Sunteți pe pagina 1din 13

Best Practice & Research Clinical Gastroenterology Vol. 15, No. 5, pp.

691703, 2001

doi:10.1053/bega.2001.0229, available online at http://www.idealibrary.com on

1 Pathogenesis of NSAID-induced gastroduodenal mucosal injury


John L. Wallace*
PhD

Professor Mucosal Inammation Research Group, Faculty of Medicine, University of Calgary, Calgary, Alberta, Canada

The use of non-steroidal anti-inammatory drugs (NSAIDs), even in the era of selective COX-2 inhibitors, remains limited by the ability of these agents to cause gastroduodenal ulceration and bleeding. This damage is caused mainly through the ability of these agents to inhibit prostaglandin synthesis, which has a negative impact on several components of mucosal defence. Many NSAIDs also have topical irritant eects on the epithelium which may be particularly important in the production of small intestinal injury. While the presence of acid in the lumen of the stomach may not be a primary factor in the pathogenesis of NSAIDinduced gastroenteropathy it can make an important contribution to the chronicity of these lesions and to bleeding by impairing the restitution process, interfering with haemostasis and inactivating several growth factors that are important in mucosal defence and repair. Through better understanding of the pathogenesis of ulcers induced by NSAIDs, some new approaches to the development of more eective and safer anti-inammatory drugs have been taken in recent years. Key words: ulcer; non-steroidal anti-inammatory drug; nitric oxide; acid; neutrophils; cyclooxygenase.

The ability of non-steroidal anti-inammatory drugs (NSAIDs) to cause ulceration and bleeding in the upper gastrointestinal tract has been recognized for over 60 years1, and has more recently been referred to as an `epidemic'.2,3 Approximately 20% of regular users of NSAIDs will develop an ulcer of the stomach or duodenum. NSAIDs are very widely used, particularly for relieving the symptoms of osteoarthritis and rheumatoid arthritis. The ulcerogenic eects of NSAIDs have become more evident to the public in recent years as a result of the aggressive marketing of selective inhibitors of cyclooxygenase (COX)-2. These agents have been proposed to exhibit the benecial eects of conventional NSAIDs without causing gastrointestinal damage. In this chapter, the mechanisms through which NSAIDs induce ulcers, cause bleeding and impair ulcer healing are reviewed. The contributions of acid and Helicobacter pylori to NSAIDassociated ulceration are also reviewed.
*Correspondence to: Department of Pharmacology & Therapeutics, University of Calgary, 3330 Hospital Drive NW, Calgary, Alberta, T2N 4N1, Canada. 15216918/01/050691 13 $35.00/00 c 2001 Harcourt Publishers Ltd. *

692 J. L. Wallace

INHIBITION OF GASTRIC PROSTAGLANDIN SYNTHESIS The discovery that NSAIDs exert their anti-inammatory eects via inhibition of prostaglandin synthesis led directly to the proposal that prostaglandins play an important role in gastric mucosal defence.4 This was followed by the demonstration that microgram quantities of exogenous prostaglandins could protect the gastrointestinal tract from injury induced by noxious agents and NSAIDs.5 The importance of prostaglandins in mucosal defence is underscored by the evidence that the ability of an NSAID to induce gastric damage correlates well with its ability to suppress gastric prostaglandin synthesis.69 Agents that are weak inhibitors of gastric prostaglandin synthesis, including selective COX-2 inhibitors, are less ulcerogenic.810 There is also a good correlation between the time- and dose-dependency of suppression of gastric prostaglandin synthesis by NSAIDs and their ability to cause gastric ulcers.68 While the importance of prostaglandins in mucosal defence has been well recognized for nearly 30 years, a question that remained largely unanswered is why suppression of gastric prostaglandin synthesis resulted in mucosal injury. The simple answer is that virtually every component of mucosal defence is to some extent prostaglandin-dependent, so inhibition of prostaglandin synthesis by NSAIDs leads to an increased susceptibility of the stomach and duodenum to injury induced by luminal irritants. Moreover, suppression of prostaglandin synthesis leads to marked alterations in the microcirculation of the stomach and intestine that appear to be critical and to early events in the pathogenesis of ulceration (discussed in more detail below). Suppression of prostaglandin synthesis can also have a negative impact upon the secretion of mucus and bicarbonate by the gastric and duodenal epithelium, the proliferation of epithelial cells, the function of immunocytes (e.g. mast cells) within the lamina propria and upon repair processes (discussed below).11,12 It is important to bear in mind that inhibition of prostaglandin synthesis does not always result in mucosal injury. Other endogenous mediators, such as nitric oxide, also play an important role in mediating mucosal defence. Indeed, nitric oxide carries out many of the same actions as prostaglandins, and there appears to be co-operative modulation of prostaglandin and nitric oxide synthesis.13,14 Specically, in situations of prolonged suppression of prostaglandin synthesis, the production of and/or reactivity to nitric oxide has been shown to be increased.15 COX-1 versus COX-2: which is more important in mucosal defence? In 1991, several groups demonstrated that the primary enzyme responsible for prostaglandin synthesis, cyclo-oxygenase, existed in at least two forms. COX-1 is the predominant form expressed in the normal gastrointestinal tract1618 but COX-2 can be detected and has been shown to be rapidly up-regulated in response to a number of stimuli. For example, in the rat's stomach, COX-2 was found to be induced within an hour of administration of aspirin or indomethacin16, or following a period of ischaemia and re-perfusion.19 There is mounting evidence that even a very small contribution of COX-2 to total gastric prostaglandin synthesis can be very important. Gretzer et al20 demonstrated that the increased resistance of the rat stomach to injury following exposure to a mild irritant was prevented by pre-treatment with highly selective COX-2 inhibitors. Similarly, ischaemiare-perfusion injury in the rat's stomach was signicantly exacerbated by pre-treatment with highly selective COX-2 inhibitors.19 The most likely explanation for these ndings is that COX-2 is up-regulated in response to tissue hypoxia in an eort to restore blood ow through the production

