12

Click here to load reader

Chirality and Nonsteroidal Anti-Inflammatory Drugs

Embed Size (px)

Citation preview

Page 1: Chirality and Nonsteroidal Anti-Inflammatory Drugs

Drugs 1996; 52 Suppl. 5: 47-58 0012-6667/96/=5-0047/$06.00/0

© Adis International limited. All rights reserved.

Chirality and Nonsteroidal Anti-Inflammatory Drugs Peter J. Hayball Pharmacy Department, Repatriation General Hospital, Daw Park, South Australia, Australia

Summary The nonsteroidal anti-inflammatory drugs (NSAIDs) are of significant clinical importance and include congeners of many chemical classes, some of which incorporate an asymmetric or chiral carbon atom_ With very few exceptions, chiral NSAIDs have been marketed for clinical use as racemates_ However, it is apparent that differences, sometimes major, exist between enantiomers in terms of their pharmacological and toxicological properties_ With regard to the ability of chiral NSAIDs to inhibit cyclo-oxygenase, their chief mechanism of action, major or exclusive activity is confined to enantiomers of the S-stereoconfigura­tion. Accordingly, it is questionable whether the R-antipodes should be included in the final drug product for use in the clinic. In addition to differences between enantiomers in terms of their pharmacodynamic properties, pharmacokinetic differences are possible for chiral NSAID isomers, and these may modulate pre­existing enantioselectivities at the site of action of such compounds. As a conse­quence, a considerably simpler pharmacological profile is likely to result from the use of single enantiomers versus racemic mixtures.

Nonsteroidal anti-inflammatory drugs (NSAIDs) are widely prescribed agents for alleviating pain and inflammation associated with tissue injury. NSAIDs are a structurally diverse group of thera­peutic agents and include a subclass of compounds possessing molecular asymmetric or chiral ele­ments. Such chiral NSAIDs are the focus of this review and currently include congeners from 3 distinct chemical groups: (a) the 2-arylpropionic acids (2-APAs) or 'profens'; (b) other arylalkanoic acids (ketorolac, indobufen, etodolac, etc.); and (c) miscellaneous compounds including agents such as oxyphenbutazone and azapropazone. With the exception of sulindac, which contains a chiral sulphur tricoordinate centre, the molecular asym­metry of these chiral NSAIDs results from the presence of a chiral carbon atom, i.e. a carbon co­valently attached to 4 different chemical groups.

Figure 1 depicts this for the most commonly used subcategory of chiral NSAIDs, the 2-APAs, a group that also encompasses the greatest number of clinically used congeners.

It was recognised early in the development of chiral NSAIDs that pharmacological differences existed between the optical isomers or enantiomers of these drugs. In terms of the classically defined pharmacological activity of NSAIDs, mediated via inhibition of cyclo-oxygenase,[ll major or exclu­sive activity was afforded to 2-APA enantiomers of the S-stereoconfigurationPl Indeed, this work prompted the development by Syntex of the highly successful isomerically pure agent, S-naproxen, in the early 1970s. However, enantioselectivity (tech­nically defined below) slid into the unconscious until the mid-1980s, a time that heralded the advent of more widely available enantioselective syn-

Page 2: Chirality and Nonsteroidal Anti-Inflammatory Drugs

48

thetic and enantiospecific bioanalytical methodol­ogies. This apparent neglect of the important role that molecular asymmetry played in defining the pharmacological and toxicological properties of chiral NSAIDs led to the majority of such com­pounds being used clinically as the more readily synthesised racemic mixture containing an equal proportion of the 2 enantiomers.

This review discusses the ramifications associ­ated with the use of chiral NSAIDs, most notably when administered as the racemate. Accordingly, an evaluation of literature pertaining to the phar­macodynamics, pharmacokinetics and toxicology of the individual enantiomers of representative agents is discussed below. Particular emphasis has been placed on the clinical applicability of acquired knowledge on this therapeutically and economically important class of drugs. In terms of the latter, the worldwide market for combined chiral and achiral NSAIDs in 1995 has been esti­mated to exceed £1000 million.!3]

1. Definition of Enantioselectivity

Biologically active agents often show a high de­gree of selectivity as a result of the discriminatory ability of molecular sites of action. This discrimi­natory ability requires complementarity between

R-2-Arylpropionic acid S-2-Arylpropionic acid

Fig. 1. Stereochemical depictions of the R- and 8-enantiomers of congeners of 2-arylpropionic acid. The chiral carbon is indi­cated by an asterisk and all chemical bonds are oriented in the plane of the page unless indicated by a solid or dashed wedge, which denotes substituents lying above and below the plane of the page, respectively. The identity and attachment site of the aryl substituent (X) varies according to the particular 2-APA. For example, in the case of ketoprofen, X is a benzoyl group attached to the 3-position on the aromatic nucleus. The arrows illustrate the unidirectional nature of the metabolic chiral inversion reaction.

© Adis International Limited. All rights reserved.

Hayball

the bioactive agent and the molecular site of action or receptor, most commonly a protein, which in tum is constructed of chiral amino acid subunits.

Inherent enantioselectivity in action can be accounted for on the basis of as few as 3 groups in the active molecule participating in a structure­specific fashion with the effector site (the so-called '3-point' interaction). Accordingly, for chiral drugs, it is hardly surprising to find enantio­selectivity the rule rather than the exception in nature, since the multitude of enantiomerically pure asymmetric interaction sites present in organ­isms make them a veritable 'chiral jungle' with which drugs must come into contact.

Differences between enantiomers in the phar­macological response to a given dose may arise from differences in drug delivery to the receptor site(s) [enantioselective pharmacokinetics], most notably in terms of unbound drug (drug not bound to protein). Alternatively, differences may exist between enantiomers in terms of their relative complementarity and subsequent activity at effector site( s) [enantioselecti ve pharmacodynamics]. Both enantioselective pharmacodynamics and pharma­cokinetics are important in describing the influence of optical isomerism on clinical pharmacology; the relative contribution of each is a complex function of many factors. The term 'enantiospecific' has usually been used in the literature to confer exclu­sivity to one enantiomer over its optical antipode, i.e. total enantioselectivity. Typically, the synthesis of enantiomers of high optical purity and exclusive activity of one enantiomer (the other isomer being devoid of activity) are referred to as enantio­specific processes.

