Dose-Dependent Inhibition of Platelet Cyclooxygenase-1 and Monocyte Cyclooxygenase-2 by Meloxicam in Healthy Subjects 1

Similar documents
Meloxicam vs etodolac cox 2 inhibition

Meloxicam: a review of its pharmacokinetics, efficacy and tolerability following intramuscular administration

Metacam 1.5 mg/ml oral suspension for dogs

Start of new generation of NSAIDs?

Irish Greyhound Board. Scientific Advisory Committee on Doping and Medication Control. Opinion on Carprofen

BIOEQUIVALENCE STUDY OF TWO BRANDS OF MELOXICAM TABLETS IN HEALTHY HUMAN PAKISTANI MALE SUBJECTS

SUMMARY OF PRODUCT CHARACTERISTICS. 1. NAME OF THE VETERINARY MEDICINAL PRODUCT Emdocam 20 mg/ml solution for injection for cattle, pigs and horses

SUMMARY OF PRODUCT CHARACTERISTICS

Non-steroidal anti-inflammatory drugs (NSAIDs) are used widely to relieve pain, with or without

Summary of Product Characteristics

SUMMARY OF PRODUCT CHARACTERISTICS. Animeloxan 1.5 mg/ml oral suspension for dogs. Active substance: Meloxicam 1.5 mg (equivalent to 0.

Just where it s needed.

Summary of Product Characteristics

NSAID Toxicity in Dogs & Cats Beware of Ibuprofen!

Pharmacokinetics and tolerability of meloxicam after i.m. administration

ANNEX I SUMMARY OF PRODUCT CHARACTERISTICS. Medicinal product no longer authorised

SUMMARY OF PRODUCT CHARACTERISTICS 1. NAME OF THE VETERINARY MEDICINAL PRODUCT

Meloxicam, an inhibitor of cyclooxygenase-2, increases the level of serum G-CSF and might

N.C. A and T List of Approved Analgesics 1 of 5

ANNEX I SUMMARY OF PRODUCT CHARACTERISTICS

ANNEX I SUMMARY OF PRODUCT CHARACTERISTICS

Caution: Federal law restricts this drug to use by or on the order of a licensed veterinarian.

SUMMARY OF PRODUCT CHARACTERISTICS 1. NAME OF THE VETERINARY MEDICINAL PRODUCT

Sea Turtle Analgesics Selection - NSAIDS. Loggerhead Coquina (postoperative ketorolac)

Research update - medicines for koalas

Veterinary Medicinal Product

ANNEX I SUMMARY OF PRODUCT CHARACTERISTICS

USA Product Label CLINTABS TABLETS. Virbac. brand of clindamycin hydrochloride tablets. ANADA # , Approved by FDA DESCRIPTION

ANNEX I SUMMARY OF PRODUCT CHARACTERISTICS 1/127

ANNEX I SUMMARY OF PRODUCT CHARACTERISTICS

Data Sheet. Arrow - Meloxicam 7.5 mg tablets are light yellow, round bevelled-edged, biconvex tablet with ML. on one side and > on the reverse.

Adverse Effects of Anti-Inflammatory Drugs:

- Federal (USA) law restricts this drug to use by or on the order of a licensed veterinarian.

SZENT ISTVÁN UNIVERSITY. Doctoral School of Veterinary Science

Copyright is owned by the Author of the thesis. Permission is given for a copy to be downloaded by an individual for the purpose of research and

COMMITTEE FOR VETERINARY MEDICINAL PRODUCTS

Development of Analytical Methods for the Determination of Flunixin and Phenylbutazone Drug Residues in Edible Bovine Tissues

The Pharmacokinetics of Firocoxib after Multiple Oral Doses to Neonatal Foals. Natasha Hovanessian

COMMITTEE FOR VETERINARY MEDICINAL PRODUCTS

NSAIDs Are You Following the Rules?

Final Report. Project code: P.PSH.0653 Prepared by: Fiona Cotter Troy Laboratories Pty Ltd Date published: July 2014

The new meloxicam range for cattle, pigs & horses

[09/2018] ANNEX I SUMMARY OF PRODUCT CHARACTERISTICS

COMMITTEE FOR VETERINARY MEDICINAL PRODUCTS

COMMITTEE FOR VETERINARY MEDICINAL PRODUCTS

Only for Intravenous Use in Beef and Dairy Cattle. Not for Use in Dry Dairy Cows and Veal Calves. For Intravenous or Intramuscular Use in Horses.

JMSCR Vol 05 Issue 03 Page March 2017

EPAR type II variation for Metacam

Metacam. The Only NSAID Approved for Cats in the US. John G. Pantalo, VMD Professional Services Veterinarian. Think easy. Think cat. Think METACAM.

COMMITTEE FOR MEDICINAL PRODUCTS FOR VETERINARY USE

ANNEX I SUMMARY OF PRODUCT CHARACTERISTICS

Effect of CYP2C9*3 mutant variants on meloxicam pharmacokinetics in a healthy Chinese population

Principles of Antimicrobial therapy

Summary of Product Characteristics

Tolerance and safety of enalapril

SUMMARY OF PRODUCT CHARACTERISTICS

Irish Medicines Board

PHARMA SCIENCE MONITOR AN INTERNATIONAL JOURNAL OF PHARMACEUTICAL SCIENCES PHARMACOKINETIC INTERACTION OF MOXIFLOXACIN AND

SUMMARY OF PRODUCT CHARACTERISTICS. KELAPRIL 2.5 mg, film coated tablets for dogs and cats [FR] KELAPRIL 2,5 film coated tablets for dogs and cats

Meloxicam withdrawal time veterinarian bovine

Dr. Omar S. Tabbouche, M.Sc, D.Sc, Pharm.D Head of Pharmacy Department New Mazloum Hospital Tripoli, Lebanon

Specific and Simple HPLC Assay of Ecofriendly Meloxicam in Pharmaceutical Formulations K.T. Mahmood 1, B.Khan 2, M. Ashraf 3 and I. U.