Pathogenesis of NSAID-ulcer 693

of vasodilatory prostaglandins. COX-2 has been shown to play a very important role in ulcer healing21,22 such that its inhibition leads to a signicant impairment of healing processes and, more specically, of angiogenesis. Moreover, in situations of gastrointestinal inammation, COX-2-derived prostaglandins serve an important function of promoting healing and down-regulating the inammatory response. Thus, inhibition of COX-2 in experimental colitis resulted in a marked exacerbation of the colonic injury, leading to perforation and death.23 There is also mounting evidence to suggest that the gastric damage induced by conventional NSAIDs does not occur because of the ability of these agents to inhibit COX-1; rather, dual suppression of COX-1 and COX-2 is necessary for damage to be produced.10 We found that selective inhibition of COX-1, while greatly reducing gastric prostaglandin synthesis, did not produce gastric damage.10 Selective inhibition of COX-2 did not cause any detectable suppression of gastric prostaglandin synthesis and did not cause gastric damage. However, when selective inhibitors of COX-1 and COX-2 were administered, gastric damage was consistently produced. These ndings were substantiated by the demonstration that a conventional NSAID with greater potency for inhibition of COX-1 than COX-224, ketorolac, did not produce gastric damage in the rat when given in a dose that markedly suppressed COX-1 activity but had no eect on COX-2.10 However, when doses were used that produced inhibition of both COX-1 and COX-2, gastric damage was consistently observed. These observations are consistent with studies of mice in which the gene for COX-1 had been disrupted. These mice exhibit negligible gastric prostaglandin synthesis, but no `spontaneous' gastric damage.25 Administration of a conventional NSAID (indomethacin), however, consistently resulted in the production of gastric damage. Taken together, these studies clearly show that both COX-1 and COX-2 signicantly contribute to gastrointestinal mucosal defence. Clinically, there is also evidence that suppression of both COX-1 and COX-2 is necessary for production of gastroduodenal ulcers. In a recent outcomes study, the use of low-dose aspirin by patients taking a selective COX-2 inhibitor (celecoxib), increased the incidence of ulcer complications such that there was no longer any benet, in terms of safety, as compared to the use of conventional NSAIDs plus lowdose aspirin.26 EFFECTS ON THE MICROCIRCULATION The component of mucosal defence that appears to be most profoundly altered by NSAIDs is the gastric microcirculatory response to injury. Normally, exposure of the mucosa to an irritant results in a rapid increase in mucosal blood ow. This is mediated via sensory aerent neurons and is probably aimed at removing any toxins or bacterial products that enter the lamina propria, neutralizing back-diusing acid and contributing to the formation of a microenvironment over sites of supercial mucosal injury that is conducive to repair.27 Impairment of this hyperaemic response has been shown to greatly increase the susceptibility of the mucosa to injury28 and to impair rapid repair of epithelial damage (restitution).27 NSAIDs have a very well characterized ability to reduce gastric mucosal blood ow.2931 Prostaglandins of the E and I series are potent vasodilators that are continuously produced by the vascular endothelium. Inhibition of their synthesis by an NSAID leads to an increase in vascular tone. In addition to these eects, NSAIDs also produce damage to the vascular endothelium.32,33 Indeed, this is a very early event

694 J. L. Wallace

following administration of NSAIDs to experimental animals32,33, and one that is also seen in ischaemiare-perfusion-associated damage in the gastrointestinal tract.34 In the latter, neutrophils have been implicated as playing a crucial role in the induction of endothelial injury.35 We therefore examined the role of the neutrophil in the pathogenesis of experimental NSAID-gastropathy. NSAID administration to rats resulted in a rapid and signicant increase in the number of neutrophils adhering to the vascular endothelium in both gastric and mesenteric venules.9,3638 This occurred within 1560 minutes of administration of the NSAID, consistent with the period of time required for suppression of prostaglandin synthesis by these drugs.27 The adherence of neutrophils to the endothelium was dependent on the expression of the b2 integrins (CD11/CD18) on the neutrophil and intercellular adhesion molecule-1 (ICAM1) on the vascular endothelium.38 Up-regulation of ICAM-1 on the vascular endothelium in the gastric microcirculation was shown to occur within 1530 minutes of administration of an NSAID to rats, and this could be prevented by administration of exogenous prostaglandins.39 Of course, the fact that NSAIDs caused neutrophils to adhere to the vascular endothelium does not necessarily mean that these cells contributed to the ensuing mucosal injury. To test this hypothesis, we examined the eects of NSAIDs in rats in which circulating neutrophils had been depleted, either with an anti-neutrophil serum or with methotrexate. These neutropenic rats developed little if any gastric damage or endothelial injury following administration of an NSAID.40 We next demonstrated that treating rabbits or rats with monoclonal antibodies that blocked NSAID-induced neutrophil adherence to the vascular endothelium also markedly reduced the extent of NSAID-induced gastric damage.33,38 A role for neutrophils in the pathogenesis of NSAID-gastropathy was further supported by the observation that administration of prostaglandins at doses previously shown to prevent gastric injury also prevented NSAID-induced leukocyte adherence.36,37 There are a number of ways in which adherence of neutrophils to the vascular endothelium contributes to the formation of gastroduodenal ulcers. Neutrophil adherence to the vascular endothelium is accompanied by the release of proteases (e.g. elastase, collagenase) and oxygen-derived free radicals (e.g. superoxide anion) from these cells. These substances may mediate much of the endothelial and epithelial injury caused by NSAIDs. Indeed, NSAID-induced mucosal injury can be markedly reduced by scavengers of oxygen-derived free radicals41,42 and by inhibitors of neutrophilderived proteases.43,44 Neutrophil adherence to the endothelium could also contribute to NSAID-induced mucosal injury by obstructing normal ow through capillaries, thereby reducing gastric mucosal blood ow. In this respect, it is noteworthy that the well-characterized ability of NSAIDs to reduce gastric blood ow2931 has been shown to occur subsequent to the appearance of `white thrombi' in the gastric microcirculation.30 Role of tumour necrosis factor-a Taken together, the studies described above suggest that, in response to suppression of prostaglandin synthesis by NSAIDs, there is a rapid up-regulation of expression of ICAM-1 on the vascular endothelium and an up-regulation of b2 integrin expression on circulating neutrophils, resulting in increased adherence of neutrophils to the endothelium. However, it remained to be determined whether the increase in expression of adhesion molecules occurred as a consequence of the reduction of prostaglandin synthesis, or was triggered by another chemical signal produced in response to NSAID administration. Santucci, Fiorucci and co-workers suggested that tumour necrosis

