• Non ci sono risultati.

Cinnamyl-3,4-Dihydroxy-{alpha}-Cyanocinnamate Is a Potent Inhibitor of 5-Lipoxygenase.

N/A
N/A
Protected

Academic year: 2021

Condividi "Cinnamyl-3,4-Dihydroxy-{alpha}-Cyanocinnamate Is a Potent Inhibitor of 5-Lipoxygenase."

Copied!
9
0
0

Testo completo

(1)

Cinnamyl-3,4-Dihydroxy-

␣-Cyanocinnamate Is a Potent

Inhibitor of 5-Lipoxygenase

Carlo Pergola, Bianca Jazzar, Antonietta Rossi, Ulrike Buehring, Susann Luderer,

Friederike Dehm, Hinnak Northoff, Lidia Sautebin, and Oliver Werz

Department of Pharmaceutical/Medicinal Chemistry, Institute of Pharmacy, University of Jena, Jena, Germany (C.P., F.D., O.W.); Department for Pharmaceutical Analytics, Pharmaceutical Institute, University of Tuebingen, Tuebingen, Germany (B.J., U.B., S.L.); Department of Experimental Pharmacology, University of Naples Federico II, Naples, Italy (A.R., L.S.); and Institute for Clinical and Experimental Transfusion Medicine, University Medical Center Tuebingen, Tuebingen, Germany (H.N.)

Received February 18, 2011; accepted March 28, 2011

ABSTRACT

Lipoxygenases (LOs) are iron-containing enzymes that catalyze the conversion of arachidonic acid into hydroperoxyeicosatet-raenoic acids (HPETEs) and other bioactive lipid mediators. In mammals, 5-LO, 15-LO, and 12-LO enzymes seem to have distinct roles in pathophysiological contexts, which have em-phasized the need for selective inhibitors. Cinnamyl-3,4-dihy-droxy-␣-cyanocinnamate (CDC) has been proposed as potent and selective inhibitor of platelet-type 12-LO (p12-LO). Here, we re-evaluated the selectivity profile of CDC on LOs, and we show that CDC is a potent and direct inhibitor of 5-LO. CDC reduced 5-LO activity in cell-free assays (purified human re-combinant enzyme or leukocyte homogenates), with IC50 val-ues in the low nanomolar range (9 –25 nM) and a selectivity index of approximately 35 and 15 over p12-LO and 15-LO1,

respectively. Likewise, CDC inhibited 5-LO product formation in intact human polymorphonuclear leukocytes and monocytes (IC50⫽ 0.45–0.8␮M). A lower potency was observed for 15-LO1, whereas p12-LO activity in platelets was hardly affected. In human whole blood, CDC efficiently reduced the formation of 5-LO products, and similar effects were observed for 12(S)-H(P)ETE and 15(S)-12(S)-H(P)ETE. Finally, CDC (3.5 and 7 mg/kg i.p.) was effective in vivo in the platelet-activating factor-induced shock in mice and reduced formation of the 5-LO product leukotriene B4in the rat carrageenan-induced pleurisy after a single oral dose of 10 mg/kg. Together, our data demonstrate that CDC is a potent inhibitor of 5-LO with efficacy in vivo and encourage further development of CDC as the lead compound.

Introduction

Lipoxygenases (LOs) are structurally related dioxygenases that catalyze the incorporation of molecular oxygen into poly-unsaturated fatty acids containing a (Z,Z)-1,4-pentadiene system (Brash, 1999). In mammals, arachidonic acid (AA) is predominantly used as substrate, and the formed hydro(pero)-xyeicosatetraenoic acids [H(P)ETEs] can then be reduced to the corresponding hydroxy compounds or be further metab-olized [e.g., to leukotrienes (LTs)]. Therefore, LOs represent important sources of lipid mediators, with critical roles in both physiological and pathological contexts (Yoshimoto and

Takahashi, 2002; Ku¨ hn and O’Donnell, 2006; Peters-Golden and Henderson, 2007). The LO reactions are region-specific and, according to the positional specificity on AA, 5-LO, 12-LO, and 15-LO have been recognized in humans.

5-LO is highly expressed in polymorphonuclear leukocytes (PMNL) and monocytes (Rådmark et al., 2007). Its products, 5(S)-H(P)ETE and LTs, are mediators of immune and inflam-matory responses, and inhibition of 5-LO is a strategy to intervene with allergic and inflammatory disorders (e.g., asthma) (Pergola and Werz, 2010) and also possibly with atherosclerosis (Ba¨ck, 2009). Human blood PMNL and mono-cytes may also contain 15-LO1 (Nadel et al., 1991; Kim et al., 1995), which is closely related to the murine leukocyte-type 12-LO (also known as 12/15-LO). These enzymes can metab-olize AA to 12(S)-H(P)ETE and to 15(S)-H(P)ETE and may directly modify the organization of membranes, and their products can interact with several inflammation-related This work was supported by the Deutsche Forschungsgemeinschaft [Grants

WE 2260/9-11, WE 2260/9-2]. C.P. received a Carl-Zeiss stipend.

Article, publication date, and citation information can be found at http://jpet.aspetjournals.org.

doi:10.1124/jpet.111.180794.

ABBREVIATIONS: LO, lipoxygenase; AA, arachidonic acid; CDC, cinnamyl-3,4-dihydroxy-␣-cyanocinnamate; DPPH, diphenylpicrylhydrazyl radical; H(P)ETE, hydro(pero)xyeicosatetraenoic acid; HPLC, high-performance liquid chromatography; LT, leukotriene; p12-LO, platelet-type 12-LO; PAF, platelet-activating factor; DMSO, dimethyl sulfoxide; PBMC, peripheral blood mononuclear cells; PBS, phosphate-buffered saline; PGC, phosphate-buffered saline plus 1 mg/ml glucose and 1 mM CaCl2; A23187, calcimycin; PMNL, polymorphonuclear leukocytes; PMSF, phenylmethylsulfonyl fluoride; ANOVA, analysis of variance; 13-HpODE, 13-hydroperoxy-9,11-octadecadienoic acid.

THEJOURNAL OFPHARMACOLOGY ANDEXPERIMENTALTHERAPEUTICS Vol. 338, No. 1

Copyright © 2011 by The American Society for Pharmacology and Experimental Therapeutics 180794/3693337

JPET 338:205–213, 2011 Printed in U.S.A.

(2)

pathways, with roles in atherogenesis, chemotaxis, and can-cer growth (Ku¨ hn and O’Donnell, 2006). The platelet-type 12-LO (p12-LO) instead metabolizes AA predominantly to 12(S)-H(P)ETE (Yoshimoto and Takahashi, 2002). Despite possible roles of p12-LO in platelet aggregation, atheroscle-rosis, angiogenesis, cancer metastasis, and cell migration, the exact contribution of p12-LO to human physiology and pathophysiology still remains elusive.

For this reason, research has extensively focused on the identification of selective p12-LO inhibitors as tools to iden-tify specific actions of p12-LO and lead compounds in drug development. However, only few p12-LO inhibitors with a selectivity index⬎10 over 5-LO and 15-LO have been iden-tified. Among the different strategies used, structural modi-fication of known 5-LO inhibitors to gain selectivity on p12-LO has been reported to be successful (Cho et al., 1991). Thus, a series of 3,4-dihydroxy-␣-cyanocinnamoyl esters de-rived from the redox-type 5-LO inhibitor caffeic acid have been claimed as potent and selective p12-LO inhibitors. In particular, the synthetic compound, cinnamyl-3,4-dihydroxy-␣-cyanocinnamate (CDC) (Fig. 1), was reported to inhibit p12-LO with an IC50of 63 nM and to be selective

approxi-mately 30 and 53 times over 5- and 12/15-LO, respectively. Therefore, this compound has been used as a pharmacologi-cal tool to identify 12-LO-mediated action and is currently considered as a rather selective 12-LO inhibitor (Stern et al., 1996; Wen et al., 1996; Kuwata et al., 1998; Ka¨lvegren et al., 2007; de Carvalho et al., 2008; Guo et al., 2011).