NSAIDs: the Past, Present, and Future

PHARMACOLOGY OF MELOXICAM, A NEW NON-STEROIDAL ANTI-INFLAMMATORY DRUG WITH AN IMPROVED SAFETY PROFILE THROUGH PREFERENTIAL INfflBITION OF COX-2

SUMMARY OF PRODUCT CHARACTERISTICS

ANNEX I SUMMARY OF PRODUCT CHARACTERISTICS

WITHDRAWAL TIME RECOMMENDATIONS ARIZONA RACE TRACKS MEET

COMMITTEE FOR MEDICINAL PRODUCTS FOR VETERINARY USE

ANNEX I SUMMARY OF PRODUCT CHARACTERISTICS 1/33

Is Robenacoxib Superior to Meloxicam in Improving Patient Comfort in Dog Diagnosed With a Degenerative Joint Process?

Pain Management in Racing Greyhounds

ANNEX I SUMMARY OF PRODUCT CHARACTERISTICS

Faculty of Veterinary Science, Szent Istvan University Department of Pharmacology and Toxicology. Nora Line

Scientific Discussion post-authorisation update for Rheumocam extension X/007

VOL. XXIII NO. II THE JOURNAL OF ANTIBIOTICS 559. ANTIBIOTIC 6640.* Ill

ANNEX I SUMMARY OF PRODUCT CHARACTERISTICS

Fluralaner (mg) for small cats kg for medium-sized cats > kg for large cats > kg 1.

SUMMARY OF PRODUCT CHARACTERISTICS

Providing Constant Analgesia with OROS Ò Hydromorphone

Part 1 : General multiple-choice questions

Assessment of gentamiciminduced nephrotoxicity in rats treated with low doses of ibuprofen and diclofenac sodium

SPECTROPHOTOMETRIC ESTIMATION OF MELOXICAM IN BULK AND ITS PHARMACEUTICAL FORMULATIONS

VETU (Derwent Veterinary Drug File)

Procedure # IBT IACUC Approval: December 11, 2017

Non-steroidal anti-inflammatory drugs in dogs and cats what s new?

Commonly Used Analgesics

ABSTRACT. The effect of exogenous oxytocin injected on different days of estrus cycle and pregnancy on uterine PGF 2a

ETX2514SUL (sulbactam/etx2514) for the treatment of Acinetobacter baumannii infections

Withdrawal period: 93 days Milk: Not authorised for use in animals producing milk for human consumption.

single intravenous and oral doses and after 14 repeated oral

The Institutional Animal Care and Use Committee (IACUC) Aquatic Animals: Analgesia and Anesthesia formulary

Welcome to. Who Wants to be a Millionaire 50:50

Voriconazole Increases while Itraconazole Decreases Plasma Meloxicam Concentrations

European Public MRL assessment report (EPMAR)

Principles of Anti-Microbial Therapy Assistant Professor Naza M. Ali. Lec 1

Critical appraisal Randomised controlled trial questions

SCIENTIFIC DISCUSSION

Scientific discussion

Synopsis. Takeda Pharmaceutical Company Limited Name of the finished product UNISIA Combination Tablets LD, UNISIA Combination Tablets

Bioconversion of meloxicam by bacteria

Transcription:

