Bioscience Reports, VoL !2, No. 5, I992

REVIEW

Pharmacological Control of Gastric Acid Secretion" Molecular and Cellular Aspects Ladislav Mirossay, ~ Yolande Di Gioia, Eric Chastre, Shahin Emami, and Christian Gespach 2 Received ls~ version December 6, 1991; revised version May 20, 1992 KEY WORDS: gastric acid secretion; growth factor; harmone; receptor-signal transduction.

INTRODUCTION Chronic ulceration of the gastroduodenal mucosa is one of the most common, so called "civilization diseases" of the gastrointestinal tract in industrially developed countries. Although the primary cause of gastric and duodenal peptic ulcers is unknown, gastric acid secretion is believed to play an important role in this major pathology. Gastric peptic ulcers, which represent 20% of all peptic ulcers (472) are mostly located in the antrum adjacent to the border of the acid producing fundic mucosa. Epidemiologically, gastric cancer is not reported to be increased in gastric ulcer disease. However, gastric endocrine cell hyperplasia and gastric carcinoid tumors have been described in two naturally occurring hypergastrinemic states, gastric atrophy with achlorhydria and Zollinger~Ellison syndrome (44, 45, 64, 363,408). Many hormones, growth factors, and local bioregulators are able to influence gastric secretions and to induce the ulcerogenic cascade or to preserve the integrity of the mucosa. This paper analyses different aspects of the molecular and genetic mechanisms involved in the pathophysiology of gastric secretions. At the same time it describes the complex interconnections between the signal transduction systems implied in the regulation of the secretory activity of parietal cells, and adjacent mucosal cells. Particular reference is focused on the development of the histamine H 2 receptor antagonists and the H +, K+-ATPase inhibitors in the treatment of peptic ulcer disease. Institut National de ia Sant6 et de la Recherche M6dicale INSERM U. 55, Unit6 de Recherches stir les Peptides Neurodigestifs et le Diab~te, H6pital Saint-Antoine, 75571 Paris Cedex 12. France. Present address: Department of Pharmacology, Medical Faculty, P. J. Safarik University, Tr. SNP 1, 04001 Kosice, Czechoslovakia. z To whom all correspondence should be addressed, at INSERM U55. 319

0144-8463/92/1000-0319506,50/0 9 i992 PlenumPublishingCorporation

320

Mirossay, Di Gioia, Chastre, Emami and Gespach PATHOPHYSIOLOGICAL BASIS

Gastric glands as a part of the lamina propria, which constitutes a vascularized and innervated conjunctive network, are divided into three sections. The crypt is composed of mucus cells in the antrum and of chief cells (pepsin-secreting cells), mucous and acid- secreting parietal cells in the fundus. The collar neck, also named isthmus, is composed of immature cells which can differentiate and rapidly renew the gastric mucosa (205,238-241). The pit cells or surface mucous cells, originate from this zone and migrate upwards, towards the mucosal surface while the other three types of cells migrate towards the base of the gland. The time required for the newly formed cells to reach the lower end of the gland varies between one hundred and three hundred days. On the other hand, several authors have reported evidence that the chief cells are able to divide, indicating that its renewal is assured by its own mitotic activity. In the mouse, the turnover time for the isthmal cells was estimated to be 16 h, while the renewal time of pit cells was assessed at 3 days in the pyloric antrum (238). The origin of gastric endocrine cells is still a matter of debate. The original A P U D hypothesis (amine precursor uptake and decarboxylation) postulated that some mucosal cells might constitute a separate lineage, derived from the neural crest after migration of the endocrine precursor cells. In this precedent terminology, A P U D cells have the capacity to capture histamine and serotonin precursors, (i.e. histidine and 5-hydroxytryptophane, respectively) and to convert them to bioamines by the action of histidine decarboxylase and DOPAdecarboxylase (228, 445). However, observations of differentiation patterns in cancer and embryological studies indicate that the gut endocrine cells are of endodermal origin, as are the other gastrointestinal epithelial cells. Recent studies have clearly demonstrated the proliferative activity of gastrin and somatostatin cells in the stomach (242). The cells constituting the gastric mucosa are, therefore, a fairly heterogeneous population in terms of proliferation activity and functional differentiation. The gastric epithelium assume a vital protective function since the repair of chemically induced gastric epithelial injury is completed in 30-60 rain by active migration of surviving cells (205). The growth of the gastric mucosa in hypophysectomized rats is partially restored by growth hormone (GH), suggesting atrophic role for G H in the upper gastrointestinal tract and regulation of pepsinogen secretion in gastric glands (254). While it is now recognized that histamine plays a central role in the regulation of acid secretion, the diversity of effectors that can elevate or diminish intragastric pH must also be considered. In man and rodents, histamine of the gastric mucosa is stored in ECL-type enterochromaffin endocrine cells, previously designated histaminocytes (338). In various animal species and human stomach, histamine also accumulates in the mastocytes (mast cells) of the lamina propria, which lie close to the tubular Structures of the mucosa (26,228). The localization of histamine in the human gastric mucosa occurs in two separate cell populations: endocrine ECL cells, restricted to the oxyntic gland area, and mast cells, disseminated throughout the wall of the antrum as welt as in the body of the stomach (255,256). Mucosal cells with hormonal or paracine activity, the

Pharmacological Control of Gastric Acid Secretion

32'.[

myenteric plexus, and the nerve endings in the smooth muscle layer and submucosa contain somatostatin, gastrin releasing peptide (GRP), vasoactive intestinal peptide (VIP), the human peptide HM with N-terminal histidine and a C-terminal methionine amide (PHM), serotonin in gastric EC cells, tachykinins, and enkephalins (107, 120, 201, 256, 416, 461). Corresponding to these bioamines, hormones and neuropeptides a set of specific receptors regulate exocrine and endocrine secretions in the stomach, as well as gastric emptying (151,160, 359,360). Digestive secretions are regulated by numerous substances, hormones and neurotransmitters, which act through the circulatory system or which are secreted in the interstitial fluid. Paracrine and neuroparacrine regulations involve a single cell or nerve fiber which release their contents in the vicinity of the neighboring effector cell (Fig. 1). For example, the somatostatin-producing D cells have long,

? H+

-U-

EPSIN B I

'*'

A T E

Fig. 1. Hormonal, paracrine and autocrine regulations of gastric secretion. Mast cells (MC) and ECL histaminocytes induce the release of gastrin by antral and duodenal G ceils. Acidification of the intestinal lumen results in secretin release by duodenal secretin S cells and induction of somatostatin release by gastric D cells. Adrenergic (ADR), cholinergic (ACH) or VIPergic neurones regulate endocrine and exocrine secretions by the gastric mucosa. Autocrine activity of prostaglandins (PG) in parietal cells (H+), chief cells (pepsin), and mucous cells participates on this process. (+) and ( - ) : stimulatory and inhibitory effects, respectively.

322

Mirossay, Di Gioia, Chastre, Emami and Gespach

nonluminal cytoplasmic processes which establish direct contacts with parietal cells, gastrin (G) cells, and histaminocytes in human and rat gastric epithelium (468). In the fundus, somatostatin secretion is regulated indirectly by gastrinreleasing peptide (GRP), via stimulatory noncholinergic neurons and inhibitory cholinergic neurons (389). In the antrum, GRP acts indirectly on somatostatin cells by stimulating gastrin release and by activating inhibitory cholinergic neurons. In autocrine regulation, the effector cell itself synthesizes or sequesters a product capable of modulating its own function. This is an auto-regulation phenomenon: the prostaglandins (PGs) and the prostacyclins (PGI) are synthesized by parietal cells and surface muciparous cells (283, 332, 405). These metabolites of arachidonic acid contribute to the protection of the mucosa by inhibiting both acid secretion and the activity of the He receptor, which is coupled to adenylate cyclase (AC) in the parietal cells. PGs exert an antisecretory effect via specific PG receptors coupled to Gi (19, 121,268,395). The G proteins serve as intermediaries between activated membrane receptors and their effector enzymes and/or ion channels. The o: subunit of Gi mediates inhibition of adenyl cyclase (vide infra). However, the protective effect of PG on gastric mucosa was observed not to be strictly dependent on the inhibition of gastric acid secretion. It was named cytoprotection, as the property of certain agents to protect the gastric mucosa and the epithelial lining, independently of an anti-secretory effect: mucus secretion, bicarbonate ions, renewal of the surface epithelium, repair, and control of the retrodiffusion of H § ions. The PGs are thus likely to regulate acid secretion, as well as the secretion of bicarbonates and mucus, by acting on the autocrine and paracrine pathways and peripheral circulation. An inhibitory role of PG on acid secretion has been demonstrated after intracisternal injection of PGE2 or its stable analogue 16,16-dimethyl PGE2 in the rat brain (374). By contrast, centrally administered neuropeptide Y (NPY) stimulated gastric acid and pepsin secretions in anesthetized rats (281). While the effect of PGs on gastric acid secretion is quite well defined, the action of leukotrienes (LTs) on the secretory processes in gastric mucosa is controversial. In isolated rabbit fundic glands LTC4 inhibited acid formation in response to histamine and dibutyryl cAMP (264). In contrast, LTC4, LTD4 and LTE4 increased basal, as well as histamine- and carbachol-stimulated acid production in isolated rabbit nmcosal cells (264). In preparations of enriched rat parietal cells, Schepp et al. (1989) observed no effect of either LTC4 or LTD4 on basal acid production. However, they found significant increase in H § formation in prestimulated rat parietal cells (380). The occurrence of platelet activating factor (PAF, paf-acether, 1-O-alkyl-2-acetyl-sn-glycero-3-phosphocholine) in glandular gastric mucosa also indicated autocrine and paracrine actions for this phospholipid mediator in the gastrointestinal tract, a family including prostaglandins, leukotrienes and lipoxins (124,421). It has been demonstrated that PAF is formed from membrane phospholipids by phospholipase Ae and causes the release of other mediators such as tumor necrosis factor a; (TNFo:), leukotriene C4 and norepinephrine during bowel necrosis (198,423). In the gastrointestinal tract, PAF is a potent ulcerogenic factor and stimulates aminopyrine uptake, as an index of H + accumulation, by isolated guinea pig parietal cells (315). The PAF

Pharmacological Control of Gastric Acid Secretion

3~

receptor from guinea pig lung belongs to the G protein-coupled receptors, vide infra (195). Several products are secreted into the lumen by the salivary glands and the gastric mucosa or are derived from milk and digested food. Examples include epidermal growth factor (EGF), transforming growth factor o~ (TGF~), bradykinin, kininogens, triiodothyronine, parathyroid hormone-related peptide, neurotensin, pro-,/-melanotropin, colony stimulating factor, calcitonin- and gastrin releasing peptide (GRP)-like peptides, prolactin, gastrin, somatostatin, PG, morphiceptin, fi-endorphin, c~-melanocyte stimulating hormone (c~-MSH), and caseinomacropeptide (109, 125, 314, 316, 398, 404, 426, 433,458, 463). Milk products, digestive products and mucosal secretions can regulate acid secretion, gastrointestinal motility or exert their trophic effects on different segments of the digestive tract via the luminal fluid and specific mucosal receptors, for instance opioid receptors activated by fl-casomorphins (70, 90, 94, 151, 171). The microvascnlarization system in gastric gland also participates in the control of stomach secretory activity and gastric mucosal protection (136, 137, 191). Histamine receptors are involved in the microvascular permeability to macromolecules, the reduction of blood flow by vasoconstrictio~ via the H~ receptors, and the increase in blood flow and acid secretion via the H2 receptors (69, 85, 191, 276, 303, 304). The increase of gastric mucosal blood flow may be the principal mechanism of protective effect of some gastroprotective agents. For example, intragastric administration of capsaicin protects against ethanol-induced rat gastric mucosal lesions (330). The effect of capsaicin seems to be mediated by stimulation of afferent sensory nerve endings followed by local release of neuropeptides (193). Since it was shown that substance P induces endotheliumdependent vasodilation (132) it is likely that these neuropeptides could in turn release nitric oxide (NO). Finally, it is possible that NO is involved in capsaicin-induced gastric protection by increasing blood flow (194). The catcitonin gene-related peptide is also abundantly distrubuted in the capsaicin-sensitive afferent innervation of the rat stomach, and may contribute to the protective effects produced by acute administration of capsaicin on gastric acid secretion. As shown in Fig. 1 and 2, the secretory activity of the parietal celt is controlled by three principal effectors which can act in synergy (57, 160). 1) Histamine activates Ha receptors, which are coupled to adenylate cyclase via G proteins (25, 103, 172, 268) and to intracellular caldum (309, 310) 2) Cholecystokinin (CCK) and gastrin may share the same receptor (409) or may activate two distinct receptors (266) 3) Acetylcholine, which activates cholinergic M2 receptors. These last effectors mobilize cGMP, the inositol cycle, PKC and Ca 2+ in the gastric epithelium (29, 68, 76, 83). Histamine and cAMP activate protein kinases in parietal cells (74, 268). A protein of 80kD, situated at the apical pole, is then phosphorylated and H+/K§ the proton pump, is activated (261,270, 431,442, 443, 464). In the resting celt the proton pump is localized in the membrane of intracellular tubovesicular structures, which undergo remarkable rearrangements when acid secretion is stimulated and may fuse with the apical membrane to form secretory canaliculi (28, 205,288). Histamine and cAMP are also involved in the regulation

324

Mirossay, Di Gioia, Chastre, Emami and Gespach

ion

Fig. 2. Membrane receptors and signaling pathways involved in the regulation of acid secretion by parietal cells.

of glycoprotein synthesis and release of mucous glycoproteins from gastric nonparietal cells (181). This observation is also consistent with the functional expression of H2 receptors in gastric cell fractions enriched in 87% of mucous cells (145). The histaminergic pathway involves a variety of pathophysiological and pharmacological mechanisms in the digestive tract: 1) histamine secretion; 2) its degradation by histamine methyltransferase (HMT); 3) the activation of the H2 receptor transducer system; 4) H2 receptor desensitization and the capture of bioamines by the histaminocyte or the effector cell, 5) control of receptor activation either by drugs or biological regulators of the histaminergic signaling pathway, such as somatostatins and PG (113, 116, 150, 160), and 6) inhibition of gastric acid secretion by both neural and hormonal mechanisms during acidification in the upper small intestine. The hormonal mechanism is mediated by somatostatin and CCK (321). Histamine receptors have been characterized, according to their pharmacological specificity, vis ~ vis selective agonists and antagonists and the transducers that they mobilize (7-9, 36, 104, 156). In the central nervous system, the H3 receptors are autoreceptors which inhibit the biosynthesis and the secretion of histamine by the histaminergic neurones (7). In the periphery, these receptors inhibit acetylcholine secretion in the trachea (199), the tonus of the sympathetic nervous system, causing the vasodilation of the mesenteric artery and the contraction of the guinea pig ileum (438). H~ receptors are coupled to the IP3-Ca 2+ cascade, while the H2 and H2h receptors are coupled to cAMP and to

PharmacologicalControl of Gastric Acid Secretion

325

cGMP, respectively. In contrast, the H 3 receptors are not coupled to adenytate cyclase (153, 156, 160, 204). The localization of the H2 receptors in parietal cells does not exclude a role for histamine in cAMP synthesis nor in the secretion of chief cells and mucus cells at the mucosal surface in the antrum and the fundus (145, 155,366, 386). In the gastrointestinal mucosa the secretion of histamine by mastocytes (MC) (Fig. 1) is induced by pentagastrin, acetylcholine, phorbol esters, calcium and ethanol, whereas it is inhibited by somatostatin, PG and /3-adrenergic receptors (100, 351, 365, 410-413). The histaminergic Ha receptors and the adrenergic 13-receptors are involved in the regulation of both acid and pepsin secretion (41, 98, 185). In the gastric mucosa, tetragastrin induces the secretion of histamine by the histaminocytes (364). Histamine can exert a negative feedback on its own secretion from ECL cells via the histamine H3 autoreceptors (17, 185). Alternatively, the H3 receptor located on parietal cells may regulate in a negative manner the acid secretory processes (17). Although it is now well known that gastrin has its own receptors on the surface of parietal cells, which stimulate acid secretion (352, 412, 413), there is some evidence that this peptide also partly controls the metabolism of histamine in gastric mucosa. It has been shown that gastrin injection increases histidine decarboxylase activity in the guinea pig (30) and pentagastrin induced decrease of histamine content in dog gastric mucosa, accompanied by an increase of histidine decarboxylase activity (259). Pentagastrin administration in humans induces a significant decrease of histamine content in the oxyntic part of the gastric mucosa. This effect if most pronounced in patients with duodenal and gastric ulcers (257). The effect of gastrin and CCK on the secretory activity of the parietal cell seems to be transmitted by two distinct receptors. In vitro, when the agonists act directly on parietal cell receptors, both gastrin and CCK-8 (octapeptide of cholecystokinin) stimulate the acid secretion with the same potency and efficacy in canine parietal cells (409). The same effect was observed on (14C) aminopyrine uptake (an index of H + accumulation in intracellular, intravesicular spaces), and the formation of inositol phosphates in isolated rabbit parietal cells (357). On the other hand, gastrin is a potent stimulator of H + production in vivo, whereas CCK-8 appears as a weak stimulator capable of inhibiting pentagastrin-stimulated acid secretion (91). These effects influence indirectly parietal cell activity and may be explained by the existance of "gastrin-type" CCK receptors on the gastric "non parietal cell" population. Histamine or somatostatin release could be induced after the stimulation of different receptors by gastrin or CCK, respectively (354). By using specific CCK receptor and gastrin receptor antagonists it was concluded that both gastrin (1-17) and CCK (1-33) stimulate histamine and acid secretion by gastrin receptors, and that gastrin stimulates acid secretion by releasing histamine (354,373). On the other hand, the release of somatostatin, which can be stimulated by CCK-8 or gastrin, is very likely controted by the CCK-type receptor (354) (Fig. 1). Some substances like GRP, secretin, carbachol, and cAMP stimulate the secretion of gastrin in normal antral mucosa as well as in the ZoUinger-Ellison syndrome (110,447). Somatostatin inhibits the transcription of the gene encoding gastrin in the mucosa of the gastric antrum (217).

326

Mirossay, Di Gioia, Chastre, Emami and Gespach

Recent biochemical studies, using isolated rabbit parietal cells, suggest that the muscarinic receptor which controls acid secretion in parietal cells appears to be of M3 subtype (244). Of the five muscarinic receptors that have been cloned, the ml, m3 and m5 subtypes are closely related in sequence and appear to be functionally equivalent (42,349). The M3 receptors potently activate PtdIns hydrolysis and stimulate large, rapid, and transient chloride currents by a pertussis toxin- insensitive G-protein pathway (236). The importance of C1currents in gastric acid secretion (414) was confirmed in a recent study using the patch-clamp technique. In apical membranes of isolated rabbit parietal cells, chloride channels were observed in histamine-stimulated acid-secreting cells but not in resting parietal cells treated by cimetidine (369). M2 muscarinic receptors and nicotinic receptors in the gastric epithelium have been previously characterized (105,265,312, 331). Different second messengers effects were observed in their dependence of ligand concentration. Stimulation of M2-muscarinic receptors by 0.1/~M carbachol 1) inhibits the formation of cAMP by activating the Gi subunit of adenylate cyclase and 2) activates phospholipase C ( P L C ) when the carbachot concentration is raised to 3 #M (10, 417). It was shown that the M2 receptors, which couple to a pertussis toxin-sensitive G protein, weakly activatephosphatidylinositol (PtDlns) hydrolysis and stimulate small, delayed and oscillatory chloride channels. Muscarinic acetylcholine receptors possess sequence homology with the mas oncogene (206,475), and may induce oncogenic progression when expressed in immature cells with proliferating activity (166). Interleukin-1, gastric inhibitory peptide (GIP) and the peptides related to pancreatic glucagon (GLP), including glicentin, truncated glucagon-like peptide (TGLP-1), oxyntomodulin and its C-terminal octapeptide, all inhibit acid secretion induced by histamine, pentagastrin, or by a meal in humans (12, 21, 95, 160, 221, 383, 451). Pancreatic glucagons (G-29), oxyntomodulin, and TGLP-1 stimulate the production of cAMP in the human gastric cell line HGT-1 (Fig. 2) and in isolated rat gastric glands (21, 143, 147, 152, 170). As a function of the ligand concentration (0.25 or 6 nM), G-29 can successively mobilize PLC and the inositol cycle, by the intermediary of glucagon GR1 receptors, and AC by the glucagon GR2 receptors (450). Glucagons thus have a twofold cytoprotective role. They inhibit acid secretion, by a mechanism which controls the gastrindependent activation of PLC and of IP3 in the parietal cell (Fig. 2), and also cause the release of mucus from surface epithelial cells in the gastric mucosa, via cAMP. In humans and in rat, G-29 activates adenylate cyclase in partially purified parietal cells. This paradoxical effect (143,291,382) may be explained by a role played glucagons on the metabolism of the parietal cell, which consumes both ATP and oxygen in large quantities (143). This last mechanism is under the control of somatostatin. Similar results were obtained by Schmidtler et al. (384) who showed that TGLP-1 and its variants exerted a direct stimulatory effect on H + production in isolated rat parietal cells. The effect of these peptides was mediated by a cAMP-dependent regulatory pathway, but independent of H2 receptors. The stimulation of acid secreting cells by TGLP-1 was not observed in vivo. T h e direct effect of this peptide might be counterbalanced by indirect inhibitory mechanisms which are excluded under in vitro conditions (384).