However, the selectivity profile of CDC was analyzed so far only in cell-free assays using crude homogenates of rat plate-lets and rat PMNL as systems to test the effects on p12-LO and on 5- and 12/15-LO, respectively. Furthermore, the assay conditions for the different enzymes (e.g., AA concentrations, presence of glutathione) were inconsistent and did not con-sider pharmacologically relevant factors that may be rele-vant in intact cells and/or in vivo. In this study, we provide a detailed evaluation of the effectiveness of CDC to inhibit human 5-LO, 15-LO1, and p12-LO in selected well estab-lished and defined test systems, taking into account essential parameters that affect the efficiency in vivo.

Materials and Methods

Materials. CDC was from Enzo Life Sciences (Lo¨rrach, Ger-many). Zileuton [N-(1-benzo[b]thien-2-ylethyl)-N-hydroxyurea] was from Sequoia Research Products (Oxford, UK). Platelet-activating factor (PAF, C-16) was from Cayman Chemical (Ann Arbor, MI). ␭-Carrageenan type IV (isolated from Gigartina aciculaire and

Gigartina pistillata) was from Sigma-Aldrich (Milan, Italy). All other

chemicals were purchased from Sigma-Aldrich, unless stated other-wise. HPLC solvents were from Merck (Darmstadt, Germany).

Expression and Purification of 5-LO, p12-LO, and 15-LO1. 5-LO was expressed in Escherichia coli Bl21 (DE3) cells transformed with pT3-5LO, and purification of 5-LO was performed as described previously (Fischer et al., 2003). In brief, E. coli were harvested by

centrifugation (7700g for 15 min) and lysed in 50 mM triethanol-amine/HCl, pH 8.0, 5 mM EDTA, 60␮g/ml soybean trypsin inhibitor, 1 mM phenylmethylsulfonyl fluoride, 1 mM dithiothreitol, and 1 mg/ml lysozyme; homogenized by sonication (3⫻ 15 s); and centri-fuged at 10,000g for 15 min followed by centrifugation at 40,000g for 70 min at 4°C. The supernatant was then applied to an ATP-agarose column (Sigma-Aldrich), and the column was eluted as described previously (Brungs et al., 1995). Partially purified 5-LO was imme-diately used for activity assays.

p12-LO was expressed in E. coli Bl21 (DE3) cells transformed with pT3-his12-LO (provided by Dr. Colin Funk, Kingston, Canada), and 15-LO1 protein was expressed in E. coli Bl21 (DE3) cells transformed with pQE-9-his15-LO1 (Dr. Hartmut Ku¨ hn, Berlin, Germany). Puri-fication of his-p12-LO and his-15-LO1 was performed using nickel-affinity chromatography. In brief, E. coli were harvested and lysed as described for 5-LO. The resulting supernatant (S40) was then mixed with nickel-nitrilotriacetic-agarose beads (QIAGEN GmbH, Hilden, Germany) in a ratio of 1:200 (for p12-LO) or 1:40 (for 15-LO1) and incubated overnight at 4°C under rotation. Beads were then washed three times for 5 min on ice with 50 mM NaH2PO4and 300 mM NaCl containing 50 mM imidazole (for p12-LO) or 20 mM imidazole (for 15-LO1) followed by wash three times for 5 min with elution buffer (50 mM NaH2PO4, 300 mM NaCl, and 250 mM imidazole). Centrif-ugation at 1200g was performed after each step. Eluted fractions were pooled and used immediately for p12-LO and 15-LO1 activity assays, respectively.

Diphenylpicrylhydrazyl Assay. The antioxidant activity of CDC was assessed by the method of Blois (1958), with slight modi-fications. In brief, 100␮l of 1, 2.5, 5, 10, 25, 50, or 100 ␮M CDC in ethanol (corresponding to 0.1, 0.25, 0.5, 1, 2.5, 5, or 10 nmol) were added to 100␮l of a solution of the stable free radical diphenylpic-rylhydrazyl (DPPH) in ethanol (100␮M, corresponding to 10 nmol) buffered with acetate to pH 5.5 in a 96-well plate. The absorbance was read at 520 nm after a 30-min incubation under gentle shaking in the dark. Ascorbic acid andL-cysteine were used as reference compounds and reduced the DPPH radical with a 1:2 (ascorbic acid/ DPPH) and 1:1 (L-Cys/DPPH) apparent stoichiometry (data not shown). All of the analyses were performed in triplicates.

Cells. Human PMNL, platelets, and monocytes were freshly iso-lated from leukocyte concentrates obtained at the Blood Center, University Hospital (Tuebingen, Germany). In brief, venous blood was collected from fasted (12 h) adult female healthy volunteers with consent. The subjects had no apparent inflammatory conditions and had not taken anti-inflammatory drugs or anticoagulants for at least 10 days before blood collection. Venous blood was subjected to cen-trifugation at 4000g, 20 min, 20°C for preparation of leukocyte con-centrates. PMNL, platelet-rich plasma, and peripheral blood mono-nuclear cells (PBMC) were separated by dextran sedimentation and centrifugation on Nycoprep cushions (PAA Laboratories, Linz, Aus-tria). For isolation of PMNL, pelleted cells were collected, and hypo-tonic lysis of erythrocytes was performed as described previously (Werz et al., 2002). For isolation of platelets, platelet-rich plasma was obtained from the supernatants, mixed with phosphate-buffered saline (PBS), pH 5.9 (3:2 v/v), and centrifuged (2100g, 15 min, room temperature). The pelleted platelets were resuspended in PBS, pH 5.9/0.9% NaCl (1:1, v/v), centrifuged again, and resuspended in PBS pH 7.4. For isolation of human monocytes, the PBMC fractions were collected and washed three times with ice-cold PBS, and then mono-cytes were separated by adherence for 1.5 h at 37°C to culture flasks (Greiner, Nuertingen, Germany; cell density was 2⫻ 107 cells/ml RPMI 1640 medium containing 2 mML-glutamine and 50 ␮g/ml penicillin/streptomycin), which finally gave a purity of approxi-mately 80%, defined by forward- and side-light scatter properties and detection of the CD14 surface molecule by flow cytometry (BD FACSCalibur; BD Biosciences, Franklin Lakes, NJ).

Determination of LO Product Formation. For assays of intact cells, 107freshly isolated PMNL or monocytes and 108freshly iso-lated platelets were finally resuspended in 1 ml of PBS plus 1 mg/ml

(3)

glucose and 1 mM CaCl2(PGC) buffer. After preincubation with CDC at 37°C for 10 min, LO product formation was started by the addition of the respective stimuli as indicated. The reaction was stopped with 1 ml of methanol, and 30␮l of 1 N-HCl, 200 ng of prostaglandin B1, and 500␮l of PBS were added.