Pathogenesis of NSAID-ulcer 695

factor-a (TNFa) might be such a signal for NSAID-induced neutrophil adherence within the gastric microcirculation.45 Prostaglandins have well characterized inhibitory eects on the release of TNFa from macrophages and mast cells46,47, and TNFa is a wellcharacterized stimulus for the expression of adhesion molecules.48 The levels of TNFa in the plasma of rats have been shown to increase signicantly following administration of indomethacin, in parallel with the accumulation of neutrophils in the gastric microcirculation and with the development of gastric injury.45 Moreover, pre-treatment with pentoxifylline, an inhibitor of TNFa synthesis, dose-dependently reduced neutrophil accumulation in the gastric microcirculation and gastric damage.45 Further evidence supporting a role for TNFa in the pathogenesis of NSAID-gastropathy comes from the studies of Appleyard et al49 who demonstrated protective eects against NSAID-induced gastric damage in the rat of a number of structurally unrelated inhibitors of TNFa synthesis and of an anti-TNFa antibody.49

Role of leukotrienes Leukotrienes are another group of mediators that might contribute to the increase in neutrophil adherence and the mucosal injury that occurs after NSAID administration. Leukotrienes, like prostaglandins, are derived from arachidonic acid. They have been shown to increase the susceptibility of the gastric mucosa to injury36,5052, at least in part through stimulatory eects on neutrophil adherence to the vascular endothelium.36 Inhibitors of leukotriene synthesis and leukotriene receptor antagonists have been shown to exert protective eects in experimental NSAID-gastropathy.5052 There is also evidence of elevated leukotriene B4 production following NSAID administration to laboratory animals36 and humans.53 Inhibitors of leukotriene synthesis and an LTB4 receptor antagonist have been shown to prevent NSAIDinduced neutrophil adherence to the vascular endothelium in vivo36 and in vitro.54

Microcirculatory eects of inhibition of COX-1 versus COX-2 As mentioned above, suppression of prostaglandin synthesis in the gastrointestinal tract results in an impairment of mucosal defence. An important question in the light of the contribution of both COX-1 and COX-2 to mucosal defence, therefore, is: what is the mechanistic basis for the impairment of mucosal defence by suppression of COX-1 versus suppression of COX-2? Given recent evidence that prostacyclin synthesis in humans occurs primarily via COX-2, we postulated that suppression of COX-2 by NSAIDs may underlie at least some of the disturbances in the gastrointestinal microcirculation that occur after administration of these drugs.10 To do this, we employed selective inhibitors of COX-1 and COX-2. The reduction of gastric blood ow following NSAID administration appears to be due primarily to the inhibition of COX-1 as it could be mimicked by administration of a selective COX-1 inhibitor but not by a selective COX-2 inhibitor (Figure 1). On the other hand, the neutrophil adherence that is observed following NSAID administration appears to be due to suppression of COX-2. Administration of a selective COX-2 inhibitor (celecoxib) caused signicant adherence of neutrophils to the vascular endothelium of post-capillary mesenteric venules in rats, similar in magnitude to what was observed following administration of a conventional NSAID. In contrast, a selective COX-1 inhibitor had no eect on neutrophil adherence within mesenteric venules.10

696 J. L. Wallace
NSAID

COX-1 inhibition

Topical irritation

COX-2 inhibition

Decreased blood Flow

Epithelial damage

Neutrophil adherence

Mucosal injury

Figure 1. Schematic diagram illustrating three of the main components of the pathogenesis of NSAIDinduced gastrointestinal mucosal injury. Suppression of COX-1 causes a profound reduction of mucosal prostaglandin synthesis which probably impairs many components of mucosal defence. Among these, the reduction of mucosal blood ow may be most important. Suppression of COX-2 leads to an increase in the number of neutrophils adhering to the vascular endothelium in the gastrointestinal microcirculation, which has been shown to be an important factor in ulcer development. Many NSAIDs, particularly acidic ones, also exert topical irritant eects on the mucosa, leading to epithelial injury.

TOPICAL IRRITANT EFFECTS OF NSAIDS While NSAIDs are capable of causing gastroduodenal ulcers when administered parenterally5557, it is likely that the topical irritant properties of these drugs still make an important contribution to their overall ulcerogenic properties (Figure 1). The ability of aspirin directly to damage the epithelial lining of the stomach was rst suggested by Davenport in the 1960s.58 What he referred to as `breaking of the barrier' resulted in back-diusion of acid into the mucosa, leading to the rupture of mucosal blood vessels. These topical irritant properties are predominantly associated with acidic NSAIDs. The unionized forms of these drugs can enter epithelial cells in the stomach and duodenum where, within the neutral intracellular environment, they are converted to an ionized state. This has been referred to as `ion trapping'.59 As the drug accumulates within the epithelial cell, osmotic movement of water results in cellular swelling, eventually to the point of lysis.59,60 NSAIDs may also damage the gastroduodenal epithelium via uncoupling of mitochondrial respiration in epithelial cells.60,61 Various NSAIDs have been shown to uncouple oxidative phosphorylation60,61, leading to a depletion of ATP and therefore a reduced ability to regulate normal cellular functions, such as maintenance of intracellular pH. The ability of NSAIDs to uncouple oxidative phosphorylation may also be related to some extent to acidic moieties (such as carboxylic acid residues) because substitutions at these sites interfere with the ability of the compounds to act as uncouplers.61 A third mechanism through which NSAIDs may exert topical irritant eects on the mucosa is through their ability to decrease the hydrophobicity of the mucus gel layer in the stomach. Lichtenberger and co-workers have proposed that this layer is a