2. Enantioselective Pharmacodynamics of Chiral NSAIDs

It has been generally thought that NSAIDs exert their predominant pharmacological effects by in­hibiting the synthesis of prostaglandins (mediators of pain and inflammation) at one or more steps in the endoperoxide biosynthetic pathway.!4] A variety of enzymes catalyse the conversion of cell membrane-derived arachidonic acid to prosta-

Drugs 1996; 52 Suppl. 5

Page 3: Chirality and Nonsteroidal Anti-Inflammatory Drugs

Chirality and NSAIDs 49

Table I. Comparative in vitro data on the potencies olthe separate enantiomers and racemic forms of the 2-arylpropionic acid NSAIDs derived from human eicosanoid-dependent test systems

Chiral NSAID Test system SIR potency ratio RSIS potency ratio Reference

Carprofen Inhibition of platelet aggregation > 24 0.58 9

Fenoprofen Inhibition of platelet aggregation 35 0.50 10

Ibuprofen Inhibition of platelet thromboxane production 0.50 11

Indobufen Inhibition of platelet thromboxane production 16 0.50 12

Ketoprofen Inhibition of platelet thromboxane production 0.50 13

Ketoprofen Inhibition of polymorph thromboxane production 3000 0.50 14

a After accounting for optical contamination in the 'pure' R-isomer, no inhibitory activity was detected for R-ketoprofen, nor did it modulate the activity of the active 8-enantiomer when present together in the test system.

glandins, thromboxanes, leukotrienes and hydroxy acids.l5] The cyclo-oxygenase subunit of prosta­glandin synthetase is the primary specific site at which NSAIDs bind and block. In terms of chiral NSAIDs, it is well characterised that for 2-APA congeners and other arylalkanoic acids,[6-8] major or exclusive in vitro inhibition of prostaglandin synthesis is elicited by enantiomers of the S­stereoconfiguration; corresponding pharmaco­dynamic data for azapropazone and oxyphenbuta­zone are currently lacking. The situation in vivo needs to be tempered in light of the unique, uni­directional metabolic biotransformation of the R­enantiomers of 2-APAs to their corresponding pharmacologically active S-antipodes (see section 3). Accordingly, depending on the particular 2-APA congener and the animal species (which together determine the extent of R to S conversion), in vivo activity may be falsely attributed to the R-enantiomer.

Representative in vitro data pertaining to human­derived test systems are surveyed in table I. High eudismic ratios have been observed for each chiral NSAID studied. Furthermore, inhibitory activity of the racemate was very close to 50% of the cor­responding pure S-isomer value, indicating a lack of interaction between enantiomers. Hayball et aU13] have suggested that, in most cases, activity attributed to isomers of the R-stereoconfiguration is likely to be due to optical contamination of this enantiomeric test compound. Clearly, caution is needed when interpreting the data in table I, since they are indices rather than direct measures of in­hibitory activity on cyclo-oxygenase. Of particular note is the variability of the concentrations of iso-

© Adis International limited. All rights reserved.

mers required to elicit a 50% inhibition (IC50) of human platelet thromboxane production.D5] For example, even after taking into account fluctua­tions in protein binding of S-ketoprofen, the esti­mated in vitro IC50 value for unbound S-ketoprofen ranged from 0.044 to 0.508 J!g/L in a hetero­geneous group of 14 elderly patients with rheuma­toid arthritis (mean ± SD = 0.216 ± 0.143 J!glL) compared with an IC50 range of 0.262 to 0.402 gIL (mean ± SD = 0.320 ± 0.062 J!glL) in 4 young healthy drug-free volunteers.[l3,16]

Recently, cyclo-oxygenase has been shown to be encoded by 2 genes giving rise to 2 distinct forms of this enzyme (COX-l and COX-2).[17] COX-I is expressed in most tissues under normo­physiological conditions and is involved in cellular homeostatic or 'housekeeping' functions, such as maintaining the microvascular integrity of the gastrointestinal tract. In contrast, COX-2 is chiefly produced in inflammatory cells in response to tissue trauma and inflammation. Accordingly, COX-2-derived prostanoids are more likely to be the autacoids responsible for tissue damage and pain.[18]

It has been suggested that selective COX-2 in­hibition, at the expense of COX-I blockade, would be a logical therapeutic goal, thereby limiting the well recognised gastrointestinal adverse effects of NSAIDs.l17,19] Hayllar and Bjarnason[20] have surveyed data on the selectivity of in vitro COX-l and COX-2 inhibition versus relative risks of seri­ous upper gastrointestinal adverse effects. Limited data have shown that ligands with less inhibitory activity for COX-I, such as S-naproxen and indo-

Drugs 1996; 52 Suppl. 5

Page 4: Chirality and Nonsteroidal Anti-Inflammatory Drugs

50

methacin, are less likely to cause gastrointestinal complications than NSAIDs such as piroxicam, which are more selective COX-I antagonists. How­ever, the multifactorial nature of the pathophysiol­ogy of NSAID-induced adverse effectsl21] poten­tially limits the clinical utility of such in vitro observations, particularly in light of the discrepan­cies in magnitude between COX-lICOX-2 block­ade and relative gastrointestinal risks; agents with vastly different in vitro ratios have similar gastro­intestinal toxicity.

The active metabolite of the achiral NSAID pro­drug, nabumetone, has been shown in in vitro ani­mal studies to be a 7-fold more selective inhibitor of COX-2 than of COX-I ,[22] yet significantly less selectivity was observed in vivo and ex vivo in healthy volunteers receiving nabumetone. l23] In vitro human studies with the enantiomers of indo­bufenl24] and ketoprofenl25] have shown greater COX-I inhibition compared with COX-2 for each drug enantiomer; similar degrees of enantio­selectivity (S > R) were noted in these studies for the inhibition of each COX isoform. Such in vitro data, while not necessarily mimicking the situation in vivo, provide useful screening information for lead compound selection in the search for better tolerated NSAIDs.

Evidence of NSAID-induced cyclo-oxygenase inhibition as the means by which these drugs exert their pharmacological and toxicological actions is a reasonable rank-order correlation between inhi­bition of prostaglandin synthesis in vitro or ex vivo and anti-inflammatory or analgesic effects in vivoP6-28] However, notable exceptions to this generalisation have prompted speCUlation that NSAID pharmacological activity is elicited, at least in part, by mechanisms independent of pros­taglandin synthesis.l29-31] Brune and co-workersl32] established that the analgesic activity of the en­antiomers of flurbiprofen were equal in an in vivo rat model of nociception, yet R-flurbiprofen, in contrast to its optical antipode, had imperceptible effects on prostaglandin synthesis. These results could not be explained in terms of potential chiral inversion (see above and section 3), and since this

© Adis International Limited. All rights reserved.

Hayball

is a negligible metabolic pathway for flurbiprofen in the rat, the investigators suggested that there were additional molecular mechanisms of anal­gesia that remained to be elucidated.l32] The exist­ence of a third, centrally located isoform of COX, selective for paracetamol, has been postulatedJI7] In the case of the oldest NSAID, salicylic acid, which does not inhibit cyclo-oxygenase at analge­sic concentrationsl33] or dosages,l34] Vanel17] has suggested that its mechanism of action may include suppression of the induction of COX. Clearly, there remains much to be elucidated regarding COX isoforms and their pharmacological modulation by NSAIDs.