For assays in whole blood, aliquots of freshly withdrawn blood (2 ml) were preincubated with CDC or vehicle (0.1% DMSO) for 10 min at 37°C, and formation of LO products was started by the addition of calcimycin (A23187) (30␮M) in the presence or absence of 100 ␮M AA (Pergola et al., 2008). Samples were incubated for 10 min at 37°C, and the reaction was stopped on ice. The samples were centrifuged (600g, 10 min, 4°C), aliquots of the resulting plasma (500␮l) were then mixed with 2 ml of methanol, and 200 ng of prostaglandin B1 were added as internal standard. The samples were placed at⫺20°C for 2 h and centrifuged again (600g, 15 min, 4°C). The supernatants were collected and diluted with 2.5 ml of PBS and 75␮l of 1 N HCl. Formed 5-LO metabolites, 12(S)-H(P)ETE and 15(S)-H(P)ETE, were extracted and analyzed by HPLC as described previously (Werz et al., 2002). 5-LO products include LTB4and its all-trans-isomers and 5(S)-H(P)ETE. The cysteinyl-LTs were not included in the anal-yses, since previous studies showed that the amount of cysteinyl-LTs formed in A23187 ionophore-stimulated blood leukocytes and whole blood is negligible (approximately 1–2% of the total amount of 5-LO products produced; Pergola et al., 2008).

For determination of LO product formation in homogenates, 1 mM EDTA was added to cells resuspended in PBS. Samples were cooled on ice (5 min) and sonicated (3⫻ 10 s) at 4°C. For 5-LO, 1 mM ATP was added. Samples of either partially purified enzymes (1 ml) or cell homogenates (corresponding to 2⫻ 106PMNL or monocytes and 108 platelets in 1 ml) were incubated 10 min at 4°C with vehicle (0.1% DMSO, control) or CDC and prewarmed for 30 s at 37°C and 2 mM CaCl2, and the indicated concentrations of AA were added. The reaction was stopped after 10 min at 37°C by the addition of 1 ml of ice-cold methanol, and 200 ng of prostaglandin B1 were added. Formed metabolites were extracted and analyzed by HPLC as de-scribed previously (Werz et al., 2002).

Animals. Female Wistar Han rats (220 –230 g; Harlan, Milan, Italy) and female CD-1 mice (25–30 g; Harlan) were housed in a controlled environment and provided with standard rodent chow and water. All animals were allowed to acclimate for 4 days before experiments and were subjected to 12-h light/dark schedule. Exper-iments were conducted during the light phase. Animal care was in compliance with Italian regulations on protection of animals used for experimental and other scientific purpose (Ministerial Decree 116192) as well as with the European Economic Community regula-tions (Official Journal of E.C. L 358/1 12/18/1986).

Mouse PAF-Induced Shock. PAF C-16 was dissolved in chloro-form and stored at⫺20°C. The PAF working solution was freshly prepared directly before use. To this aim, chloroform was evaporated under N2, and PAF was dissolved in 0.9% saline solution containing 0.25% bovine serum albumin. Mice were challenged with 200␮g/kg PAF in a volume of 200␮l via a tail vein injection 30 min after an intraperitoneal injection of either vehicle (0.9% saline solution con-taining 4 or 2% DMSO as control for CDC and zileuton, respectively), CDC or zileuton, at the indicated doses. Death (determined by ces-sation of breathing) was then recorded over a period of 2 h. All animals surviving the 2-h test session were euthanized by CO2 inhalation.

Carrageenan-Induced Pleurisy in Rats. CDC (10 mg/kg) or vehicle (0.5 ml of 0.5% carboxymethylcellulose and 10% Tween 20) were given orally 1 h before carrageenan. Rats were anesthetized with 4% enflurane mixed with 0.5 l/min O2, 0.5 l/min N2O and submitted to a skin incision at the level of the left sixth intercostal space. The underlying muscle was dissected, and saline (0.2 ml) or 1% (w/v)␭-carrageenan type IV (0.2 ml) was injected into the pleural cavity. The skin incision was closed with a suture, and the animals were allowed to recover. Animals were killed after 2 h by inhalation of CO2. The chest was carefully opened, and the pleural cavity was

rinsed with 2 ml of saline solution containing heparin (5 U/ml). The exudate and washing solution were removed by aspiration, and any exudate that was contaminated with blood was discarded. The amounts of LTB4in the supernatant of centrifuged exudate (800g for 10 min) were assayed by enzyme immunoassay (Cayman Chemical, Ann Arbor, MI) according to manufacturer’s protocol. Results are expressed as the total amount of LTB4 measured in the pleural exudate of one rat (nanograms per rat).

Statistics. Results are expressed as mean⫹ S.E.M. of the mean of n observations, where n represents the number of experiments performed on different days in duplicates or the number of animals as indicated. The IC50values were determined by linear interpola-tion and validated with GraphPad InStat program (GraphPad Soft-ware, Inc., San Diego, CA). Data fit was obtained using the sigmoidal dose-response equation (variable slope) (Prism software; GraphPad Software, Inc.). Statistical evaluation of the data was performed by one-way analysis of variance (ANOVA) for independent or correlated samples followed by Tukey honest significant difference post hoc tests. Where appropriate, Student’s t test for paired observations was applied. Contingency tables were analyzed with one-sided Fish-er’s exact test. A p value⬍ 0.05 was considered significant.

Results

Inhibition of Partially Purified LOs by CDC. We at-tempted to evaluate the molecular pharmacological profile of CDC as an inhibitor of human LOs. To this aim, we first analyzed the effect of CDC on partially purified human 5-LO, p12-LO, and 15-LO1 under standardized assay conditions (20 ␮M AA as substrate and 1 mM Ca2⫹as supplement). CDC

exhibited an extremely potent inhibitory activity on human 5-LO. We determined the IC50at 15⫾ 5 nM (Fig. 2A), and

total suppression was achieved already at 100 nM CDC. It is noteworthy that CDC was significantly less effective as a p12-LO inhibitor, and the IC50value and the concentration

needed to achieve total suppression of p12-LO (500 nM and 3 ␮M, respectively) were approximately 30 times higher than those for 5-LO (Fig. 2A). It is noteworthy that CDC also inhibited 15-LO1 (IC50⫽ 300 nM) with a similar potency as

the p12-LO (Fig. 2A). Variations of the enzymatic activities of the LOs from batch to batch did not affect the inhibitory potencies of CDC (data not shown).

Enzymatic oxidation of AA by LOs involves the generation of free radical intermediates, and redox-active compounds can scavenge radicals and/or reduce the active site iron and thus uncouple the catalytic redox-cycle of LOs. CDC was originally developed from the redox-type 5-LO inhibitor caf-feic acid, sharing the catechol moiety with the parental com-pound. To evaluate the redox potential of CDC, we analyzed its radical scavenging properties in the well recognized DPPH assay. CDC reduced the DPPH radicals with an ap-parent 1:4 stoichiometry (Fig. 2B). Typically, LO inhibitors act in a reversible manner; thus, we tested whether the effects of CDC are reversible. Partially purified 5-LO enzyme was preincubated with 0.1 ␮M CDC. After 10-fold dilution with assay buffer (0.1– 0.01␮M CDC), 5-LO product forma-tion was almost at the same level as that after treatment with 0.01 ␮M CDC without precedent dilution (Fig. 2C). Similar results were observed for p12-LO and 15-LO1 (CDC concentrations of 1 and 0.1␮M). Taken together, CDC is a direct, redox-active, reversible LO inhibitor and exhibits the following order of potency: 5-LO⬎ 15-LO1 ⫽ p12-LO.