PharmacologicalControl of Gastric Acid Secretion

327

Certain natural substances, such as the PGs, somatostatins 14 and 28, oxyntomodulin, glicentin, EGF, and the tachykinins inhibit acid secretion by blocking the activation of the histaminergic or gastrin-dependent pathways (21, 119, 120, 148, 223, 329). The calcitonin gene-related peptide (CGRP), as well as vasoactive intestinal peptide (VIP), secretin, acetylcholine, pentagastrin, ~.eukotrienes and EGF induce the secretion and biosynthesis of these natural substances (160, 176,202,224, 320). Somatostatin receptors are coupled to the Gi subunit of adenylate cyclase or to other transduction systems (2, 11, 148, 154, 290, 387). Some evidence exists for the expression of cytosolic somatostatin binding sites activating phosphoprotein phosphatases in the digestive tract (345,346). The molecular cloning of genes encoding two different somatostatin receptors with higher affinity for somatostatin-14 than somatostatin-28, and their tissue-specific distribution has been recently reported by Yamada and his colleagues (469). Still unknown are the internalization and intracellular transport mechanisms for this peptide as well as the physiological roles for the membranous, cytosolic and nuclear somatostatin receptors (46, 345, 346) (Fig. 2). The possibility that somatostatin autoregulates the secretion of gastric epithelium D cells has been suggested by in v i v o experiments in dogs. In humans, the half-life of somatostatin-14 is three rain, whereas that of the somatostatin analogue SMS 201-995 (Sandostatin R) is 60 rain. This drug has thus a potential importance in ulcer treatment when anti-H2 medication is found to be insufficient. Unfortunately, this analogue does not appear to be effective in treating the bleeding that occurs subsequently to a peptic ulcer (78, 148, 151, 154, 155,370). A recent discovery established that VIP causes a weak, transient increase in basal and histamine-stimulated acid secretion and sustained increase in somatostatin secretion. This sustained increase in somatostatin, despite the return of acid secretion to basal levels, indicated that VIP directly exerts a stimulatory and an indirectly inhibitory effect on acid secretion. The indirect effect was mediated by released somatostatin in response to VIP stimulation of nonparietal cells (388), as shown in Fig. 1. The physiological role of catecholamines in acid secretion is still poorly understood. There is a human fi2 adrenergic receptor localized preferentially in the gastric fundus, which is under the control of somatostatin-14 (41). fi adrenergic receptor activation of the parietal cell is coupled to H + secretion in isolated rat gastric glands (60, 160). The stimulation of presynaptic o:2 adrenergic receptors, localized in the vagus, lowers basal gastric acid secretion by inhibiting the release of acetylcholine (72,209,473). This is confirmed by the results of Savola et al. (375) with the highly specific a~2-adrenoceptor agonist, medetomidine. Its action was blocked by atipamezol, a novel 0:~-adrenergic receptor antagonist. The EGF receptors were also identified on isolated gastric glands of guinea pig (126) and basolateral membranes of parietal cells (297). Increase in EGF immunoreactivity and the inhibition of gastric acid secretion during the healing of gastric ulcers, argue for a possible role of EGF in the ulcer healing process (171). PGs and certain peptides when injected into the central nervous system, may indirectly mediate acid secretion, gastric contractibility and alimentary supply

328

Mirossay, Di Gioia, Chastre, Emami and Gespach

(183,243,424). This is well established for thyrotropin-releasing hormone (TRH), corticotropin-releasing factor (CRF), and basic fibroblast growth factor (bFGF). Injection of TRH into the central nervous system increases the gastric acid secretion and induces the development of gastric lesions in the rat. Contrarily, intracisternally injected bFGF inhibits the secretion of both gastric acid and pepsin and has a mucosal protective effect through a mechanism involving the central nervous system (318,319). The central application of CRF inhibits gastric acid secretion, motility, and emptying, augments bicarbonate content and confers protection against cold restraint-induced gastric mucosal damage in rats (165). On the other hand, peripheral administration of CRF does not enhance alkaline secretion and is not able to inhibit acid-induced villous damage in duodenum (246). Acid secretion is thus integrated into an extremely complex regulatory network (Fig. 1 and 2), in which stimulations, inhibitions, and potentializations all intervene (35,289). The occurrence of these processes depends on chronological factors and on the coordination between agressive digestive secretions (acid, proteases, lipases) and those that are protective (water, bicarbonate, mucus). The integrity of the mucosa is directly linked to the repair systems of the protective barrier and to the healing and rapid renewal of gastric epithelial cells, which are themselves dependent on growth factors and on genes controlling cellular proliferation and differentiation.

RECEPTOR-SIGNAL TRANSDUCTION Biological substances regulate cellular activity via cell surface, cytoplasmic, and nuclear membrane receptors. There are at least five principal pathways which allow information to be communicated by the receptor- signal transduction systems: 1) adenylate cyclase and guanylate cyclase (GC); 2) the IP3-calciumdiacylglycerol-protein kinase C; 3) phospholipase A2; 4) the tyrosine kinases; and 5) the calcium channel-coupled receptors.

Cyclic AMP- and Cyclic GMP-Dependent Regulatory Pathways Three types of mammalian particulate adenylate cyclases have been biochemically distinguished and isolated from various tissues: a brain-specific, calmodulin-sensitive form (type I), a calmodulin-insensitive form identified in brain and lung (type II), and an olfactory-specific, calmodulin-activated cyclase (18). Particulate adenylate cyclases are regulated by G proteins, but the soluble form is independent of these regulatory proteins (427). The general structure of adenylate cyclase resembles those of membrane channels, leading to the speculation that these transmembrane proteins might have dual roles, as both enzymes and transporters. The particulate, calmodulin-sensitive adenylate cyclase from bovine brain has been found to have 1134 aminoacid residues with both the N- and C-terminal ends being intracellular. Its sequence consists of two large hydrophobic and two large hydrophilic domains. Each hydrophobic domain

Pharmacological Control of Gastric Acid Secretion

329

consists of 6 putative transmembrane spans which are interconnected by short, alternatingly extracellular and intracellular, hydrophobic sequences. The hydrophobic domains themselves are connected at the intracellular side, by one of the large (43 kD) hydrophilic domains (226). The cloning and characterization of a type IV adenylate cyclase has been recently described. It is insensitive to calmodulin and appears to be widely distributed (140). Adenylate cyclase activity is regulated after receptor stimulation by two regulatory subunits (Fig. 3). One of the regulatory subunits has a stimulatory effect (Gs) on the formation of cyclic AMP (cAMP), while the other has an inhibitory effect (Gi) on this process (187,358). The binding of ligand to its receptor exerts a stimulatory effect on the coupling reaction between the receptor cytoplasmic domain and Gs, a heterotrimer composed of o6 fi and 7 subunits. This coupling process is followed by increased exchange of bound GDP for GTP on the G protein. The excess of G-GTP protein complex allows the Gcr-GTP subunit to dissociate from the fi, 7 dimer. The Gsoc-GTP complex subsequently activates the catalytic subunit of adenylate cyclase, while the inhibitory activity of

) ~GDP \

b----I

/

\ //

~ A AMPc TP

Fig. 3. Activation of adenylate cyclase (AC) via the histamine H2 receptor and the Gs subunits ols, fi, and y.

330

Mirossay, Di Gioia, Chastre, Emami and Gespach

Gia~ is weakened. The /~, y dimer of Gi is predominantly responsible for adenylate cyclase inhibition (227). There exist three different y subunits isoforms: y1, y2, y3, each encoded by three distinct genes. The o~subunits can be grouped into four distinct classes (248,420). The activity of adenylate cyclase catalytic subunit remains unchanged as long as GTP is not hydrolyzed. The intrinsic GTP-ase activity terminates the signal. Breakdown of GTP to GDP at the Gc~ nucleotide binding site causes the release of Gol-GDP, restoration of the or, /3, y trimer, and restitution of basal level of adenylate cyclase activity. This cycle is generally followed by all naturally occurring ligands, except the diterpene forskolin, which directly activates adenylate cyclase in the absence of Gs (394). Cell surface receptor activation and adenylate cyclase stimulation induce subsequent cAMP accumulation and cAMP-dependent protein kinase (PKA) activation (74). In the resting cell, PKA forms a heterotetramer which is catalytically inactive. The regulatory subunit (PKA-R) liberates two catalytic subunits (PKA-C) after cAMP-binding. Two PKA-C monomers are then capable of phosphorylating a large number of cellular proteins, whereas PKA-R conserves its dimeric form. Mammalian tissues contain two categories of PKA regulatory subunits: type I and type II. Several isoforms of each PKA-R have been cloned. PKA-RIo: and PKA-RIIol are expressed in most cells (237,393), while the expression of PKA-RI/3 and PKA-RII/3 is more tissue-dependent (79,207). An inverse relationship was observed between types I and II isoform expression during ontogenic development, cell differentiation, and neoplastic transformation. It has been proposed that PKA-RI overexpression is responsible for the augmentation of cellular growth and malignancy, whereas the elevation of cellular levels of PKA-RH isoenzyme is associated with tissue differentiation and growth arrest (77). The activation and inhibition of PKA by histamine and its antagonists mimic the functional characteristics and the pharmacological specificity of H2 receptors (151). Cyclic AMP is hydrolyzed to adenosine monophosphate by a phosphodiesterase (PDE), localized either in the cytosol or in the membrane. Multiple forms of PDE with distinct properties are present in mammalian tissues. The unique among them is the PDE II isoenzyme family, because of its cGMP-induced ability to hydrolyse cAMP (459). Guanylate cyclases exists both in cytosolic and particulate forms (141,390). The membrane-bound guanylate cyclase appears to span the membrane bilayer as a single polypeptide chain, with an intracellular catalytic domain and an extracellular binding domain for the series of atrial natriuretic peptides. Recently, an endogenous activator of the intestinal guanylate cyclase has been recently purified and designated guanylin (84). Guanylin appears to act in a manner similar to the heat-stable enterotoxins in stimulating cGMP levels in intestine and other tissues, through the same extracellular binding region of the particulate enzyme. At least five independent membrane guanylate cyclases, showing a similar pattern have been cloned (162). Detergent- insoluble forms of guanylate cyclase appear to predominate in gastrointestinal mucosa (391). The cytosolic form of guanylate cyclase is found in most mammalian tissues. The soluble guanylate cyclase is a heterodimer of 70 and 82 kD subunits containing catalyticlike domains (Fig. 4). They must interact with each other in order to generate the

Pharmacological Control of Gastric Acid Secretion

331 Cellutar

Responses I

Protein

Phosphorylation

Target Cell -,~

Photoreceptor

cAMP - PDE

i

Na+/Ca2*channels

Ca2LCaM

/

! PGI 2 H starn ne Rece I PAF ~Aeh

eGMP-~ membrane

GTPJ~ L-Arg ~ - - - ~

L-Arg

,NO / EDRF

~ ~

SINP ~

Effeetor Cell

Cycl~se

~Transceilular

~

Gap

t

Fig. 4. Intracellular and transcellular signaling via the NO synthase and the soluble guanylate cyclase pathways. Activation of membrane receptors by prostacyclins (PGIs), histamine, platelet activating factor (PA) or acetylcholine (ACh) results in Ca2 + uptake by target cells, stimulation of cytosolic NO synthase (nitric oxide synthase: NOS) by the Ca2+-calmodutin complex (Ca2+-CaM), production of nitric oxide (NO), i.e. the endothelium-derived relaxing factor (NO/EDRF). The release of NO, derived from the oxidative metebolism of one of the guanidinonitrogens of L-arginine, or produced after the uptake of sodium nitroprusside (SNP) induces activation of the soluble form of guanylate cyclase and mobilization of the cGMP cascade: membrane channels, cGMP-dependent protein kinases (cGMP-PK), cAMP phosphodiesterases (cAMP-PDE). NO is a gaseous oxide of nitrogen that diffuses through cell membranes and reacts with adjacent effector cells.

basal and stimulated guanylate cyclase activity (51). It contains heme, which is thought to be the recognition site of nitric oxide (NO) and other NO-generating agents (nitroprusside, nitroglycerin). NO is formed from arginine in a reaction catalyzed by the enzyme NO synthase, the activity of which is regulated by calmodulin and calcium (Ca2+-Cam). Thus, agonist-induced cGMP accumulation presumably results from agonist-stimulated Ca 2+ influx through calcium channels in the plasma membrane, activation of NO synthase (48), and stimulation of the soluble guanylate cyclase by NO. Various substances may activate specific receptors and induce the formation of NO. This found to be the case for substances such as acetylcholine (ACh), adrenaline, noradrenaline, histamine, prostacyclins (PG12) arginine-vasopressin, 5-hydroxytryptamine, ADP, thrombin, PAF, and bradykinin (444). As mentioned above, the stimulation of sensory nerves by some agents, with subsequent release of substance P, may be one of the protective components in the gastric mucosa mediated by NO-dependent vasodilatation (330). The remarkable difference between cAMP-dependent and cGMP-dependent

332

Mirossay, Di Gioia, Chastre, Emamiand Gespach

transduction systems is that the molecular sites of action of cGMP are not exclusively the protein kinases, but also cAMP-PDE, ion channels, and possibly other target proteins involved in cellular responses, as shown in Fig. 4.

Phosphatidylinositol / PKC Pathway Receptor-mediated phosphoinositide breakdown is catalyzed by agonist activation of PLC (Fig. 2). There are at least four or five types of PLC in mammalian tissues (or, /3, y, 6 and e) activated by the Gp-linked receptors, such as muscarinic M2 receptors, and gastrin/CCK receptors (80, 347, 348, 353,368). Three types of Gp proteins are involved in this process, and have been described according to their sensitivity to pertussis and cholera toxins (47, 279, 302, 385, 400, 406, 440). In parietal cells, at least one cholera toxin-insensitive Gp protein(s) appeared to be involved in gastrin receptor coupling to PLC (353). Each PLC has similar catalytic properties and hydrolyzes three, common phosphoinositides: phosphatidylinositol (PI), phosphatidylinositol 4-phosphate (PIP) and phosphatidylinositol 4,5-bisphosphate (PIPz). PLC transforms PIP2 into two messengers, inositol 1,4,5-trisphosphate (IP3) and diacylglycerol (DG), the latter is also the product of phosphatidylcholine hydrolysis (33,277). When IP3 is released into the cytoplasm, it mobilizes Ca 2+ stores from the endoplasmic reticulum (ER), as shown in Fig. 2. The transitory rise in free calcium and/or its binding to calmodulin induces the activation of specific protein kinases. In parietal cells, both gastrin and CCK receptors mediate phosphoinositide breakdown accompanied by a rise in free intracellular calcium (Ca 2+) (357). In isolated rabbit parietal cells, gastrin induces a rapid rise in IP3 concomitant with acid formation (356). The kinetics of gastrin-induced rise in IP3 is biphasic with a rapid increase occurring within the first 15 seconds, followed by a sustained increase. While the first event is insensitive to pertussis toxin and extracellular calcium the latter is affected by these two factors. The physiological role of the sustained increased level of IP3 is not clear (355). Recent studies have also indicated that the CaZ+-calmodulin pathway may have a critical role in the secretion of pepsinogen induced by carbachol and CCK in rabbit mucosa maintained in organ culture (295). Evidence is now accumulating to indicate that other membrane phospholipids are also degraded following receptor activation. This is the case for phosphatidyl-choline which is probably hydrolyzed by both a phospholipase C and a receptor-coupled phospholipase D (PLD), producing phosphatidic acid or phosphatidylbutanol (432). In various tissues, PLD might be activated by: 1) G protein-coupled receptors, 2) phosphorylation by a receptor tyrosine kinase, 3) the DG/PKC cascade, or 4) the IP3/Ca 2+ cascade. Lipidsoluble diacylglycerol allows protein kinase C (PKC), localized in the cytoplasm in its inactive form, to translocate towards the plasma membrane, where it is activated (Fig. 2). PKC is a Ca 2+ and phospholipid-dependent serine-threonine protein kinase, which is ubiquitously distributed in tissues and organs, with the highest level of activity observed in brain cells (229,292). The family of PKC enzymes consists of at least 7 isotypes o~, /31,/311, y, 6, e, ~, which exhibit distinct tissue-specific patterns of expression (219). The 7 isotype appears

PharmacologicalControl of Gastric Acid Secretion

333

to be expressed predominantly in the central nervous tissue, whereas the 0~and isotypes are differentially expressed in a wide variety of cell types (219). The presence of elevated levels of Ca 2+ and phosphatidyl serine are required for the stabilization of the binding of PKC to the membrane. Phorbol esters and certain mutagens directly stimulate the C kinases in the presence of calcium in a fashion similar to diacylglycerol (311). The PKC contains 2 phorbol ester binding domains which may have interactive functions (55). The hydrolysis of diacylglycerol produces arachidonic acid (AA), which is metabol~ed by the lipo-oxygenase pathway (L-ox) into leukotrienes and by the cyclo-oxygenase pathway (C-ox) into thromboxane A 2 and the series of prostaglandins D2, E2 and F2 (Fig. 2). It is now generally accepted that the PKC enzymatic complex plays a crucial role in a number of cellular responses related to differentiation, proliferation, tumor promotion and membrane protein functions (190,313). Several other G proteins, possessing a GTPase activity, couple a large number of other cell surface receptors to other transducers and ion channels (308). The following proteins may be cited as examples: transducin Gt, which activates cGMP-dependent phosphodiesterases in the retina during the photoreceptor cascade, the Go proteins isolated in the central nervous system and the heart, and the pertussis-insensitive Goal6 protein specifically expressed in hematopoietic cells (3,269, 308,418, 428).

Phospholipases A2 and Phospholipases D Phospholipases A2 (PLA2) are enzymes responsible for the release of arachidonic acid (AA) from membrane phospholipids (PL) and hydrolyze the 2-acyl ester bond of phosphoglycerides. Investigation of the subcellular distribution of these enzymes revealed the existence of both cytosolic and membranebound forms. In cytosol of various cells and tissues including blood cells and kidney, a cytosolic PLA2 preferentially hydrolyses phospholipids containing arachidonic acid esterified in the 2-position. The activity of cytoslic PLAa is calcium-dependent and may be reulated by physiological Ca 2+ intracellular levels (99). Because simultaneous increase in Ca2+-dependent activity and PLA2 membrane translocation have been observed, it was suggested that membranebound and cytosolic PLA2 activities represent one enzyme and that the membrane-bound form is the biologically active PLA2 (343). Recent molecular cloning of the human Ca2+-sensitive cytosolic PLA2 gene demonstrates that the corresponding cDNA encodes the functional 85 kD cytosolic PLA2. The cytosolic PLA2 protein contains a structural element homologous with the conserved region C2 of the regulatory domain of PKC (339). In rabbit platelets, the histamine Ha receptor is coupled to PLA2 via pertussis toxin-sensitive G proteins (301). Taken together, these results suggest that PLA2 is coupled to receptors through G proteins and is stimulated by an activated PKC in concomitance with intracellular Ca a+ elevation (43,167,325). In addition, a source of diglycerides different from phosphatidylinositol has been detected in various ce~l types. The role of phosphatidylcholine as a

334

Mirossay, Di Gioia, Chastre, Emami and Gespach

precursor for second messengers in signal transduction was suggested to be mediated via phospholipase C or phospholipase D.