Pathogenesis of NSAID-ulcer 697

primary barrier to acid-induced damage in the stomach.62,63 They have demonstrated that gastric mucosal surface is hydrophobic. Moreover, this hydrophobicity can be reduced by various pharmacological agents, including NSAIDs6365, and this eect persisted for long periods of time after stopping treatment with NSAIDs.62,64 Many attempts have been made to modify NSAIDs such that their topical irritant eects are reduced. These include formulation of the drugs in slow-release or enteric coated tablets, and preparation of the drug as a pro-drug that requires hepatic metabolism in order to be active (e.g. sulindac). However, the incidence of gastroduodenal ulceration with pro-drugs is comparable to what is seen with standard NSAIDs6668, an observation which has been used to support the view that topical irritant properties of NSAIDs are not paramount in terms of their ability to induce frank ulceration in the stomach. However, this possibility cannot be completely excluded because most NSAIDs are excreted in bile. Thus, even with parenteral administration of the NSAID, reux of duodenal contents into the stomach would result in topical exposure of the gastric mucosa to these drugs. IMPAIRMENT OF REPAIR PROCESSES In addition to being able to induce ulcers in the stomach and duodenum, NSAIDs can interfere with the healing of pre-existing ulcers. Indeed, NSAIDs are able to interfere with the rapid repair processes that come into play when the surface epithelium is damaged (i.e. restitution), as well as interfering with the more prolonged process of healing when a penetrating ulcer forms in the stomach or duodenum. Restitution Damage to the gastric epithelium is not an usual event, but progression of this type of damage to an ulcer is rare. One of the reasons for this is that the stomach and duodenum are able to repair epithelial injury very rapidly (i.e. within minutes) via the process of restitution. Restitution involves the rapid migration of healthy cells from the gastric pits over areas which have been denuded; thus, an intact epithelial barrier is rapidly reestablished.69 The cells move along the basement membrane, which has been shown to be crucial to the restitution process.6972 The basement membrane can be damaged by acid, which results in impairment of restitution and progressive erosion of deeper layers of the mucosa.27,73,74 Damage to the basement membrane is prevented, in most situations, by the formation over sites of injury of a microenvironment in which the pH is maintained at near neutral.27,73 A `mucoid cap' consisting of mucus, cellular debris and plasma proteins forms within seconds of gastric epithelial injury. This layer traps the plasma that leaks from the underlying microcirculation.73 It is the plasma that accounts for the near-neutral pH within the protective mucoid cap because even a very brief interruption of mucosal blood ow results in a rapid decrease in the pH within the mucoid cap and this, in turn, results in the formation of haemorrhagic erosions.33 As reviewed above, NSAIDs decrease mucosal blood ow. In doing so, NSAIDs can interfere with the appropriate functioning of the mucoid cap and therefore with the process of restitution. Indeed, we observed that following systemic administration of an NSAID to a rat in which supercial epithelial damage had been induced, the pH within the mucoid cap began to decline in parallel (temporally) with the inhibition of prostaglandin synthesis .33 Within minutes, the formation of haemorrhagic erosions was clearly evident. These eects could be prevented by administration of prostaglandins.33

698 J. L. Wallace

Ulcer healing The ability of NSAIDs to delay the healing of ulcers and promote their bleeding is well characterized.75,76 The eects on ulcer healing are probably related to the ability of NSAIDs to suppress prostaglandin synthesis. At a site of ulceration, and particularly around the ulcer margin, cyclo-oxygenase-2 is the primary contributor to prostaglandin synthesis. Studies in mice and rats initially demonstrated the marked up-regulation of cyclo-oxygenase-2 in ulcerated gastric tissue21,22, and this has been conrmed in humans.77 Treatment of laboratory animals with selective inhibitors of cyclo-oxygenase2 resulted in signicant delaying of ulcer healing.21,22 These observations, and reports of selective cyclo-oxygenase-2 inhibitors exacerbating intestinal inammation and ulceration23, suggest that caution should be exercised in regarding the new cyclooxygenase-2 inhibitors as `GI-safe'.78,79 Ulcer bleeding is generally believed to be due to suppression of platelet aggregation by NSAIDs. Such an eect is due to inhibition of COX-1 within the platelet. However, recent data suggest that ulcer bleeding may be due to other factors. While the use of COX-2 inhibitors is associated with less ulcer formation than that which occurs with conventional NSAIDs, the proportion of those ulcers that bleed is the same.80

ROLE OF ACID NSAID-induced ulcers have been observed in patients who are achlorhydric81,82, suggesting that gastric acid secretion does not play a role in the pathogenesis of NSAIDinduced gastroduodenal ulcers. Early studies of histamine H2 receptor antagonists for prevention of NSAID-induced ulcers yielded unpromising results, further suggesting that acid was not an important contributor to these ulcers.8385 More recently, however, studies employing higher doses of H2 antagonists or using proton pump inhibitors have clearly demonstrated signicant benecial eects in terms of preventing NSAID-induced ulcers.86,87 The results of these studies suggest that profound suppression of acid secretion is necessary in order to have a signicant impact on the incidence of NSAID-induced ulcers. There are a number of ways in which acid could contribute to NSAID-induced ulcer formation. Acid can certainly exacerbate damage to the gastric mucosa induced by other agents.69 Acid can also interfere with haemostasis by limiting the ability of platelets to aggregate.88 As outlined above, acid can interfere with the process of restitution, resulting in the conversion of supercial injury to deeper mucosal necrosis. Finally, as several growth factors (e.g. broblast growth factor) that are important for maintenance of mucosal integrity and for repair of supercial injury are acid-labile, the presence of acid in the lumen can limit the contributions of these factors to mucosal defence and repair.89 It is important to note that NSAIDs can increase gastric acid secretion, although it is not clear whether such eects have any impact on ulcer formation or healing. Prostaglandins exert inhibitory eects on parietal cells90, so the inhibition of their synthesis by NSAIDs can result in an increase in gastric acid secretion.91 It was recently reported that this eect of NSAIDs is attributable to inhibition of COX-1 rather than COX-2.92 Indeed, even in the inamed stomach, COX-2-derived prostaglandins do not appear to exert any signicant eect on gastric acid secretion.92