0022-3565/99/2901-0276$03.00/0 THE JOURNAL OF PHARMACOLOGY AND EXPERIMENTAL THERAPEUTICS Vol. 290, No. 1 Copyright 1999 by The American Society for Pharmacology and Experimental Therapeutics Printed in U.S.A. JPET 290:276 280, 1999 Dose-Dependent Inhibition of Platelet Cyclooxygenase-1 and Monocyte Cyclooxygenase-2 by Meloxicam in Healthy Subjects 1 MARIA R. PANARA, GIULIA RENDA, MARIA G. SCIULLI, GIOVANNA SANTINI, MARIA DI GIAMBERARDINO, MARIA T. ROTONDO, STEFANIA TACCONELLI, FRANCESCA SETA, CARLO PATRONO, and PAOLA PATRIGNANI Department of Medicine and Aging, Division of Pharmacology, University of Chieti G. D Annunzio School of Medicine, Chieti, Italy Accepted for publication March 5, 1999 This paper is available online at http://www.jpet.org ABSTRACT We evaluated whether therapeutic blood levels of meloxicam are associated with selective inhibition of monocyte cyclooxygenase (COX)-2 in vitro and ex vivo. Concentration-response curves for the inhibition of monocyte COX-2 and platelet COX-1 were obtained in vitro after the incubation of meloxicam with whole blood samples. Moreover, 11 healthy volunteers received placebo or 7.5 or 15 mg/day meloxicam, each treatment for 7 consecutive days, according to a randomized, doubleblind, crossover design. Before dosing and 24 h after the seventh dose of each regimen, heparinized whole blood samples were incubated with lipopolysaccharide (10 g/ml) for 24 h at 37 C, and prostaglandin E 2 was measured in plasma as an index of monocyte COX-2 activity. The production of thromboxane B 2 in whole blood allowed to clot at 37 C for 60 min The adverse effects of nonsteroidal anti-inflammatory drugs (NSAIDs) on the upper gastrointestinal tract of humans have been documented for nearly 50 years (Douthwaite and Lintott, 1938). The gastroduodenal complications range from dyspepsia to fatal upper gastrointestinal tract bleeding and perforation (Allison et al., 1992). In 1971, Vane proposed that the inhibition of prostaglandin (PG) formation by NSAIDs is the basis for their therapeutic actions as well as for their side effects. This theory has gained general acceptance. The key enzyme in the synthesis of prostanoids is prostaglandin endoperoxide synthase (PGHS), which is endowed with two distinct catalytic activities: a cyclooxygenase (COX) activity converts arachidonic acid (AA) to the endoperoxide PGG 2 and a peroxidase activity converts PGG 2 to PGH 2 (Smith et al., 1996). PGH 2 is further metabolized by specific synthases and isomerases to various prostanoids. Except for aspirin that inhibits prostanoid biosynthesis by Received for publication November 30, 1998. 1 This work was supported by a grant from Boehringer Ingelheim Italia spa. was assessed as an index of platelet COX-1 activity. The administration of placebo did not significantly affect plasma prostaglandin E 2 (21.3 7.5 versus 19.1 4 ng/ml, mean S.D., n 11) or serum thromboxane B 2 (426 167 versus 425 150 ng/ml) levels. In contrast, the administration of 7.5 and 15 mg of meloxicam caused dose-dependent reductions in monocyte COX-2 activity by 51% and 70%, respectively, and in platelet COX-1 activity by 25% and 35%, respectively. Although the IC 50 value of meloxicam for inhibition of COX-1 was 10-fold higher than the IC 50 value of COX-2 in vitro, this biochemical selectivity was inadequate to clearly separate the effects of meloxicam on the two isozymes after oral dosing as a function of the daily dose and interindividual variation in steady-state plasma levels. irreversible blockade of the COX channel of PGHS, the other NSAIDs, such as ibuprofen and indomethacin, produce reversible or time-dependent irreversible COX inhibition by competing with the substrate, AA, for a common binding site (Smith et al., 1996). Recent epidemiological studies have shown that comparable therapeutic doses of NSAIDs are associated with a similar risk of serious gastrointestinal bleeding complications (Langman et al., 1994; Henry et al., 1996). The discovery of two isoforms of PGHS (reviewed by Smith et al., 1996), a constitutive PGHS-1 or COX-1 present in almost all cell types, and a form induced in a more restricted cell-specific fashion by mitogenic and inflammatory stimuli, PGHS-2 or COX-2, has led to the suggestion that the anti-inflammatory action of NSAIDs is due to inhibition of COX-2, whereas the toxic effects on the stomach and bleeding complications are due to inhibition of COX-1 (reviewed by Vane, 1994). Because COX-1 and COX-2 are structurally distinct proteins with only 60% homology (Smith et al., 1996), the development of drugs that selectively inhibit the activity of COX-2 might lead to a new generation of anti-inflammatory drugs ABBREVIATIONS: PGHS, prostaglandin endoperoxide synthase; COX, cyclooxygenase; LPS, lipopolysaccharide; AA, arachidonic acid; PGE 2, prostaglandin E 2 ; TXB 2, thromboxane B 2 ; NSAID, nonsteroidal anti-inflammatory drug. 276