Tyrosine Kinase Transduction Systems The fourth type of communication may be illustrated by two types of protein-tyrosine kinases: 1) the receptor type tyrosine kinase complex of insulin, EGF, or platelet-derived growth factor (PDGF), and 2) the non-receptor types of protein-tyrosine kinases encoded by different (proto)oncogenes, which are membrane associated enzymes without any ligand binding domain: pp60 ..... , pp160 gag-abl, pp130 gag-fps and pp59 c-fyn (63). The receptor tyrosine kinases have a similar molecular topology. All possess a large glycosylated, extracellular ligand binding domain, a single hydrophobic transmembrane region, and a cytosolic domain that contains a tyrosine kinase catalytic unit (168, 462, 471). It is possible to classify these receptors into 4 subclasses according to their distinct structural characteristics (441). The activity of the receptor type protein-tyrosine kinase is regulated by ligand binding and the subsequent conformational alteration at the extracellular domain (371). This process induces receptor oligomerization, which stabilizes interactions between adjacent cytoplasmic domains and leads to activation of kinase function by molecular interaction. Receptor oligomerization is an universal phenomenon among growth factor receptors. Receptor tyrosine kinases catalyze the phosphorylation of exogenous substrates as well as tyrosine residues within their own polypeptide chain. The ligand-stimulated autophosphorylation (or cross-phosphorylation) on tyrosine not only enhances the tyrosine kinase activity of the receptors but also creates binding sites for recruitment of specific cellular enzymes (reviewed in 441). These enzymes then transduce signals to the cell interior. For example, PLCy1 and phosphatidylinositol 3-kinase (PtdIns 3-kinase), the cytosolic enzymes in resting cell, are recruited by the PDGF receptor after its activation (216,446). The stimulation of PLC-~I, an isoenzyme of PLC, is responsible for PIP2 turnover after EGF and PDGF receptor stimulation. This activation of PLC-y1 is accompanied by tyrosine phosphorylation (348). Activation of PtdIns-specific PLC-1 would stimulate the classical PtdIns turnover pathway, leading to elevation of cytosolic Ca 2+ and stimulation of PKC with subsequent activation of a multitude of cellular events including changes in cytosolic pH, potassium levels, and transcription. A newly discovered enzyme implicated in the cross-talk of the transduction pathways, is PtdIns-3 kinase. This enzyme was shown to be composed of two subunits of 85 and 110 kD (460). The 85 kD subunit is a direct substrate of the PDGF receptor and pp60 ..... . PtdIns 3-kinase was the first enzyme found to be associated with tyrosine kinase. Upon activation, PtdIns 3-kinase phosphorylates the D-3 position of the inositol ring. Until recently, phosphatidylinositol was thought to be phosphorylated only at the D-4 and D-5 positions of the inositol ring. The discovery of PtdIns 3-kinase uncovered a family of D-3-phosphorylated phosphoinositides (65,460). The phosphoinositides that are phosphorylated at the 3 position are not substrates for any of the known PtdIns-specific PLCs (253),

Pharmacological Control of Gastric Add Secretion

335

indicating that they are not converted to inositol phosphates in response to activators of the classical pathway. It appears to be possible that they act in a new signalling pathway distinct from the clasical Ptdins turnover pathway (13, 66). Some other enzymes, such as the ras GTPase-activating protein (GAP) or c-raf proto-oncogene product, are direct substrates for receptor tyrosine kinases (296, 299).

Calcium Channel-Coupled Receptors The increase of Ca 2+ concentration after receptor stimulation, which is coupled to PLC, occurs most commonly in two phases. The first elevation of Ca~ + originates from intracellular stores and is believed to be mediated by IP3 (vide supra). The second component is due to Ca 2+ uptake from extracellular space, although its control by PLC is still unclear. The mechanisms regulating the calcium channel-coupled receptors are not yet elucidated (reviewed in 286). The entry of extracellular Ca 2+ after depletion of rapidly exchanging intracellular Ca 2+ stores may be the first mechanism involved in the control of plasmalemma Ca 2+ channels. Depletion of intracellular Ca 2+ stores appears to induce the activation of plasma membrane Ca 2+ channels via unidentified signal. The same channels may be directly activated by way of membrane receptors, coupled to them through G-proteins. Recently, in pituitary cells and adrenocortical cells~ it was shown that a pertussis toxin-sensitive G-protein is implicated in the mediation of hormone stimulated inward Ca 2+ currents (178). The existence of two different channels controled by the two above mentioned mechanisms, or a third channel, operating under influence of an unknown second messenger, are possible (286).

Molecular Structure of Membrane Receptors The classical techniques of solubilization and purification of membrane proteins have revealed the primary sequences of a series of receptors that are coupled to G proteins: adrenergic, muscarinic and serotoninergic receptors, receptors of substance K and angiotensin (102, 174, 222, 417), and many others. All these receptors share structural analogies with opsin and rhodopsin in the retina. This family of receptors possesses an extracellular domain binding the natural ligand or its analogues, a transmembrane domain of seven segments anchored in the hydrophobic part of the membrane and a cytoplasmic domain, which is in contact with both the transducer and the proteins regulating the desensitization and the internalization of the receptor (457). The transmembrane domain is the most highly conserved region of these receptors coupled to G proteins. The extracellular loops, localized between segments 2, 3 and 4, 5 contain cysteine residues, suggesting the presence of disulfide bonds between these loops (127). The extracellular, N-terminal domain of the protein contains asparagine residues, which link oligosaccharides. The seven serine or threonine residues at the extracellular C-terminal are potential substrates for phosphorylation by specific kinases, which are involved in the process of desensitization. The domain associated with the G proteins is as yet poorly characterized. The

336

Mirossay, Di Gioia, Chastre, Emami and Gespach

intracellular loop between segments 5 and 6 of the receptor is in contact with G proteins (222). Various experimental strategies, including gene cloning, sitedirected mutagenesis, low-stringency hybridization polymerase chain reaction (PCR) amplification, and gene tranfer in heterologous systems, such as microinjected Xenopus laevis oocytes, and have been employed to establish the structural differences and analogies between these membrane proteins, ionic channels, membrane transporters and adenylate cyclases in animal species (174, 180, 226, 467). These techniques have been also used to detect new receptor subtypes, as in the case of the adrenergic/33 receptor (118). In brain and peripheral tissues, histamine actions are mediated via cytosolic and cell surface receptors, previously classified as the H~, HIC, H2, H2h and H 3 histamine receptors, according to pharmacological, biochemical characterization and gene cloning. The stimulation of G protein-coupled Ha receptors promotes phospholipase C and A2 activation, Hz receptors induce cAMP formation, while the signal transduction system activated by the recently discovered histamine H 3 receptors was responsive to GTP (9). In platelets, intracellular histamine has been shown to promote platelet aggregation, through binding to a novel intracellular histamine H~c receptor of #M affinity, suggesting intracrine regulation by histamine in this sytem (377). Among the histaminergic receptors, the Ha sub-type has been well characterized using radioactive ligands and photo-activated probes, which allow the specific pharmacological properties of this sub-type to be reproduced by histamine, mepyramine, iodobolpyramine, iodoazidophenpyramine, arylazide histamine, and 7-azidoketanserine (122, 130, 225, 361, 455,466). The binding domain of the H~-receptor lies within a protein possessing disulfide bridges and having an apparent molecular weight of 56 kD in the brain, 50 kD in the thymus (322,361), and 670-800 kD in the liver (130). However, the identity of the Ha-site in liver is disputed (247). A molecular form of 160 kD was observed in the cerebral cortex following radio-inactivation. For the Hz-receptors, a non-specific binding component and the capture of bioamine are observed with histamine, cimetidine, ranitidine, impromidine, tiotidine, and SKF 93479 in purified membranes and in cellular systems (22, 113, 116, 133, 157, 456). These limitations have considerably restricted research progress on the solubilization and identification of Hz-receptors using conventional receptor purification and identification techniques. Cloning and functional characterization of 2 complementary DNAs encoding the histamine H1 and H2 receptors was recently achieved using an electrophysiolocal assay in oocytes after injection of polyadenylated RNAs (470) and by PCR amplification using degenerate oligonucleotides primers based upon homology to other G protein-coupled receptors (138, 139). RECEPTOR-SIGNAL INTEGRATION IN CELLULAR SYSTEMS Cross-Talk and Desensitization

It is now well established that the principal types of signaling pathways, mentioned above, are also interconnected in a complex network of functional

Pharmacological Control of Gastric Acid Secretion

337

interactions capable of modulating their respective activities. These functional relationships among membrane transducers are directly implicated in the inhibition, potentialisation, and desensitization of cellular activity regulated by cell surface receptors. In the gastric mucosa, these receptor-signal transduction pathways act in an interdependent manner to integrate such cellular activities as proliferation, differentiation and digestive functions in the normal and transformed gastric epithelium (70,151,290,324). Use of 12-O-tetradecanoylphorbol-13-acetate (TPA) in the direct activation of PKC in rat parietal cells showed the reduction of histamine-stimulated cAMP content and aminopyrine accumulation (177). Furthermore, the 1-oleoyl-2-acetyl-glycerol (DG), also inhibited histaminestimulated aminopyrine accumulation (5). Similarly, EGF inhibits histaminestimulated cAMP production (176) and aminopyrine accumulation in rat parietal cells (400). EGF exerts this effect by stimulation of PDE, followed by the diminution of the cellular cAMP content since its activity is abolished in the presence of isobutylmethylxantine (IBMX) (400). Cooperative effects have been suggested between cAMP and gastrin-induced PtdIns breakdown on the acid secretory activity in parietal ceils (353). This was documented by an effect of cholera toxin on rabbit parietal cells. The ADP-ribosylation of the o: subunit of the Gs protein, which leads to a persistent stimulation of adenylate cyclase, induced a pronounced potentialization of gastrin-stimulated aminopyrine uptake

(353). In other tissues (e.g. in heart), the activation of adenylate cyclase by /~-adrenergic agents leads to phosphorylation of voltage-dependent Ca 2+ channels (16). A subsequent effect on Ca 2+ influx has been proposed. Even adenylate cyclase itself may be regulated by PKA; the direct phosphorylation of adenylate cyclase by PKA may be partialJy responsible for the heterologous desensitization of glucagon-stimulated adenylate cyciase activity in hepatocytes (334). In frog erythrocytes, a 130kD protein copurified with adenylate cyclase activity was phosphorylated after treatment with the PKC activator TPA (474). The increase of Ca~+ may also influence adenylate cyclase activity by complexing with calmodulin and activating other pathways which can modulate adenylate cyclase activity (123), or adenylate cyclase may be stimulated directly by Ca2+-calmodulin in some tissues (293). The desensitization of the receptor-transducer systems follows upon a primary exposure of the target tissue either to the natural effector or to a drug. The importance of this phenomenon, designated "tachyphylaxis" or "refractoriness", is to be considered in view of the protection of the tissue from secretory dysfunction after chronic stimulation or drug treatment. This mechanism has been demonstrated for hormones, neuromediators, paracrine agents, and drugs (151). The desensitization 1) is followed by the internalization of the receptor, which "disappears" from the cell surface, 2) is related to the alteration of the receptor which is phosphorylated by a specific kinase or 3) to decoupling of the receptor to transducer, or 4) to a distal regulation by the "second messenger" (67, 151, 160, 182, 335). Vasopressin can short-circuit both the mitogenic activity and the induction of the oncogene c-fos. Glucocorticoids increase the transcription of the

338

Mirossay, Di Gioia, Chastre, Eraamiand Gespach

fl-adrenergic receptor gene, while isoproterenol destabilizes the receptor gene transcript by diminishing the half-life of messenger RNA. Recovery of the cell's sensitivity to the receptor agonist is generally slower and involves either the recycling (towards the membrane) of the internalized receptors or the synthesis of new receptor molecules. For gastric He receptors, we have analyzed the pharmacological, biochemical and functional properties of the desensitization process following brief or long-term exposure of human gastric HGT-1 cells to histamine and its H1 and H2 receptor specific analogues (113, 114, 335). Desensitization process is rapid (half-life of 20 minutes). It is also temperaturedependent and sensitive to histamine dose 20 times lower than that required to activate the H2 receptor. Histamine H2 receptor desensitization manifests itself by a progressive diminution of stimulation. It is maximal for histamine at a concentration of 10-SM (114). This phenomenon is controlled by the action of histamine and cimetidine on the H2 receptors and results from the uncoupling of the Gs subunit to AC or from the persistent activation of Gi (114,335). However, the fate of histamine and its H2 receptors in the parietal cell is, at present unknown. Identical results were obtained when human HGT-1 gastric cells were cultivated in the presence of 10-4M histamine for six days. Chronic treatment of HGT-1 cells by histamine produces a homologous desensitization, since the functional activity of peptidergic receptors activated by glucagon, VIP and GIP, was not affected in this cell line. In heterologous desensitization, the activation of PKC induces the internalization, as well as a considerable reduction, of both the muscarinic receptors and the gastrin receptors in parietal cells purified to 95% (76). Regulation of gastric histamine receptors, PG receptors, and gastrin receptors have been observed in vivo, in relationship with suckling and weaning in the rat (147, 150, 415). We have also observed the down-regulation of the secretin binding sites and its associated heterologous desensitization in rat gastric glands cAMP system (24). The desensitization of the H1 receptor of histamine involves a diminution of the number of binding sites for tritiated mepyramine, a selective antagonist (151,160).

MODELS AND MECHANISMS OF ACID SECRETION

Faced with the complexity of the cell populations composing the gastric mucosa and the parameters regulating acid secretion, cellular preparations and tissue culture allow the direct analysis of histamine H2 receptors and the activation of the H+/K§ proton pump. These methods utilize chemical agents, such as EDTA, and dispersive enzymes, such as pronase, dispase and collagenase, which act on everted gastric pouches or mucosal scrappings. Certain preparations combine the sequential exposure of cells to enzymes and to EDTA, followed by a specific cell type enrichment, either by Percoll gradient separation or centfifugation i n an elutriation rotor (160). Other authors have developed primary cultures of isolated gastric cells or gastric cells in organotypic culture, which produce mucus, pepsin, PG, gastrin or acid (75, 175, 188, 231, 283, 317, 372). However, these cells proliferate poorly, restricting experimentation. T h e

PharmacologicalControl of Gastric Acid Secretion

339

production of immortal gastric epithelial cell lines by genomic transfection with the oncogenes myc, mutated p53, E1A, SV40 or polyoma virus Jarge T could provide alternative models for studying acid secretion, instead of those involving laboratory animals, normal surviving cells or nonfunctional tumor cells (70, 117, 422). Establishing cell lines derived from sporadic gastric cancers, notably the HGT-1 cell line (230, 337, 396), has greatly improved our knowledge on the expression and pharmacological properties of histaminergic and peptidergic receptors in humans. The study of structure-activity relationships and the short-term and long-term effects of H2 antagonists in clinical experimentation have been developed in our laboratory for HGT-1 cells, in parallel with isolated human gastric glands (112-114, 149, 151, 154). HGT-1 cells are relatively undifferentiated morphologically, but possess a number of peptidergic receptors characteristic of the human gastric epithelium, such as those activated by VIP, gastric inhibitory peptide (GIP), glucagon-related peptides and somatostatins (21, 111, 144, 151, 152, 170, 344, 359, 360). Another approach developed in our laboratory was the establishment of gastric cell lines after genomic transfection of rat gastric glands with recombinant ecotropic retroviruses carrying the immortalizing viral oncogene SV40 large T. After selection we obtained five RGC cell lines possessing poorly differentiated epithelial cells morphology, as detected by phase-contrast microscopy and electron microscopy (71). The next model is represented by the poorly differentiated, human gastric adenocarcinoma cell line MKN-45. This cell line is considered to be derived from the germ cells of the gastric glands and possesses an intracellular canaliculus-like structure, a characteristic feature of parietal cells. Treatment of MKN-45 cells with retinoic acid enhanced cAMP production in response to histamine and increased the number of H2-receptors visualized by (3H) tiotidine binding. It is possible that this effect reflects the differentiation towards more mature gastric epithelial cells (6).

PHARMACOLOGICAL CONTROL OF GASTRIC ACID SECRETION Cimetidine, Ranitidine and Famotidine: The Progress of the Antagonism at H2-geceptors The synthesis by Black of a series of imidazote derivatives including burimamide, the first H2-antagonist, opened the way to a new and efficient treatment of gastroduodenal ulcers (36,189). This work was recognized with the awarding of the Nobel prize in physiology and medicine in 1988. The discovery lies at the origin of the development, by several laboratories, of a series of specific and powerful H2 receptor antagonists possessing prolonged activity and no undesirable side effects. The H2 antagonists have considerable structural variety (38, t13, 135, 149, 287, 392, 425) but possess three elements in common (Fig. 5): 1) an aromatic core, which is either an imidazole (cimetidine), aminomethyl furan (ranitidine, SKF 93479), guanidine thiazote YM-11170, i.e. famotidine or triazole (AH 22216), 2) a flexible chain CH2-S-CH2-CH~ and 3) a cyanoguanidine polar group.