Postulated mechanisms of action of NSAIDs independent of their effects on prostanoid produc­tion include the following: (a) the ability of NSAIDs (including aspirin and salicylate) to lower sulfated glycosaminoglycan synthesis in articular cartilage in vitro,l35] although this mechanism is unlikely to contribute to the typical analgesic anti­inflammatory actions of these drugs; (b) suppres­sion of neutrophil aggregation, chemotaxis, de­granulation and resultant superoxide (020-) generationl36] and inhibition of inflammatory oedema by an action on polymorphonuclear leuco­cytes;l37] and (c) inhibition of mitochondrial ~-oxidation offatty acidsl38-40] (a mechanism more likely to convey toxicity than efficacy). The clinical significance of most of these alternative postulated mechanisms of action remains to be fully elucidated. However, it has been suggestedl29] that the effects of NSAIDs on stimulus-response coupling in neutrophils may explain the anti­inflammatory activity of salicylic acid. Further­more, a recent study showed that NSAIDs could be divided into 3 distinct categories based on their effects on stimulus-induced neutrophil oxidative burst. It was suggested that some NSAIDs, while providing temporary relief of arthritic symptoms, could exacerbate the underlying inflammatory condition as a result of 020--mediated tissue damage.l41 ]

The prostaglandin-independent effects of NSAIDs are thought to result from their ability to

Drugs 1996; 52 Suppl. 5

Page 5: Chirality and Nonsteroidal Anti-Inflammatory Drugs

Chirality and NSAIDs

partition (or dissolve) into cellular phospholipid bilayers, whereupon they physically disrupt inter­cellular signalling and subcellular protein-protein interactions.[Z9] This mode of action would be ex­pected to lack the degree of structural specificity inherent in the cyclo-oxygenase mechanism of NSAID action and, as suggested by Evans,[6] is likely to occur without the same degree of en­antioselectivity. Physical properties such as lipid or aqueous solubility are identical for individual enantiomers of a chiral drug, and thus the process of membrane partitioning should be nonenantio­selective, assuming that bilayer sequestered (carrier) proteins are not involved. This appears to be supported by the lack of enantioselectivity ob­served with the uncoupling of oxidative phos­phorylation in rat liver mitochondria and Ozo­production in macrophages induced by clindanac enantiomers,[4Z] and a recent report of equipotent inhibition by R-, S- and RS-ibuprofen of human polymorphonuclear cell function in vitro.l43 ]

In summary, it is likely that inhibition of prosta­glandin synthesis is the primary mechanism by which NSAIDs exert their pharmacological and toxicological actions. More recent reports of prostaglandin-independent biological properties of these drugs are beginning to contribute to a more detailed understanding of their clinical pharma­cology. However, it needs to be recognised that, to date, there is no proven connection between any of these alternative mechanisms and the clinical or toxic effects of NSAIDs.

3. Enantioselective Pharmacokinetics

Interactions of chiral drugs with chiral macro­molecules may occur during drug absorption, dis­tribution, metabolism and excretion, and hence each of these pharmacokinetic processes may be enantioselective.l44-46] Enantioselectivity at the dispositional level is often clouded by the com­plexity of the biological system. Discerning the identity of the enantioselective process in opera­tion is not always possible because of interfering 'chiral noise'. Pharmacokinetic differences be­tween enantiomers may alter pre-existing differ-

© Adis Internafional Limited, All rights reserved,

51

ences at the effector site, either tempering or ac­centuating pharmacodynamic enantioselectivity. An exhaustive review of enantioselective pharma­cokinetics of chiral NSAIDs (see Evans[6]) is be­yond the scope of this paper. Accordingly, a brief discussion of representative data is provided for clinically important pharmacokinetic processes and the way in which they impact on the therapeu­tic efficacy of enantiomeric NSAIDs.

Pharmacokinetic studies of chiral NSAIDs must be performed using enantioselective analytical methods if meaningful dispositional and drug concentration-effect data are to be obtained.l47,48]

Much of the earlier work in this area was per­formed using nonstereoselective (achiral) assays, which quantified plasma and urinary drug concen­trations of chiral NSAIDs as their unresolved en­antiomers (i.e. total drug); therefore, pharmaco­kinetic data from these studies appear to reflect administration of a single compound rather than a mixture. Accordingly, the pharmacokinetic para­meters of chiral NSAIDs administered as race­mates, generated using nonstereoselective analy­tical methods, are unlikely to reflect the true pharmacokinetic parameters of the individual enantiomers.

Examples of dramatic differences between enantiomers in terms of their respective pharma­cokinetic parameters include etodolac[49] and ketorolac.l50] For instance, following administra­tion of a single intramuscular bolus dose of RS­ketorolac to healthy volunteers, the systemic expo­sure to S-ketorolac [as measured by area under the plasma concentration-time profile (AUC)] was ap­proximately 50% of that of its optical antipode.l50] Similarly, both the distribution volume and elimi­nation half-life of the separate enantiomers were different. Taken in conjunction with the probable involvement ofhigbly enantioselective prostaglandin­dependent processes in both the activity and toxic­ity of this drug,l51] plasma concentrations of indi­vidual isomers are required to correlate plasma levels with therapeutic effect.

Stereo selectivity in drug metabolism is an im­portant determinant of possible differences in the

Drugs 1996: 52 Suppl. 5

Page 6: Chirality and Nonsteroidal Anti-Inflammatory Drugs

52

disposition of pairs of enantiomers,[52-54) and the literature is replete with examples of enantio­selectivity in metabolic transformations, including the chiral NSAIDs. In general, it appears that while drug absorption and excretion do not show signifi­cant enantioselectivity (except for actively trans­ported drugs), and plasma protein binding exhibits differences for some compounds,[16,55,56) the largest enantiomeric differences in disposition are the result of hepatic-metabolising activity. In the case of the 2-APA class of chiral NSAIDs, the R-enantiomer may be subject to a unique, highly enantioselective biotransformation process, leading to chiral inversion to the corresponding S-enantiomer, the extent of which is both substrate­and species-dependent (fig. 1).[7) For example, fenoprofen undergoes significant chiral inversion in man,[IO) which contributes to the substantially greater AVC observed for the pharmacologically active S-isomer. In contrast, ketoprofen is mini­mally inverted in man[5?) and its disposition is largely nonstereoselective; the AVC profiles for the individual isomers are virtually superimpos­able after racemic drug administration. [1 6)

The mechanism of the inversion reaction is thought to involve the stereospecific formation of a coenzyme A thioester [catalysed by acyl co­enzyme A ligase(s)] from the R-enantiomer of the 2-APA analogue, which then undergoes a number of alternative fates (see CaldweW58) and Shirley et al.l59) for detailed surveys): (a) racemisation of the chiral centre in the 2-APA moiety, followed by hydrolysis to yield a mixture of enantiomers of the parent drug; (b) hydrolysis with retention of stereoconfiguration yielding the original R-2-APA; or (c) acyl transfer of the 2-APA moiety to glycerol, resulting in the formation of a hybrid triglyceride. Interestingly, recent studies in dogs given the archetypal 2-APA compound, 2-phenylpropionic acid, have demonstrated that both enantiomers are probable substrates for canine hepatic acyl coen­zyme A ligase(s) and thus undergo bidirectional inversion.l60) Both R- and S-2-phenylpropionic acid were present in plasma after dosing with either antipode.