Pathogenesis of NSAID-ulcer 699

Role of Helicobacter pylori The two greatest risk factors for peptic ulcer disease are use of NSAIDs and colonization of the upper gastrointestinal tract by Helicobacter pylori. Because these two factors produce gastroduodenal injury through dierent mechanisms, they may have additive eects in terms of producing ulcers. This is quite a controversial issue, with conicting results having been reported as to the presence or absence of such an interaction. H. pylori has been shown to enhance gastric mucosal prostaglandin synthesis, while NSAIDs suppress it.93 Thus, it has been suggested that infection with H. pylori could actually be protective against NSAID-induced ulceration. Hawkey has suggested that `as a broad generalisation, in therapeutic studies of NSAID users, those who have no ulcer at trial entry are more prone to ulcer development if they are H. pylori-positive'.93 However, in patients who have ulcers at baseline, H. pylori-positive individuals are less likely to develop ulcers, particularly if taking acid-suppressive therapy. In a study of patients who were starting NSAIDs for the rst time, with no history of a previous ulcer and no ulcer at the start of the study, the use of bismuth-based eradication therapy was associated with a much lower incidence of gastric ulcer after 2 months.94 On the other hand, eradication of H. pylori has been associated, paradoxically, with a slower healing of gastric ulcers when compared to patients who did not undergo eradication therapy. In the case of duodenal ulcers, there appears to be more clear evidence that eradication of H. pylori is benecial (whether or not the patient continues to take NSAIDs).93

SUMMARY Substantial progress has been made over the past three decades towards better understanding of the pathogenesis of gastroduodenal ulcers induced by NSAIDs. This had led to the development of several new types of NSAID that have greatly reduced gastrointestinal toxicity relative to conventional NSAIDs. For example, selective inhibitors of cyclo-oxygenase-2, which spare the major form of cyclo-oxygenase in the gastrointestinal tract, produce fewer gastric and duodenal ulcers than do the conventional NSAIDs that inhibit both cyclo-oxygenase-1 and -2. Some questions remain about the eectiveness of selective COX-2 inhibitors versus conventional NSAIDs, particularly with respect to producing analgesia, and with respect to toxicity in other organs (e.g. cardiovascular system). Nitric oxide-releasing NSAIDs represent another approach to reducing the adverse eects of this class of drugs.95,96 There is also considerable evidence that concomitant use of proton-pump inhibitors can greatly reduce the gastrointestinal toxicity of conventional NSAIDs. When cost issues are taken into consideration, this approach may have advantages over the use of some of the newer NSAIDs, such as selective COX-2 inhibitors.

Practice points . where an NSAID is needed, choose the NSAID with the lowest toxicity prole . in high-risk patients, co-prescribe a proton-pump inhibitor with a conventional NSAID or with a COX-2 inhibitor . caution should be exercised in use of COX-2 inhibitors in patients taking lowdose aspirin

700 J. L. Wallace

Research agenda . further dene the critical events which lead to gastroduodenal injury following NSAID administration (at a molecular level) . better understand the contribution of prostaglandins derived from COX-2 in mucosal defence and repair . design better anti-inammatory drugs that lack GI toxicity but have an ecacy similar to, or better than, conventional NSAIDs

Acknowledgements
Dr Wallace is a Medical Research Council of Canada Senior Scientist and an Alberta Heritage Foundation for Medical Research Scientist. He holds the Crohn's and Colitis Foundation of Canada Chair in Intestinal Disease Research.

REFERENCES
1. Douthwaite AH & Lintott SAM. Gastroscopic observation of the eect of aspirin and certain other substances on the stomach. Lancet 1938; 2: 12221225. 2. Gabriel SE & Bombardier C. NSAID induced ulcers. An emerging epidemic? Journal of Rheumatology 1990; 17: 14. 3. Singh G, Ramey DR, Morfeld D et al. Gastrointestinal tract complications of nonsteroidal antiinammatory drug treatment in rheumatoid arthritis. A prospective observational cohort study. Archives of Internal Medicine 1996; 156: 15301536. 4. Vane JR. Inhibition of prostaglandin synthesis as a mechanism of action for aspirin-like drugs. Nature 1971; 231: 232235. 5. Robert A, Schultz JR, Nezamis JE & Lancaster C. Gastric antisecretory and antiulcer properties of PGE2 , 15-methyl PGE2 , and 16,16-dimethyl PGE2. Gastroenterology 1976; 70: 359370. 6. Lanza FL. A review of gastric ulcer and gastroduodenal injury in normal volunteers receiving aspirin and other nonsteroidal anti-inammatory drugs. Scandinavian Journal of Gastroenterology 1989; 24 (supplement 163): 2431. 7. Rainsford KD & Willis C. Relationship of gastric mucosal damage induced in pigs by antiinammatory drugs to their eects on prostaglandin production. Digestive Diseases and Sciences 1982; 27: 624635. 8. Whittle BJR. Temporal relationship between cyclooxygenase inhibition, as measured by prostacyclin biosynthesis, and the gastrointestinal damage induced by indomethacin in the rat. Gastroenterology 1981; 80: 9498. 9. Wallace JL, McCaerty D-M, Carter L et al. Tissue-selective inhibition of prostaglandin synthesis in rat by tepoxalin: anti-inammatory without gastropathy. Gastroenterology 1993; 105: 16301636. *10. Wallace JL, McKnight W, Reuter BK & Vergnolle N. NSAID-induced gastric damage in the rat: requirement for inhibition of both cyclooxygenase-1 and -2. Gastroenterology 2000; 119: 706714. 11. Wallace JL. Nonsteroidal anti-inammatory drugs and gastroenteropathy: the second hundred years. Gastroenterology 1997; 112: 10001016. 12. Wallace JL & Granger DN. The cellular and molecular basis for gastroduodenal mucosal defense. FASEB Journal 1996; 10: 731740. 13. Wallace JL. Cooperative modulation of gastrointestinal mucosal defence by prostaglandins and nitric oxide. Clinical and Investigative Medicine 1996; 19: 346351. 14. Wallace JL & Miller MJS. Nitric oxide and mucosal defence: a little goes a long way. Gastroenterology 2000; 119: 512520. 15. Ferraz JG, McKnight W, Sharkey KA & Wallace JL. Impaired vasodilatory responses in the gastric microcirculation of cirrhotic rats. Gastroenterology 1995; 108: 11831191. 16. Kargman S, Charleson S, Cartwright M et al. Characterization of prostaglandin G/H synthase 1 and 2 in rat, dog, monkey, and human gastrointestinal tracts. Gastroenterology 1996; 111: 448454. 17. Davies NM, Sharkey KA, Asfaha S et al. Aspirin induces a rapid up-regulation of cyclooxygenase-2 expression in the rat stomach. Alimentary Pharmacology and Therapeutics 1997; 11: 11011108.