340

Mirossay, Di Gioia, Chastre, Emami and Gespach

.,c

cimetidine

NN~N

ranitidine

s•CH•SCH2CH•CNH2 NSOzNHz

f amotidine

N / C\ HzN NH~

OCH3 H3 C , , , ~ ' ~ CH 3 0 N - - ~ / ~ O C

H3

omeprazole H

Fig. 5. Molecularstructure of three He antagonists and of omeprazole, a specificH § K+-ATPase inhibitor. These drugs possess remarkable specificity of action on membrane receptors. The interaction site of a competitive inhibitor may be located in one or several extracellular loops of the binding site. It may also lie within the intra-membrane of intracellular domains of the receptor (vide infra). In the gastric epithelium, at therapeutic doses, cimetidine, ranitidine, famotidine and the compounds A H 22216 and SKF 93479 have no effect on the /3-adrenergic, acetylcholine, secretin, glucagon, VIP, GIP or PGE2 receptors (113, 143, 153, 158, 272, 287, 291). The criteria used to evaluate the potency of the competitive antagonists (IC~0) includes the antagonist dose resulting in 50% inhibition of induced receptor activity at a defined concentration (S) of histamine. If the Ka is equal to the histamine dose that produces half-maximal receptor stimulation, then, the inhibition constant Ki---ICs0/(1 + S / K a ) allows a direct comparison between the potency of the drugs studied in the different systems. The second criteria is the Schild equation (116), in which antagonism is expressed by the histamine dose ratios (DR) required to produce half-maximal responses in the presence of different antagonist concentrations: log (DR - 1) = n log (antagonist) - log Kb. For a simple competitive antagonism, the Schild plot yields a straight line with a slope of unity. The intercept with the abscissa (DR = 2), is the pA2 value ( - l o g Kb), i.e., the negative log of the receptorantagonist apparent dissociation constant (113,142). The competitive antagonists can thus be classed according to their pA2 (37, 73, 87, 169, 173, 245): tiotidine, famotidine (7.5-7.8)> ranitidine (6.9-7.2)> cimetidine (5.7-6.7). Famotidine is thus 3.5 to 16 times more potent than ranitidine and cimetidine in isolated human gastric glands and in HGT-1 cells (153). These values agree with results obtained in vitro and in vivo for adenylate cyclase stimulation and acid secretion in

PharmacologicalControl of Gastric Acid Secretion

341

humans, rat, and dogs (34, 163, 197, 328, 378, 381,407, 4.25). For example, acid secretion is 50% inhibited when the plasma concentration is 20-30 ng/ml for famotidine and 100 and 500 ng/ml for ranitidine and cimetidine, respectively (49, 327, 392). In the treatment of Zollinger-Ellison syndrome, acid secretion is blocked by a dose of famotidine of 60 mg administered every six hours. The same effect is obtained with ranitidine only at a dose of 500 mg or with cimetidine at a dose of 1.9 g (197). The H2 antagonists are capable of interacting with H1 and H 3 receptors, but do so only at very high doses that exceed the concentrations measured in the peripheral blood of patients treated for gastroduodenal ulcers. Likewise, the histamine H2 receptors in human gastric glands are selective vis-a-vis the H~, H2, and H3 agonists and antagonists (149,153). Famotidine is 330 and 6,7000 times more potent on the H2 receptors than is triprolidine or thioperamide, antagonists of H1 and H3 receptors, respectively: H2> H~ > H3 (153). In accordance with the observations of Green and Maayani on the central nervous system (164), the anti-depressants amitriptyline and imipramine block the activity of the H2 receptor in gastric mucosa of guinea pig (22, 23,439). This action does not seem to occur with acid secretion in vivo, during the treatment of depression (40). However, another tricyclic antidepressant, trimipramine, has been found to promote ulcer healing in certain patients unresponsive to cimetidine and anti-acid treatments (271,350). The pharmacokinetics and drug metabolism (163, 285, 342, 392), as well as the specificity and the reversibility of the interaction between the H2 antagonist and the receptor-transducer system, influences the effectiveness and handling of peptic ulcers. These parameters determine the dose and the timing of daily intake of medication. For example, cimetidine, which has a specific action on the H2 receptor, does not desensitize its H2 antagonism and does not produce a rebound effect when the treatment of cultured human gastric HGT-I cells by cimetidine is interrupted (115). When HGT-1 cells are exposed to 10-5M cimetidine for 6 days, the activity of the He receptor is not altered after the antagonist is eliminated from the incubation medium by successive washes. Under the same conditions, cimetidine does not modify its own Ha receptor antagonism, which confirms that desensitization occurs subsequently to receptor activation (115). The gastric cells are not sensitized to histamine following a prolonged blockade by cimetidine (113, 114, 212, 275). The hypersecretion of acid, observed by several authors following the cessation of treatment by H2-receptor antagonists is thus open to question. There is no pharmacological evidence, from experimental protocols using isolated gastric cells which support a secretory rebound with H2 antagonists (212, 275). Cimetidine protects the H2 receptor against the desensitization induced by histamine (115). This protective effect occurs at elevated. concentrations of cimetidine (10 .5 to 10 -4 M). Those concentrations are in the range of values measured from peripheral blood (4 to 8 x 10-6 M) during the first eight hours following a standard administration of cimetidine (39). The interaction between cimetidine and the H2 receptor is thus rapidly reversible and the same situation is observed in vivo (189). Compound SKF 93479 produces, to the contrary, a complete loss of H2 receptor activity without affecting the peptidergic receptors in the human gastric HGT-1 cells (115). This drug has a very slow

342

Mirossay, Di Gioia, Chastre, Emamiand Gespach

dissociation rate (about 45 rain). It is 50 times more potent if there is a 120 rain pre-incubation with the HGT-1 cells, as compared to simultaneous addition with histamine (113). The compound, SKF 93479, belongs to a class of H2 antagonists which includes loxtidine (a derivative of phenoxytriazole), ranitidine (furan), famotidine (thiazole), AH 22216 (triazole) and L-643, 441 (thiadiazole). These antagonists have a prolonged, partially irreversible activity (38, 50, 101, 113, 153, 287, 328, 437). This apparent effect of famotidine on the H2 receptor in human gastric HGT-1 cells is in accordance with Schild's representation, which yields a slope of n =0.64, different from unity, thus indicating a non-competititve antagonism (153). These results are equally compatible with the prolonged in vivo action of famotidine, both in animals and man (88, 161, 197, 294, 328, 397, 407, 425). In dogs with a gastric fistula, acid secretion is suppressed by 100%, 59% or 18% by the oral administration of 2mg/kg famotidine at 4, 24 and 48 hours preceeding the histamine test (328). In comparison, ranitidine, at a concentration of 15 mg/kg, exerts an inhibitory effect of 87% at four hours but only a marginal reduction of 9% when administered 24 hours before histamine. There is, however, some controversy concerning the relative irreversibility and the non-competitive inhibition by famotidine on the H2 receptor in gastric ceils isolated from animals. In particular, famotidine exerts a reversible antagonism in the rabbit and the unweaned rat (34,401) and a partially reversible antagonism in parietal cells isolated from the adult guinea pig (169). A linear Schild plot representation, with a pA2 of 7.5 for acid secretion, has been reported in mice (6.03 and 5.35 for ranitidine and cimetidine, respectively) and a pA2 of 7.65 for gastric adenylate cyclase was measured in guinea-pig (6.33 for cimetidine) (37, 173). The histaminergic inhibition by famotidine is constant as a function of time in the mouse and the rabbit (37,401). These differences may be attributed to the diversity of tests used to evaluate the activity of H2 receptor and acid secretion in vivo and in vitro (89). The H2 receptor as well as the H + pump (vide infra) possess functional and pharmacological properties that differ with respect to their post-natal development in animal species: man, guinea pig or rabbit (1, 14, 149, 245). In respect to the maturation of the gastric epithelium during ontogenesis, and nutritional factors, such as milk, which can modify the activity of the H2 receptor (112, 147, 150, 151, 160). Lastly, the use of dispersive enzymes and prolonged incubation periods (3 to 5 hours) at 37~ necessary to isolate gastric cells or purified membranes (vide supra). These techniques may also result in a degradation or a desensitization of histaminergic and peptidergic receptors (113, 115, 366). Finally, certain Hz antagonists indirectly control the activity of the histaminergic receptor by inhibiting the degradation of histamine both in vivo and in vitro (430). Pharmacological Control of the Gastric H+/K+-ATPase

The H+/K+-ATPase (EC 3.6.1.36) is a membrane-associated and ouabaininsensitive enzyme that provides the driving force for HCI secretion into the gastric lumen, resulting a pH difference greater than 6units between the cytoplasm and the gastric lumen. The parietal cells undergo extensive mot-

Pharmacological Controm of Gastric Acid Secretion

343

phological transformation which results in H + secretion. These changes include an extensive reduction in the tubulovesicular system in the cytoplasm together with the appearance of intracellular secretory canaliculi (265). Several agents including histamine, gastrin, and carbachol induce the fusion of the vesicles with the luminal membrane or expansion of the apical plasma membrane. Each secretagogue induced rapid increases in steady-state levels of mRNAs encoding carbonic anhydrase II and H+/K+-ATPase in isolated canine gastric parietal cells (58). The H+/K+-ATPase catalyzes the electroneutral exchange of cytoplasmic H + and external K +, coupled with ATP hydrolysis. The proton pump is composed of 2 subunits, the 114kD catalytic ~ subunit concerned with ion translocation, and the/3 subunit is a 60-80 kD glycoprotein with a 35 kD protein core appearing to have no transport function. The enzyme is composed of 1034 residues. The Asp-386 residue (phosphorylation site), the Lys-497 residue (pyridoxal 5'-phosphate-binding site) and the Lys-518 residue (fluoresceinisothiocyanate-binding site) are located near the catalytic ATP-binding site. Omeprazole binds the cysteine residue of the el subunit. Both subunits span the plasma membrane, the a; subunit probably 8 times and the /3 subunit only once. The proton pump belongs to the family of P-type ATPases, which include Na+/K +ATPase and Ca2+-ATPase that forms the aspartyl phosphate intermediate during ATP hydrolysis. The recently cloned H+/K+-ATPase/3 subunits from the human, mouse, rat and bovine stomach show a number of structural similarities with the Na+/K+-ATPase/3 subnit. Both contain a large extracytosolic domain containing 3 pairs of disulfide-linked cysteine residues (61, 62, 260, 336, 402). The ol and/3 subunits of the Na+/K+-ATPase are both required for ATPase activity and for binding the inhibitor ouabain. Moreover, the H+/K+-ATPase/3 isoform can act as a substitute for the Na+/K+-ATPase /3 subunit. It can assemble to the oc Na+/K+-ATPase subunit and support the structural maturation, intracellular transport, and functional activity of the crNaK//3HK hybrid chimera, including Na+/K+-pump transport capabilities in Xenopus oocytes (196). Both pumps are expressed in parietal cells but at opposite poles of the cell. Kinetic enzyme studies have defined a complex succession of catalytic and transport reactions, involving K+/K + exchange, and enzyme phosphorylation in the presence of M g 2+ (249, 250, 336, 419). The results using isolated vesicles from the apical membrane of the oxyntic cells and permeabilized gastric glands suggest that the H+/K +ATPase exist in 2 functionally different compartments (184, 465). Stimulation promotes the recruitment of the enzyme from an inactive, K+-restricted, to an active, K+-permeable compartment. This electroneutral exchange of extracellular K + and intracellular H + generates an extremely high transmembrane pH gradient. Northern and Western blot analysis has shown that the H+/K+-ATPase and/3 subunits are expressed in the stomach of a variety of animal species, and specifically localized in the tubulovesicular network of the parietal cell (61,402). Both the ol and /3 subunits have been reported as major autoantigens in autoimmune gastritis associated with the pernicious anemia (54, 213, 273, 435). Pernicious anamia is the final stage of autoimmune gastritis and the resultant gastric lesions shows severe mucosal atrophy with ensuring failure of intrinsic factor secretion by the parietal cell. The nucleotide sequences of the 0~subunit of

344

Mirossay, Di Gioia, Chastre, Emami and Gespach

human, rat, and pig H+/K+-ATPase have been reported (262). The mouse /3 subunit gene is split into 7 exons, the intracellular amino-terminal and putative transmembrane domains are encoded by individual exons, the extracellular domain is encoded by 5 exons (62). The genes encoding the H+/K+-ATPase ol and /3 subunits appear to be expressed exclusively in parietal cells. The availability of cosmid clones containing 28 kb of sequences flanking the 5' end of the H+/K+-ATPase/3 subunit gene should allow others to study factors involved in the tissue-specific expression of this gene in parietal cells and the response elements involved in transcription (62). The upstream region of the rat H+/K+-ATPase /3 subunit gene contains repeat sequences and palindromes, potential binding sites for RNA polymerase II and E4TF1, and CACCC box sequences (263). The stomach, but not other tissues, has nuclear proteins capable of binding to the regions upstream of the potential RNA polymerase II binding sites (TATA box). Sequence analysis of the human H+/K+-ATPase ~ subunit promoter revealed the presence of sequence motifs that were also present in the promoter region of the human and rat pepsinogen genes (179, 203, 214, 262, 323). The 3 tandem GATAGC sequences in the rat gene may be important for controlled expression of the H+/K+-ATPase /3 subunit gene in gastric parietal cells (263). These observations suggest the potential role for stomach-specific transcriptional response elements. Further characterization and isolation of these cis-acting elements should allow the construction of efficient vectors containing fusion genes for appropriate regional, temproal and cell-specific expression in gastric epithelia of transgenic animals. Maturative changes, biochemical, and genetic aspects of the H+/K+-ATPase ontogeny have been studied during fetal and postnatal development in various animal species (82, 186, 274, 340). The H+/K+-ATPase can be directly blocked by a number of benzimidazolie derivatives, including omeprazole (Fig. 5) and picoprazole (59, 160, 249, 452-454). Consequently there exists different drugs that can control acid secretion by acting either on the basal pole of the parietal cell (cholinergic, gastrinic and H2 receptor antagonists) or the apical pole (omeprazole, trifluoperazine, verapamil and trimethoxybenzoate). Other molecules possessing a tertiary amine functional group inhibit the K + site of proton ATPase: trifluoperazine, an inhibitor of calmodulin, verapamil and diethylaminooctyl trimethoxybenzoate (160, 200, 307). Other compounds also inhibit gastric acid secretion as non-competitive inhibitors of the H+/K+-ATPase (339, 341), polyamines and fenoctimine (a substituted piperidine), or as competitive inhibitors of the H+/K+-ATPase, sofalcone (an isoprenyl chacone competitive at the ATP site), SCH28080 (a competitor of the K + site of the enzyme), the sulfoxide reagent Ro 18-5364 and the substituted benzimidazole AG-1749 containing a trifluoroethoxy group (27,300, 305,403). Bafilomycin, on the other hand, was much more effective in inhibiting proton transport mediated by the H+-ATPases in the kidney and bone (ICs0=2nM) than in inhibiting the H+/K+-ATPase in gastric membrane vesicles, ICs0 = 50 #M (282). Omeprazole consists of a benzimidazole core and of a pyridine core substituted with methyl and oxymethyl groups and irreversibly inhibits the proton pump Fig. 5. When acid secretion is stimulated, the half-life of omeprazole is 3 min; when secretion is

Pharmacological Control of Gastric Acid Secretion

345

inhibited, the half-life increases to 73 rain. Omeprazole is a prodrug. It is activated by acid and converted in the intravesicular space and in secretory channels to its active form, a sulphenamide that interacts with the luminally accessible SH-groups of the enzyme (128, 218, 258, 298, 452-454). The compound fi-Mercaptoethanol prevented binding of the inhibitor and inhibition of the enzyme. Omeprazole was also shown to interact with the Na+/K+-ATPase from dog kidney (218), at a lower affinity (ECso = 19-186/~m) than the H+/K+-ATPase (ECso = 5.2-36 ~m). Omeprazole is a potent, long-lasting inhibitor of gastric acid secretion in animals and humans with gastric and duodenal ulcers, gastroesophageal reflux disease, and Zollinger-Ellison syndrome (129, t92, 235,448). Basal H + secretion was inhibited by 50%, 3 h after a single 60 mg dose of omeprazole, and by 78%, 4 h after administration of the drug (284). Omeprazole is absorbed from small intestine and its half-life in plasma is about 60 min. The duration of action following a single dose of omeprazole exceeds 24 h because of the stable complex between the enzyme and inhibitor. Side Effects of Antiacid Drugs The control of gastric acid secretion can be achieved by the pharmacological control of histamine H2 receptors, cholinergic receptors, H+/K+-ATPase, and combined with the introduction of other possible therapeutic agents, such as PG or new drugs with antiulcer and antisecretory activities (93). With the exception of ranitidine and famotidine, most of the anti-secretory agents that possess a prolonged or irreversible action have been withdrawn from preclinical experimentation (clinical trials) following the demonstration of their toxic effects, including the appearance of gastric cancers in the rat for compound SKF 93479, tiotidine and loxtidine. The toxic and secondary effects of the H 2 antagonists presently prescribed for peptic ulcer and the Zollinger-Ellison syndrome treatment (376) are compared with the toxicity of omeprazole and the prostaglandin analogues (enprostil and misoprostol, for example)~ This comparison is presented in Table 1. These effects are variable and depend on whether the administration is an oral or parenteral route. The administration of loxtidine, tiotidine, as well as of the H+/K+-ATPase inhibitors omeprazole and the substituted benzimidazole analogue BY308, induce achlorhydria and hypergastrinemia followed by neoplastic transformation and hyperplasia of the enterochromaffin cells into rat oxyntic mucosa (50, 92, 106, 108,234, 242,278,333,367, 434). There exists also the evidence that omeprazole may have a possible genotoxic effect on gastric mucosal cells. Its administration caused a significant increase in the uptake of radiolabelled thymidine into DNA in gastric mucosal cells, while the effect of loxtidine was negative. This effect of omeprazole can not be explained on the basis of increased gastrin levels which is identical after loxtidine and omeprazole treatment (53). A potential role for the anti-secretory drugs on carcinogenic N-nitrosylated compounds produced by the bacterial flora, was demonstrated in the rat. The same situation has been observed only in 39 treated patients and three controls out of 9,504 patients taking cimetidine and 8,994 matched controls (278). It is therefore, difficult

Mirossay, Di Gioia, Chastre, Emami and Gespach

346 T a b l e 1.

TISSUES

Toxic and side-effects produced by antiacid drugs during the treatment of gastric ulcers and the ZoUinger-EUison syndrome CIM

STOMACH Cancers/carcinomas Hypergastrinemia + Proliferation, hyperplasia Mucosa ECL Somatostatin P+ Serotonine Gastrin P Endogenous PG INTESTINE Trophic effects + Constipation, diarrhea + LIVER Cytochrome P450 ++ CARDIO-VASCULAR SYSTEM Inotropic effects, hypotension +/Endocrine system Anti-androgene + PRL, LH, FSH, TSH, PTH, + T4, E BLOOD, HEMATOPOIESIS Cytopenia + Coagulation + CENTRAL NERVOUS SYSTEM Confusion, headache, attention +

RA

FA

OME

+

-

rat: + ++

+

P P+ P+/P+

+

+ +

+

PG

p+/H + p+/H + p+/+ p+/H +

+ p+/H + p+ p-

++

+

++

+/+/-

+/-

+/-

+/-

+ +

+/-?

+

--?

cimetidine; RA: ranitidine; FA: famotidine; OME: omeprazole; PG: prostaglandins. (+1) and ( - ) : presence and absence of adverse effects; P+ and H+: stimulation of cell proliferation (P) and induction of hyperplasia (H) CIM:

to conceive that these antisecretory agents would have a significant carcinogenic effect in man. In agreement, omeprazole does not alter the morphological characteristics of the endocrine cells in the oxyntic mucosa of patients given omeprazole for duodenal ulcer of Zollinger-Ellison syndrome (86, 280). Patients with Zollinger-Ellison syndrome given omeprazole for up to 3 years developed no significant changes in percentage of enterochromaffin ECL cells, no carcinoid tumors, and no changes in gastrin serum concentrations. The increase in serum gastrin levels observed in duodenal ulcer patients during drug therapy has a very low trophic effect, if any, on human gastric ECL cells. ECL cells are currently regarded as a major source of histamine in the human oxyntic mucosa. The rat oncogenicity studies are an inadequate risk model for human due to the sharp differences observed (32). Certain antisecretory agents, including cimetidine and omeprazole, interfere with the metabolism of certain drugs by cytochrome P450, a constituent of the microsomal oxidative system in the liver (20, 96, 134, 233,251). The presence of an imidazole core in cimetidine, a benzimidazole moiety in omeprazole and a furan ring in ranitidine enables these drugs to fix to the heme iron of cytochromone P450 (96). Famotidine, which possesses a thiazole ring, is

PharmacologicalControl of Gastric Acid Secretion

347

prescribed at weaker therapeutic doses and does not interfere with the oxidative metabolism of the drugs. It appears to be established that treatment with cimetidine or ranitidine can affect cardiac rhythm and blood pressure whereas other secondary effects are generally minor (267, 429). It is still controversial whether famotidine has the same side-effects. Hematological, anti-androgen (impotence, gynecomastia) and endocrine effects have been described, especially for cimetidine (210). At therapeutic doses, this antagonist inhibits the binding of dihydrotestosterone to androgen receptors and increases the concentration of circulating testosterone in marl. Ranitidine and famotidine do not produce this non-specific effect on the androgen receptors (81, 131, 252, 326). Cimetidine crosses the placental barrier by passive transport, is concentrated in milk and produces a sexual dysfunction in the adult rat (4,379). This antagonist should thus be used with caution in women who are pregnant or nursing. The adverse effects of H2 antagonists on the brain, the cardiovascular system, and hematopoiesis are due to the presence of histaminergic receptors in the cerebral structures, the cardiac muscle, the endothelial cells in brain and peripheral tissue, in both stem cells and mature blood cell (31, 89, 97, 160, 208, 211, 220, 232). These cells, whether normal or leukemic, possess H2-type receptors, coupled to AC (totipotent CFUs stem cells, granulocytes, lymphocytes, monocytes) or posses H2h-type receptors, coupled to guanylate cyclase as in the case of the blood platelets (56, 142, 146, 156, 157, 159, 160). The H2 receptors induce the transition from Go to S of the cell cycle in the precursors FU gm and stimulate the proliferation and differentiation of granulocytes (t5, 215, 306). Histamine, via the H2 receptors inhibits the secretion of free radicals by the monocytes/macrophages lignages, as well as chemotaxis, phagocytosis and the secretion of lysosomal enzymes (52, 146). Thus, cimetidine, ranitidine, and the other antagonists at histamine H2 receptors are likely to affect allergic, immune and inflammatory reactions, whereas they would probably have an inhibitory effect on hematopoiesis. On the other hand, the undesirable effects observed during the treatment of peptic ulcers may be beneficial in certain pathological situations in which the histaminergic receptors are implicated (52).