1999 Inhibition of Cyclooxygenase-1 and -2 by Meloxicam 277 with improved gastrointestinal safety (Vane, 1994). A number of compounds have been described that selectively inhibit COX-2 versus COX-1; these include substances that were initially selected for development by drug companies based on an improved pharmacological profile in animal models and were later shown to preferentially inhibit COX-2, as well as newly designed specific COX-2 inhibitors. The former group of compounds includes meloxicam and nimesulide. Meloxicam, a NSAID derived from enolic acid, has shown anti-inflammatory activity in a rat model of adjuvant arthritis at doses 20-fold lower than those causing ulcerogenic effects (Engelhardt et al., 1995). In vitro, meloxicam was 3- to 300-fold more potent in inhibiting COX-2 than COX-1 (Churchill et al., 1996; Engelhardt et al., 1996; Patrignani et al., 1997; Riendeau et al., 1997; Pairet et al., 1998). However, major limitations in expressing the degree of biochemical selectivity based on COX-1/COX-2 IC 50 ratios obtained in vitro are related to variable results obtained with different assay methods, on the one hand, and to the lack of information on the actual extent of COX-1 inhibition achieved at therapeutic plasma levels of the inhibitor, on the other hand. The induction of COX-2 in circulating monocytes in response to lipopolysaccharide (LPS) is a relatively simple model that is suitable for evaluating the extent of COX-2 inhibition both in vitro (Patrignani et al., 1994, 1997; Panara et al. 1995) and ex vivo after oral dosing of NSAIDs in humans (Patrignani et al., 1994; Cipollone et al., 1995; Panara et al., 1998). Thus, the aim of our study was to test the biochemical selectivity of meloxicam toward monocyte COX-2 versus platelet COX-1 in vitro as well as ex vivo after the oral administration of two therapeutic doses (7.5 and 15 mg/day for 7 days) versus placebo in healthy subjects. Materials and Methods In Vitro Study. The effects of meloxicam on platelet COX-1 and monocyte COX-2 activities of whole blood were assessed by incubating increasing concentrations of meloxicam with peripheral whole blood samples drawn from four healthy volunteers (two female and two male subjects; age range, 25 42 years). Ex Vivo Study. Fourteen healthy volunteers who gave their informed consent were screened for the inclusion and exclusion criteria of the study. One volunteer was excluded due to abnormal laboratory values, and another volunteer dropped out because of an adverse event (the onset of influenza) that required antibiotic treatment. Twelve healthy volunteers (seven female and five male subjects; age range, 20 41 years) were included in the study. One subject completed the treatments but was excluded from the final evaluation of the results because naproxen was detected in his plasma samples and he subsequently admitted taking the nonstudy drug. Placebo and 7.5 and 15 mg of meloxicam were administered over 6 weeks according to a randomized, double-blind, crossover design. Each treatment was administered for 7 days. After completion of each treatment period, there was a wash-out period of 7 days. Peripheral venous blood samples were drawn before starting the treatment and 24 h after the seventh dose of each regimen for the assessment of whole blood PGE 2 and TXB 2 production and the measurement of meloxicam plasma levels. COX-2 Assay. Aliquots of peripheral blood samples (1 ml) containing 10 IU of sodium heparin were incubated in both the absence and the presence of LPS (10 g/ml) for 24 h at 37 C as described in detail elsewhere (Patrignani et al., 1994). Plasma was separated by centrifugation (10 min at 2000 rpm) and kept at 80 C until assayed for PGE 2. COX-1 Assay. Peripheral blood samples were drawn from the same subjects, and 1-ml aliquots of whole blood were immediately transferred into glass tubes and allowed to clot at 37 C for 60 min. Serum was separated by centrifugation (10 min at 3000 rpm) and kept at 30 C until assayed for TXB 2, as previously described (Patrono et al., 1980; Patrignani et al., 1982). Effects of Meloxicam on Monocyte COX-2 Activity and Platelet COX-1 Activity In Vitro. Meloxicam was dissolved in dimethyl sulfoxide (10 10,000 g/ml), and 2- l aliquots of the solutions were transferred directly by means of a pipette into test tubes to give final concentrations of 0.02 to 20 g/ml. Heparinized 1-ml whole blood samples were drawn from healthy volunteers who received pretreatment with 300 mg of aspirin 48 h before sampling to suppress the activity of platelet COX-1. These samples were incubated at 37 C for 24 h with increasing concentrations of meloxicam in the presence of 10 g/ml LPS, and PGE 2 levels were assayed in plasma as an index of monocyte COX-2 activity. Meloxicam was also incubated with 1-ml whole blood samples (drawn from the same donors when they had not taken any NSAIDs during the 2 weeks before the study) that were allowed to clot for1hat37 C, and serum TXB 2 levels were measured as a reflection of platelet COX-1 activity. Analyses of PGE 2 and TXB 2. PGE 2 and TXB 2 were measured by previously described and validated radioimmunoassays (Ciabattoni et al., 1979; Patrono et al., 1980; Patrignani et al., 1982). Unextracted plasma and serum samples were diluted in the standard diluent of the assay (0.02 M phosphate buffer, ph 7.4) and assayed in a volume of 1.5 ml at a final dilution of 1:50 to 1:20,000. The least detectable concentration was 1 to 2 pg/ml for both prostanoids. Plasma Meloxicam Concentrations. Plasma concentrations of meloxicam were determined by HPLC as described previously (Turck et al., 1997). Reagents. [ 3 H]PGE 2 and [ 3 H]TXB 2 (200 250 Ci/mmol) were from Du Pont de Nemours GmbH (Bad Homburg, Germany). Authentic PGE 2 and TXB 2 were from Cayman Chemical Co. (Ann Arbor, MI). Anti-PGE 2 and anti-txb 2 sera were obtained in our laboratory, and their characteristics have been described previously (Ciabattoni et al., 1979; Patrono et al., 1980). Heparin, LPS derived from Escherichia coli 026:B6, acetylsalicylic acid, and dimethyl sulfoxide were purchased from Sigma Chemical Co. (St. Louis, MO). Statistical Analysis. Statistical comparisons were made by ANOVA, and significant differences between treatments were determined by Student-Newman-Keuls test. Having excluded a carry-over effect (by comparing the inhibition of COX-1 and COX-2 activities and meloxicam plasma levels measured after a 1-week treatment with 7.5 mg of meloxicam given first with measurements obtained when 7.5 mg of meloxicam was given after the 15-mg dose or placebo), the data were combined regardless of the treatment sequence. The post-treatment measurements were compared with the basal measurements and values of p.05 were considered statistically significant. The sigmoidal dose-response curves were analyzed with ALLFIT, a basic computer program for simultaneous curve-fitting based on a four-parameter logistic equation (De Lean et al., 1978). Results We measured PGE 2 production in heparinized whole blood samples, incubated with LPS for 24 h, as a reflection of the inducible COX activity of monocyte COX-2 (Patrignani et al., 1994). Plasma immunoreactive PGE 2 averaged 21.3 7.5 ng/ml (mean S.D., n 11) in samples obtained before treatment. Moreover, we measured TXB 2 production during whole blood clotting as a reflection of the constitutive COX activity of platelet COX-1 (Patrono et al., 1980; Patrignani et al., 1982). Serum immunoreactive TXB 2 averaged 426 167 ng/ml (mean S.D., n 11) at baseline. We also assessed the intrasubject variability of these indexes of COX-1 and COX-2 activity on three different occasions over a 14- to 42-day