Pathogenesis of NSAID-ulcer 701 18. Ferraz JGP, Sharkey KA, Reuter BK et al. Induction of cyclooxygenase-1 and -2 in the rat stomach during endotoxemia: role in resistance to damage. Gastroenterology 1997; 113: 195204. 19. Maricic N, Ehrlich K, Gretzer B et al. Selective cyclo-oxygenase-2 inhibitors aggravate ischaemiareperfusion injury in the rat stomach. British Journal of Pharmacology 1999; 128: 16591666. *20. Gretzer B, Ehrlich K, Maricic N et al. Selective cyclo-oxygenase-2 inhibitors and their inuence on the protective eect of a mild irritant in the rat stomach. British Journal of Pharmacology 1998; 123: 927935. 21. Mizuno H, Sakamoto C & Matsuda K. Induction of cyclooxygenase-2 in gastric mucosal lesions and its inhibition by the specic antagonist delays healing in mice. Gastroenterology 1997; 112: 387397. 22. Schmassmann A, Peskar BM, Stettler C et al. Eects of inhibition of prostaglandin endoperoxide synthase-2 in chronic gastro-intestinal ulcer models in rats. British Journal of Pharmacology 1998; 123: 795804. *23. Reuter BK, Asfaha S, Buret A et al. Exacerbation of inammation-associated colonic injury in rat through inhibition of cyclooxygenase-2. Journal of Clinical Investigation 1996; 98: 20762085. 24. Warner TD, Giuliano F, Vojnovic I et al. Nonsteroid drug selectivities for cyclo-oxygenase-1 rather than cyclo-oxygenase-2 are associated with human gastrointestinal toxicity: a full in vitro analysis. Proceedings of the National Academy of Sciences of the USA 1999; 96: 75637568. 25. Langenbach R, Morham SG, Tiano HF et al. Prostaglandin synthase 1 gene disruption in mice reduces arachidonic acid-induced inammation and indomethacin-induced gastric ulceration. Cell 1995; 83: 483492. *26. Silverstein FE, Faich G, Goldstein JL et al. Gastrointestinal toxicity with celecoxib vs nonsteroidal antiinammatory drugs for osteoarthritis and rheumatoid arthritis. Journal of the American Medical Association 2000; 284: 12471255. 27. Wallace JL & McKnight GW. The mucoid cap over supercial gastric damage in the rat. A highpH microenvironment dissipated by nonsteroidal antiinammatory drugs and endothelin. Gastroenterology 1990; 99: 295304. 28. Holzer P, Livingston EH & Guth PH. Sensory neurons signal for an increase in rat gastric mucosal blood ow in the face of pending acid injury. Gastroenterology 1991; 101: 416423. 29. Ashley SW, Sonnenschein LA & Cheung LY. Focal gastric mucosal blood ow at the site of aspirininduced ulceration. American Journal of Surgery 1985; 149: 5359. 30. Kitahora T & Guth PH. Eect of aspirin plus hydrochloric acid on the gastric mucosal microcirculation. Gastroenterology 1987; 93: 810817. 31. Gana TJ, Huhlewych R & Koo J. Focal gastric mucosal blood ow in aspirin-induced ulceration. Annals of Surgery 1987; 205: 399403. 32. Rainsford KD. Microvascular injury during gastric damage by anti-inammatory drugs in pigs and rats. Agents and Actions 1983; 13: 457460. 33. Wallace JL, Keenan CM & Granger DN. Gastric ulceration induced by nonsteroidal anti-inammatory drugs is a neutrophil-dependent process. American Journal of Physiology 1990; 259: G462G467. 34. Granger DN, Kvietys PR & Perry MA. Leukocyte-endothelial cell adhesion caused by ischemia and reperfusion. Canadian Journal of Physiology and Pharmacology 1993; 71: 6775. 35. Hernandez LA, Grisham MB, Twohig B et al. Granulocytes: the culprit in ischemic damage to the intestine. American Journal of Physiology 1987; 253: H669H703. 36. Asako H, Kubes P, Wallace JL et al. Indomethacin-induced leukocyte adhesion in mesenteric venules: role of lipoxygenase products. American Journal of Physiology 1992; 262: G903G908. 37. Asako H, Kubes P, Wallace JL et al. Modulation of leukocyte adhesion to rat mesenteric venules by aspirin and salicylate. Gastroenterology 1992; 103: 146152. 38. Wallace JL, McKnight W, Miyasaka M et al. Role of endothelial adhesion molecules in NSAID-induced gastric mucosal injury. American Journal of Physiology 1993; 265: G993G998. 39. Andrews FJ, Malcontenti-Wilson C & O'Brien PE. Eect of nonsteroidal anti-inammatory drugs on LFA1 and ICAM-1 expression in gastric mucosa. American Journal of Physiology 1994; 266: G657G664. 40. Wallace JL, Arfors K-E & McKnight GW. A monoclonal antibody against the CD18 leukocyte adhesion molecule prevents indomethacin-induced gastric damage in the rabbit. Gastroenterology 1991; 100: 878883. 41. Del Soldato P, Foschi D, Benoni G & Scarpignato C. Oxygen free radicals interact with indomethacin to cause gastrointestinal injury. Agents and Actions 1985; 17: 484488. 42. Vaananen PM, Meddings JB & Wallace JL. Role of oxygen-derived free radicals in indomethacin-induced gastric injury. American Journal of Physiology 1991; 261: G470G475. 43. Yoshida N, Cepinskas G, Granger DN et al. Aspirin-induced, neutrophil-mediated injury to vascular endothelium. Inammation 1995; 19: 297312. 44. Murakami K, Okajima K, Harada N et al. Plaunotol prevents indomethacin-induced gastric mucosal injury in rats by inhibiting neutrophil activation. Alimentary Pharmacology and Therapeutics 1999; 13: 521520.