© Adis International limited. All rights reserved.

Hayball

In addition to chiral inversion, enantioselect­ivity may be observed for other routes of meta­bolism, for example, oxidation and conjugation with glucuronic acid. In the case of glucuroni­dation, chiral NSAIDs such as ketoprofen,[61) naproxen,[61) ketorolac[62) and ximoprofen[63) are subject to significant metabolism by hepatic uri­dine diphosphoglucuronosyltransferase yielding acyl glucuronide metabolites. Interestingly, acyl glucuronides are intrinsically reactive species, which are capable of undergoing a variety of chem­ical reactions under physiological pH and temper­ature conditions.l64) These reactions include hy­drolysis of the conjugate back to the parent drug (aglycone), intramolecular rearrangement yielding isomeric (positional) glucuronide conjugates, and formation of covalent adducts with biological macromolecules.

The facile hydrolysis of acyl-linked glucuro­nides of NSAIDs has potential clinical implica­tions, particularly in situations where such metabo­lites accumulate systemically. Although such NSAIDs have negligible percentages of their doses excreted into urine as unchanged drug, paradoxi­cally a number of them have diminished clearance in patients with renal dysfunction, in elderly pa­tients in whom renal function is expected to be di­minished, and in subjects receiving concurrent probenecid.l65-6?) It has been hypothesised that such observations are consistent with a decrease in the renal excretion of acyl glucuronides, leading to the systemic accumulation of these labile conju­gates followed by their facile hydrolysis back to the parent drug (fig. 2).l68) As discussed in a recent review, [69) the systemic regeneration of parent NSAID in the above scenario will have unique stereochemical implications for 2-APA congeners subject to metabolic chiral inversion. In such cases after racemic drug administration, a greatly exaggerated exposure to the pharmacologically active deconjugated S-isomer would be predicted, as has been shown in animal studies with the archetypal 2-APA, 2-phenylpropionic acid.l68) Consistent with this systemic cycling hypothesis, delayed clearance of R- and S-ketoprofen in

Drugs 1996; 52 Suppl. 5

Page 7: Chirality and Nonsteroidal Anti-Inflammatory Drugs

Chirality and NSAIDs 53

R-Aglycone 5-Aglycone

H fH3 0

-()'t-c~ ~OH -:?' " HO OH x~1 0 0 COO-

R-Glucuronide S-Glucuronide

! ! Urine Urine

Fig. 2. The systemic cycle originally proposed by Metlin et al.l68] for the formation, hydrolysis and renal clearance of epimeric acyl glucuronides of 2-arylpropionic acid NSAIDs that undergo metabolic chiral inversion. Not shown in this simplified model are: (a) additional biotransformation pathways for both aglycones and glucuronides; (b) regeneration of aglycone from rearranged isomers of biosynthetic glucuronides and, to a lesser extent (quantitatively), from irreversibly bound drug-protein adduct; and (c) biliary excretion of epimeric glucuronides[69] (reproduced with the permission of John Wiley & Sons, New York).

rheumatoid arthritis patients with diminished creatinine clearance has been observed.l161

The ability of acyl glucuronide metabolites of both chiral and achiral NSAIDs to form covalent attachments with protein is also likely to have clinical consequences, particularly when the clear­ance of these metabolites is delayed. The putative toxicological sequelae of adduct formation are dis­cussed below.

4. Toxicology of Chiral NSAID Enantiomers

Inhibition of arachidonic acid metabolism via the cyclo-oxygenase pathway is the cause of a range of well documented adverse effects of con­geners of this drug class. Gastropathy induced by NSAIDs is the principal adverse effect and is thought to result from blockade of COX-I, the en­zyme subtype responsible for generating cyto-

© Adls International Limited. All rights reserved.

protective 'housekeeping' prostaglandins such as PGE2 and PGIz, which increase gastric mucosa blood flow and mucin production.[21) It would be expected that S-enantiomers are primarily respons­ible for eliciting the gastrointestinal adverse effects of these drugs. Using the enantiomers of flurbi­profen, animal studies have demonstrated that R-flurbiprofen caused minimal gastrointestinal toxicity, which in tum correlated with its marginal inhibitory effect on local prostaglandin product­ion. [32.70) After long term administration of either racemic or enantiomerically pure flurbiprofen to rats, Wechter and co-workers[71) found that the R­enantiomer alone did not elicit gastrointestinal lesions, while the racemate proved to be 2 to 4 times more ulcerogenic in enantiomerically equivalent doses than the S-enantiomer. This suggests a possible attenuation of NSAID-induced gastropathy from the prostaglandin-independent

Drugs 1996; 52 Suppl. 5

Page 8: Chirality and Nonsteroidal Anti-Inflammatory Drugs

54

actions of R-flurbiprofen. It has been speculated that the R-enantiomer might exacerbate gastro­intestinal toxicity by inducing leucocyte adherence in postcapillary venules, and thereby exacerbating the tissue destruction resulting from local ischaemia induced by cyclo-oxygenase antagonismPl,72]

Epidemiological studies have revealed a dose­response relationship for gastrointestinal toxicity, in addition to a relationship between the duration of exposure and toxicity, so that the risks are great­est for patients who receive between 2 and 4 pre­scriptions or therapy lasting 2 to 4 monthsP3,74] The toxicity referred to in these studies is serious upper gastrointestinal bleeding or perforation. Additional risk factors include old age, where it is generally accepted that NSAID-induced gastro­pathy operates on an increased basal risk of peptic ulceration, thereby giving an overall increased risk of complications of peptic ulceration in the elderly. It remains to be elucidated whether systemic cyc­ling of NSAIDs metabolised to renally eliminated acyl glucuronides (see section 3) enhances the risk of gastrointestinal toxicity in elderly, renally com­promised patients.