278 Panara et al. Vol. 290 period. The intrasubject coefficients of variation averaged 16 9% and 23 11% for repeated measurements of serum TXB 2 and plasma PGE 2, respectively. Of the 12 healthy volunteers who entered a randomized, double-blind, three-period crossover trial of placebo and 7.5 and 15 mg/day meloxicam, 11 complied with the protocol for efficacy analysis. As shown in Fig. 1, the administration of placebo did not affect PGE 2 or TXB 2 production to any statistically significant extent (PGE 2, 19.1 4 ng/ml; TXB 2, 425 150 ng/ml). The administration of 7.5 and 15 mg/day meloxicam caused a statistically significant (p.01) dosedependent reduction in both monocyte COX-2 activity by 51% and 70%, respectively, and in platelet COX-1 activity by 25% and 35%, respectively. However, the inhibition of monocyte COX-2 activity was significantly (p.025 and.002, respectively) higher than that of platelet COX-1 activity at both dose levels. Plasma concentrations of meloxicam measured 24 h after the oral administration of 7.5 and 15 mg of meloxicam for 7 days averaged 0.69 0.38 and 1.48 0.84 g/ml, respectively (mean S.D., n 11, p.01). To explore the meloxicam concentration-response relationship for inhibition of COX isozyme activity, plasma drug concentrations were corrected for the individual hematocrit values. As shown in Fig. 2, the relationship between meloxicam concentrations added in vitro and percentage inhibition of COX activities fitted Fig. 2. Comparison of meloxicam concentrations and inhibition of monocyte COX-2 and platelet COX-1 activities assessed in vitro and ex vivo. In the in vitro study, increasing concentrations of meloxicam (0.02 20 g/ ml) were incubated with 1 ml of heparinized whole blood samples (drawn from healthy volunteers pretreated with 300 mg of aspirin 48 h before sampling to suppress the activity of platelet COX-1) in the presence of 10 g/ml LPS for 24 h and plasma PGE 2 levels were assayed as an index of monocyte COX-2 activity (Œ). Meloxicam was also incubated with 1-ml whole blood samples (drawn from the same donors when they had not taken any NSAIDs during the 2 weeks before the study) that were allowed to clot for 60 min, and serum TXB 2 levels were measured as an index of platelet COX-1 activity (F). Mean S.E.M. values of COX-1 and COX-2 inhibition from four different experiments are depicted. Sigmoidal concentration-response curves fitting the experimental data were generated by ALLFIT analysis. In the ex vivo study, steady-state meloxicam whole blood concentrations measured in 11 healthy subjects 24 h after dosing with 7.5 and 15 mg/day are plotted on the same graph versus the percentage inhibition of monocyte COX-2 ( ) and platelet COX-1 (E) activity evaluated ex vivo at the same time. sigmoidal dose-response curves. Using ALLFIT, we determined IC 50 values for monocyte COX-2 and platelet COX-1 inhibition in vitro of 0.17 0.09 and 1.94 0.58 g/ml (mean S.E.M., n 4), respectively, that were different from each other in a statistically significant fashion (p.024). When individual drug levels and corresponding COX isozyme inhibition measured ex vivo are fitted onto these sigmoidal dose-response curves obtained in vitro, it becomes apparent that there is a substantial degree of overlapping of COX-2 and COX-1 inhibition within an approximately 10- fold range of steady-state trough meloxicam plasma levels. However, although individual ex vivo measurements of COX-2 inhibition tended to distribute fairly evenly on both sides of the COX-2 concentration-response curve obtained in vitro, measurements of COX-1 inhibition were unevenly distributed to the left of the COX-1 concentration-response curve. Four subjects, in the 7.5 mg meloxicam phase, and two of the same four subjects, in the 15 mg meloxicam phase of the study, failed to show any detectable inhibition of COX activities, despite drug plasma levels in the therapeutic range. Fig. 1. Effects of 1-week administration of meloxicam (7.5 and 15 mg/day) and placebo on COX isozyme activities ex vivo. Eleven healthy volunteers received placebo or 7.5 and 15 mg/day meloxicam, each treatment for 7 consecutive days, over 6 weeks according to a double-blind, crossover design. After completion of each treatment period, there was a wash-out period of 7 days. Peripheral venous blood samples were drawn before and 24 h after the seventh dose of each treatment for the assessment of PGE 2 production in heparinized whole blood incubated for 24 h with LPS (10 g/ml) (an index of the COX activity of monocyte COX-2) in the presence of 50 g/ml acetylsalicylic acid to suppress the activity of platelet COX-1 (A) and TXB 2 production in whole blood allowed to clot for 60 min (an index of the COX activity of platelet COX-1) (B). Measurements are expressed as mean 1 SD values. **p.01 for the comparison of post-treatment versus basal and placebo values. Discussion A variety of methods have been developed to study the effects of NSAIDs on the COX activity of PGHS isozymes in vitro. Thus, inhibition of COX-1 and COX-2 has been evaluated in purified enzyme systems (Mitchell et al., 1993), microsomal membranes from cells transfected with the murine (Meade et al., 1993) or human isozymes (Laneuville et al., 1994), cultured cells selectively expressing COX-1 and COX-2 as well as in cultured intact cells, such as bovine aortic endothelial cells and endotoxin-activated macrophages (Mitchell et al., 1993). The affinity of many COX inhibitors