Inhibition of LOs by CDC in Cell Homogenates. The efficiency of LO inhibitors often depends on the assay

(4)

condi-tions, and cellular components (e.g., in cell homogenates) and/or the cellular environment (e.g., in intact cells) may strongly influence their inhibitory activity. Thus, we next analyzed the effects of CDC on the activity of LO enzymes in homogenates of human PMNL (5-LO, 15-LO1), monocytes (5-LO, leukocyte-type 12-LO), and platelets (p12-LO) under optimized AA concentrations (20, 10, and 3␮M AA, respec-tively). In accordance with data obtained for partially puri-fied LOs, CDC potently inhibited 5-LO activity in homoge-nates of both PMNL and monocytes. In fact, the respective IC50values were in the low nanomolar range (25⫾ 5 and 9 ⫾

5 nM; Fig. 3, A and B). However, CDC reduced 15(S)-H(P)ETE synthesis by 15-LO1 in PMNL homogenates at significantly higher concentrations, with an IC50of 2⫾ 0.5

␮M (Fig. 3A). For 15-LO1-derived formation of 12(S)-H(P)ETE, the IC50was higher than 3␮M (Fig. 3A). Similar

results were obtained for inhibition of the leukocyte-type 12-LO in monocyte homogenates; the resulting IC50 values

were 1.5⫾ 1 ␮M for 12(S)-H(P)ETE and 0.5 ⫾ 0.2 ␮M for 15(S)-H(P)ETE (Fig. 3B). Finally, inhibition of p12-LO in platelet homogenates was also achieved at higher concentra-tions (IC50⫽ 1 ⫾ 0.3 ␮M; Fig. 3C) than inhibition of 5-LO.

Thus, the potency of CDC in cell homogenates is similar to that observed for the isolated enzymes.

Inhibition of LO Product Formation by CDC in In-tact Cells. We next evaluated the effect of CDC on LOs in

intact cells. After stimulation of PMNL with A23187 iono-phore, 5-LO products and 12(S)-H(P)ETE are synthesized. CDC concentration-dependently inhibited 5-LO product for-mation, with an IC50of 0.5⫾ 0.1 ␮M (Fig. 4A, left), and thus

resulted to be more potent than the 5-LO inhibitor zileuton under the same assay condition (IC50⫽ 1.7 ⫾ 0.7 ␮M; Fig.

4A, right). The formation of the 5-LO metabolites LTB4and

5(S)-H(P)ETE was inhibited with a similar potency by CDC (data not shown). A similar effect was observed for inhibition of 12(S)-H(P)ETE (IC50⫽ 0.8 ⫾ 0.2 ␮M). In the presence of

exogenous AA, CDC still inhibited 5-LO (IC50 ⫽ 0.8 ⫾ 0.1

␮M) but was significant less effective in reducing 12(S)-H(P)ETE formation (IC50, approximately 2 ␮M) and only

hardly reduced 15(S)-H(P)ETE production (Fig. 4B). Analo-gous to PMNL, CDC efficiently reduced 5-LO product forma-tion in A23187-stimulated monocytes in the presence or ab-sence of exogenous AA (IC50⫽ 0.45 ⫾ 0.1 ␮M, both; Fig. 5, A

and B). A similar effect was observed in A23187-stimulated cells on 12(S)-H(P)ETE formation (IC50⫽ 1.1 ⫾ 0.2 ␮M; Fig.

5A). It is noteworthy that the addition of exogenous AA resulted in a significant improvement of the potency on 12(S)-H(P)ETE formation by the leukocyte-type 12-LO (IC50

⫽ 0.1 ⫾ 0.05 ␮M), whereas the reduction of 15(S)-H(P)ETE was observed only at higher concentrations (IC50 ⫽ 6 ⫾ 1

␮M) (Fig. 5B). In A23187-stimulated platelets, CDC only moderately inhibited p12-LO (IC50 ⬎ 10 ␮M), whereas a

Fig. 2. Inhibition of partially purified lipoxygenases by

CDC. A, partially purified recombinant human 5-LO, p12-LO, and 15-LO1 were preincubated with CDC or vehicle (DMSO, 0.1%) at 4°C in 1 ml of PBS pH 7.4 containing 1 mM EDTA and 1 mM ATP. After 10 min, samples were prewarmed at 37°C for 30 s, 2 mM CaCl2and 20␮M AA

were added, and after another 10 min, LO product forma-tion was determined. Data are expressed as percentage of control, means⫹ S.E.M.; n ⫽ 3, duplicates. LO products in 100% controls were (in nanograms per milliliter): 5-LO, 705.43⫾ 84.51; 12(S)-H(P)ETE, 1539.96 ⫾ 53.91; 15(S)-H(P)ETE, 636.64⫾ 100.4. 5-LO, ⴱ, p ⬍ 0.05, ⴱⴱⴱ, p ⬍ 0.001 versus 100% control; 12-LO and 15-LO, †, p⬍ 0.05, ††, p ⬍ 0.01, †††, p⬍ 0.001 versus 100% control; ANOVA ⫹ Tukey. B, radical scavenging properties. CDC was incubated with 10 nmol of DPPH in 100␮l of EtOH for 30 min at room temperature, and the absorbance at 520 nm was measured. Data are means⫾ S.D.; triplicates. C, partially purified recombinant human 5-LO, p12-LO, and 15-LO1 were pre-incubated with CDC at the concentrations of 0.1␮M (5-LO, left) or 1 ␮M (12- and 15-LO, middle and right, respec-tively) for 10 min at 4°C. One aliquot then was diluted with assay buffer 10-fold, whereas the other one was not altered, and 20␮M AA plus 1 mM Ca2⫹were added to start the reaction. For comparison, 5-LO was preincubated with 0.01 ␮M CDC and 12- and 15-LO1 with 0.1 ␮M CDC, and then 20␮M AA plus 1 mM Ca2⫹were added (no dilution). All of

the samples then were incubated for 10 min at 37°C, and 5-LO product formation was analyzed as described. Values are given as means⫹ S.E.M.; n ⫽ 3. ⴱ, p ⬍ 0.05, ⴱⴱⴱ, p ⬍ 0.001, Student’s t test.

(5)

paradoxical increase of 12(S)-H(P)ETE formation was ob-served when cells were stimulated by exogenous AA (Fig. 5C). Finally, 12(S)-H(P)ETE formation also was evaluated after activation of platelets with more pathophysiological relevant stimuli, i.e., thrombin and collagen. Under these

conditions, the inhibitory potential of CDC increased (IC50⫽

0.8⫾ 0.1 and 1 ⫾ 0.1 ␮M for thrombin and collagen, respec-tively; Fig. 5C).

Inhibition of LO Product Formation by CDC in Whole Blood. Cell-free assays and test systems based on isolated cells represent suitable models to evaluate the inter-ference of test compounds with selected targets. On the other hand, to take into account the influence of in vivo relevant parameters on the efficiency of a LO inhibitor, whole-blood assays include critical pharmacokinetic and pharmacody-namic variables and are considered relevant test systems for translation of drug actions in vivo. CDC concentration-de-pendently inhibited 5-LO product formation in A23187-stim-ulated whole blood (Fig. 6A, left). Compared with isolated neutrophils and monocytes, the potency was slightly reduced (IC50⫽ 3.2 ⫾ 1.5 ␮M), but still CDC was essentially

equipo-tent to the 5-LO inhibitor zileuton (IC50⫽ 1.5 ⫾ 1 ␮M; Fig.

