. (2012) A new series of the SMTP plasminogen modulators with a phenylamine-based side-chain. J. Antibiot. 65, 361–367 13. Hu, W., Ohyama, S., and Hasumi, K. (2000) Activation of fibrinolysis by SMTP-7 and -8, novel staplabin analogs with a pseudosymmetric structure. J. Antibiot. 53, 241–247 14. Hu, W., Narasaki, R., Nishimura, N., and Hasumi, K. (2012) SMTP (Stachybotrys microspora triprenyl phenol) enhances clot clearance in a pulmonary embolism model in rats. Thromb. J. 10, 2 15. Koyanagi, K., Narasaki, R., Yamamichi, S., Suzuki, E., and Hasumi, K.

(2014) Mechanism of the action of SMTP-7, a novel small-molecule modulator of plasminogen activation. Blood Coagul. Fibrinolysis 25, 316 –321 Miyazaki, T., Kimura, Y., Ohata, H., Hashimoto, T., Shibata, K., Hasumi, K., and Honda, K. (2011) Distinct effects of tissue-type plasminogen activator and SMTP-7 on cerebrovascular inflammation following thrombolytic reperfusion. Stroke 42, 1097–1104 Hashimoto, T., Shibata, K., Nobe, K., Hasumi, K., and Honda, K. (2010) A novel embolic model of cerebral infarction and evaluation of Stachybotrys microspora triprenyl phenol-7 (SMTP-7), a novel fungal triprenyl phenol metabolite. J. Pharmacol. Sci. 114, 41– 49 Shibata, K., Hashimoto, T., Nobe, K., Hasumi, K., and Honda, K. (2010) A novel finding of a low-molecular-weight compound, SMTP-7, having thrombolytic and anti-inflammatory effects in cerebral infarction of mice. Naunyn Schmiedebergs Arch. Pharmacol. 382, 245–253 Akamatsu, Y., Saito, A., Fujimura, M., Shimizu, H., Mekawy, M., Hasumi, K., and Tominaga, T. (2011) Stachybotrys microspora triprenyl phenol-7, a novel fibrinolytic agent, suppresses superoxide production, matrix metalloproteinase-9 expression, and thereby attenuates ischemia/reperfusion injury in rat brain. Neurosci. Lett. 503, 110 –114 Sawada, H., Nishimura, N., Suzuki, E., Zhuang, J., Hasegawa, K., Takamatsu, H., Honda, K., and Hasumi, K. (2014) SMTP-7, a novel smallmolecule thrombolytic for ischemic stroke: a study in rodents and primates. J. Cereb. Blood Flow Metab. 34, 235–241 Shibata, K., Hashimoto, T., Nobe, K., Hasumi, K., and Honda, K. (2011) Neuroprotective mechanisms of SMTP-7 in cerebral infarction model in mice. Naunyn Schmiedebergs Arch. Pharmacol. 