1999 Inhibition of Cyclooxygenase-1 and -2 by Meloxicam 279 for COX-1 decreases in the presence of micromolar concentrations of exogenous arachidonate, whereas the affinity for COX-2 is unaffected by changes in substrate concentration (Swinney et al., 1997). This results in dramatic differences in COX-1/COX-2 selectivity of NSAIDs as a function of the arachidonate concentration employed. Therefore, measuring eicosanoid production in intact cells that selectively express COX-1 or COX-2 and utilizing arachidonate released from endogenous lipid stores is considered to be a more reliable screening method than measuring instantaneous inhibition of isozyme activity in a recombinant system utilizing exogenous substrate (Meade et al., 1993; Laneuville et al., 1994). We described a relatively simple model of human COX-2 expression in LPS-stimulated whole blood (Patrignani et al., 1994). This method is suitable for evaluating the extent of COX-2 inhibition both in vitro and ex vivo, after oral dosing of NSAIDs in humans. We demonstrated that LPS dosedependently stimulated blood cells to produce easily detectable amounts of PGE 2 in a time-dependent fashion that correlated with the accumulation of a protein doublet of approximately 72 kda in monocytes but not in other blood cells (Patrignani et al., 1994). In contrast, the constitutive level of expression of COX-1 in unstimulated monocytes was not affected by the incubation with LPS up to 24 h (Patrignani et al., 1994). Dexamethasone, an inhibitor of COX-2 expression in monocytes/macrophages (O Banion et al., 1992), as well as several highly selective COX-2 inhibitors, largely suppressed PGE 2 production in response to LPS (Patrignani et al., 1994, 1997; Panara et al., 1995). For the assessment of COX-1/COX-2 selectivity, this whole blood assay of COX-2 activity was integrated with a previously described whole blood assay for platelet COX activity (Patrono et al., 1980; Patrignani et al., 1982). The latter has been extensively used for characterizing the time and dose dependence of platelet COX inhibition by aspirin in humans (Patrignani et al., 1982; Patrono et al., 1985). This methodological approach has allowed the characterization of the biochemical selectivity of recently approved NSAIDs, including nabumetone (Patrignani et al., 1994; Cipollone et al., 1995) and nimesulide (Panara et al., 1998). Interestingly, the inhibitory effects of a large number of NSAIDs on human gastric PGE 2 synthesis correlated with COX-1 inhibitory potency in whole blood (Cryer and Feldman, 1998). Meloxicam is a novel NSAID derived from enolic acid. The drug shows prolonged and almost complete absorption after oral dosing and is 99.5% bound to plasma proteins (Turck et al., 1996). Meloxicam is metabolized to four biologically inactive main metabolites, which are excreted in both urine and feces (Turck et al., 1996). The elimination half-life of meloxicam is approximately 20 h, and steady-state plasma concentrations are achieved within 3 to 5 days (Turck et al., 1996). In rats, meloxicam inhibited the hind paw swelling to adjuvant injection with an ID 50 value that was approximately 20-fold lower than the ED 50 value for its ulcerogenic effect on the stomach (Engelhardt et al., 1995). In vitro, in different assays performed in the presence of exogenous arachidonate, meloxicam showed variable COX-1/COX-2 IC 50 ratios that ranged from 3, in guinea pig macrophages (Engelhardt et al., 1996), to 300, in Chinese hamster ovary cells stably transfected with human COX-1 and COX-2 (Riendeau et al., 1997). In the present study, we set out to reassess the biochemical selectivity of meloxicam in vitro, to evaluate the dose dependence and selectivity of monocyte COX-2 inhibition associated with two commonly used therapeutic regimens of meloxicam, and to relate the extent of COX inhibition to steady-state plasma levels of the drug administered to healthy volunteers. When evaluated in vitro, the IC 50 value for meloxicam for platelet COX-1 inhibition was 1 order of magnitude higher than the corresponding IC 50 value for monocyte COX-2. However, this modest degree of biochemical selectivity was inadequate to clearly separate the effects of meloxicam on the two isozymes after oral dosing in healthy subjects as a function of the daily dose and interindividual variation in steady-state trough levels of the drug (Fig. 2). There is no obvious explanation for the consistently higher level of inhibition of platelet COX-1 detected ex vivo than that predicted by the in vitro concentration-response relationship. Because we did not perform this type of assessment after single oral dosing, we cannot exclude some degree of cumulative inhibition of platelet COX-1 by meloxicam on repeated daily dosing. The present findings underscore the limitations of assessing the biochemical selectivity of COX-2 inhibitors based on COX-1/ COX-2 IC 50 ratios obtained in vitro and emphasize the need to evaluate the actual extent of COX-1 inhibition achieved throughout the therapeutic range of plasma inhibitor concentrations after oral dosing in humans. Thus, 24 h after the seventh daily administration of 7.5 and 15 mg of meloxicam, both platelet COX-1 and monocyte COX-2 activities were inhibited in a dose-dependent fashion. However, the inhibition of monocyte COX-2 activity was significantly higher than that of platelet COX-1 activity at both dose levels. Serum TXB 2 was only marginally, although significantly, reduced by the administration of 7.5 mg of meloxicam. This 25% reduction in COX-1 activity was slightly higher than the intrasubject coefficient of variation of this biochemical index based on repeated measurements of serum TXB 2 over a 2- to 6-week period. In contrast, Stichtenoth et al. (1997) reported that the administration of 7.5 mg of meloxicam to healthy subjects for 6 days did not significantly affect platelet aggregation and TXB 2 production in plateletrich plasma in response to 1 mm AA. This discrepancy is likely due to a reduction in the affinity of meloxicam for the COX-1 active site in the presence of very high concentrations of exogenous AA. Moreover, it should be emphasized that assessment of platelet COX-1 and monocyte COX-2 inhibition at meloxicam trough levels will underestimate maximal inhibition of the enzymes at C max when plasma levels of the drug are approximately 2.4-fold higher. Information about the gastrointestinal safety of meloxicam and other NSAIDs in clinical use has been obtained from a multinational meloxicam clinical trial program conducted in rheumatoid arthritis (RA), osteoarthritis (OA), and other rheumatic diseases (Distel et al., 1996; Dequeker et al., 1998; Hawkey et al., 1998). In double-blind studies in RA and OA, 7.5 and 15 mg of meloxicam caused significantly (p.05) less gastrointestinal adverse events (serious and nonserious) than 20 mg of piroxicam, 100 mg of slow-release diclofenac and 750 to 1000 mg of naproxen. Although both 7.5 and 15 mg/day regimens of meloxicam showed a detectable advantage over other NSAIDs in their gastrointestinal safety profile, there was a trend toward a dose-effect relationship with