6A, right). However, no total suppression of 5-LO product synthesis was achieved up to 30␮M CDC. The 5-LO metab-olites LTB4 and 5(S)-H(P)ETE were inhibited to a similar

extent (data not shown). The addition of exogenous AA fur-ther reduced the inhibitory potential of CDC (IC50⫽ 8 ⫾ 1

␮M) (Fig. 6B). It is noteworthy that the potency of CDC to suppress formation of 12(S)-H(P)ETE (IC50⫽ 3–8 ␮M) and

15(S)-H(P)ETE (IC50⫽ 23 ⫾ 13 ␮M; Fig. 6, A and B) was

rather similar as that for 5-LO. Note that for inhibition of all three LOs, a plateau was reached at approximately 40 to 50% of remaining LO activity.

CDC Protects Mice against PAF-Induced Shock and Reduces LTB4 in Carrageenan-Treated Rats. To test whether 5-LO inhibition by CDC translates into significant functional effects in vivo, we used the model of the

PAF-Fig. 3. Inhibition of lipoxygenases by CDC in cell homogenates.

Homog-enates of PMNL (A), monocytes (B), or platelets (C) were preincubated with the indicated concentrations of CDC or vehicle (DMSO, 0.1%) at 4°C in 1 ml of PBS, pH 7.4, containing 1 mM EDTA and 1 mM ATP. After 10 min, samples were prewarmed at 37°C for 30 s; 2 mM CaCl2and 20␮M

(PMNL), 10␮M (monocytes), or 3 ␮M (platelets) AA were added, and after another 10 min, LO product formation was determined. Data are expressed as percentage of control, means⫹ S.E.M.; n ⫽ 3, duplicates. 5-LO products,ⴱ, p ⬍ 0.05, ⴱⴱⴱ, p ⬍ 0.001 versus 100% control; 12(S)-H(P)ETE, †, p⬍ 0.05, ††, p ⬍ 0.01, †††, p ⬍ 0.001 versus 100% control; 15(S)-H(P)ETE, ‡, p⬍ 0.05, ‡‡, p ⬍ 0.01, ‡‡‡, p ⬍ 0.001 versus 100% control; ANOVA⫹ Tukey. LO products in 100% controls were (in nano-grams per milliliter): A, PMNL (2⫻ 106cells/ml): 5-LO, 421.11⫾ 70.39;

12(S)-H(P)ETE, 87.22⫾ 8.03; 15(S)-H(P)ETE, 157.77 ⫾ 58.77. B, mono-cytes (2⫻ 106cells/ml): 5-LO, 74.7⫾ 70.39; 12(S)-H(P)ETE, 168.69 ⫾

30.52; 15(S)-H(P)ETE, 15.3 ⫾ 4.2. C, platelets (108 cells/ml):

12(S)-H(P)ETE, 491.02⫾ 191.3.

Fig. 4. Inhibition of lipoxygenase product formation by CDC in intact

PMNL. A and B, freshly isolated PMNL were preincubated (10 min, 37°C) with CDC, zileuton, or vehicle (DMSO, 0.1%) in PGC buffer and stimu-lated with 2.5␮M A23187 (A) or 2.5 ␮M A23187 plus 20 ␮M AA (10 min, 37°C) (B). Data are expressed as percentage of control, means⫹ S.E.M.;

n⫽ 3, duplicates. 5-LO products, ⴱ, p ⬍ 0.05, ⴱⴱⴱ, p ⬍ 0.001 versus 100% control; 12(S)-H(P)ETE, ††, p⬍ 0.01, †††, p ⬍ 0.001 versus 100% control; ANOVA⫹ Tukey. LO products in 100% controls were (in nanograms per milliliter, 107cells/ml): A23187: 5-LO, 200.5⫾ 6; 12(S)-H(P)ETE, 56.5 ⫾

13. A23187⫹ AA: 5-LO, 952.1 ⫾ 45.2; 12(S)-H(P)ETE, 82.94 ⫾ 20.07; 15(S)-H(P)ETE, 218.42⫾ 81.87.

(6)

induced lethal shock in mice. In fact, previous studies have established a critical role of 5-LO products in the acute PAF toxicity (Chen et al., 1994; Byrum et al., 1997) and represents a suitable model to test the in vivo efficacy of LT-synthesis inhibitors. Administration of 200 ␮g/kg PAF to vehicle-treated mice caused mortality in 10 of 10 animals (within 30 min). CDC at the dose of 3.5 mg/kg, but not at 1 mg/kg i.p., led to a survival of 40% of the animals (Table 1). The survival rate was not further improved by doubling the dose (7 mg/kg i.p.). Treatment with the 5-LO inhibitor zileuton at the dose of 10 mg/kg i.p. before PAF injection also resulted in improve-ment in survival of 40% (Table 1). Finally, we tested the effect of CDC on the levels of the 5-LO product LTB4in an in

vivo model of carrageenan-induced pleurisy in rats. The ad-ministration of a single oral dose of 10 mg/kg CDC to rats 1 h

Fig. 5. Inhibition of lipoxygenase product formation by CDC in intact

monocytes and platelets. LO product formation in intact monocytes after preincubation (10 min, 37°C) with CDC or vehicle (DMSO, 0.1%) in PGC buffer and stimulation with 2.5 ␮M A23187 (A) or 2.5 ␮M A23187 plus 20 ␮M AA (B) (10 min, 37°C). Data are expressed as percentage of control, means⫹ S.E.M.; n ⫽ 3, duplicates. 5-LO prod-ucts,ⴱ, p ⬍ 0.05, ⴱⴱ, p ⬍ 0.01, ⴱⴱⴱ, p ⬍ 0.001 versus 100% control; 12(S)-H(P)ETE, ††, p⬍ 0.01, †††, p ⬍ 0.001 versus 100% control; 15(S)-H(P)ETE, ‡, p⬍ 0.05, ‡‡, p ⬍ 0.01 versus 100% control; ANOVA ⫹ Tukey. LO products in 100% controls were (nanograms per milliliter, 2⫻ 106

cells/ml): A23187: 5-LO, 65.2⫾ 30; 12(S)-H(P)ETE, 189.9 ⫾ 68. A23187⫹ AA: 5-LO, 44.6 ⫾ 18; 12(S)-H(P)ETE, 289.9 ⫾ 148; 15(S)-H(P)ETE, 11.9⫾ 3.9. C, p12-LO product formation in intact platelets after preincubation (10 min, 37°C) with the indicated concentrations of CDC or vehicle (DMSO) in PGC buffer and stimulation with 2.5␮M A23187, 3␮M AA, 8 ␮g/ml collagen, or 1 U/ml thrombin (10 min, 37°C). Data are expressed as percentage of control, means⫹ S.E.M.; n ⫽ 3, duplicates. Thrombin, collagen,ⴱⴱ, p ⬍ 0.01, ⴱⴱⴱ, p ⬍ 0.001 versus 100% control; A23187, ††, p⬍ 0.01 versus 100% control; ANOVA ⫹ Tukey. 12(S)-H(P)ETE in 100% controls was (nanograms per milliliter, 108 cells/ml): A23187, 408.9 ⫾ 77.4; AA, 311.6 ⫾ 55.7; collagen,

120.6⫾ 34.4; thrombin, 241.1 ⫾ 31.4.