CONCLUSIONS The recent progress in the pharmacology of drugs directed against peptic ulcers shows that this major pathology can now be managed by a series of compounds that control either the activity of membrane receptors activated by histamine (H2 receptors), gastrin and acetylcholine, or, alternatively, the functioning of the gastric H+/K+-ATPase. The antagonism at H2 receptors has been substantiated through the development of new, powerful, selective drugs having prolonged activity and ability to regulate acid secretion without blocking it irreversibly. These characteristics are at the origin of the reduction of certain secondary effects observed during treatment with omeprazole and cimetidine: hypergastrinemia, hyperplasia of the gastric epithelial cells and rearrangement of the relative density of gastrin- and somatostatin-secreting cells, interference with

348

Mirossay, Di Gioia, Chastre, Emami and Gespach

the liver cytochrome P450, action on cardiac rhythm and hematologic and anti-androgen effects. The success in molecular cloning of the genes encoding the histamine H1 and H2 receptors (138, 139, 362, 470) might be very helpful in ulcer therapy based on the control of gastric acid secretion by H2 receptor antagonists and cloning of the H 3 receptor gene. The evidence that the 4.5 kb DNA sequence cloned is that of the H2 receptor gene was based on the following data. Histamine increased in a dose-dependent manner cellular cAMP content in L cells permanently transfected with the H2 receptor gene. The H2-selective antagonist cimetidine shifted the histamine dose-response curve to the right. The H2 receptor-selective ligand (3H)tiotidine binding was inhibited by cimetidine. Expression of H2 receptor gene by Northern blot analysis was found in parietal cells of the fundic mucosa but not in chief cells, liver, ileum, heart, or adrenal glands (139). These elegant studies by Gantz et al. should provide essential information regarding the expression of the histamine H2 receptor gene using polyadenylated RNAs or PCR in tissues suspected to express the H2 receptor gene and those that accumulate the messenger at low levels. The comparison of the aligned sequences and conformations of the human~ canine and rat histamine H2 receptors should elucidate the molecular domains and the amino acid sequences of the H2 receptor structure serving as flexible elements responsible for the conformational variants that are involved in the acquisition of active/inactive forms of the transmembrane binding receptor and its coupling to the corresponding Gs protein complex. In addition these analyses should also identify the molecular domains of the H2 receptor that are essential for its three dimensional architecture. Molecular rearrangements between the receptor and the ligand may therefore play a crucial role in the acquisition of the activated forms of the receptor binding domain and the resulting coupling of the transmembrane protein to the G protein. In order to improve our actual knowledge on the tripartite interaction among histamine, the antagonist and the H2 receptor, the search for better H2 antagonists should be coordinated in view of the molecular modelling of the H2 histaminergic receptor structure, and the design of test compounds with potentially specific biological activity (436). Previous investigations on structureactivity of agonists and antagonists at H2 receptors (116) may be considered together with site-directed mutagenesis of the H2 receptor gene and analysis of the histamine conformers in biological fluids (449). In this connection, it should be stressed that deletions and mutations in the receptor structure, its putative ligand recognition domains and c~s protein interaction domains may influence the functional targeting of the proreceptor in the plasma membrane after gene transfer, to translocate aberrant forms of the binding site that are not able to undergo the activated receptor conformations after ligand interaction. These limitations are extended by the possibility that the active sites necessary to induce these receptor/ligand conformational changes are not necessarily located in adjacent aminoacids. Similar ligand structures in peptide hormones- and histamine-may therefore activate very different membrane receptors by their aminoacid sequence and molecular architecture. We have presented here a brief overview of the complexity and of the diversity of the regulatory networks implicated in acid secretion and the

Pharmacological Control of Gastric Acid Secretion

349

p r e s e r v a t i o n o f t h e g a s t r o - d u o d e n a l m u c o s a . This s t a t u s q u o is i n t e g r a t e d in t h e i m m e d i a t e r e s p o n s e o f t h e c e l l u l a r m a c h i n e r y to t h e s i g n a l i n g s y s t e m s , a n d in t h e m e m o r i z a t i o n of this signal b y a d a p t a t i o n to a n o t h e r s t i m u l u s , a n d finally, in a late r e s p o n s e which i n v o l v e s a g e n e t i c e x p r e s s i o n , t h e r e n e w a l a n d t h e f u n c t i o n a l d i f f e r e n t i a t i o n o f t h e cells o f t h e g a s t r i c e p i t h e l i u m . F u r t h e r i d e n t i f i c a t i o n o f t h e e x t e r n a l a n d n u c l e a r factors with p o s i t i v e o r n e g a t i v e i n f l u e n c e o n t h e p r o m o t e r / e n h a n c e r r e s p o n s e e l e m e n t s o f t h e g e n e s e n c o d i n g t h e H2 r e c e p t o r o r the H + / K + - A T P a s e s u b u n i t s will p r o v i d e n e w insights in t h e p h y s i o l o g i c a l , m o l e c u l a r a n d p h a r m a c o l o g i c a l c o n t r o l o f gastric s e c r e t i o n . I m p r o v i n g o u r k n o w l e d g e o f h i s t a m i n e s e c r e t i o n a n d t h e m e c h a n i s m s of gastric s e c r e t i o n b y p h y s i o l o g i c a l r e g u l a t o r s , as well as s t u d y i n g t h e r e p a i r s y s t e m s o f t h e gastric m u c o s a , s h o u l d l e a d to a r e f i n e m e n t in t h e use o f m e d i c a t i o n for t h e t r e a t m e n t o f p e p t i c ulcers.

ACKNOWLEDGEMENTS A i d e d by a G r a n t f r o m M e r c k S h a r p a n d D o h m e C h i b r e t to C. G . (1992)

REFERENCES I. Ackerman, S. H. (1982) Ontogeny of gastric acid secretion in the rat: Evidence for multiple response systems. Science 217: 75. 2. Aktories, A., Schultz, G. and Jakobs, K. H. (1983) Somatostatin-induced stimulation of a high-affinity GTPase in membranes of $49 lymphoma cyc- and H21a variants. Molecular Pharm. 24:183. 3. Amatruda III, T. T., Steele, D., Slepak, V. Z. and Simon, M. (1991) Goal6, a G protein 0; subunit specifically expressed in hematopoietic cells. Proc. Natl. Acad. Sci. USA 88: 5587. 4. Anand, S. and Van Thiel, D. H. (1982) Prenatal and neonatal exposure to cimetidine results in gonadal and sexual dysfunction in adult males. Science 218:493. 5. Anderson, N. G. and Hanson, P. J. (1985) Involvement of calcium-sensitive phospholipiddependent protein kinase in control of acid secretion by isolated rat parietal cells. Biochem. J. 232: 609. 6. Arima, N., Yamashita, Y., Nakata, H., Nakamura, A., Kinoshita, Y. and Chiba, T. (1991). Presence of histamine H2-receptors on human gastric carcinoma cell line MKN-45 and their increase by retinoic acid treatment. Biochem. Biophys. Res. Commun. 176: 1027. 7. Arrang, J. M., Garbarg, M., Lancelot, J. C., Lecomte, J. M., Pollard, H., Robba, M., Schunack, W. and Schwartz, J. C. (1987) Highly potent and selective ligands for histamine H3-receptors. Nature 327:117. 8. Arrang, J. M., Garbarg, M. and Schwartz, J. C. (1983) Auto-inhiNtion of brain histamine release mediated by a novel class (H3) of histamine receptor. Nature 302: 832. 9. Arrang, J. M., Roy, J., Morgat, J. L., Schunack, W. and Schwartz, J. C. (1990) Histamine H3 receptor binding sites in rat brain membranes. Modulations by guanyl nucleotides and divalent cations. Eur. J. Pharmacol. 188:219. 10. Ashkenazi, A., Winslow, J. W., Peralta, E. G., Peterson, G. L., Schimerlik, M. I., Capon, D. J. and Ramachandran, J. (1987) An M2 muscarinic receptor subtype coupled to both adenylyl and phosphoinositide turnover. Science 238: 672. 11. Atwell, M. M. and Hanson, P. J. (1988) Effect of pertussis toxin on the inhibition of secretory activity by prostaglandin E2, somatostatin, epidermal growth factor and 12-Otetradecanoiylphorbol 13-acetate in parietal cells from rat stomach. Biochim. Biophys. Acta 971: 282.

350

Mirossay, Di Gio~a, Chastre, Emami and Gespach

12. Audousset-Puech, M. P., Jarrousse, C., Dubrasquet, M., Aumelas, A., Castro, B., Bataille, D. and Martinez, J. (1985) Synthesis of the C-terminal octapeptide of pig oxyntomodulin Lys-Arg-Asn-Lys-Asn-Asn-Ile-Ala: A potent inhibitor of pentagastrin-induced acid secretion. Y. Med. Chem 28:1529. 13. Auger, K. R., Serunian, L. A., Soltoff, S. P., Libby, P. and Cantley, L. C. (1989) PDGF-dependent tyrosine phosphorylation stimulates production of novel polyphosphoinositides in intact cells. Cell 57:167. 14. Avril, P., Ducroc, R., Garzon, B., Moreau, E., Hervatin, F., Millet, P. and Geloso, J. P. (1986) D6veloppement de la sensibilit6 gastrique ~t l'histamine chez le rat. C. R. Acad. Sci. 18:739. 15. Aymard, J. P., Aymard, B., Netter, P., Bannwarth, B., Trechot, P. and Streiff, F. (1988) Haematological adverse effects of histamine H2-receptor antagonists. Medical Toxicology 3: 430. 16. Azuma, J., Sawamura, A., Harada, H., Tanimoto, T., Ishiyama, T., Morita, Y., Yamamura, Y. and Sperelakis, N. (1981) Cyclic adenosine monophosphate modulation of contractility via slow Ca 2+ channels in chick heart. J. Mol. Cell. Cardiol. 13:577. 17. Bado, A., Moizo, L., Laigneau, J. P. and Lewin, M. J- M. (1992) Pharmacological characterization of histamine H 3 receptors in isolated rabbit gastric glands. Am. J. Physiol. 262: G56. 18. Bakalyar, H. A. and Reed, R. R. (1990) Identification of a Specialized Adenylyl Cyclase that May Mediate Odorant Detection. Science 250:1403. 19. Barr, D. B., Duncan, J. A., Kiernan, J. A., Soper, B. D. and Tepperman, B. L. (1988) Binding and biological actions of prostaglandin E2 and 12 in cells isolated from rabbit gastric mucosa. J. Physiol. 405: 39. 20. Bast, A., Smid, K. and Timmerman, H. (1989). The effects of cimetidine, ranitine and famotidine on rat hepatic microsomal cytochrome P-450 activities. Agents and Actions 27: 188. 21. Bataille, D. (1987)Ent6roglucagons: 6tat actuel. Gastroenter. Clin. Biol. 11:21% 22. Batzri, S., Harmon, J. W., Dyer, J. and Thompson, W. F. (1982) Interaction of histamine with gastric mucosal cells. Effect of histamine H2 antagonists on binding and biological response. Molec. Pharmacol. 22: 41. 23. Batzri, S. (1984) Action of tricyclic antidepressants on cyclic AMP in mucosal cells from guinea pig stomach. Eur. J. Pharmac. 105: 167. 24. Bawab, W., Chastre, E. and Gespach, C. (1991) Functional and structural characterization of the secretin receptors in rat gastric glands: Desensitization and glycoprotein nature. Biosci. Rep. 11:33. 25. Bearer, C. F., Chang, L. K., Rosenfeld, G. C. and Thompson, W. J. (1981) Histamine stimulation of rat gastric parietal cell adenylyl cyclase: modulation by guanine nucleotides. Arch. Biochem. Biophys. 207:325. 26. Beaven, M. A., Soil, A. H. and Lewin, K. J. (1982) Histamine synthesis by intact mast cells from canine fundic mucosa and liver. Gastroenterol. 82:254. 27. Beil, W., Staar, U. and Sewing, K. F. (1987) SCH 28080 is a more selective inhibitor than SCH 32651 at the K + site of gastric K§ Eur. J. Pharm. 139:349. 28. Berglindh, T., Helander, H. and Obrink, K. (1976) Effects of secretagogues on oxygen consumption, aminopyrine accumulation and morphology in isolated gastric glands. Acta Physiol. Scand. 97: 401. 29. Berglindh, T., Sachs, G. and Takeguchi, N. (1980) Ca2+-dependent secretagogue stimulation in isolated gastric glands. Am. J. Physiol. 239:G90. 30. Bergmark, J., Gran6rus, G., Henningsson, S., Lundell, L. and Rosengren, E. (1976) Histamine metabolism of the guinea pig gastric mucosa. J. Physiol. 257:419. 31. Berlin, R. G. (1987) Famotidine does not have a negative inotropic effect. The Lancet 2" 1468. 32. Berlin, R. G. (1991) Omeprazole: Gastrin and gastric endocrine cell data from clinical studies. Dig. Dis. Sci. 36:129. 33. Berridge, M. J. and Irvine, R. F. (1984) Inositol trisphosphate, a novel second messenger in cellular signal transduction. Nature 312: 315. 34. Bertaccini, G., Coruzzi, G., Poli, E. and Adami, M. (1986) Pharmacology of the novel H2 antagonist famotidine: in vitro studies. Agents and Actions 19:180. 35. Black, E. W., Cornwell, T. L., Lincoln, T. M., Strada, S. J., Thompson, W. J. (1989) Fura 2 analysis of cytosolic calcium regulation in elutriated rat gastric parietal cells. J. Cell. Physiol. 139: 632. 36. Black, J., Duncan, W., Durant, C., Ganellin, C. and Parsons, M. (1972) Definition and antagonism of histamine H2 receptors. Nature 236: 385. 37. Black, J. W., Left, P. and Shankley, N. P. (1985) Further analysis of anomalous pKB values for histamine H2-receptor antagonists on the mouse isolated stomach assay. Br. J. Pharmac. 86:581.

Pharmacological Control of Gastric Acid Secretion

351

38. Blakemore, R. C., Brown, T. H., Durant, G. J., Ganelin, C. R., Parsons, M. E., Rasmussen, A. C. and Rawlings, D. A. (1981) SKF 93479, a potent and long acting histamine H2-receptor antagonist. Proc. o f the B.P.S., April 200P. 39. Bodenmar, G., Norlander, B., Fransson, L. and Walan, A. (1979) The absorption of cimetidine before and during maintenance treatment with cimetidine and the influence of a meal on the absorption of cimetidine: studies in patients with peptic ulcer disease. Br. J. CIin. Pharmacot. 7:23. 40. Bohman, T., Myren, 3., Flaten, O. and Schrumpf, E. (1990) The effect of trimipramine, cimetidine and atropine on gastric secretion, Scand. J. Gastroenterol 15:177. 41. Boige, N., Dupont, C., Chrenut, B., Gespach, C. and Rosselin, G. (1984) Beta-adrenergic regulation of cyclic adenosine 3',5'monophosphate accumulation in human gastric epithelial glands. Inhibitory effect of somatostatin. Eur. J. Clin. Invest. 14:42. 42. Bonner, T. I. (1992) Domains of muscarinic acetylcholine receptors that confer specificity of G protein coupling. TIPS 13: 48. 43. Bonventre, J. V. and Swidler, M. (1988) Calcium dependency of prostaglandin E2 production in rat glomerular mesangial cells. J. Clin. Invest. 82: 168. 44. Borch, K., Renvall, H. and Liedberg, G. (1985) Gastric endocrine cell hyperplasia and carcinoid tumors in pernicious anemia. Gastroenterol. 88: 638. 45. Bordi, C. D'Adda, T., Pilato, F. P. and Ferrari, C. (1987) Carcinoid (ECL cell) tumor of the oxyntic mucosa of the stomach. A hormone-dependent neoplasm? Prog. Surg. Pathot. 8:177. 46. Bottari, S. (I989) Nuclear somatostatin receptors. The gastrointestinal epithelium, International conference, Riom September 24, France, abstr. 47. Brass, L. F., Woolkalis, M. J. and Manning, D. R. (1988) Interactions in platelets between G proteins and the agonists that stimulate phospholipase C and inhibit adenyly~ cyclase. J. Biol. Chem. 263: 5348. 48. Bredt, D. S., Hwang, P. M., Glatt, C. E., Lowenstein, C., Reed, R. R. and Snyder, S. H. (1991) Cloned and expressed nitric oxide synthase structurally resembles cytochrome P-450 reductase. Nature 351: 714. 49. Brimblecombe, R. W. and Duncan, W. A. M. (1977) The relevance to man of preclinical data for cimetidine. In: Burland WL, Simkins MA, eds. Cimetidine, Proceedings of the Second International Symposium on Histamine H2 receptor antagonists. Amsterdam: Exerpta Medica. 112. 50. Brittain, R. T., Jack, D., Reeves, J. J. and Stables, R. (1985) Pharmacological basis for the induction of gastric carcinoid tumours in the rat by loxtidine, an unsurmountable histamine H2-receptor blocking drug. Br. J. Pharmac. 85:843. 51. Buechler, W. A., Nakane, M. and Murad, F. (1991) Expression of soluble guanylate cyclase activity requires both enzyme subunits. Biochem. Biophys. Res. Commun. 174:351. 52. Burland, W. L. and Mills, J. G. (1982) The pathophysiological role of histamine and potential therapeutic uses of H1 and H2 anthihistamines. In: Wright PSG~ ed. Pharmacology of histamine receptors. The stone bridge press, Bristol, London, Boston, p 436. 53. Burlinson, B., Morriss, S., Gatehouse, D. G., Tweats, D. J. and Jackson, M. R. (1991) Uptake of tritiated thymidine by cells of the rat gastric mucosa after exposure to toxtidine or omeprazole. Mutagenesis 6:11. 54. Burman, P , Mardh, S., Norberg, L. and Karlsson, A. (1989) Parietal Cell Antibodies in Pernicious Anemia Inhibit H +, K+-Adenosine Triphosphatase, the Proton Pump of the Stomach. Gastroenterol. 96: 1434. 55. Burns, D. J. and Bell, R. M. (1991) Protein Kinase C Contains Two PhorboI Ester Binding Domains. J. Biol. Chem. 266: 18330. 56. Byron, J. W. (1980) Pharmacodynamic basis for the interaction of cimetidine with the bone marrow stem cells (CFUs). Exp. Hematol. 8:256. 57. Cabero, J. L., Li, Z.-Q. and Mardh, S. (1991) Gastrin potentiates histamine-stimulated aminopyrine accumulation in isolated rat parietal ceils. Am. J. Physiol. 261:G621. 58. Campbell, V. W. and Yamada, T. (1989) Acid Secretagogue-i•duced Stimulation of Gastric Parietal cell gene expression. J. Biol. Chem. 264: 11381. 59. Campbell, V. W. and Yamada, T. (1991) Effect of omeprazole on gene expression in canine gastric parietal cells. Am. J. Physiol. 260:G434. 60~ Canfield, S. P., Curwain, B. P. and Price, C. (1978)/3-adrenoreceptor agonist stimulation of acid secretion in rat isolated stomach. Proc. o f the B.P.S. 425P. 61. Canfield, V. A., Okamoto, C. T., Chow, D., Dorfman, J., Gros, P., Forte, J. G. and Levenson, R. (1990) Cloning of the H, K-ATPase /3 Subunit. Tissue-specific expression, chromosomal assignment, and relationship to Na, K-ATPase/3 bubunits. J. BioL Chem. 265:8.