280 Panara et al. Vol. 290 respect to gastrointestinal side effects (Distel et al., 1996) that needs to be confirmed in larger studies. It should be emphasized that none of the published studies had the statistical power to detect realistic differences in the risk of the most serious upper gastrointestinal bleeding complications between meloxicam and other NSAIDs, given the very low incidence of such complications. Factors other than COX-1/COX-2 selectivity, such as the daily dose and pharmacokinetic profile, are likely to represent the major determinants of the occurrence of serious adverse events associated with the administration of NSAIDs with relatively narrow biochemical selectivity (COX- 1/COX-2 ratios falling within a range of 0.5 20). Recent epidemiological observations are consistent with this hypothesis by reporting a very similar relative risk of serious gastrointestinal bleeding complications associated with nimesulide (COX-1/COX-2 IC 50 ratio 18; Panara et al., 1998) as with conventional NSAIDs (Garcia Rodriguez et al., 1998). We conclude that meloxicam is a more potent inhibitor of monocyte COX-2 than platelet COX-1 activity; however, the extent of COX-1 sparing is largely a function of the daily dose and interindividual variability in drug levels. Whether this modest degree of biochemical selectivity is responsible for an improved gastrointestinal safety profile remains to be determined in population-based observational studies with hospitalization due to upper gastrointestinal bleeding as the primary outcome measure. Acknowledgments We thank the medical students of the University of Chieti G. D Annunzio School of Medicine for their generous cooperation throughout the studies, Dr. Michel Pairet (Boehringer Ingelheim Pharma KG, Germany) for helpful discussions, Dr. Dietrich Turck (Boehringer Ingelheim Pharma KG, Germany) for the measurement of meloxicam plasma levels, and Daniela Basilico for expert editorial assistance. References Allison MC, Howatson AG, Torrance CJ, Lee FD and Russell R (1992) Gastrointestinal damage associated with the use of nonsteroidal antiinflammatory drugs. N Engl J Med 327:749 754. Churchill L, Graham AG, Shih CK, Pauletti D, Farina PR and Grob PM (1996) Selective inhibition of human cyclo-oxygenase-2 by meloxicam. Inflammopharmacology 4:125 135. Ciabattoni G, Pugliese F, Spaldi M, Cinotti GA and Patrono C (1979) Radioimmunological measurement of prostaglandins E 2 and F 2 in human urine. J Endocrinol Invest 2:173 182. Cipollone F, Ganci A, Panara MR, Greco A, Cuccurullo F, Patrono C and Patrignani P (1995) Effects of nabumetone on prostanoid biosynthesis in humans. Clin Pharmacol Ther 58:335 341. Cryer B and Feldman M (1998) Cyclooxygenase-1 and cyclooxygenase-2 selectivity of widely used non steroidal anti-inflammatory drugs. Am J Med 104:413 421. De Lean A, Munson PJ and Rodbard D (1978) Simultaneous analysis of families of sigmoidal curves: Application to bioassay, radioligand assay and physiological dose-response curves. Am J Physiol 235:E97 E102. Dequeker J, Hawkey C, Kahan A, Steinbruck K, Alegre C, Baumelou E, Bégaud B, Isomaki H, Littlejohn G, Mau J and Papazoglou S, on behalf of the SELECT study group (1998) Improvement in gastrointestinal tolerability of the selective COX-2 inhibitor, meloxicam, compared with piroxicam: Results of the Safety and Efficacy Large Scale Evaluation of COX Inhibiting Therapies (SELECT) trial in osteoarthritis. Br J Rheumatol 37:946 951. Distel M, Mueller C, Bluhmki E and Fries J (1996) Safety of meloxicam: A global analysis of clinical trials. Br J Rheumatol 35(Suppl 1):68 77. Douthwaite AH and Lintott GAM (1938) Gastroscopic observation of the effect of aspirin and certain other substances on the stomach. Lancet 2:1222 1224. Engelhardt G, Bogel R, Schnitzler C and Utzmann R (1996) Meloxicam: Influence on arachidonic acid metabolism (part I, in vitro findings). Biochem Pharmacol 51:21 28. Engelhardt G, Homma D, Schlegel K, Utzmann R and Schnitzler C (1995) Antiinflammatory, analgesic, antipyretic and related properties of meloxicam, a new non-steroidal anti-inflammatory agent with favorable gastrointestinal tolerance. Inflamm Res 44:423 433. Garcia Rodriguez LA, Cattaruzzi C, Troncon MG and Agostinis L (1998) Risk of hospitalization for upper gastrointestinal bleeding associated with ketorolac, other NSAIDs, calcium antagonists and other antihypertensive drugs. Arch Intern Med 158:33 39. Hawkey C, Kahan A, Steinbruck K, Alegre C, Baumelou E, Bégaud B, Dequeker J, Isomaki H, Littlejohn G, Mau J, Papazoglou S, and the International Melissa Study Group (1998) Gastrointestinal tolerability of meloxicam compared to diclofenac in osteoarthritis patients. Br J Rheumatol 37:937 945. Henry D, Lim L L-Y, Garcia Rodriguez LA, Perez Gutthann S, Carson JL, Griffin M, Savage R, Logan R, Moride Y, Hawkey C, Hill S and