Fig. 6. Inhibition of LO product formation by CDC in human whole blood

and in carrageenan-treated rats. LO product formation in human whole blood, after preincubation (10 min, 37°C) with CDC, zileuton, or vehicle (DMSO, 0.1%) in PGC buffer and stimulation with 30␮M A23187 (A) or 30␮M A23187 ⫹ 100 ␮M AA (10 min, 37°C) (B). Data are expressed as percentage of control, means⫹ S.E.M.; n ⫽ 3, duplicates. 5-LO products, ⴱ, p ⬍ 0.05, ⴱⴱ, p ⬍ 0.01, ⴱⴱⴱ, p ⬍ 0.001 versus 100% control; 12(S)-H(P)ETE, †, p⬍ 0.05, ††, p ⬍ 0.01 versus 100% control; 15(S)-H(P)ETE, ‡, p⬍ 0.05, ‡‡, p ⬍ 0.01 versus 100% control; ANOVA ⫹ Tukey. LO products in 100% controls were (nanograms per milliliter plasma): A23187: 5-LO, 171.8⫾ 43.2; 12(S)-H(P)ETE, 359.4 ⫾ 106. A23187 ⫹ AA: 5-LO, 355.7 ⫾ 33.7; 12(S)-H(P)ETE, 3101.8 ⫾ 585; 15(S)-H(P)ETE, 44.6 ⫾ 8.4. C, LTB4pleural levels measured 2 h after intrapleural

injection of carrageenan in rats treated orally with vehicle (veh, 0.5 ml of 0.5% carboxymethylcellulose and 10% Tween 20) or CDC (10 mg/kg) 1 h before carrageenan injection. Data are expressed as percentage of control, means⫹ S.E.M.; n ⫽ 7, each. ⴱⴱⴱ, p ⬍ 0.001, Student’s t test.

(7)

before carrageenan significantly reduced LTB4pleural levels

by 34% (Fig. 6C).

Discussion

In this study, we analyzed the inhibitory actions and the selectivity profile of CDC on human 5-LO, 15-LO1, and p12-LO. On the basis of previous results by others obtained in rat cell homogenates, CDC so far has been considered as a potent and selective p12-LO inhibitor and as such used as a phar-macological tool to establish roles of p12-LO in biological processes. Our data confirm that CDC is a direct inhibitor of human LOs but demonstrate that 5-LO and not p12-LO is the preferred target of CDC in both cell-free and cell-based assays. It is noteworthy that inhibition of 5-LO by CDC was also evident in whole blood and that CDC was effective in vivo in the PAF-induced shock in mice and reduced LTB4

levels in carrageenan-treated rats. On the basis of its ability to inhibit 5-LO in various biological assays and the efficacy in animal models, we suggest further analysis to exploit the pharmacological potential of CDC. On the other hand, the potent inhibition of 5-LO by CDC together with its redox features suggest caution in interpreting results of pharma-cological studies where CDC is used as a tool to tie roles of p12-LO in biological processes.

LOs are AA-metabolizing enzymes implicated in several dis-eases, including inflammation and cancers (Yoshimoto and Takahashi, 2002; Ku¨ hn and O’Donnell, 2006; Peters-Golden and Henderson, 2007). To explore the respective contributions of 5-LO, 15-LO, and p12-LO in physiological and pathophysio-logical states and to pursue rational drug design selective in-hibitors would be valuable but many developed compounds act on more than one LO isoform. LOs are in fact all iron-containing redox-active enzymes and share AA as substrate, which may bind to a common, more or less conserved AA-binding pocket at the active site. With regard to inhibition of p12-LO, the fla-vonoid baicalein, which classically has been considered a selec-tive p12-LO inhibitor (IC50 ⫽ 0.64 ␮M) (Sekiya and Okuda,

1982), displays a similar inhibitory potency on 15-LO1 (IC50⫽

0.63 ␮M) (Deschamps et al., 2006) and also inhibits cellular 5-LO product formation (IC50⫽ 7.13 ␮M) (Kimura et al., 1985).

Likewise, gallocatechin gallate, which is selective for p12-LO compared with 15-LO1 (IC50 ⫽ 0.14 and ⬎100 ␮M,

respec-tively), effectively also inhibits 5-LO activity (IC50 ⫽ 2 ␮M)

(Yamamoto et al., 2005). In recent work from Suzuki et al. (2000), the tropolone hinokitiol was identified as a potent and selective p12-LO inhibitor (IC50⫽ 0.1 ␮M for p12-LO, 17 ␮M for

5-LO, 50␮M for leukocyte-type 12-LO, and ⬎100 ␮M for 15-LO1). However, this compound undergoes a fast photochemical degradation by sunlight, which hampered its use as a lead compound. In addition, five novel 12-LO inhibitors were recently identified by high-throughput screening of diverse libraries (Deschamps et al., 2007), but only the selectivity toward 15-LOs was analyzed and only on isolated enzymes, although no data are available from cell-based assays or in vivo studies.

CDC was developed from the 5-LO inhibitor caffeic acid to gain selectivity on p12-LO. Initially, CDC was found to be potent and rather selective for p12-LO over 5-LO (5-LO/ p12-LO IC50 ratio ⫽ 30) and 12/15-LO (12/15-LO/p12-LO

IC50 ratio ⫽ 53). In this initial study, the compound was

evaluated in homogenates of platelets and PMNL from rats in the presence of 0.05␮Ci of [14C]AA and 1 mM glutathione

for p12-LO and the presence of 0.2␮Ci of [14C]AA for 5-LO

and 12/15-LO. The formed AA metabolites were quantified by thin-layer chromatography and autoradiography. In our study, we analyzed CDC using human partially purified LO enzymes and homogenates of human cells under consistent assay conditions and monitored the metabolites via HPLC. It is noteworthy that we observed a higher inhibitory potential of CDC for 5-LO (IC50⫽ 9–25 nM) than initially reported for

rat neutrophil homogenates (IC50 ⫽ 1.89 ␮M) (Cho et al.,

1991), whereas similar effects were found for human 15-LO1 and rat 12/15-LO (IC50⫽ 2 and 3.33 ␮M, in human and rat

neutrophil homogenates, respectively). On the contrary, a significant lower potency against human p12-LO (IC50 ⫽

0.5–1 ␮M) was evident in comparison to previous work in homogenates of rat platelets (IC50 ⫽ 63 nM) (Cho et al.,

1991). Thus, our data indicate a divergent selectivity profile of CDC than initially reported and define 5-LO as a prefer-ential target. The variations observed in the potency of CDC are not readily understood, although both species-related differences (homogenates of human and rat cells) and/or the different experimental conditions (e.g., amount of AA as sub-strate, presence of glutathione for p12-LO incubations) might be relevant. In accordance with our findings, Bu¨ rger et al. (2000) reported an IC50⫽ 0.3 ␮M for inhibition of isolated

p12-LO by CDC in the presence of 100␮M AA. In addition, Tornhamre et al. (2000) determined an IC50 of 50 ␮M for

inhibition of p12-LO in the cytosolic fraction of human plate-lets in the presence of 25␮M AA and 10 ␮M 13-hydroperoxy-9,11-octadecadienoic acid (13-HpODE). Under these condi-tions, however, the oxidative capacity of 13-HpODE might have strongly impaired the potency of CDC.