It has become apparent that an alternative ara­chidonic acid catabolic pathway generates prod­ucts that profoundly affect cellular and vascular reactions associated with inflammationP5] The lipoxygenase pathway generates hydroperoxy­eicosatetraenoic acids (HPETEs), which are re­duced by peroxidase to the corresponding hydroxy acids. One of the hydroperoxy derivatives,5-HPETE, is the precursor of a group of intensely acti ve auta­coids referred to as leukotrienes. Accordingly, NSAID-induced blockade of cyclo-oxygenase activity, while inhibiting production of prosta­glandins and thromboxanes, may putatively in­crease the catabolism of arachidonic acid via lipoxygenase pathways. This could lead to over­production of leukotrienes, hydroxy-fatty acids and tissue-destructive free oxygen radicals.

Theoretical advantages might therefore be conferred upon the clinical use of agents that block both cyclo-oxygenase and lipoxygenase pathways of arachidonate catabolism,[75] such as

© Adis Intematlonal Limited. All rights reserved.

Hayball

ketoprofenP6] However, this hypothesis has been challenged by Somasundaram et al.,[21] who argue that metabolic diversion of arachidonic acid would be unlikely, since cyclo-oxygenase and lipoxy­genase have different subcellular locations (cyto­sol and endoplasmic reticulum, respectively[22]). In addition to drug toxicity predicted from inhibition of eicosanoid production, the participation of 2-APA analogues in the metabolic chiral inversion process, as well as the generation of covalent drug-protein adducts from those acidic NSAIDs that are meta­bolised to reactive acyl glucuronides, have each been hypothesised to lead to toxic sequelae.

The generation of R-2-arylpropionyl coenzyme A thioesters, as obligate intermediates in the chiral inversion reaction (section 3), has been suggested as having toxicological significance.£58]It has been shown by use of racemic mixtures of a number of 2-APA analogues that these drugs are capable of inhibiting cholesterogenesis and fatty acid synthe­sis in vitro, and that this activity is correlated with their ability to form hybrid triacylglycerolsP7] As suggested by Caldwell and Marsh,[78] the forma­tion of such hybrid lipids and their incorporation into cell membranes have the potential to alter membrane structure and function. Accordingly, the involvement of R-profens in coenzyme A thioester intermediate formation might confer an advantage for the use of the S-enantiomers alone in clinical practice. However, there is no direct clinical evi­dence of corresponding toxicity to date. Indeed, as argued by Williams and Day,[79] there is little evi­dence to suggest that S-naproxen is significantly less toxic than those 2-APAs used clinically as the racemate. Moreover, Evans[6] makes the point that the efficacy and safety of racemic ibuprofen, which is subject to significant chiral inversion in man, is noteworthy.

In humans, many chiral NSAIDs are metabo­lised to renally eliminated acyl glucuronides to a significant extent (section 3). It has been argued that the ability of such drugs (without exception) to form covalent drug-protein adducts, via these reactive metabolites, may lead to immunotoxi­cological responses.£64] A significant proportion of

Drugs 1996; 52 Suppl. 5

Page 9: Chirality and Nonsteroidal Anti-Inflammatory Drugs

Chirality and NSAIDs

those drugs withdrawn from the US and UK mar­kets over the last 2 decades because of induction of severe hypersensitivity reactions were carboxyl­ic acids subject to significant glucuronide conjuga­tion.l80] Ding et aU81] have recently identified Lys-199 as the most prominent covalent binding site for tolmetin glucuronide on the human serum albumin molecule; Lys-199 is a lysine E-amino group located in a hydrophobic region notable as a target for covalent binding of recognised immuno­gens such as penicillin-derived benzylpenicilloyl radicals. [82]

It has been further argued that the ubiquitous involvement of the glucuronic acid moiety in acyl glucuronide-mediated adduct formation may ex­plain the apparent immunological cross-reactivity with several NSAIDs in susceptible patients.[81] Clearly, the ability of both enantiomers of chiral NSAIDs to participate in adduct formation[69] suggests that marketing of drugs as the sole en­antiomer will not negate covalent drug-protein generation, although adduct density is likely to be diminished. It should also be noted that increased systemic exposure to acyl glucuronides, due to conjugate recycling, may result in higher adduct formation, as has been shown in humans after inhibition of the renal elimination of zomepirac glucuronide by probenecid.l83] However, the link between adduct formation in humans and immuno­toxicity remains to be conclusively proven.

5. Conclusions

Chiral NSAIDs, in particular 2-APA analogues, exhibit well characterised enantioselective phar­macokinetic, pharmacodynamic and toxicological properties, the degree of enantioselectivity being dependent on the particular pharmacological pro­cess under investigation. Accordingly, a need for a 3-dimensional understanding of the pharmacolog­ical properties is implicit in understanding the in­teraction between chiral NSAIDs and biological environments. It is clear, given the existence of the chiral inversion reaction for some 2-APAs, that the activity and disposition of such drugs (when ad­ministered as racemates) are unlikely to be the

© Adis International Llm~ed. All rights reserved.

55

simple sum of activities and disposition of the in­dividual enantiomers.

There has been considerable debate as to whether all chiral NSAIDs should be marketed as the pure S-enantiomer for use in the treatment of inflammatory conditions, since it is generally ac­cepted that the activity of such drugs in this context is confined to isomers. of the S-stereoconfiguration. Benefits of using the S-enantiomer alone might be the avoidance of so-called 'isomeric ballast'[47] and a reduction in the metabolic load to the patient. Furthermore, for those agents that are administered as race mates and undergo extensive chiral in­version (e.g. ibuprofen[67] and fenoprofen[lO]), the effective dose of the active agent is unknown, which is clearly an undesirable situation. In addi­tion, to the extent that a knowledge of clinical pharmacology facilitates dosage selection, single isomers might be preferred because of less complex pharmacokinetic profiles, drug interactions and concentration-effect relationships; however, very few studies have correlated plasma concentrations of NSAIDs with assessments of therapeutic re­sponse.l84] Moreover, the interindividual varia­tions in the pharmacokinetics of the S-isomer of the metabolically complex 2-APA congener, ibuprofen, were similar irrespective of whether the drug was administered as the racemate or as the separate S-enantiomer.[85] As commented upon by Williams,P5] the limited clinical studies of 2-APA analogues administered as the separate S-isomers, while demonstrating efficacy as anti-inflammatory agents, have not yet facilitated attempts to establish concentration-effect relationships for these drugs. Furthermore, the possible role ascribed to R­enantiomers of chiral NSAIDs in eliciting prosta­glandin-independent actions in blunting pain and inflammation complicate the 'enantiomer versus racemate' debate for this chiral drug class.