384, 103–108 Newman, J. W., Morisseau, C., and Hammock, B. D. (2005) Epoxide hydrolases: their roles and interactions with lipid metabolism. Prog. Lipid Res. 44, 1–51 Gomez, G. A., Morisseau, C., Hammock, B. D., and Christianson, D. W. (2004) Structure of human epoxide hydrolase reveals mechanistic inferences on bifunctional catalysis in epoxide and phosphate ester hydrolysis. Biochemistry 43, 4716 – 4723 Michaelis, U. R., and Fleming, I. (2006) From endothelium-derived hyperpolarizing factor (EDHF) to angiogenesis: epoxyeicosatrienoic acids (EETs) and cell signaling. Pharmacol. Ther. 111, 584 –595 Spector, A. A., and Norris, A. W. (2007) Action of epoxyeicosatrienoic acids on cellular function. Am. J. Physiol. Cell Physiol. 292, C996 –1012 Iliff, J.J., Jia, J., Nelson, J., Goyagi, T., Klaus, J., and Alkayed, N. J. (2010) Epoxyeicosanoid signaling in CNS function and disease. Prostaglandins Other Lipid Mediat. 91, 68 – 84 Inceoglu, B., Jinks, S. L., Ulu, A., Hegedus, C. M., Georgi, K., Schmelzer, K. R., Wagner, K., Jones, P. D., Morisseau, C., and Hammock, B. D. (2008) Soluble epoxide hydrolase and epoxyeicosatrienoic acids modulate two distinct analgesic pathways. Proc. Natl. Acad. Sci. U.S.A. 105, 18901–18906 Tran, K. L., Aronov, P. A., Tanaka, H., Newman, J. W., Hammock, B. D., and Morisseau, C. (2005) Lipid sulfates and sulfonates are allosteric competitive inhibitors of the N-terminal phosphatase activity of the mammalian soluble epoxide hydrolase. Biochemistry 44, 12179 –12187 Morisseau, C., Schebb, N. H., Dong, H., Ulu, A., Aronov, P. A., and Hammock, B. D. (2012) Role of soluble epoxide hydrolase phosphatase activity in the metabolism of lysophosphatidic acids. Biochem. Biophys. Res. Commun. 419, 796 – 800 Oguro, A., and Imaoka, S. (2012) Lysophosphatidic acids are new substrates for the phosphatase domain of soluble epoxide hydrolase. J. Lipid Res. 53, 505–512 Hasumi, K., Hasegawa, K., and Kitano, Y. (2007) Isolation and absolute configuration of SMTP-0, a simplest congener of the SMTP family nonlysine-analog plasminogen modulators. J. Antibiot. 60, 463– 468 Hasegawa, K., Koide, H., Hu, W., Nishimura, N., Narasaki, R., Kitano, Y., and Hasumi, K. (2010) Structure-activity relationships of 11 new congeners of the SMTP plasminogen modulator. J. Antibiot. 63, 589 –593 Nishimura, Y., Suzuki, E., Hasegawa, K., Nishimura, N., Kitano, Y., and Hasumi, K. (2012) Pre-SMTP, a key precursor for the biosynthesis of the SMTP plasminogen modulators. J. Antibiot. 65, 483– 485 Hu, W., Kitano, Y., and Hasumi, K. (2003) SMTP-4D, -5D, -6D, -7D and -8D, a new series of the non-lysine-analog plasminogen modulators with a