LOs contain a non-heme iron, and the catalytic reaction proceeds by a radical-based mechanism where the iron cycles between the ferrous and ferric state (Rådmark et al., 2007). Agents containing a catechol moiety may reduce the active site iron and uncouple the catalytic cycle of LOs (Pergola and Werz, 2010). We confirmed the radical scavenging properties of CDC and thus suggest the reducing feature as the mech-anism responsible for LO inhibition. Note that the catechol structure is assumed to combine antioxidant and iron-chelat-ing properties, resultiron-chelat-ing in highly efficient inhibition of LO enzymes, but electronic and steric features as well as the hydrophobicity of the respective catechol derivative may also determine potency and selectivity (Whitman et al., 2002; Pontiki and Hadjipavlou-Litina, 2008). CDC possesses a 3,4-dihydroxystyrene structure and has a computed log P of 3.63,

TABLE 1

CDC protects mice against PAF-induced shock

PAF was injected into the tail vein of female mice at a dose of 200␮g/kg. CDC (1, 3.5, and 7 mg/kg), zileuton (10 mg/kg), or vehicle control (0.9% saline solution containing 4% DMSO or 2% DMSO, as control for CDC and zileuton, respectively) was admin-istered intraperitoneally 30 min before PAF. The number of surviving animals and the total number of animals tested for each group are given.

Survivors/Total CDC 0 0/10 1 mg/kg 0/5 3.5 mg/kg 2/5 7 mg/kg 4/10* Zileuton 0 0/5 10 mg/kg 2/5

(8)

and both the styrene double bound on catechol derivatives and log P values close to 4 have been reported as favorable for inhibition of 5-LO (Naito et al., 1991). It is noteworthy that the recent elucidation of the structure of human 5-LO has revealed a distinct active site compared with other LOs (Gil-bert et al., 2011), which may be relevant for the selectivity profile of LO inhibitors.

Cellular LO product formation is subject to regulation by diverse proteins (e.g., 5-LO-activating protein), signaling molecules (e.g., mitogen-activated protein kinases and Ca2⫹),

and lipids (e.g., phosphatidylcholine, diacylglycerol), which may in turn also influence the efficiency of drugs (Pergola and Werz, 2010). CDC inhibited 5-LO in both PMNL and monocytes and showed similar efficiencies as the 5-LO inhib-itor zileuton. A significant inhibition by CDC was also ob-served for the leukocyte-type 12-LO, whereas 12(S)-H(P)ETE formation in platelets depended on the stimulation. Thus, p12-LO activity was inhibited by CDC when thrombin and collagen (and to a lesser extent ionophore) were used as stimuli, but not in presence of exogenous AA. The addition of exogenous AA also partially modified inhibition of 12(S)-H(P)ETE by CDC in neutrophils and monocytes, suggesting a possible influence of CDC on cellular AA release. CDC also significantly reduced 5-LO product synthesis in whole blood, which implicates 5-LO as a pharmacological relevant target. In this model, CDC also inhibited 12(S)-H(P)ETE and 15(S)-H(P)ETE production, with a similar potency as that for 5-LO, indicating that blood component(s) might have a stronger influence on the inhibitory activity on 5-LO than on other LOs. Finally, CDC was effective in animal models where 5-LO products play critical roles. Although the protective effect of CDC also may ultimately be related to inhibition of other LOs than 5-LO, CDC was equipotent to the 5-LO in-hibitor zileuton. It should be noted that such consistent effi-ciency for a 5-LO inhibitor is rare, because several com-pounds, which potently inhibited 5-LO in isolated cells, failed in whole blood or in vivo because of high plasma protein binding or competition with blood components (e.g., fatty acids), and zileuton currently represents the only 5-LO in-hibitor that could reach the market (Werz, 2002). Further-more, redox-active polyphenols are often intensively oxidized during absorption after oral administration, but CDC was orally active in reducing LTB4levels in carrageenan-treated

rats. These data support the use of CDC as a potential lead candidate for further developments. However, it should be pointed out that the clinical systemic use of redox type 5-LO inhibitors have often been hampered by interference with other redox systems or by production of reactive radical spe-cies and consequent side effects (e.g., hemolysis or methemo-globin formation) (McMillan and Walker, 1992). Although polyphenolics with catechol rings have been shown to cause erythrocyte hemolysis and methemoglobin formation less readily than those containing phenol rings (Galati et al., 2002), an accurate analysis of the potential toxicity of CDC should be performed to justify further developments for sys-temic use. On the other hand, the characteristic of CDC to combine 5-LO inhibition with redox properties might be use-ful for topical applications, e.g., in dermatological diseases involving LTs (Wedi and Kapp, 2001; Kaplan, 2002). Acknowledgments

We thank Daniela Mu¨ ller for expert technical assistance.

Authorship Contributions

Participated in research design: Pergola, Jazzar, Rossi, Buehring,

Dehm, Sautebin, and Werz.

Conducted experiments: Pergola, Jazzar, Rossi, Buehring, Dehm,

and Luderer.

Performed data analysis: Pergola, Jazzar, Rossi, Buehring,

Lud-erer, Dehm, Sautebin, and Werz.

Wrote or contributed to the writing of the manuscript: Pergola and

Werz.

Other: Northoff contributed human blood and leukocyte concentrates.

References

Ba¨ck M (2009) Inhibitors of the 5-lipoxygenase pathway in atherosclerosis. Curr

Pharm Des 15:3116 –3132.

Blois MS (1958) Antioxidant determinations by the use of a stable free radical.

Nature 181:1199 –1200.

Brash AR (1999) Lipoxygenases: occurrence, functions, catalysis, and acquisition of substrate. J Biol Chem 274:23679 –23682.

Brungs M, Rådmark O, Samuelsson B, and Steinhilber D (1995) Sequential induc-tion of 5-lipoxygenase gene expression and activity in Mono Mac 6 cells by trans-forming growth factor beta and 1,25-dihydroxyvitamin D3. Proc Natl Acad Sci

USA 92:107–111.

Bu¨ rger F, Krieg P, Marks F, and Fu¨ rstenberger G (2000) Positional- and stereo-selectivity of fatty acid oxygenation catalysed by mouse (12S)-lipoxygenase isoen-zymes. Biochem J 348:329 –335.

Byrum RS, Goulet JL, Griffiths RJ, and Koller BH (1997) Role of the 5-lipoxygenase-activating protein (FLAP) in murine acute inflammatory responses. J Exp Med

185:1065–1075.

Chen XS, Sheller JR, Johnson EN, and Funk CD (1994) Role of leukotrienes revealed by targeted disruption of the 5-lipoxygenase gene. Nature 372:179 –182. Cho H, Ueda M, Tamaoka M, Hamaguchi M, Aisaka K, Kiso Y, Inoue T, Ogino R,

Tatsuoka T, and Ishihara T (1991) Novel caffeic acid derivatives: extremely potent inhibitors of 12-lipoxygenase. J Med Chem 34:1503–1505.

de Carvalho DD, Sadok A, Bourgarel-Rey V, Gattacceca F, Penel C, Lehmann M, and Kovacic H (2008) Nox1 downstream of 12-lipoxygenase controls cell proliferation but not cell spreading of colon cancer cells. Int J Cancer 122:1757–1764. Deschamps JD, Gautschi JT, Whitman S, Johnson TA, Gassner NC, Crews P, and

Holman TR (2007) Discovery of platelet-type 12-human lipoxygenase selective inhibitors by high-throughput screening of structurally diverse libraries. Bioorg

Med Chem 15:6900 – 6908.

Deschamps JD, Kenyon VA, and Holman TR (2006) Baicalein is a potent in vitro inhibitor against both reticulocyte 15-human and platelet 12-human lipoxyge-nases. Bioorg Med Chem 14:4295– 4301.

Fischer L, Szellas D, Rådmark O, Steinhilber D, and Werz O (2003) Phosphorylation-and stimulus-dependent inhibition of cellular 5-lipoxygenase activity by nonredox-type inhibitors. FASEB J 17:949 –951.