It would appear that there are at least theoretical advantages to marketing chiral NSAIDs as en­antiomerically pure drug products, if only to min­imise the potential additional toxicity mediated by the optical antipode. Birkett,[86] arguing from a drug regulatory standpoint, has concluded that the

Drugs 1996; 52 Suppl. 5

Page 10: Chirality and Nonsteroidal Anti-Inflammatory Drugs

56

balance of the argument is overwhelmingly in favour of the use of separate enantiomers. Indeed, the pharmaceutical industry appears to be moving away from developing racemates in favour of en­antiomerically pure or achiral drugs)87J While ob­viously benefiting the economies of pharmaceuti­cal companies, the reworking of traditionally safe and efficacious racemic NSAIDs into patentable S-isomeric formulations is likely to contribute significantly to our understanding of the basic and clinical pharmacology of these therapeutically important drugs. Modern marketing applications for enantiomerically 'cleaned-up' drugs must, by necessity, be accompanied by extensive basic and clinical pharmacological data on both the safety and efficacy of these drugs, irrespective of what has been previously learnt about the racemate. This additional knowledge should enhance more rational prescribing of chiral NSAIDs and, in the process, enable us to learn more about pain and inflammatory processes and their response to pharmacological intervention.

References 1. Vane JR. Inhibition of prostaglandin synthesis as a mechanism

of action for aspirin-like drugs. Nature 1971; 231: 232-5 2. Takeguchi C, Sih CJ. A rapid spectrophotometric assay for pros­

taglandin synthetase: application to the study of non-steroidal anti-inflammatory agents. Prostaglandins 1972; 2: 169-84

3. Lewis AJ, Furst DE. Non-steroidal anti-inflammatory drugs: mechanisms and clinical uses. 2nd ed. New York: Marcel Dekker, 1994: 1-2

4. Vane JR, Botting RM. The mode of action of anti-inflammatory drugs. Postgrad Med J 1990; 66 Suppl. 4: S2-17

5. Cashman J, McAnulty G. Nonsteroidal anti-inflammatory drugs in peri surgical pain management. Drugs 1995; 49: 51-70

6. Evans AM. Enantioselective pharmacodynamics and pharma­cokinetics of chiral non-steroidal anti-inflammatory drugs. Eur J Clin Pharmacol1992; 42: 237-56

7. Hutt AJ, Caldwell J. The metabolic chiral inversion of 2-arylpropionic acids - a novel route with pharmacological con­sequences. J Pharm Pharmacol1983; 35: 693-704

8. Williams KM. Enantiomers in arthritic disorders. Pharmacol Ther 1990; 46: 273-95

9. Gaut ZN, Baruth H, Randall LO, et al. Stereoisomeric relation­ships among anti-inflammatory activity, inhibition of platelet aggregation, and inhibition of prostaglandin synthetase. Pros­taglandins 1975; 10: 59-66

10. Rubin A, Knadler MP, Ho PPK, et al. Stereoselective inversion of (R)-fenoprofen to (S)-fenoprofen in humans. J Pharm Sci 1985; 74: 82-4

11. Evans AM, Nation RL, Sansom LN, et aI. Effect of racentic ibuprofen dose on the magnitude and duration of platelet cyclo-oxygenase inhibition: relationship between inhibition of thromboxane production and the plasma unbound con-

© Adis International limited. All rights reserved.

Haybal/

centration of S( + )-ibuprofen. Br J Clin Pharmacol 1991; 31: 131-8

12. Cerletti C, Manarini S, Colombo M, et al. The (+)-enantiomer is responsible for the antiplatelet and anti-inflammatory ac­tivity of (±)-indobufen. J Pharm Pharmacol1990; 42: 885-7

13. Hayball PJ, Nation RL, Bochner F. Enantioselective pharmaco­dynamics of the nonsteroidal antiinflammatory drug ketoprofen: in vitro inhibition of human platelet cyclo­oxygenase activity. Chirality 1992; 4: 484-7

14. Suesa N, Fernandez MR, Gutierrez M, et aI. Stereoselective cyclooxygenase inhibition in cellular models by the en­antiomers of keto prof en. Chirality 1993; 5: 589-95

15. Williams KM. Chiral NSAIDs: so what? Agents Actions Suppl 1993; 44: 15-22

16. Hayball PJ, Nation RL, Bochner F, et al. The influence of renal function on the enantioselective pharmacokinetics and phar­macodynantics of ketoprofen in patients with rheumatoid ar­thritis. Br J Clin Pharmacol1993; 36: 185-93

17. Vane JR. Towards a better aspirin. Nature 1994; 367: 215-6 18. Dray A, Bevan S. Inflammation and hyperalgesia: the team ef­

fort. Trends Pharm Sci 1993; 14: 287-90 19. Aslanian R, Carruthers NI, Kaminski JJ. Cyclo-oxygenase-2: a

novel target for therapeutic intervention. Exp Opin Invest Drugs 1994; 3: 1323-5

20. Hayllar J, Bjarnason I. NSAIDs, COX-2 inhibitors, and the gut. Lancet 1995;346:521-2

21. Somasundaram S, Hayllar H, Rafi S, et al. The biochentical basis of non-steroidal anti-inflammatory drug-induced dam­age to the gastrointestinal tract: a review and a hypothesis. Scand J Gastroenterol 1995; 30: 289-99

22. Meade EA, Smith WL, DeWitt DL. Differential inhibition of prostaglandin endoperoxide synthase (cyclooxygenase) iso­zymes by aspirin and other nonsteroidal anti-inflammatory drugs. J Bioi Chern 1993; 268: 6610-14

23. Cipollone F, Ganci A, Panara MR, et aI. Effects of nabumetone on prostanoid biosynthesis in humans. Clin Pharmacol Ther 1995; 58: 335-41

24. Patrignani P, Panara MR, Greco A, et al. Biochemical and phar­macological characterization of the cyclooxygenase activity of human blood prostaglandin endoperoxide synthases. J Pharmacol Exp Ther 1994; 271: 1705-12

25. Panara MR, Greco A, Santini G, et al. Effects of the novel anti-inflammatory compounds, N-[2-(cyclohexyloxy)-4-nitrophenyl] methanesuphonamide (NS-398) and 5-methane­sui phonamido-6-( 2,4-difluorothiopheny 1)-l-indanone (L-745,337), on the cyclo-oxygenase activity of human blood prostaglandin endoperoxide synthases. Br J Pharmacol1995; 116: 2429-34

26. Matsuda K, Tanaka Y, Ushiyama S, et al. Inhibition of prosta­glandin synthesis by sodium 2-[4-(2-oxocyclopentylmethyl) phenyl]-propionate dihydrate (CS-600), a new anti-inflam­matory drug, and its active metabolite in vitro and in vivo. Biochem Pharmacol1984; 15: 2473-8