352

Mirossay, Di Gioia, Chastre, Emami and Gespach

62. Canfield, V. A. and Levenson, R. (1991) Structural organization and transcription of the mouse gastric H § K+-ATPase/3 subunit gene. Proc. Natl. Acad. Sci. USA 88:8247. 63. Cantley, L. C., Auger, K. R., Carpenter, C., Duckworth, B., Graziani, A., Kapeller, R. and Soltoff, S. (1990) Oncogenes and signal transduction. Cell. 64:281. 64. Carney, J. A., Go, V. L. W., Fairbanks, V. F., Moore, B. A., Alport, E. C. and Nora, F. E. (1983) The syndrome of gastric argyrophil carcinoid tumors and non-antral gastric atrophy. Ann. intern. Med. 99:761+ 65. Carpenter, C. L. and Cantley, L. C. (1990) Phosphoinositide kinases. Biochemistry 29:11147. 66. Carpenter, C. L., Duckworth, B. C+, Auger, K. R., Cohen, B., Schaffhausen, B. S. and Cantley, L. C. (1990) Purification and characterization of phosphoinositide 3-kinase from rat liver. J. Biol. Chem. 265:19704. 67. Carpentier, J. L., Gorden, P., Robert, A. and Orci, L. (1986) Internalization of polypeptide hormones and receptor recycling. Experientia 42:734. 68. Chang, R. S. L., Lotti, V. J. and Chen, T. B. (1985) Cholecystokinin receptor mediated hydrolysis of inositol phospholipids in guinea pig gastric glands. Life Sci. 36:965. 69. Charbon, G. A., Brouwers, H+ A. and Sala, A. (1980) Histamine H1- and H2-receptors in the gastrointestinal circulation. Naunyn-Schemiedeberg's Arch. Pharmacol., 312:123. 70. Chastre, E., Emami, S. and Gespach, C. (1989) Expression of membrane receptors and (proto)oncogenes during the oncogenic development and neoplasic transformation of the intestinal mucosa. Life Sci. 44:1721. 71. Chastre, E., Emami, S. and Gespach, C. (1991) Imrnortalisation et transformation tumorale de l'6pith61ium gastro-intestinal chez l'homme et le rat: applications en canc6rologie et dans la mucoviscidose. M3d. Sci. 7:17. 72. Cheng, H. C., Gleason, E. M., Nathan, B. A., Lachman, P. J. and Woodward, J. K. (1981) Effects of clonidine on gastric acid secretion in the rat. J. Pharmacol. Exp. Ther. 217: 121. 73. Cheret, A. M., Pignal, F. and Lewin, J. M. (1981) Effects of H2-receptor antagonists cimetidine, ranitidine and ICI 125,211 on histamine-stimulated adenylate cyclase activity in guinea pig gastric mucosa. Molec. Pharmac. 20: 326. 74. Chew, C. (1985) Parietal cell protein kinases selective activation of type I cAMP-dependent protein kinase by histamine. J. Biol. Chem. 260:7540. 75. Chew, C. S., Ljungstr6m, M., Smolka, A. and Brown, M. R. (1989) Primary culture of secretagogue-responsive parietal cells from rabbit gastric mucosa. Am. J. Physiol. 256: G254. 76. Chiba, T., Fisher, S. K., Agranoff, B. W. and Yamada, T. (1989) Autoregulation of muscarinic and gastrin receptors on gastric cells. Am. J. Physiol. 256: G356. 77. Cho-Chung, Y. S. (1990) Role of cyclic AMP receptor protein in growth, differentiation, and suppression of malignancy: New approaches to therapy. Canc. Res. 50:7093. 78. Christiansen, J., Ottenjann, R., Von Arx and the study group (1989). Placebo-controlled trial with the somatostatin analogue SMS 201-995 in peptic ulcer bleeding. Gastroenterol. 97: 568. 79. Clegg, C. H., Codd, G. G. and McKnight, G. S. (1988) Genetic characterization of a brain specific form of the type I regulatory subunit of cAMP-dependent protein kinase. Proc. Natl. Acad. Sci. USA, 85: 3703. 80. Cockcroft, S. (1987) Polyphosphoinositide phosphodiesterase: regulation by a novel guanine nucleotide binding protein, Gp. Trends in Biochem. Sci. 12:75. 81. Corinaldesi, D. R., Pasquali, R., Paternico, A., Stanghellini, V., Paparo, G. F., Ricci, Maccarini, M. and Barbara, L. (1987) Effects of short- and long-term administrations of famotidine and ranitidine on some pituitary, sexual and thyroid hormones. Drugs Explt. Clin. Res. 10: 647. 82. Crothers, J. M., Reenstra, W. and Forte, J. G. (1990) Ontogeny of gastric H+-K+-ATPase in suckling rabbits. Am. J. Physiol. 257: G913. 83. Culp, D. J., Wolosin, J. M., Soil, A. H. and Forte, J. G. (1983) Muscarinic receptors and guanylate cyclase in mammalian gastric glandular cells. Am. J. Physiol. 245:G760. 84. Currie, M. G., Fok, K. F., Kato, J., Moore, R. J., Hamra, F, K., Duffin, K. L. and Smith, C. E. (1992) Guanylin: An endogenous activator of intestinal guanylate cyclase. Proc. Natl. Acad. Sci. USA 89: 947. 85. Curwain, B. P. and Turner, N. C. (1981) The involvement of histamine receptor subtypes in gastric acid secretion and mucosal blood flow in the anaesthetized rabbit. J. Physiol. 311:431. 86. D'Adda, T., Pilato, F. P., Lazzaroni, M., Robutti, F., Bianchi-Porro, G. and Bordi, C. (1991) Ultrastructural Morphometry of Gastric Endocrine Cells Before and After Omeprazole. A study in the oxyntic mucosa of duodenal ulcer patients. Gastroenterol. 100:1563. 87. Daly, M. J., Humphray, J. M. and Stables, R. (198l) Some in vitro and in vivo actions of the new histamine H2-receptor antagonist, ranitidine. Br. J. Pharmac. 72:49.

Pharmacological Control of Gastric Acid Secretion

353

88. Dammann, H. G., Walter, T. A., Hentschel, E., Muller, P. and Simon, B. (1987) Famotidine: proven once-a-day treatment for gastric ulcer. Scand. J. Gastroenterol. 22:29. 89. Dam Trung Tuong, M., Garbarg, M. and Schwartz, J. C. (1980) Pharmacological specificity of brain histamine H2-reeeptors differs in intact cells and cell-free preparations. Nature 28:548. 90. Daniel, H., Vohwinkel, M. and Rehner, G. (1990) Effect of casein and A6-casomorphins on gastrointestinal motility in rats. J. Nutr. 120: 252. 91. Davison, J. S. and Najafi-Farashah, A. (1982) The effect of dibutyryl cyclic GMP, a competitive antagonist to cholecystokinin-pancreozymin, on gastric acid secretion in the isolated mouse stomach. Life Sci. 31:355. 92. Decktor, D. L., Pendleton, R. G., Kellner, A. T. and Davis, M. A. (1989) Acut effects of ranitidine, famotidine and omeprazole on plasma gastrin in the rat. J. Pharmac. Exp. Ther. 249:1.

93. Dedieu-Chaufour, C., Hertz, F., Caussade, F. and Cloarec, A. (1991) Pharmacological Profile of UP 5145-52, an Original Antiulcer and Antisecretory Agent. J. Pharmacol. Exp. Ther. 259:190. 94. Dembinski, A., Gregory, H., Konturek, S. J. and Polanski, M. (1982) Trophic action of epidermal growth factor on the pancreas and gastrointestinal mucosa in rats. J. Physiol. 325: 35. 95. Depigny, C., Lupo, B., Kervran, A. and Bataille, D. (1984) Mise en 6vidence d'un site r6cepteur sp6cifique du glucagon-37 (Oxyntomodulin/Enteroglucagon bioactif) dans les glandes oxyntiques de rat. C. R. Acad. Sci. Paris 299:677. 96. Dhumeaux, D. and Delehier, J. C. (1988) Interactions des inhibiteurs de la s6er6tion acide gastrique avec le syst~me microsomal h6patique d'oxydatJon des m6dicaments. In: Contr61e de la s6cr6tion acide, Mignon, M., Galmiche, J. P., Ruszniewski, P. and Pappo, M. John Libbey Eurotext, Paris, 129. 97. Dickey, W. and Symington, M. (1987) Broad-complex tachycardia after intravenous cimetidine. The Lancet i 8523: 99. 98. Dickinson, K. E. J., Anderson, W., Matsumoto, H. and Uemura, Hirschowitz, B. I. (1989) ~25I-iodopindolol binding to frog esophageal peptic cells. Detection of amine uptake and ~-adrenergic receptors coupled to pepsinogen secretion. Biochem. Pharm. 38:1481. 99. Diez, E. and Mong, S. (1990) Purification of a phospholipase A2 from human monocytic leukemic U937 cells. J. Biol. Chem. 265: 14654. 100. Dinda, P. K., Leddin, D. J. and Beck, I. T. (1988) Histamine is involved in ethanol-induced jejunal microvascular injury in rabbits. Gastroenterol. 95:1227. 101. Dobrilla, G., De Pretis, G., Piazzi, L., Boero, A , Camarri, E., Crespi, M., Fontana, G., Ideo, G., Manenti, F., Marenco, G., Morettini, A., Palasciano, G., Rossini, F. and Tittobello, -A. (1988) Comparison of once-daily bedtime administration of famotidine and ranitidine in the short-time treatment of duodenal ulcer. A multicenter, double-blind, controlled study. Scand. J. Gastroenterol. 22: 21. 102. Dohlman, H. G., Caron, M. G. and Lefkowitz, R. J. (1987) A family of receptors coupled to guanine nucleotide regulatory proteins. Biochem. 26: 2657. 103. Dousa, T. P. and Code, C. F. (1974) Effect of histamine and its methyl derivatives on cyclic AMP metabolism in gastric mucosa and its blockade by an H2 receptor antagonist. J. Ctin. Invest. 53:334. 104. Durant, G. J., Ganellin, C. R. and Parsons, M. E. (1975) Chemical differentiation of histamine H1- and H2-receptor agonists. J. Med. Chem. 18:905. 105. Ecknauer, R., Thompson, W. J., Johnson, L. R. and Rosenfeld, G. C. (1980) Isolated parietal cells: [3H] QNB binding to putative cholinergic receptors. Am. J. PhysioL 239: G204. 106. Eissele, R., RoBkopf, B., Koop, H., Adler, G. and Arnold, R. (1991) Proliferation of Endocrine Cells in the Rat Stomach Caused by Drug-Induced Achlorhydria. Gastroenterol. 101: 70. 107. Ekblad, E., Ekelund, M., Graffner, H., Hakanson, R. and Sundler, F. (1985) Peptidecontaining nerve fibers in the stomach wall of rat and mouse. Gastroenterol. 89:73. 108. Ekman, L., Hansson, E., Havu, N., Carlsson, E. and Lundberg, C. (1985) Toxicological studies on omeprazole. Scan& J. Gastroenter. 20:53. 109. Ekman, R., Ivarsson, S. and Jansson, L. (1985) Bombesin, neurotensin and pro-y-melanotropin immunoreactants in human milk. Regulatory Peptides 10:99. 130. Elouaer-Blanc, L., Sobhani, I., Ruszniewski, P., Duet, M., Lehy, T., Mignon, M.. Bonfils, S. and Lewin, M. J. M. (1988) Gastrin secretion by gastrinoma cells in long-term culture. Am. J. Physiol. 255: G596. 111. Emami, S., Chastre, E., Bod6r6, H., Gespach, C., Bataille, D. and Rosselin, G. (1986) Functional receptors for VIP, GIP, glucagon-29 and 37- in the HGT-1 human gastric cancer cell line. Peptides 7:121.

354

Mirossay, Di Gioia, Chastre, Emami and Gespach

112. Emami, S., Chastre, E., MuHiez, N., Gonzales, M. and Gespach, C. (1986) VIP and histamine H2 receptor activity in human fetal gastric glands. Experientia 42:423. 113. Emami, S., Gespach, C. and Bod6r6, H. (1985) Selective disappearance of histamine H2-receptor activity in the human gastric cell line HGT-1 after short-term or chronic treatment by histamine or its H2-antagonists. Agents and Actions 16:195. 114. Emami, S., Gespach, C., Forgue-Lafitte, M. E., Broer, Y. and Rosselin, G. (1983). Histamine and VIP interactions with receptor-cyclic AMP systems in the human gastric cancer cell line HGT-1. Life Sci. 33:415. 115. Emami, S. and Gespach, C. (1986) Desensitization by histamine of H2 receptor activity in HGT-1 human cancerous gastric cells. Agents and Actions 18: 129. 116. Emami, S. and Gespach, C. (1986) Pharmacology of histamine H2 receptor antagonists in the human gastric cancer cell line HGT-1. Structure-activity relationship of isocytosine-furan and imidazole derivatives related to cimetidine. Biochem. Pharmac. 35:1825. 117. Emami, S., Mir, L., Gespach, C. and Rosselin, G. (1989) Transfection of fetal rat intestinal epithelial cells by viral oncogenes: Establishment and characterization of the E1A-immortalized SLC-11 cell line. Proc. Natl. Acad. Sci. USA. 86:3194. 118. Emorine, L., Marullo, S., Briend-Sutren, M., Patey, G., YTate, K., Delavier-Klutchko, C. and Strosberg, A. D. (1989) Molecular characterization of the human /33 adrenergic receptor. Science 245:1118. H9. Euler, A. R., Bailey, R. J., Zinny, M. A., Brandon, M. L., Rousseau, B., Ferguson, J. P., Wood, D. R. and Le, V. H. (1989) Arbaprostil [15(R)-15-methyl prostagladin E2] in a single nighttime dose of either 50 or 100 mg in acute duodenal ulcer. Gastroenterology 97: 98. 120. Evangelista, S., Lippe, I. T., Rovero, P., Maggi, C. A. and Meli, A. (1989) Tachykinin protect against ethanol-induced gastric lesions in rats. Peptides 10:79. 121. Fanning, J. C., Tyler, M. J. and Shearman, D. J. C. (1982) Converting a stomach to a uterus: The microscopic structure of the stomach of the gastric brooding frog rheobatrachus silus. Gastroenterol. 82: 62. 122. Fedan, J. S., Hogaboom, G. K., Westfall, D. P. and O'DonneU, J. P. (1983) Photoaffinity labeling of P2-purinergic and HI-histamine receptors in intact smooth muscle. Feder. Proc. 42: 2846. 123. Felder, C. C., Kanterman, R. Y., Ma, A. L. and Axelrod, J. (1989) A transfected M1 muscarinic acetylcholine receptor stimulates adenylate cyclase via phosphatidylinositol hydrolysis. J. Biol. Chem. 264:20356. 124. Fernandez-Gallardo, S., Gijon, M. A., Garcia, M. del C., Cano, E. and Sanchez Crespo, M. (1988) Biosysthesis of platelet activating-factor in glandular gastric mucosa. Evidence for the involvement of the "de novo" pathway and modulation by fatty acids. Biochem. J. 254:707. 125. Ferrado, T., Rainero, I., De Gennaro, T., Ogerro, R., Mostert, M., Dottola, P. and Pinessi, L. (1990)/3-endorphin-like and a-MSH-like immunoreactivities in human milk. Life Sci. 47: 633. 126. Forgue-Laffitte, M. E., Kobari, L., Gespach, C., Chamblier, M. C. and Rosselin, G. (1984) Characterization and repartition of epidermal growth factor-urogastrone receptors in gastric glands isolated from young and adult guinea pigs. Biochim. Biophys. Acta 798: 192. 127. Fraser, C. M. (1989) Site-directed mutagenesis of /3-adrenergic receptors. Identification of conserved cysteine residues that independently affect ligand binding and receptor activation. J. Biol. Chem. 264: 9266. 128. Fryklund, J. (1988) Specific labelling of gastric H§ K§ by omeprazole. Biochem. Pharm. 37: 2543. 129. Frucht, H., Maton, P. N. and Jensen, R.T. (1991) Use of Omeprazole in Patients with Zollinger-Ellison Syndrome. Digestive Diseases and Sciences 36:394. 130. Fukui, H., Wang, N. P., Watanabe, T. and Wada, H. (1988) Solubilization, characterization and partial purification of [3H]mepyramine-binding protein, a possible histamine H1 receptor, from rat liver membrane. Jap. J. Pharmac. 46:127. 131. Funder, J. W. and Mercer, J. E. (1979) Cimetidine, a histamine H2 receptor antagonist, occupies androgen receptors. J. Clin. Endocrinol. Metab. 48: 189. 132. Furchgott, R. F. (1983) Role of endothelium in responses of vascular smooth muscle. Circ. Res. 53: 557. 133. Gajtkowski, G. A., Norris, D. B., Rising, T. J. and Wood, T. P. (1983). Specific binding of 3H-tiotidine to histamine H2 receptors in guinea pig cerebral cortex. Nature 304:65. 134. Galbraith, R. A. and Jellinck, P. H. (1989) Differential effects of cimetidine, ranitidine and famotidine on the hepatic metabolism of estrogen and testosterone in male rats. Biochem. Pharmac. 38: 2046. 135. Ganellin, C. R. (1985) A survey of recently described H2-receptor histamine antagonists. In:

Pharmacological Control of Gastric Acid Secretion

355

Frontiers in Histamine Research. A tribute to Heinz Schild. Advances in the Biosciences, 51, Ed Ganellin, C. R. and Schwartz, J. C. Pergamon Press Ltd, England, pp 47. 136. Gannon, B., Browning, J., O'Brien, P. and Rogers, P. (1984) Mucosal microvascular architecture of the fundus and body of human stomach. Gastroenterol. 86:866. 137. Gannon, B., Browning, J. and O'Brien, P. (1982) The microvascular architecture of the glandular mucosa of rat stomach. J. Anat. 135:667. 138. Gantz, I., Munzert, G., Tashiro, T., Sch~iffer, M., Wang, L., Del Valle, J. and Yamada, T. (1991) Molecular cloning of the human histamine H2 receptor. Biochem. Biophys. Res. Commun. 178:1386. 139. Gantz, I., Schgffer, M., Del Valle, J., Logsdon, C., Campbell, V., Uhter, M. and Yamada, T. (1991) Molecular cloning of a gene encoding the histamine H2 receptor. Proc. Natl. Acad. Sci. USA 88:429. 140. Gao, B. and Gilman, A. G. (1991) Cloning and expression of a widely distributed (type IV) adenylyl cvclase. Proc. Natl. Acad. Sci. USA 88:10178. 141. Garbers, D. L. (1989) Guanylate cyclase, a cell surface receptor. J. Biol. Chem. 264:9103. 142. Gespach, C., Abita, J. P. (i982) Human polymorphonuclear neutrophils. Pharmacological characterization of histamine receptors mediating the elevation of cyclic AMP. Molec. Pharmac. 21:78. 143. Gespach, C., Bataille, D., Dutrillaux, M. C. and Rosselin, G. (1982) The interaction of glucagon, gastric inhibitory peptide and somatostatin with cyclic AMP production systems present in rat gastric glands. Biochim. Biophys. Acta "/20:7. 144. Gespach, C., Bawab, W., Chastre, E., Emami, S., Yanaihara~ N. and Rossetin, G. (1988) Pharmacology and molecular identification of vasoactive intestinal peptide (VIP) receptors in normal and cancerous gastric mucosa in man. Biochem. Biophys. Res. Comm. 151:939. 145. Gespach, C., Bouhours D., Bouhours, J. F. and Rosselin, G. (1982) Histamine interaction on surface recognition sites of H2-type in parietal and non-parietal cells isolated from the guinea pig stomach. FEBS Lett., 149, 85. 146. Gespach, C., Chedeville, A., Hurbain-Kosmath, I., Housset, B., Derenne, J. P. and Abita, J. P. (1989) Adenylate cyclase activation by VIP-helodermin and histamine H2 receptors in human THP-1 monocytes/macrophages: A possible role in the regulation of superoxide anions production. Regulatory Peptides 26:158 (abstr.). 147. Gespach, C., Cherel, Y. and Rosselin, G. (1984) Development of sensitivity to cAMP-inducing hormones in the rat stomach. Am. J. Physiol. 247:G231. 148. Gespach, C., Dupont, C., Bataille, D. and Rosselin, G. (1980) Selective inhibition by somatostatin of cyclic AMP production in rat gastric glands~ Demonstration of a direct effect on the parietal cell function. FEBS Lett. 114:247. 149. Gespach, C., Emami, S., Boige, N. and Rosselin, G. (1983) Chemical differentiation of histamine receptors coupled to cyclic AMP generation in human and guinea pig fundic glands, tn Gut Peptides and Ulcer, Miyoshi, A., ed., Biomed. Res. Foundation, Tokyo, 55. 150. Gespach, C., Emami, S., Chastre, E., Launay, J. M. and Rosselin, G. (1987) Up- and down-regulation of membrane receptors as possible mechanisms related to the antiutcer actions of milk in rat gastric mucosa. Biosci. Rep. 7:135-t42. 151. Gespach, C., Emami, S. and Chastre, E. (1988) Membrane receptors in the gastroin~estina! tract. Biosci. Rep. 8:199. 152. Gespach, C., Emami, S. and Rosselin, G. (!984) Gastric inhibitory peptide (GIP), pancreatic glucagon and vasoactive intestinal peptide (VIP) are cAMP-inducing hormones in the human gastric cancer cell line HGT-1. Homologous desensitization of VIP receptor activity. Biochem. Biophys. Res. Cornmun. 120:641. 153. Gespach, C., Fagot, D. and Emami, S. (1988) Pharmacological control of the human gastric histamine H2 receptor by famotidine: comparison with H1, H2 and H3 receptor agonists and antagonists. Eur. J. Clin. Invest. 19:1. 154. Gespach, C., Hansen, A. and Hoist, J. (t989) Differential regu!ation of membrane receptors sensitive to histamine (H2-type), isoproterenol (/32-type) and glucagon-like peptides by the somatostatin analogue Sandostatin in rat gastric glands. Agents and Actions 27:169 155. Gespach, C., Hui, Bon Hoa, D. and Rosselin, G. (1983) Regulation by vasoac~ive intestinal peptide, histamine, somatostatin-14 and-28 of cyclic adenosine monophosphate levels in gastric glands isolated from the guinea pig fundus or antrum. Endocrinology 112:1597. 156. Gespach, C., Launay, J. M., Emami, S., Bondoux, D. and Dreux, C. (1986) Biochemical and pharmacological characterization of histamine-mediated up-regulation of human platelet serotonin uptake. Evidence for a subclass of l~stamine H2 receptors (H2h) highly sensitive to H2 receptor antagonists. Agents and Actions 18:115.