Anti-inflammatory Target of a Novel Thrombolytic Stroke Drug

35838 JOURNAL OF BIOLOGICAL CHEMISTRY

ONE 7, e39165 51. Luo, P., Chang, H. H., Zhou, Y., Zhang, S., Hwang, S. H., Morisseau, C., Wang, C. Y., Inscho, E. W., Hammock, B. D., and Wang, M. H. (2010) Inhibition or deletion of soluble epoxide hydrolase prevents hyperglycemia, promotes insulin secretion, and reduces islet apoptosis. J. Pharmacol. Exp. Ther. 334, 430 – 438 52. Qiu, H., Li, N., Liu, J. Y., Harris, T. R., Hammock, B. D., and Chiamvimonvat, N. (2011) Soluble epoxide hydrolase inhibitors and heart failure. Cardiovasc. Ther. 29, 99 –111 53. Ulu, A., Davis, B. B., Tsai, H. J., Kim, I. H., Morisseau, C., Inceoglu, B., Fiehn, O., Hammock, B. D., and Weiss, R. H. (2008) Soluble epoxide hydrolase inhibitors reduce the development of atherosclerosis in apolipoprotein e-knockout mouse model. J. Cardiovasc. Pharmacol. 52, 314 –323 54. Motoki, A., Merkel, M. J., Packwood, W. H., Cao, Z., Liu, L., Iliff, J., Alkayed, N. J., and Van Winkle, D. M. (2008) Soluble epoxide hydrolase inhibition and gene deletion are protective against myocardial ischemiareperfusion injury in vivo. Am. J. Physiol. Heart. Circ. Physiol. 295, H2128 –H2134 55. Morisseau, C., and Hammock, B. D. (2013) Impact of soluble epoxide hydrolase and epoxyeicosanoids on human health. Annu. Rev. Pharmacol. Toxicol. 53, 37–58 56. Sanders, W. G., Morisseau, C., Hammock, B. D., Cheung, A. K., and Terry, C. M. (2012) Soluble epoxide hydrolase expression in a porcine model of arteriovenous graft stenosis and anti-inflammatory effects of a soluble epoxide hydrolase inhibitor. Am. J. Physiol. Cell Physiol. 303, C278 –C290 57. Ulu, A., Harris, T. R., Morisseau, C., Miyabe, C., Inoue, H., Schuster, G., Dong, H., Iosif, A. M., Liu, J. Y., Weiss, R. H., Chiamvimonvat, N., Imig, J. D., and Hammock, B. D. (2013) Anti-inflammatory effects of ␻-3 polyunsaturated fatty acids and soluble epoxide hydrolase inhibitors in angiotensin-II-dependent hypertension. J. Cardiovasc. Pharmacol. 62, 285–297 58. Thomson, S. J., Askari, A., and Bishop-Bailey, D. (2012) Anti-inflammatory effects of epoxyeicosatrienoic acids. Int. J. Vasc. Med. 2012, 605101 59. Li, L., Li, N., Pang, W., Zhang, X., Hammock, B. D., Ai, D., and Zhu, Y. (2014) Opposite effects of gene deficiency and pharmacological inhibition of soluble epoxide hydrolase on cardiac fibrosis. Plos ONE 9, e94092 60. Zeldin, D. C., Moomaw, C. R., Jesse, N., Tomer, K. B., Beetham, J., Hammock, B. D., and Wu, S. (1996) Biochemical characterization of the human liver cytochrome P450 arachidonic acid epoxygenase pathway. Arch. Biochem. Biophys. 330, 87–96 61. Jiang, H., Zhu, A. G., Mamczur, M., Morisseau, C., Hammock, B. D., Falck, J. R., and McGiff, J. C. (2008) Hydrolysis of cis- and trans-epoxyeicosatrienoic acids by rat red blood cells. J. Pharmacol. Exp. Ther. 326, 330 –337 62. Zhang, W., Yang, A. L., Liao, J., Li, H., Dong, H., Chung, Y. T., Bai, H., Matkowskyj, K. A., Hammock, B. D., and Yang, G. Y. (2012) Soluble epoxide hydrolase gene deficiency or inhibition attenuates chronic active inflammatory bowel disease in IL-10⫺/⫺ mice. Dig. Dis. Sci. 57, 2580 –2591 63. Zhang, W., Koerner, I. P., Noppens, R., Grafe, M., Tsai, H. J., Morisseau, C., Luria, A., Hammock, B. D., Falck, J. R., and Alkayed, N. J. (2007) Soluble epoxide hydrolase: a novel therapeutic target in stroke. J. Cereb. Blood Flow Metab. 27, 1931–1940 64. Zhang, W., Otsuka, T., Sugo, N., Ardeshiri, A., Alhadid, Y. K., Iliff, J. J., DeBarber, A. E., Koop, D. R., and Alkayed, N. J. (2008) Soluble epoxide hydrolase gene deletion is protective against experimental cerebral ischemia. Stroke 39, 2073–2078 65. Jouihan, S. A., Zuloaga, K. L., Zhang, W., Shangraw, R. E., Krasnow, S. M., Marks, D. L., and Alkayed, N. J. (2013) Role of soluble epoxide hydrolase in exacerbation of stroke by streptozotocin-induced type 1 diabetes mellitus. J. Cereb. Blood Flow Metab. 33, 1650 –1656