Fries JT (1996) Variability in risk of gastrointestinal complications with individual non-steroidal antiinflammatory drugs: Results of a collaborative meta-analysis. BMJ 312:1563 1566. Laneuville O, Breuer DK, De Witt DL, Funk CD and Smith WL (1994) Differential inhibition of human prostaglandin endoperoxide H synthase-1 and -2 by nonsteroidal anti-inflammatory drugs. J Pharmacol Exp Ther 271:927 934. Langman MJS, Well J, Wainwright P, Lawson DH, Rawlins MD, Logan RF, Murphy M, Vessey MP and Colin-Jones DG (1994) Risks of bleeding peptic ulcer associated with individual non-steroidal anti-inflammatory drugs. Lancet 343:1075 1078. Meade EA, Smith WL and DeWitt DL (1993) Differential inhibition of prostaglandin endoperoxide synthase (cyclooxygenase) isozymes by aspirin and other nonsteroidal anti-inflammatory drugs. J Biol Chem 268:6610 6614. Mitchell JA, Akarascereenont P, Theimermann C, Flower RJ and Vane JR (1993) Selectivity of nonsteroidal antiinflammatory drugs as inhibitors of constitutive and inducible cyclooxygenase. Proc Natl Acad Sci USA 90:11693 11697. O Banion MK, Winn VD and Yung DA (1992) cdna cloning and functional activity of a glucocorticoid-regulated inflammatory cyclooxygenase. Proc Natl Acad Aci USA 89:4888 4892. Pairet M, van Ryn H, Schierok H, Mauz A, Trummlitz G and Engelhardt G (1998) Differential inhibition of cyclooxygenase-1 and -2 by meloxicam and 4 -isomer. Inflamm Res 47:270 276. Panara MR, Greco A, Santini G, Sciulli MG, Rotondo MT, Padovano R, di Giamberardino M, Cipollone F, Cuccurullo F, Patrono C and Patrignani P (1995) Effects of the novel anti-inflammatory compounds, N-[2-(cyclohexyloxy)-4-nitrophenyl] methane-sulphonamide (NS-398) and 5-methanesulphonamido-6-(2,4-difluorothio-phenyl)-1-indanone (L-745,337), on the cyclo-oxygenase activity of human blood prostaglandin endoperoxide synthases. Br J Pharmacol 116:2429 2434. Panara MR, Padovano R, Sciulli MG, Santini G, Renda G, Rotondo MT, Pace A, Patrono C and Patrignani P (1998) Effects of nimesulide on constitutive and inducible prostanoid biosynthesis in man. Clin Pharmacol Ther 63:672 681. Patrignani P, Filabozzi P and Patrono C (1982) Selective cumulative inhibition of platelet thromboxane production by low-dose aspirin in healthy subjects. J Clin Invest 69:1366 1372. Patrignani P, Panara MR, Greco A, Fusco O, Natoli C, Iacobelli S, Cipollone F, Ganci A, Créminon C, Maclouf J and Patrono C (1994) Biochemical and pharmacological characterization of the cyclooxygenase activity of human whole blood prostaglandin endoperoxide synthases. J Pharmacol Exp Ther 271:1705 1712. Patrignani P, Panara MR, Sciulli MG, Santini G, Renda G and Patrono C (1997) Differential inhibition of human prostaglandin endoperoxide synthase-1 and -2 by nonsteroidal anti-inflammatory drugs. J Physiol Pharmacol 48:623 631. Patrono C, Ciabattoni G, Pinca E, Pugliese F, Castrucci G, De Salvo A, Satta MA and Peskar BA (1980) Low dose aspirin and inhibition of thromboxane B 2 production in healthy subjects. Thromb Res 17:317 327. Patrono C, Ciabattoni G, Patrignani P, Filabozzi P, Catella F, Davum G and Forni L (1985) Clinical pharmacology of platelet cyclooxygenase inhibition. Circulation 72:1177 1184. Riendeau D, Percival MD, Boyce S, Briedeau C, Charleson S, Cromlish W, Ethier D, Evans J, Falgueyret J-P, Ford-Hutchinson AW, Gordon R, Greig G, Gresser M, Guay J, Kargman S, Léger S, Mancini JA, O Neill G, Ouellet M, Rodger IW, Thérien M, Wang Z, Webb JK, Wong E, Xu L, Young RN, Zamboni R, Prasit P and Chan C-C (1997) Biochemical and pharmacological profile of a tetrasubstituted furanone as a highly selective COX-2 inhibitor. Br J Pharmacol 121:105 117. Smith WL, Garavito RM and DeWitt DL (1996) Prostaglandin endoperoxide H synthases (cyclooxygenase)-1 and -2. J Biol Chem 271:33157 33160. Stichtenoth DO, Wagner B and Frolich JC (1997) Effects of meloxicam and indomethacin on cyclooxygenase pathways in healthy volunteers. J Invest Med 45:44 49. Swinney DC, Mak AY, Barnett J and Ramesha CS (1997) Differential allosteric regulation of prostaglandin H synthase 1 and 2 by arachidonic acid. J Biol Chem 272:12393 12398. Turck D, Roth W and Busch U (1996) A review of the clinical pharmacokinetics of meloxicam. Br J Rheumatol 35 (Suppl 1):13 16. Turck D, Su C, Heinzel G, Busch U, Bluhmki E and Hoffmann J (1997) Lack of interaction between meloxicam and warfarin in healthy volunteers. Eur J Clin Pharmacol 51:421 425. Vane JR (1971) Inhibition of prostaglandins as a mechanism of action for aspirin-like drugs. Nat New Biol 231:232 235. Vane JR (1994) Towards a better aspirin. Nature (London) 367:215 216. Send reprint requests to: Paola Patrignani, Ph.D., Cattedra di Farmacologia I, Dipartimento di Medicina e Scienze dell Invecchiamento, Università di Chieti G. D Annunzio, Via dei Vestini, 31, 66013 Chieti, Italy. E-mail: ppatrignani@unich.it