356

Mirossay, Di Gioia, Chastre, Emami and Gespach

157. Gespach, C., Marrec, N. and Abita, J. P. (t986) Pharmacological characterization of the histamine uptake system in HL-60 human promyelocytic leukemia cells. Biosci. Rep. 6:403. 158. Gespach, C., Menez, I. and Emami, S. (1983) Potent and specific blockade by AH 22216 of histamine H2-receptor-mediated acid secretion in isolated rabbit gastric cells. Biosci. Rep. 3:871. 159. Gespach, C., Saal, F., Cost, H. and Abita, J. P. (1982) Identification and characterization of surface receptors for histamine in the human promyelocytic leukemia cell line HL-60. Molec. Pharmac. 22:547. 160. Gespach, C. (1986) R6cepteurs H2 histaminergiques et peptidergiques de l'6ptih61ium gastrointestinal. Pharmacologie, fonction et d6veloppement. Th~se d'Etat, Universit6 de Paris VII, volume I, 1. 161. Gitlin, N., McCullough, A. J., Smith, .I.L., Mantell, G., Berman, R. and other members of the multicenter famotidine cooperative study group (1987) A mulilticenter, double-blind, randomized, placebo-controlled comparison of nocturnal and twice-a-day famotidine in the treatment of active duodenal ulcer disease. Gastroenterol. 92: 48. 162. Goy, M. F. (1991) cGMP: The wayward child of the cyclic nucleotide family. TINS 14:293. 163. Grant, S. M., Langtry, H. D. and Brogden, R. N. (1989) Ranitidine. An updated review of its pharmacodynamic and pharmacokinetic properties and therapeutic use in peptic ulcer disease and other allied diseases. Drugs 37: 801. 164. Green, J. P. and Maayani, S. (1977) Tricyclic antidepressant drugs block histamine H2 receptor in brain. Nature 269-"163. 165. Gunion, M. W., Kauffman, G. L. and Tach6, Y. (1990) Intrahypothatamic corticotropinreleasing factor elevates gastric bicarbonate and inhibits stress ulcers in rats. Am. J. PhysioL 258: G152. 166. Gutkind, J. S., Novotny, E. A., Brann, M. R. and Robbins, K. C. (1991) Muscarinic acetylcholine receptor subtypes as agonist-dependent oncogenes. Proc. Natl. Acad. Sci. USA 88: 4703. 167. Halenda, S. P., Banga, H. S., Zavoico, G. B., Lau, L.-F. and Feinstein, M. B. (1989) Synergistic release of arachidonic acid from platelets by activators of protein kinase C and Ca2 + ionophores. Evidence for the role of protein phosphorylation in the activation of phospholipase A2 and independence from the Na§ + exchanger. Biochemistry 28:7356. 168. Hanks, S. K., Quinn, A. M. and Hunter, T. (1988) The protein kinase family: conserved features and deduced phylogeny of the catalytic domains. Science 241:42. 169. Hanneman, H. and Sewing, K. F. (1986) Ranitidine and famotidine differ in their interaction with the parietal cell histamine H2-receptor. Br. J. Pharmacol 88: 258P. 170. Hansen, A. B., Gespach, C., Rosselin, G. E. and Hoist, J. J. Effect of truncated glucagon-like peptide 1 on cAMP in rat gastric glands and HGT-1 human gastric cancer cells. FEBS Lett. 236:119. 171. Hansson, H.-A., Hong, L. and Helander, H. F. Changes in gastric EGF, EGF receptors and acidity during healing of gastric ulcer in the rat. Acta Physiol. Scand. 190 138:241. 172. Harada, M. and Maeno, H. (1982) Effect of guanine nucleotides on histamine-sensitive adenylate cyclase from the parietal cells of guinea pig. Arch. Biochem. Biophys. 213"405. 173. Harada, M., Terai, M. and Maeno, H. (1983) Effect of a new potent H2-receptor antagonist 3[[[2-[(diamin•methy•ene)amin••]-4-thiaz••y•]methy•]thi•]-N2-su•fam••ypr•pi•namidine (YM11170) on gastric mucosal histamine-sensitive adenylate cyclase from guinea pig. Biochem. Pharmacol. 32:1635. 174. Hartig, P. R. (1989) Molecular biology of 5-HT receptors. TIPS 10:64. 175. Harty, R. F., Van der Vijver, J. C_ and McGuigan, J. E. (1977) Stimulation of gastrin secretion and synthesis in antral organ culture. J. Clin. Invest. 60:51. 176. Hatt, J. F. and Hanson, P. J. (1988) Inhibition of gastric acid secretion by epidermal growth factor. Effects on cyclic AMP and on prostaglandin production in rat isolated parietal cells. Biochem J. 255: 789. 177. Hatt, J. F. and Hanson, P. J. (1989) Sites of action of protein kinase C on secretory activity in rat parietal cells. Am. J. Physiol. 256:G129. 178. Hausdorff, W. P. and Sekura, R. D., Aguilera, G. and Catt, K. J. (1987) Control of aldosterone production by angiotensin II is mediated by two guanine nucleotide regulatory proteins. Endocrinology 120:1668. 179. Hayano, T., Sogawa, K., Ichihara, Y., Fujii-Kuriyama, Y. and Takahashi, K. (1988) Primary Structure of Human Pepsinogen C Gene. J. Biol. Chem. 263: 1382. 180. Hediger, M. A., Coady, A. H., Ikeda, T. S. and Wright, E. M. (1987) Expression cloning and cDNA sequencing of the Na§ co-transporter. Nature 330:379.

Pharmacological Control of Gastric Acid Secretion

357

181. Heim, H. K., Oestmann, A. and Sewing, Fr. K. (1991) Effects of histamine and activators of the cyclic AMP system on protein synthesis in and release of high molecular weight glycoproteins from isolated gastric non-parietal cells. Br. J. Pharmacol. 104:526. 182. Hejblum, G., Gaili, P., Boissard, C., Ast6sano, A., Marie, J. C., Anteunis, A., Hui Bon Hoa, D. and Rosselin, G. (1988) Combined ultrastructurai and biochemical study of cellular processing of vasoactive intestinal peptide and its receptors in human colonic carcinoma cells in culture. Canc. Res. 48:6201. 183. Hernandez, D. E. (1989) Neurobiology of brain-gut interactions. Implications for ulcer disease. Dig. Dis. Sci. 34:1809. 184. Hersey, S. J., Perez, A., Matheravidathu, S. and Sachs, G. (1989) Gastric H+-K+-ATPase in situ. Evidence for compartmentalization. Am. J. Physiol. 257:G539. 185o Hervatin, F., Arrang, J. M., Dubrasquet, M. and Lewin, M. J. M. (1989) Le r6cepteur histaminergique H3 inhibe la s6cr6tion acide gastrique stimul6e par la pentagastrine. Gastroenterol. Clin. Biol. 13:A8. 186. Hervatin, F., Moreau, E., Ducroc, R., Garzon, B., Avril, P., Millet, P. and Geloso, J. P. (1987) Ontogeny of rat gastric H+-K+,ATPase activity. Am. J. Physiol. 252: G28. 187. Hildebrandt, J. D., Sekura, R. D., Codina, J., Iyengar, R., Manclark, C. R. and Birnbaumer, L. (1983) Stimulation and inhibition of adenylyl cyclase mediated by distinct regulatory proteins. Nature 302: 706. 188. Hiraishi, H., Terano, A., Ota, S. I., Ivey, K. J. and Sugimoto, T~ (1986) Effect of cimetidine on indomethacin-induced damage in cultured rat gastric mucosal cells. Comparison with prostaglandins. J. Lab. Clin. Med. 108:608. 189. Hirschowitz, B. I. and Molina, E. (1983) Effect of cimetidine on histamine-stimulated gastric acid and electrolytes in dogs. Am. J. Physiol. 244:G416. 190. Hocevar, B. A. and Fields, A. P. (1991) Selective translocation of fill-protein kinase C to the nucleus of human promyelocytic (HL60) leukemia cells. J. Biol. Chem. 266:28. 191. Holm-Rutili, L. and Obrink, K. J. (1985) Rat gastric microcirculation in vivo. Am. J. Physiol. 248: G741. 192. Holt, S. and Howden, C. W. (1991) Omeprazole. Overview and Opinion. Digestive Diseases and Sciences 36: 385. 193. Holzer, P., Livingston, E. H., Saria, A. and Guth, P. H. (1991) Sensory neurons mediate protective vasodilatation in rat gastric mucosa. Am. J. Physiol. 260: G363. 194. Holzer, P., Pabst, M. A., Lippe, I. T., Peskar, B. M., Peskar, B. A., Livingston, E. H. and Guth, P. H. (1990) Afferent nerve-mediated protection against deep mucosal damage in the rat stomach. Gastroenterol. 98: 838. 195. Honda, Z.-I., Nakamura, M., Miki, I., Minami, M., Watanabe, T., Seyama, Y., Okado, H., Toh, H., Ito, K., Miyamato, T. and Shimizu, T. (1991) Cloning by functional expression of platelet-activating factor receptor from guinea-pig lung. Nature 349:342. 196. Horisberger, J. D., Jaunin, P., Reuben, M. A., Lasater, L. S., Chow, D. C., Forter, J~ G., Sachs, G., Rossier, B. C. and Geering, K. (1991) The H, K-ATPase fl-Subunit Can Act as a Surrogate for the/3-Subunit of Na, K-pumps. J. Biol. Chem. 266:19131. 197. Howard, J. M., Chremos, A. N., Collen, M. J., McArthur, K. E., Cherner, J. A., Maton, P. N., Ciarleglio, C. A., Cornelius, M. J., Gardner, J. D. and Jensen, R. T. (1985) Famotidine, a new, potent, long-acting histamine H2-receptor antagonist: comparison with cimetidine and ranitidine in the treatment of Zollinger-Ellison syndrome. Gastroenterol. 88:1026. 198. Hsueh, W., Gonzalez-Crussi, F. and Arroyave, J. L. (1988) Sequential Release of Leukotrienes and Norepinephrine in Rat Bowel After Ptatelet-Activating Factor. A mechanism study of platelet-activating factor-induced bowel necrosis. Gastroenterol. 94: 1412. 199. Ichinose, M., Stretton, C. D., Schwatrz, J. C. and Barnes, P. L. (1989) Histamine H3-receptors inhibit cholinergic neurotransmission in guinea pig airways. Br. J. Pharmac. 97:13. 200. Ira, W. B., Blakeman, D. P., Mendlein, J. and Sachs, G. (1984) Inhibition of (H + +K+)ATPase and H + accumulation in hog gastric membranes by trifluoperazine, verapamil and 8-(N,N-Diethylamino)octyl-3,4,5-Trimethoxybenzoate. Bioch. Biophys. Acta. 770:65. 20i. Inokuchi, H., Kawai, K., Takeuchi, Y. and Sano, Y. (1984) Immunobistochemical study on the morphology of enterochromaffin cells in the human fundic mucosa. Cell Tiss. Res. 235:703. 202. Inui, T., Chiba, T., Okimura, Y., Morishita, T., Nakamura, A., Yamaguchi, A., Yamatani, T., Kadowaki, S., Chihara, K. and Fujita, T. (1989) Presence and release of calcitonin gene-related peptide in rat stomach. Life Sciences 45: 1199. 203. Ishihara, T., Ishihara, Y., Hayano, T., Katsura, I., Sogawa, K., Fujii-Kuriyama, Y. and Takahahsi, K. (1989) Primary Structure and Transcriptional Regulation of Rat Pepsinogen C Gene. J. Biol. Chem. 254:10193.

358

Mirossay, Di Gioia, Chastre, Emami and Gespach

204. Ishikawa, S. and Sperelakis, N. (1987) A novel class (H3) of histamine receptors on perivascular nerve terminals. Nature 327: 158. 205. Ito, S. (1987) Functional gastric morphology. In Physiology of the Gastrointestinal Tract, 2rid Edition, Johnson, L. R., ed., Raven Press, New York 817. 206. Jackson, T. R., Blair, L. A. C., Marshall, J., Goedert, M. and Hanley, M. R. (1988) The mas oncogene encodes an angiotension receptor. Nature 335: 437. 207. Jahnsen, T., Hedin, L., Kidd, V. J., Beattie, W. G., Lohmann, S. M., Walter, U., Durica, J., Schulz, T. Z., Schlitz, E., Browner, M., Lawrence, C. B., Goldman, D., Ratoosh, S. L. and Richards, J. (1986) Molecular cloning, cDNA structure, and regulation of the regulatory subunit of type II cAMP-dependent protein kinase from rat ovarian granulosa cells. J. Biol. Chem. 261:12352. 208. Jefferys, D. B. and Vale, J. A. (1978) Cimetidine and bradycardia. The Lancet i:828. 209. Jennewein, H. M. (1977) The effect of clonidine on gastric acid secretion in rats and dogs. Naunyn-Schmiedeberg's Arch. Pharmacol. 297: 85. 210. Jensen, R. T., Collen, M. J., Pandol, S. J., Allende, H. D., Raufman, J. P., Bissonnette, B. M., Duncan, W. C., Durgin, P. L., Gillin, J. C. and Gardner, J. D. (1983) Cimetidine-induced impotence and breast changes in patients with gastric hypersecretory states. New Engl. J. Med. 308: 883211. Johnson, W. S. and Miller, D. R. (1988) Ranitidine and bradycardia. Annals Internal Med. 108: 493. 212. Jones, D. B., Howden, C. W., Burget, D. W., Silletti, C. and Hunt, R. H. (1988) Alteration of H2 receptor sensitivity in duodenal ulcer patients after maintenance treatment with an H2 receptor antagonist. Gut 29: 890. 213. Jones, C. M., Toh, B. H., Pettitt, J. M., Martinelli, T. M., Humphris, D. C., Callaghan, J. M., Goldkorn, I., Mu, F. T. and Gleeson, P. A. (1991) Monoclonal antibodies specific for the core protein of the/3-subunit of the gastric proton pump (H+/K § ATPase). An autoantigen targetted in pernicious anaemia. Eur. J. Biochem. 197:49-59. 214. Kageyama, T., Tanabe, K. and Koiwai, O. (1990) Structure and Development of Rabbit Pepsinogens. stage-specific zymogens, nucleotide sequences of cDNAs, molecular evolution, and gene expression during development. J. Biol. Chem. 265:17031. 215. Kalinyak, K. A., Sawutz, D. G., Lampkin, B. C., Johnson, C. L. and Whitsett, J. A. (1985) Effects of dimaprit on growth and differentiation of human promyelocytic cell line, HL-60. Life Sciences 36:1909. 216. Kaplan, D. R., Morrison, D. K., Wong, G., McCormick, F. and Williams, L. T. (1990) PDGF /3-receptor stimulates tyrosine phosphorylation of GAP and association of GAP with a signaling complex. Cell. 61:125. 217. Karnick, P. S., Monahan, S. J. and Wolfe, M. M. (1989) Inhibition of gastrin gene expression by somatostatin. J. Clin. Invest. 83:367. 218. Keeling, D. J., Fallowfield, C., Milliner, K. J., Tingley, S. K., Ire, R. J. and Underwood, A. H. (1985) Studies on the mechanism of action of omeprazole. Bioechem. Pharmacol. 34:2967. 219. Kikkawa, U., Kishimoto, A. and Nishizuka, Y. (1989) Annu. Rev. Biochem. 58:31-44. 220. Kirch, W., Halabi, A., Linde, M., Santos, S. R. and Ohnhaus, E. E. (1989) Negative effects of famotidine on cardiac performance assessed by noninvasive hemodynamic measurements. Gastroenterol. 96:1388. 221. Kirkegaard, P., Moddy, A. J., Hoist, J. J., Loud, F. B., Olsen, P. S. and Christiansen, J. (1982) Glicentin inhibits gastric acid secretion in the rat. Nature 297:156. 222. Kobilka, B. K., Kobilka, T. S., Daniel, K., Regan, J. W., Caron, M. G. and Lefkowitz, R. J. (1988) Chimeric oc2-, fl2-adrenergic receptors: delineation of domains involved in effector coupling and ligand binding specificity. Science 240:1310. 223. Konturek, S. J., Bielanski, W., Konturek, J. W., Oleksy, J. and Yamazaki, J. (1989) Release and action of epidermal growth factor on gastric secretion in humans. Scand. J. Gastroenterol. 24: 485. 224. Konturek, S. J., Bilski, J., Dembinski, A., Warzecha, A., Beck, G. and Jendralla, H. (1987) Effects of leukotrienes on gastric acid and alkaline secretions. Gastroenterol. 92:1209. 225. Korner, M., Bouthenet, M. L., Ganellin, C. R., Garbarg, M., Gros, C., Ire, R. J., Sales, N. and Schwartz, J. C. (1986) [125]iodobolpyramine, a highly sensitive probe for histamine HI-receptors in guinea pig brain. Eur. J. Pharmac. 120: 151. 226. Krupinski, J., Coussen, F., Bakalyar, H. A., Tang, W. J., Feinstein, P. G., Orth, K., Slaughter, C., Reed, R. R. and Gilman, A. G. (1989) Adenylyl cyclase amino acid sequence: possible channel- or transporter-like structure. Science 244" 1558. 227. Krupinski, J. (1991) The adenylyl cyclase family. Mol. Cell. Biochem. 104"73-79.