VOLUME 289 • NUMBER 52 • DECEMBER 26, 2014

Downloaded from http://www.jbc.org/ at University of Michigan on March 3, 2015

D-amino acid moiety. J. Antibiot. 56, 832– 837 35. Rostami, A., Gregorian, S. K., Brown, M. J., and Pleasure, D. E. (1990) Induction of severe experimental autoimmune neuritis with a synthetic peptide corresponding to the 53–78 amino acid sequence of the myelin P2 protein. J. Neuroimmunol. 30, 145–151 36. Sasaki, M., Bharwani, S., Jordan, P., Elrod, J. W., Grisham, M. B., Jackson, T. H., Lefer, D. J., and Alexander, J. S. (2003) Increased disease activity in eNOS-deficient mice in experimental colitis. Free Radic. Biol. Med. 35, 1679 –1687 37. Schroeder, K. W., Tremaine, W. J., and Ilstrup, D. M. (1987) Coated oral 5-aminosalicylic acid therapy for mildly to moderately active ulcerative colitis. A randomized study. N. Engl. J. Med. 317, 1625–1629 38. Dieleman, L. A., Pen˜a, A. S., Meuwissen, S. G., and van Rees, E. P. (1997) Role of animal models for the pathogenesis and treatment of inflammatory bowel disease. Scand. J. Gastroenterol. Suppl. 223, 99 –104 39. Cooper, H. S., Murthy, S. N., Shah, R. S., and Sedergran, D. J. (1993) Clinicopathologic study of dextran sulfate sodium experimental murine colitis. Lab. Invest. 69, 238 –249 40. te Velde, A. A., Verstege, M. I., and Hommes, D. W. (2006) Critical appraisal of the current practice in murine TNBS-induced colitis. Inflamm. Bowel Dis. 12, 995–999 41. Morisseau, C., Goodrow, M. H., Newman, J. W., Wheelock, C. E., Dowdy, D. L., and Hammock, B. D. (2002) Structural refinement of inhibitors of urea-based soluble epoxide hydrolases. Biochem. Pharmacol. 63, 1599 –1608 42. Gomez, G. A., Morisseau, C., Hammock, B. D., and Christianson, D. W. (2006) Human soluble epoxide hydrolase: structural basis of inhibition by 4-(3-cyclohexylureido)-carboxylic acids. Protein Sci. 15, 58 – 64 43. Zeldin, D. C., Wei, S., Falck, J. R., Hammock, B. D., Snapper, J. R., and Capdevila, J. H. (1995) Metabolism of epoxyeicosatrienoic acids by cytosolic epoxide hydrolase: substrate structural determinants of asymmetric catalysis. Arch. Biochem. Biophys. 316, 443– 451 44. Marowsky, A., Burgener, J., Falck, J. R., Fritschy, J. M., and Arand, M. (2009) Distribution of soluble and microsomal epoxide hydrolase in the mouse brain and its contribution to cerebral epoxyeicosatrienoic acid metabolism. Neuroscience 163, 646 – 661 45. Argiriadi, M. A., Morisseau, C., Hammock, B. D., and Christianson, D. W. (1999) Detoxification of environmental mutagens and carcinogens: structure, mechanism, and evolution of liver epoxide hydrolase. Proc. Natl. Acad. Sci. U.S.A. 96, 10637–10642 46. Strauss, K. I., Gruzdev, A., and Zeldin, D. C. (2013) Altered behavioral phenotypes in soluble epoxide hydrolase knockout mice: effects of traumatic brain injury. Prostaglandins Other Lipid Mediat. 104, 18 –24 47. Bettaieb, A., Nagata, N., AbouBechara, D., Chahed, S., Morisseau, C., Hammock, B. D., and Haj, F. G. (2013) Soluble epoxide hydrolase deficiency or inhibition attenuates diet-induced endoplasmic reticulum stress in liver and adipose tissue. J. Biol. Chem. 288, 14189 –14199 48. Jung, O., Jansen, F., Mieth, A., Barbosa-Sicard, E., Pliquett, R. U., Babelova, A., Morisseau, C., Hwang, S. H., Tsai, C., Hammock, B. D., Schaefer, L., Geisslinger, G., Amann, K., and Brandes, R. P. (2010) Inhibition of the soluble epoxide hydrolase promotes albuminuria in mice with progressive renal disease. PLoS ONE 5, e11979 49. Luria, A., Bettaieb, A., Xi, Y., Shieh, G. J., Liu, H. C., Inoue, H., Tsai, H. J., Imig, J. D., Haj, F. G., and Hammock, B. D. (2011) Soluble epoxide hydrolase deficiency alters pancreatic islet size and improves glucose homeostasis in a model of insulin resistance. Proc. Natl. Acad. Sci. U.S.A. 108, 9038 –9043 50. Liu, Y., Dang, H., Li, D., Pang, W., Hammock, B. D., and Zhu, Y. (2012) Inhibition of soluble epoxide hydrolase attenuates high-fat-diet-induced hepatic steatosis by reduced systemic inflammatory status in mice. PLoS

Soluble epoxide hydrolase as an anti-inflammatory target of the thrombolytic stroke drug SMTP-7.

Although ischemic stroke is a major cause of death and disability worldwide, only a small fraction of patients benefit from the current thrombolytic t...
4MB Sizes 1 Downloads 7 Views