702 J. L. Wallace 45. Santucci L, Fiorucci S, Giansanti M et al. Pentoxifylline prevents indomethacin induced acute gastric mucosal damage in rats: role of tumour necrosis factor alpha. Gut 1994; 35: 909915. 46. Hogaboam CM, Bissonnette EY, Chin BC et al. Prostaglandins inhibit inammatory mediator release from rat mast cells. Gastroenterology 1993; 104: 122129. 47. Kunkel SL, Wiggins RC, Chensue SW & Larrick J. Regulation of macrophage tumor necrosis factor production by prostaglandin E2. Biochemical and Biophysics Research Communications 1986; 137: 404410. 48. Rothlein R, Czajkowski M, O'Neill MM et al. Induction of intercellular adhesion molecule 1 on primary and continuous cell lines by pro-inammatory cytokines. Regulation by pharmacologic agents and neutralizing antibodies. Journal of Immunology 1988; 141: 16651669. 49. Appleyard CB, McCaerty DM, Tigley AW et al. Tumour necrosis factor mediation of NSAID-induced gastric damage: role of leukocyte adherence. American Journal of Physiology 1996; 270: G42G48. 50. Rainsford KD. The eects of 5-lipoxygenase inhibitors and leukotriene antagonists on the development of gastric lesions induced by nonsteroidal antiinammatory drugs in mice. Agents and Actions 1987; 21: 316319. 51. Pihan G, Rogers C & Szabo S. Vascular injury in acute gastric mucosal damage: mediatory role of leukotrienes. Digestive Diseases and Sciences 1988; 33: 625632. 52. Vaananen PM, Keenan CM, Grisham MB & Wallace JL. A pharmacological investigation of the role of leukotrienes in the pathogenesis of experimental NSAID-gastropathy. Inammation 1992; 16: 227240. 53. Hudson N, Balsitis M, Everitt S & Hawkey CJ. Enhanced gastric mucosal leukotriene B4 synthesis in patients taking non-steroidal anti-inammatory drugs. Gut 1993; 34: 742747. 54. Yoshida N, Takemura T, Granger DN et al. Molecular determinants of aspirin-induced neutrophil adherence to endothelial cells. Gastroenterology 1993; 105: 715724. 55. Wallace JL & McKnight GW. Characterization of a simple animal model for nonsteroidal antiinammatory drug induced antral ulcer. Canadian Journal of Physiology and Pharmacology 1993; 71: 447452. 56. Henry D, Dobson A & Turner C. Variability in the risk of major gastrointestinal complications from nonaspirin nonsteroidal anti-inammatory drugs. Gastroenterology 1993; 10: 10781088. 57. Estes LL, Fuhs DW, Heaton AH & Butwinick CS. Gastric ulcer perforation, associated with the use of injectable ketorolac. Annals of Pharmacotherapy 1993; 27: 4243. *58. Davenport HW. Gastric mucosal hemorrhage in dogs. Eects of acid, aspirin, and alcohol. Gastroenterology 1969; 56: 439449. 59. Fromm D. How do non-steroidal anti-inammatory drugs aect gastric mucosal defenses? Clinical and Investigative Medicine 1987; 10: 251258. 60. Somasundaram S, Hayllar H, Ra S et al. The biochemical basis of non-steroidal anti-inammatory druginduced damage to the gastrointestinal tract: a review and a hypothesis. Scandinavian Journal of Gastroenterology 1995; 30: 289299. 61. Mahmud T, Wrigglesworth JM, Scott DL & Bjarnason I. Mitochondrial function and modication of NSAID carboxyl moiety. Inammopharmacology 1996; 42: 189194. *62. Lichtenberger LM. The hydrophobic barrier properties of gastrointestinal mucus. Annual Review of Physiology 1995; 57: 565583. 63. Goddard PJ, Hills BA & Lichtenberger LM. Does aspirin damage canine gastric mucosa by reducing its hydrophobicity? American Journal of Physiology 1987; 252: G421G430. 64. Lichtenberger LM, Wang ZM, Romero JJ et al. Non-steroidal anti-inammatory drugs (NSAIDs) associate with zwitterionic phospholipids: insight into the mechanism and reversal of NSAID-induced gastrointestinal injury. Nature Medicine 1995; 1: 154158. 65. Lichtenberger LM, Wang ZM, Giraud MN et al. Eect of naproxen on gastric mucosal hydrophobicity. Gastroenterology 1995; 108: A149. 66. Graham DY, Smith JL, Holmes GI & Davies RO. Nonsteroidal anti-inammatory eect of sulindac sulfoxide and sulde on gastric mucosa. Clinical Pharmacology and Therapeutics 1985; 38: 6570. 67. Carson JL, Strom BL, Soper KA et al. The relative gastrointestinal toxicity of the nonsteroidal antiinammatory drugs. Archives of Internal Medicine 1987; 147: 10541059. 68. Hawthorne AB, Mahida YR, Cole AT & Hawkey CJ. Aspirin-induced gastric mucosal damage: prevention by enteric-coating and relation to prostaglandin synthesis. British Journal of Clinical Pharmacology 1991; 32: 783. 69. Morris GP & Wallace JL. The roles of ethanol and of acid in the production of gastric mucosal erosions in rats. Virchows Archiv B (Cell Pathology) 1981; 38: 2338. 70. Svanes K, Ito S, Takeuchi K & Silen W. Restitution of the surface epithelium of the in vitro frog gastric mucosa after damage with hyperosmolar sodium chloride. Gastroenterology 1982; 82: 14091426. *71. Lacy ER & Ito S. Rapid epithelial restitution of the rat gastric mucosa after ethanol injury. Laboratory Investigation 1984; 51: 573583.