27. Ferreira SH, Vane JR. Mode of action of anti-inflammatory agents which are prostaglandin synthetase inhibitors. In: Vane JR, Ferreira SH, editors. Anti-inflammatory drugs. Berlin: Springer, 1979

28. Tomlinson RV, Ringold HJ, Qureshi MC, et al. Relationship between inhibition of prostaglandin synthesis and drug effi­cacy: support for the current theory on mode of action of aspirin-like drugs. Biochem Biophys Res Commun 1972; 46: 552-8

29. Abramson SB, Weissmann G. The mechanism of action of non­steroidal antiinflammatory drugs. Arthritis Rheum 1989; 32: 1-9

30. Day RO, Graham GG, Williams KM, et al. Clinical pharmacol­ogy of non-steroidal anti-inflammatory drugs. Pharmacol Ther 1987; 33: 383-433

31. Goodwin JS. Mechanism of action of nonsteroidal anti-inflam­matory drugs. Am J Med 1984; 77 Suppl. 1A: 57-64

Drugs 1996: 52 Suppl. 5

Page 11: Chirality and Nonsteroidal Anti-Inflammatory Drugs

Chirality and NSAIDs

32. Brune K, Beck WS, Geisslinger G, et al. Aspirin-like drugs may block pain independently of prostaglandin synthesis inhibi­tion. Experientia 1991; 47: 257-60

33. Brune K, Rainsford KD, Wagner K, et al. Inhibition by anti­inflammatory drugs of prostaglandin production in cultured macrophages. Factors influencing the apparent drug effects. Naunyn-Schmiedebergs Arch Pharmacol1981; 315: 269-76

34. Rosenkranz B, Fisher C, Meese CO, et al. Effects of salicylic and acetylsalicylic acid alone and in combination on platelet aggregation and prostanoid synthesis in man. Br J Clin Phar­maco11986; 21: 309-17

35. Brandt KD, Palmoski MJ. Effects of salicylates and other non­steroidal anti-inflammatory drugs on articular cartilage. Am J Med 1984; 77 Suppl. IA: 65-9

36. Shelly J, Hoff SF. Effects of non-steroidal anti-inflammatory drugs on isolated human polymorphonuclear leukocytes (PMN): chemotaxis, superoxide production, degranulation and N-formyl-L-methionyl-L-Ieucyl-L-phenylalanine (fMLP) receptor binding. Gen Pharmacol 1989; 20: 329-34

37. Rampart M, Williams TJ. Suppression of inflammatory oedema by ibuprofen involving a mechanism independent of cyclo­oxygenase inhibition. Biochem Pharmacol1986; 35: 581-6

38. Freneaux E, Fromenty B, Berson A, et al. Stereoselective and nonstereoselective effects of ibuprofen enantiomers on mito­chondrial f3-oxidation of fatty acids. J Pharmacol Exp Ther 1990; 255: 529-35

39. Geneve J, Hayat-Bonan B, Labbe G, et al. Inhibition of mito­chondrial f3-oxidation of fatty acids by pirprofen. Role in microvesicular steatosis due to this nonsteroidal anti-inflam­matory drug. J Pharmacol Exp Ther 1987; 242: 1133-7

40. Zhao B, Geisslinger G, Hall I, et al. The effect of the en­antiomers of ibuprofen and flurbiprofen on the f3-oxidation of palmitate in the rat. Chirality 1992; 4: 137-41

41. Twomey BM, Dale MM. Cyclooxygenase-independent effects of non-steroidal anti-inflammatory drugs on the neutrophil respiratory burst. Biochem Pharmacol 1992; 43: 413-8

42. Kawai K, Shiojiri H, Fukushima H, et al. The effect of clindanac, a potent anti-inflammatory drug, on mitochondrial respiration: a consideration of the uncoupling activity of op­tical enantiomers. Res Commun Chern Pathol Pharmacol 1984; 45: 399-406

43. Villanueva M, Heckenberger R, Strobach H, et al. Equipotent inhibition by R(-)-ibuprofen, S(+)-ibuprofen and racemic ibuprofen of human polymorphonuclear cell function in vitro. Br J Clin Pharmacol1993; 35: 235-42

44. Jamali F, Mehvar R, Pasutto FM. Enantioselective aspects of drug action and disposition: therapeutic pitfalls. J Pharm Sci 1989; 78: 695-715

45. Lee EJD, Williams KM. Chirality: clinical pharmacokinetic and pharmacodynamic considerations. Clin Pharmacokinet 1990; 18: 339-45

46. Tucker GT, Lennard MS. Enantiomer specific pharmacokinet­ics. Pharmacol Ther 1990; 45: 309-29

47. Ariens EJ. Stereochemistry, a basis for sophisticated nonsense in pharmacokinetics and clinical pharmacology. Eur J Clin Pharmacol 1984; 26: 663-8

48. Evans AM, Nation RL, Sansom LN, et al. Stereoselective drug disposition: potential for misinterpretation of drug disposi­tion data. Br J Clin Pharmacol1988; 26: 771-80

49. Jamali F, Mehvar R, Lemko C, et al. Application of a stereo­specific high-performance liquid chromatographic assay to a pharmacokinetic study of etodolac enantiomers in humans. J Pharm Sci 1988; 77: 963-6

50. Hayball PJ, Wrobel J, Tamblyn JG, et al. The pharmacokinetics of ketorolac enantiomers following intramuscular adminis­tration of the racemate. Br J Clin Pharmacol1994; 37: 75-8

51. Guzman A, Yuste F, Toscano RA, et al. Absolute configuration of (-)-5-benzoyl-l ,2-dihydro-3H-pyrrolo[ 1 ,2a Jpyrrole-I­carboxylic acid. J Med Chern 1986; 29: 589-91

© Adis International Limited, All rights reserved.