Galati G, Sabzevari O, Wilson JX, and O’Brien PJ (2002) Prooxidant activity and cellular effects of the phenoxyl radicals of dietary flavonoids and other polyphe-nolics. Toxicology 177:91–104.

Gilbert NC, Bartlett SG, Waight MT, Neau DB, Boeglin WE, Brash AR, and New-comer ME (2011) The structure of human 5-lipoxygenase. Science 331:217–219. Guo QY, Miao LN, Li B, Ma FZ, Liu N, Cai L, and Xu ZG (2011) Role of

12-lipoxygenase in decreasing P-cadherin and increasing angiotensin II type 1 recep-tor expression according to glomerular size in type 2 diabetic rats. Am J Physiol

Endocrinol Metab 300:E708 –E716.

Ka¨lvegren H, Andersson J, Grenegård M, and Bengtsson T (2007) Platelet activation triggered by Chlamydia pneumoniae is antagonized by 12-lipoxygenase inhibitors but not cyclooxygenase inhibitors. Eur J Pharmacol 566:20 –27.

Kaplan AP (2002) Clinical practice. Chronic urticaria and angioedema. N Engl J Med

346:175–179.

Kim JA, Gu JL, Natarajan R, Berliner JA, and Nadler JL (1995) A leukocyte type of 12-lipoxygenase is expressed in human vascular and mononuclear cells. Evidence for upregulation by angiotensin II. Arterioscler Thromb Vasc Biol 15:942–948. Kimura Y, Okuda H, and Arichi S (1985) Studies on scutellariae radix; XIII. Effects

of various flavonoids on arachidonate metabolism in leukocytes. Planta Med 51: 132–136.

Ku¨ hn H and O’Donnell VB (2006) Inflammation and immune regulation by 12/15-lipoxygenases. Prog Lipid Res 45:334 –356.

Kuwata H, Nakatani Y, Murakami M, and Kudo I (1998) Cytosolic phospholipase A2 is required for cytokine-induced expression of type IIA secretory phospholipase A2 that mediates optimal cyclooxygenase-2-dependent delayed prostaglandin E2 gen-eration in rat 3Y1 fibroblasts. J Biol Chem 273:1733–1740.

McMillan RM and Walker ER (1992) Designing therapeutically effective 5-lipoxy-genase inhibitors. Trends Pharmacol Sci 13:323–330.

Nadel JA, Conrad DJ, Ueki IF, Schuster A, and Sigal E (1991) Immunocytochemical localization of arachidonate 15-lipoxygenase in erythrocytes, leukocytes, and air-way cells. J Clin Invest 87:1139 –1145.

Naito Y, Sugiura M, Yamaura Y, Fukaya C, Yokoyama K, Nakagawa Y, Ikeda T, Senda M, and Fujita T (1991) Quantitative structure-activity relationship of catechol derivatives inhibiting 5-lipoxygenase. Chem Pharm Bull (Tokyo) 39: 1736 –1745.

Pergola C, Dodt G, Rossi A, Neunhoeffer E, Lawrenz B, Northoff H, Samuelsson B, Rådmark O, Sautebin L, and Werz O (2008) ERK-mediated regulation of leuko-triene biosynthesis by androgens: a molecular basis for gender differences in inflammation and asthma. Proc Natl Acad Sci USA 105:19881–19886. Pergola C and Werz O (2010) 5-Lipoxygenase inhibitors: a review of recent

(9)

Peters-Golden M and Henderson WR Jr (2007) Leukotrienes. N Engl J Med 357: 1841–1854.

Pontiki E and Hadjipavlou-Litina D (2008) Lipoxygenase inhibitors: a comparative QSAR study review and evaluation of new QSARs. Med Res Rev 28:39 –117. Rådmark O, Werz O, Steinhilber D, and Samuelsson B (2007) 5-Lipoxygenase:

regulation of expression and enzyme activity. Trends Biochem Sci 32:332–341. Sekiya K and Okuda H (1982) Selective inhibition of platelet lipoxygenase by

baica-lein. Biochem Biophys Res Commun 105:1090 –1095.

Stern N, Kisch ES, and Knoll E (1996) Platelet lipoxygenase in spontaneously hypertensive rats. Hypertension 27:1149 –1152.

Suzuki H, Ueda T, Jura´nek I, Yamamoto S, Katoh T, Node M, and Suzuki T (2000) Hinokitiol, a selective inhibitor of the platelet-type isozyme of arachidonate 12-lipoxygenase. Biochem Biophys Res Commun 275:885– 889.

Tornhamre S, Elmqvist A, and Lindgren JA (2000) 15-Lipoxygenation of leukotriene A(4). Studies Of 12- and 15-lipoxygenase efficiency to catalyze lipoxin formation.

Biochim Biophys Acta 1484:298 –306.

Wedi B and Kapp A (2001) Pathophysiological role of leukotrienes in dermatological diseases: potential therapeutic implications. BioDrugs 15:729 –743.

Wen Y, Nadler JL, Gonzales N, Scott S, Clauser E, and Natarajan R (1996)

Mech-anisms of ANG II-induced mitogenic responses: role of 12-lipoxygenase and bipha-sic MAP kinase. Am J Physiol 271:C1212–C1220.

Werz O (2002) 5-Lipoxygenase: cellular biology and molecular pharmacology. Curr

Drug Targets Inflamm Allergy 1:23– 44.

Werz O, Bu¨ rkert E, Samuelsson B, Rådmark O, and Steinhilber D (2002) Activation of 5-lipoxygenase by cell stress is calcium independent in human polymorphonu-clear leukocytes. Blood 99:1044 –1052.

Whitman S, Gezginci M, Timmermann BN, and Holman TR (2002) Structure-activity relationship studies of nordihydroguaiaretic acid inhibitors toward soy-bean, 12-human, and 15-human lipoxygenase. J Med Chem 45:2659 –2661. Yamamoto S, Katsukawa M, Nakano A, Hiraki E, Nishimura K, Jisaka M, Yokota K,

and Ueda N (2005) Arachidonate 12-lipoxygenases with reference to their selective inhibitors. Biochem Biophys Res Commun 338:122–127.

Yoshimoto T and Takahashi Y (2002) Arachidonate 12-lipoxygenases.

Prostaglan-dins Other Lipid Mediat 68 – 69:245–262.

Address correspondence to: Dr. Carlo Pergola, Chair of Pharmaceutical/ Medicinal Chemistry, Institute of Pharmacy, Friedrich-Schiller-University Jena, Philosophenweg 14, D-07743 Jena, Germany. E-mail: carlo.pergola@uni-jena.de

Riferimenti

Documenti correlati

Con l’intensificarsi del numero delle esibizioni, diventò difficile per le ragazze occuparsi sia dell’esecuzione musicale che della recitazione, così, a partire dal 1924

Quindi riteniamo che questa metodica possa fornire una misura più accurata dello stato di idratazione rispetto ai valori estrapolati dalle stime calcolate dalla metodica

We provide a linguistic and legal analysis of a particular case of modificatory provision (the efficacy suspension), show how such knowledge can be formalized in a linguistic

Being originally devised for describing the behaviour of unbounded, uniform systems, the renor- malisation group has been subsequently generalised in order to account for the

These aggressive displays have been observed both in species where male-male contests occur in non-resource based leks (male dominance polygyny), as well as in species where

The initial description of SMC heterogeneity was made in the rat carotid artery injury model, wherein 2 SMC populations were identified: a spindle-shaped phenotype,

Given the difficulty of distinguishing among mental retardation, learning disabilities and OCD and the importance of precocious diagnosis in treating OCD especially since there