Pathogenesis of NSAID-ulcer 703 72. Moore R, Carlson S & Madara JL. Rapid barrier restitution in an in vitro model of intestinal epithelial injury. Laboratory Investigation 1989; 60: 237244. 73. Wallace JL & Whittle BJR. Role of mucus in repair of gastric epithelial damage in the rat. Inhibition of epithelial recovery by mucolytic agents. Gastroenterology 1986; 91: 603611. 74. Black BA, Morris GP & Wallace JL. Eects of acid on the basal lamina of the rat stomach and duodenum. Virchows Archiv B (Cell Pathology) 1985; 50: 109118. 75. Armstrong CP & Blower AL. Non-steroidal anti-inammatory drugs and life threatening complications of peptic ulceration. Gut 1997; 28: 527532. 76. Stadler P, Armstrong D, Margalith D et al. Diclofenac delays healing of gastroduodenal mucosal lesions. Double-blind, placebo-controlled endoscopic study in healthy volunteers. Digestive Diseases and Sciences 1991; 36: 594600. 77. Jackson LM, Wu KC, Mahida YR et al. Cyclooxygenase (COX) 1 and 2 in normal, inamed, and ulcerated human gastric mucosa. Gut 2000; 47: 762770. 78. Wallace JL. Selective COX-2 inhibitors: is the water becoming muddy? Trends in Pharmacological Sciences 1999; 20: 46. 79. Wallace JL, McKnight W & Bak A. Selective inhibitors of COX-2: are they really eective? Are they really GI-safe? Joumal of Clinical Gastroenterology 1998; 27: S28S34. *80. Bombardier C, Laine L, Recin A et al. Comparison of upper gastrointestinal toxicity of rofecoxib and naproxen in patients with rheumatoid arthritis. New England Journal of Medicine 2000; 343: 15201528. 81. Rashid F & Toolon EP. Misoprostol healed a benign nonsteroidal anti-inammatory drug-induced gastric ulcer in a patient with pentagastrin-fast achlorhydria. Journal of Clinical Gastroenterology 1993; 16: 219221. 82. Janssen M, Dijkmans BAC, Vanderbroucke JP et al. Achlorhydria does not protect against benign upper gastrointestinal ulcers during NSAID use. Digestive Diseases and Sciences 1994; 39: 362365. 83. Agrawal NM. Epidemiology and prevention of non-steroidal anti-inammatory drug eects in the gastrointestinal tract. British Journal of Rheumatology 1995; 34 (supplement 1): 510. 84. Koch M, Capurso L, Dezi A et al. Prevention of NSAID-induced gastroduodenal mucosal injury: metaanalysis of clinical trials with misoprostol and H2-receptor antagonists. Digestive Diseases 1995; 13 (supplement 1): 6274. 85. Simon TJ, Berger ML, Hoover ME et al. A dose-ranging study of famotidine in prevention of gastroduodenal lesions associated with non-steroidal anti-inammatory drugs (NSAIDs): results of a U.S. multicenter trial. American Journal of Gastroenterology 1994; 89: 1644 (abstract). 86. Taha AS, Hudson N, Hawkey CJ et al. Famotidine for the prevention of gastric and duodenal ulcers caused by nonsteroidal antiinammatory drugs. New England Journal of Medicine 1996; 334: 14351439. *87. Hawkey CJ, Karrasch JA, Szczepanski L et al. Omeprazole compared with misoprostol for ulcers associated with nonsteroidal antiinammatory drugs. New England Journal of Medicine 1998; 338: 727734. 88. Green FW, Kaplan MM, Curtis LE & Levine PH. Eect of acid and pepsin on blood coagulation and platelet aggregation: a possible contributor to prolonged gastroduodenal mucosal hemorrhage. Gastroenterology 1978; 74: 3843. 89. Szabo S, Folkman J, Vattay P et al. Accelerated healing of duodenal ulcers by oral administration of a basic broblast growth factor in rats. Gastroenterology 1994; 106: 11061111. 90. Soll AH. Mechanisms of action of antisecretory drugs: studies on isolated canine fundic mucosal cells. Scandinavian Journal of Gastroenterology 1986; 21: 16. 91. Ligumsky M, Goto Y & Yamada T. Prostaglandins mediate inhibition of gastric acid secretion by somatostatin in the rat. Science 1983; 219: 301303. 92. Barnett K, Bell C, McKnight W et al. Role of cyclooxygenase-2 in modulating gastric acid secretion in the normal and inamed rat stomach. American Journal of Physiology 2000; 279: G1292G1297. 93. Hawkey CJ. What consideration should be given to Helicobacter pylori in treating nonsteroidal antiinammatory drug ulcers? European Journal of Gastroenterology and Hepatology 2000; 12 (supplement 1): S17S20. 94. Chan FK, Sung JJ, Chung SC et al. Randomised trial of eradication of Helicobacter pylori before nonsteroidal anti-inammatory drug therapy to prevent peptic ulcers. Lancet 1997; 350: 975979. *95. Wallace JL, Reuter B, Cicala C et al. Novel NSAID derivatives with markedly reduced ulcerogenic properties. Gastroenterolgy 1994; 107: 173179. 96. Wallace JL, Elliott SN, Del Soldato P et al. Gastrointestinal sparing anti-inammatory drugs: the development of nitric oxide-releasing NSAIDs. Drug Development Research 1997; 42: 144149.

S-ar putea să vă placă și