57

52. Caldwell J, Winter SM, Hutt AJ. The pharmacological and toxi­cological significance of the stereochemistry of drug disposi­tion. Xenobiotica 1988; 18 Suppl. 1: 59-70

53. Eichelbaum M. Pharmacokinetic and pharmacodynamic conse­quences of stereoselective drug metabolism in man. Biochem PharmacoI1988;37:93-6

54. Testa B. Substrate and product stereoselectivity in mono­oxygenase-mediated drug activation and inactivation. Biochem Pharmacol1988; 37: 85-92

55. Evans AM, Nation RL, Sansom LN, et al. Stereoselective plasma protein binding of ibuprofen enantiomers. Eur J Clin Pharmacol1989; 36: 283-90

56. Hayball PJ, Holman JW, Nation RL, et al. Marked enantio­selective protein binding of ketorolac in vitro: elucidation of enantiomer unbound fractions following facile synthesis and direct chiral HPLC resolution of tritium-labelled ketorolac. Chirality 1994; 6: 642-8

57. Jamali F, Russell AS, Foster RT, et al. Ketoprofen pharmacoki­netics in humans: evidence of enantiomeric inversion and lack of interaction. J Pharm Sci 1990; 79: 460-1

58. Caldwell J. Stereochemical determinants of the nature and con­sequences of drug metabolism. J Chromatogr A 1995; 694: 39-48

59. Shirley MA, Guan X, Kaiser DG, et al. Taurine conjugation of ibuprofen in humans and in rat liver in vitro. Relationship to metabolic chiral inversion. J Pharmacol Exp Ther 1994; 269: 1166-75

60. Tanaka Y, Shimomura Y, Hirota T, et al. Formation of glycine conjugate and (-)-(R)-enantiomer from(+)-(S)-2-phenyl­propionic acid suggesting the formation of the CoA thioester intermediate of (+ )-(S)-enantiomer in dogs. Chirality 1992; 4: 342-8

61. Upton RA, Buskin IN, Williams RL, et al. Negligible excretion of unchanged ketoprofen, naproxen and probenecid in urine. J Pharm Sci 1980; 69: 1254-7

62. Mroszczak E, Lee FW, Combs D, et al. Ketorolac tromethamine absorption, distribution, metabolism, excretion, and pharma­cokinetics in animals and humans. Drug Metab Dispos 1987; 15: 618-26

63. Mayo BC, Chasseaud LF, Hawkins DR, et al. The metabolic fate of 14C-ximoprofen in rats, baboons and humans. Xenobiotica 1990; 20: 233-46

64. Spahn-Langguth H, Benet LZ. Acyl glucuronides revisited: is the glucuronidation process a toxification as well as a detox­ification mechanism? Drug Metab Rev 1992; 24: 5-48

65. Advenier C, Roux A, Gobert C, et al. Pharmacokinetics of ketoprofen in the elderly. Br J Clin Pharmacol1983; 16: 65-70

66. Anttila M, Haataja M, Kasanan A. Pharmacokinetics of na­proxen in subjects with normal and impaired renal function. Eur J C1in Pharmacol1980; 18: 263-8

67. Foster RT, Jamali F, Russell AS. Pharmacokinetics of ketoprofen enantiomers in cholecystectomy patients: influ­ence of probenecid. Eur J Clin Pharmacol1989; 37: 589-94

68. Meffin PJ, Sallustio BC, Purdie YJ, et al. Enantioselective dis­position of 2-arylpropionic acid nonsteroidal anti-inflamma­tory drugs. I. 2-Phenylpropionic acid disposition. J Pharmacol Exp Ther 1986; 238: 280-7

69. Hayball PJ. Formation and reactivity of acyl glucuronides: the influence of chirality. Chirality 1995; 7: 1-9

70. Peskar BM, Kluge S, Peskar BA, et al. Effects of pure en­antiomers of flurbiprofen in comparison to racemic flurbiprofen on eicosanoid release from various rat organs ex vivo. Prostaglandins 1991; 42: 515-31

71. Wechter WJ, Bigornia AE, Murray ED, et al. Rac-flurbiprofen is more ulcerogenic than its (S)-enantiomer. Chirality 1993; 5: 492-4

72. Wallace JL, Granger DN. Pathogenesis of NSAID gastropathy: are neutrophils the culprits? Trends Pharm Sci 1992; 13: 129-31

Drugs 1996; 52 Suppl, 5

Page 12: Chirality and Nonsteroidal Anti-Inflammatory Drugs

58

73. Carson JL, Willett LR. Toxicity of nonsteroidal anti-inflamma­tory drugs. An overview of epidemiological evidence. Drugs 1993; 46 Suppl. I: 243-8

74. Willett LR, Carson JL, Strom BL. Epidemiology of gastrointes­tinal damage associated with nonsteroidal anti-inflammatory drugs. Drug Saf 1994; 10: 170-81

75. Rainsford KD. Inhibitors of eicosanoid metabolism. In: Curtis­Prior PB, editor. Prostaglandins - biology and chemistry of prostaglandins and related eicosanoids. Edinburgh: Churchill Livingstone, 1988: 52-68

76. Kubota T, Komatsu H, Kawamoto H, et al. Studies on the effects of anti-inflammatory action of benzoyl-hydratropic acid (ketoprofen) and other drugs, with special reference to pros­taglandin synthesis. Arch Int Pharmacodyn Ther 1979; 237: 169-76

77. Fears R, Richards DH. Association between lipid-lowering ac­tivity of aryl-substituted carboxylic acids and formation of substituted glycerolipids in rats. Biochem Soc Trans 1981; 9: 572-3

78. Caldwell J, Marsh MV. Interrelationship between xenobiotic metabolism and lipid biosynthesis. Biochem Pharmacol1983; 32: 1667-72

79. Williams KM, Day RO. The contribution of enantiomers to vari­ability in response to anti-inflammatory drugs. Agents Ac­tions Supp11988; 24: 76-84

80. Benet LZ, Spahn-Langguth H, Iwakawa S, et al. Predictability of the covalent binding of acidic drugs in man. Life Sci 1993; 53: PLl41-6

© Adls Interna~onal Limited. All rights reserved.

Hayball

81. Ding A, Ojingwa JC, McDonagh AF, et al. Evidence for cova­lent binding of acyl glucuronides to serum albumin via an imine mechanism as revealed by tandem mass spectrometry. Proc Nat! Acad Sci USA 1993; 90: 3797-801

82. Yvon M, Wal J-M. Identification of lysine residue 199 of human serum albumin as a binding site for benzylpenicilloyl groups. FEBS Lett 1988; 239: 237-40

83. Smith PC, McDonagh AF, Benet LZ. Irreversible binding of zomepirac to plasma protein in vitro and in vivo. J Clin Invest 1986; 77: 934-9

84. Simkin PA. Concentration-effect relationships of NSAID. J Rheumatol1988; 15 Suppl. 17: 40-3

85. Geisslinger G, Stock KP, Loew D, et al. Variability in the stereoselective disposition of ibuprofen in patients with rheu­matoid arthritis. Br J Clin Pharmacol1993; 35: 603-7

86. Birkett DJ. Racemates of enantiomers: regulatory approaches. Clin Exp Pharmacol Physiol1989; 16: 479-83

87. Lennard MS. Clinical pharmacology through the looking glass: reflections on the racemate vs enantiomer debate. Br J Clin Pharmacol1991; 31: 623-5

Correspondence and reprints: Dr Peter Haybal/, School of Pharmacy and Medical Sciences, University of South Australia, North Terrace, South Australia 5000, Australia.

Drugs 1996; 52 Suppl. 5