Pharmacological Control of Gastric Acid Secretion

359

228. Kubota, H., Taguchi, Y., Tohyama, M., Matsuura, N., Shiosaka, S., Ishihara, T., Watanabe, T., Shiotani, Y, and Wada, H. (1984) Electron microscopic identification of histidine decarboxylase-containing endocrine cells of the rat gastric mucosa. Gastroenterol. 87:496. 229. Kuo, J., Andersson, R. G. G., Wise, B. C., Mackerlova, L., Solomousson, I., Brackett, N. L., Kotoh, N., Shoji, M. and Wrenn, R. W. (1980) Calcium-dependent protein kinase: widespread occurrence in various tissues and phyla of the animal kingdom and comparison of effects of phospholipid, calmodulin and trifluoperazine. Proc. Natl. Acad. Sci. USA 77:7039. 230. Laboisse, C. L., Augeron, C., Couturier-Turpin, M. H., Gespach, C., Cheret, A. M. and Potet, F. (1982) Characterization of a newly established human gastric cancer cell line HGT-1 bearing histamine H2-receptors. Cancer Res. 42:1541. 231. Lamprecht, S. A., Schwartz, B., Krugliak, P., Odes, H. S., Guberman, R., Krawiec, J. (1986) Role of cyclic GMP in gastrin secretion from rat antral mucosae in organ culture. J. Cyclic Nucl. & Prot. Phosphor. Res. 11:177. 232. Lane, G. P. and Speed, B. (1989) Agranulocytosis after ranitidine administration. Med. J. Aust. 150: 595. 233. Langman, M. J. S. (1988) Effets des antagonistes histaminiques H2 sur le mdtabolisme des mddicaments et sur le syst~me endocrine. In: Contrdle de la sdcre6tion acide, Mignon, M., Galmiche, J. P., Ruszniewski, P., Pappo, M., John Libbey Eurotext, Paris, p. 141. 234. Larsson, H., Carlsson, E., Hakanson, R., Mattsson, H., Nilsson, G., Seensalu, R., Wallmark, B. and Sundler, F. (1988) Time-course of development and reversal of gastric endocrine cell hyperplasia after inhibition of acid secretion. Studies with omeprazole and ranitidine in intact and antrectomized rats. GastroenteroL 95:1477. 235. Larsson, H., Carlsson, E., Junggren, U., Otbe, L., Sj6strand, S. E., Skanberg, I. and SuudeU, G. (1983) Inhibition of Gastric Acid Secretion by Omeprazole in the Dog and Rat. Gastroenterol. 85: 900. 236. Lechleiter, J., Girard, S., Glapharn, D. and Peralta, E. (1991) Subcellular patterns of calcium release determined by G protein-specific residues of muscarinic receptors. Nature 350:505. 237. Lee, D. C., Carmichael, D. F., Krebs, E. G. and McKnight GS (1983) Isolation of a cDNA clone for the type I regulatory subunit of bovine cAMP-dependent protein kinase. Proc. Natl. Acad. Sci. USA 80:3608. 238. Lee, E. R. and Leblond, C. P. (1985) Dynamic histology of the antral epithelium in the mouse stomach. II Ultrastructure and renewal of isthmal cells. Am. J. Anat. 172:205. 239. Lee, Er and Leblond, C. P. (1985) Dynamic histology of the antral epithelium in the mouse stomach. IV Ultrastructural and renewal of gland cells. Am. J. Anat., 172, 241. 240. Lee, E. R. (1985) Dynamic histology of the antral epithelium in the mouse stomach. I Architecture of antral units. Am. J. Anat. 1"/2:187. 24l. Lee, E. R. (1985) Dynamic histology of the antral epithelium in the mouse stomach, iII Ultrastructure and renewal of pit cells. Am. J. Anat. 172: 225. 242. Lehy, T., Mignon, M., Cadiot, G., Elouar-Blanc, L., Ruszniewski, P., Lewin, M. J. M. and Bonfils, S. (1989) Gastric endocrine cell behavior in Zollinger-Ellison patients upon long-ter m potent antisecretory treatment. Gastroenterol. 96: 1029. 243. Lenz, H. J., Forquignon, I., Driige, G. and Greten, H. (1989) Effects of neuropeptides on gastric acid and duodenal bicarbonate secretions in freely moving rats. Regul. Peptides 24:293. 244. Leonard, A., Cuq, P., Magous, R. and Bali, J.-P. (1991) M3-subtype muscarinic receptor that controls intracellular calcium re~ease and inositot phosphate accumulation in gastric parietal cells. Biochem. Pharmacol. 42: 839. 245. Leth, R., Elander, B., Haglund, U., Olbe, L. and Fellenius, E. (1987) Histamine H2-receptor of human and rabbit parietal cells. Am. J. Physiol. 253:G497. 246. Leung, F. W. (1991) Role of blood flow and alcaline secretion in acid-induced deep duodenal villous injury in rats. Am. J. Physiol. 26:G399. 247. Leurs, R., Bast, A. and Timmerman, H. (1989) High affinity, saturable [3H]-mepyramine binding site on rat liver plasma membrane do not represent histamine HI-receptors. A warning. Biochem. Pharmac. 38:2175. 248. Levine, M. A., Smallwood, P. M., Moen, P. T., Helman, L. J. and Ahn, T. G. (1990) Molecular cloning of/~3 subunit, a third form of the G protein /3-subunit polypeptide. Proc. Natl. Acad. Sci. USA 87:2329. 249. Lewin, M. J. M. (1984) Omdprazole: un nouveau mdcanisme d'action et une efficacit6 sans prdcddent dans le traitment de l'hypersdcredtion acide gastrique. GastroenteroL Clin. Biol. 8:939. 250. Lewin, M. J. M. (1985) La sdcrdtion acide de l'estomac: un mdcanisme biologique peu connu. Mgdecine /Sciences. 85:234.

360

Mirossay, Di Gioia, Chastre, Emami and Gespach

251. Lewis, J. H. (1987) Hepatic effects of drugs used in the treatment of peptic ulcer disease. A m J. Gastroenterol. 82: 987. 252. Liang, T. (1984) Absence of androgen receptor affinity for MK-208, a new H2-receptor antagonist. Clin. Res. 32:283A. 253. Lips, D. L., Majerus, P. W., Gorga, F. R., Young, A. T. and Benjamin, T. L. (1989) Phosphatidylinositol 3-phosphate is present in normal and transformed fibroblasts and is resistant to hydrolysis by bovine brain phospholipase C II. J. Biol. Chem. 264: 8759. 254. Lobie, P. E., Breipohl, W. and Waters, M. J. (1990) Growth Hormone Receptor Expression in the Rat Gastrointestinal Tract. Endocrinol. 126:299. 255. L6nroth, H., Hakanson, R., Lundell, L. and Sundler, F. (1990) Histamine containing endocrine cells in the human stomach. Gut 31:383. 256. L6nroth, H., Lundell, L. and Rosengren, E. (1989) Histamine metabolism of the human gastric mucosa--a study on the regional distribution of the amine and enzyme activities. Scand. J. Clin. Lab. Invest. 49:23. 257. L6nroth, H., Rosengren, E., Olbe, L. and Lundell, L. (1990) Histamine metabolism in human gastric mucosa. Effect of pentagastrin stimulation. Gastroenterology 98: 921. 258. Lorentzon, P., Eklundh, B., Br~indstr6m, A. and Wallmark, B. (1985) The mechanism for inhibition of gastric (H + + K+)-ATPase by omeprazole. Biochem. Biophys. Acta. 817:25. 259. Lundell, L., Forssel, H., L6nroth, H., Rosengren, E. and Wingren, U. (1986) Histamine storage and formation in canine gastric mucosa. Effect of pentagastrin stimulation. Acta Physiol. Scand. 128:587. 260. Ma, J. Y., Song, Y. H., Sj6strand, S. E., Rask, L. and Mardh, S. (1991) cDNA cloning of fl-subunit of the human gastric H, K-ATPase. Biochem. Biophys. Research Comm. 180: 39. 261. Machen, T. E., Rutten, M. J. and Ekblad, E. B. M. (1982) Histamine, cAMP and activation of piglet gastric mucosa. Am. J. Physiol. 242:G79. 262. Maeda, M., Oshiman, K. I., Tamura, S. and Futai, M. (1990) Human Gastric ( H § ATPase Gene. Similarity to (Na+ + K+)-ATPase genes in exon/intron organization but difference in control region. J. Biol. Chem. 265:9027. 263. Maeda, M., Oshiman, K. I., Tamura, S., Kaya, S., Mahmood, S., Reuben, M. A., Lasater, L. S., Sachs, G. and Futai, M. (1991) The Rat H+/K+-ATPase/3 Subunit Gene and Recognition of Its Control Region by Gastric DNA Binding Protein. J. Biol. Chem. 266:21584. 264. Magous, R., Bali, J. P., Rossi, J. C. and Girard, J. P. (1983) Leukotrienes stimulate acid secretion from isolated gastric parietal cells. Biochem. Biophys. Res. Commun. 144:897. 265. Magous, R., Baudiere, B. and Bali, J. P. (1985) Muscarinic receptors in isolated gastric fundic mucosal cells. Biochem. Pharmac. 34: 2269. 266. Magous, R., Galleyrand, J. C. and Bali, J. P. (1989) Common or distinct receptors for gastrin and cholecystokinin in gastric mucosa? Biochim. Biophys. Acta 1010:357-362. 267. Mahon, W. A. and Kolton, M. (1978) Hypotension after intravenous cimetidine. Lancet i:828. 268. Major, J. S. and Scholes, P. (1978) The localization of a histamine H2-receptor adenylate cyclase system in canine parietal cells and its inhibition by prostaglandins. Agents and Actions 8: 324. 269. Malbon, C. C., Mangano, T. J. and Watkins, D. C. (1985) Heart contains two substrates (M = 40.000 and 41.000) for pertussis toxin-catalyzed ADP-ribosylation that copurify with Ns. Biochem. Biophys. Res. Commun. 128:899. 270. Mangeat, P., Gespach, C., Marchis-Mouren, G. and Rosselin, G. (1982) Differential effects of histamine, vasoactive intestinal polypeptide, prostaglandin E2 and somatostatin on cyclic AMP-dependent protein kinase activation in gastric glands isolated from the guinea pig fundus and antrum. Regul. Peptides 3: 155. 271. Mangla, J. C. and Pereira, M. (1982) Tricyclic antidepressants in the treatment of peptic ulcer disease. Arch. Intern. Med. 142:273. 272. Mardh, S., Song, Y. H. and Wallmark, B. (1988) Effects of some antisecretory drugs on acid production, intracellular free Ca2+, and cyclic AMP production in isolated pig parietal cells. Scand. J. Gastroenter 23: 977. 273. Mardh, S., Ma, J. Y., Song, Y. H., Aly, A. and Henriksson, K. (1991) Occurrence of autoantibodies against intrinsic factor, H, K-ATPase, and pepsingoen in atrophic gastritis and rheumatoid arthritis. Scand. J. Gastroenterol. 26:1089. 274. Marino, L., Muglia, B. H. and Yamada, T. (1990) H+-K§ and carbonic anhydrase II gene expression in the developing rat fundus. Am. J. Physiol. 259:G108. 275. Marks, I. N. (1989) Alteration of H2 receptor sensitivity in duodenal ulcer patients. Gut 30"701. 276. Marshall, I. (1984) Characterization and distribution of histamine H1- and H2-receptors in precapillary vessels. J. Cardiovasc. Pharmac. 6:$587. 277. Martinson, E. A., Goldstein, D. and Brown, J. H. (1989) Muscarinic receptor activation of phosphatidytchol_ine hydrolysis_ J. Biol. Chem. 264:14748.

Pharmacological Control of Gastric Acid Secretion

361

278. Martin, F., Martin, M. S. and Lorcerie, B. (1988) Les antis6cr6toires provoquent-ils des cancers gastriques? In: Contr61e de la s6cr6tion acide, Mignon, M., Galmiche, J. P., Ruszniewski, P. and Pappo, M., John Libbey Eurotext, Paris, 119. 279. Martin, T. F. J., Lueas, D. O., Bajjalieh, S. M. and Kowalchyk, J. A. (1986) Thyrotropinreleasing hormone activates a CaZ+-dependent polyphosphoinositide phosphodiesterase in permeable GH3 cells. J. Biol. Chem. 261:2918. 280. Maton, P. N., Lack, E. E., Collen, M. J., Cornelius, M. J., David, E., Gardner, J. D. and Jensen, R. T. (1990) The effect of Zollinger-Ellison syndrome and omeprazol therapy on gastric oxyntic endocrine cells. Gastroenterol. 99: 943. 281. Matsuda, M., Aono, M., Moriga, M. and Okuma, M. (199i) Centrally administered NPY stimulated gastric acid and pepsin secretion by a vagally mediated mechanism. Regulatory Peptides 35: 31. 282. Mattson, J. P., V~i~in~inen, K., Wallmark, B. and Lorentzon, P. (1991) Omeprazole and bafilomycin, two proton pump inhibitors. Differentiation of their effects on gastric, kidney and bone H+-translocating ATPases. Biochim. Biophys. Aeta. 1065:261. 283. Matuoka, K., Tanaka, M., Mitsui, Y. and Murota, S. I. (1983) Cultured rabbit gastric epithelial cells producing prostaglandin I2. GastroenteroL 84:498. 284. McArthur, K. E., Collen, M. J., Maton, P. N., Cherner, J. A., Howard, J. M., Ciarleglio, C. A., Cornelius, M. J., Jensen, R. T. and Gardner, J. D. (1985) Omeprazole. Effective, convenient therapy for Zollinger-Ellison syndrome. Gastroenterol. 88:939. 285. McNeil, J. J., Mihaly, G. W., Anderson, A., Marshall, A. W., Smallwood, R. A. and Louis, W. J. (1981) Pharmacokinetics of the H2-receptor antagonist ranitidine in man. Br. J. Clin. Pharmac. 12:411. 286. Meldolesi, J., Clementi, E., Fasolato, C., Zacchetti, D. and Pozzan, T. (1991) CAP.+ influx following receptor activation. TIPS 12: 289. 287. Menez, I., Gespach, C., Emami, S. and Rosselin, G. (1983) Irreversible and specific inactivation by AH 22216 of histamine H2 receptors in the human gastric cancer cell line HGT-1. Biochem. Biophys. Res. Commun. 116:251. 288. Mercier, F., Reggio, H., Devilliers, G., Bataille, D. and Mangeat, P. (1989) Membranecytoskeleton dynamics in rat parietal cells: mobilization of actin and spectrin upon stimulation of gastric acid secretion. J. Cell Biol. 108:441. 289. Michelangeli, F., Ruiz, M. C., Fernandez, E. and Ciarrocchi, A. (1989) Role of Ca 2+ in H + transport by rabbit gastric glands studied with A23187 and BAPTA, an incorporated Ca 2+ chelator. Biochim. Biophys. Acta 983:82. 290. Michelangeli, F., Ruiz, M. C., Rodriguez, C. L. and Pelacca, A. (1988) Somatostatin inhibition of secretagogue and forskolin-stimulated gastric acid secretion. Am. J. Physiol. 254:G531. 291. Miederer, S. E., Schepp, W., Dein, H. J. and Ruoff, H. J. (1986) Effect of glucagon on adenylate cyclase activity and acid production of isolated human parietal cells. Klin. Wochenschr 64: 746. 292. Minakuchi, R., Takai, Y., Yu, B. and Nishizuka, Y. (1981) Widespread occurrence of calcium-activated, phospholipid dependent protein kinase in mammalian tissues. J. Biochem. (Tokyo) 89:1651. 293. Minocherhomjee, A. M., Shattuck, R. L. and Storm, D. R. (1988) Calmodulin-stimulated adenylate cyclases. In: Cohen, P. and Klee, C. 13. (eds): Calmodulin. Elsevier Science Publishers, New York, 249. 294. Miwa, T. and Miyoshi, A. (1987) Famotidine in the treatment of gastritis. Scand. J. Gastroenter 22: 46. 295. Miyamoto, T., Itoh, M., Noguchi, Y. and Yokochi, K. (1992) Role of intracellular C a 2+ and the calmodulin messenger system in pepsinogen secretion from isolated rabbit gastric mucosa. Gut. 33:21. 296. Molloy, C. J., Bottaro, D. P., Fleming, T. P., Marshall, M. S., Gibbs, J. B. and Aaronson, S. A. (1989) PDGF induction of tyrosine phosphorylation of GTPase activating protein. Nature 342:711. 297. Mori, S., Morishita, Y., Sakai, K., Kurimoto, S., Okamoto, M., Kawamoto, T. and Kuroki, T. (1987) Electron microscopic evidence for epidermal growth factor receptor (EGF-R)-like immunoreactivity associated with the basolateral surface of gastric parietal cells. Acta Pathol. Jap. 37:1909. 298. Morri, M., Takata, H. and Takeguchi, N. (1989) Acid activation of omeprazole in isolated gastric vesicles, oxyntic cells, and gastric glands. GastroenteroL 96: 1453. 299. Morrison, D. K., Kaplan, D. R., Escobedo, J. A., Rapp, U. R., Roberts, T. M. and Williams, L. T. (1989) Direct activation of the serine/threonine kinase activity of Raf-1 through trysoine phosphorylation by the PDGF/3-receptor. Ceil. 58:649.

362

Mirossay, Di Gioia, Chastre, Emami and Gespach

300. Murakami, S., Muramatsu, M., Aihara, H. and Otomo, S. (1991) Inhibition of gastric H +, K+-ATPase by the antiulcer agent, sofalcone Biochemical Pharmacology. 42:1447. 301. Murayama, T., Kajiyama, Y. and Nomura, Y. (1990) Histamine-stimulated and GTP-binding proteins-mediated phospholipase A2 activation in rabbit platelets. J. Biol. Chem. 265:4290. 302. Murayama, T. and Ui, M. (1985) Receptor-mediated inhibition of adenylate cyclase and stimulation of arachidonic acid release in 3T3 fibroblasts. J. Biol. Chem. 260:7226. 303. Nagata, H. and Guth, P. H. (1983) Effect of histamine on microvascular permeability in the rat stomach. Am. J. Physiol. 245: G201. 304. Nagata, H. and Guth, P. H. (1984) Effect of topical histamine on mucosal microvascular permeability and acid secretion in the rat stomach. Am. J. Physiol. 246: G654. 305. Nagaya, H., Satoh, H. and Maki, Y. (1990) Possible mechanism for the inhibition of acid formation by the proton pump inhibitor AG-1749 in isolated canine parietal cells. J. Pharmacol. Exp Therap. 252:1289. 306. Nakaya, N. and Tasaka, K. (1987) The influence of histamine on precursors of granulocytic leukocytes in murine bone marrow. Life Sci. 42:999. 307. Nandi, J., King, R. L., Kaplan, D. S. and Levine, R. A. (1990) Mechanisms of gastric proton pump inhibition by calcium channel antagonists. J. Pharmacol. Exp. Therap. 252:1102. 308. Neer, E. J. and Clapham, D. E. (1988) Roles of G protein subunits in transmembrane signalling. Nature 333:129. 309. Negulescu, P. A. and Machen, T. E. (1988) Intracellular Ca regulation during secretagogue stimulation of the parietal cell. Am. J. Physiol. 254:C130. 310. Negulescu, P. A., Reenstra, W. W. and Machen, T. E. (1989) Intracellular Ca requirements for stimulus-secretion coupling in parietal cell. Am. J. Physiol. 256: C241. 311. Niedel, J. E., Kuhn, L. J. and Vandenbark, G. R. (1983) Phorbol diester receptor copurifies with protein kinase C. Proc. Natl. Acad. Sci. USA 80:36. 312. Nishikawa, H., Taniguchi, T., Ninomiya, H. and Fujiwara, M. (1988) Nicotinic acetylcholine receptors in the rat stomach: I. (-)-[3H]nicotine binding. Eur. J. Pharmac. 146: 97. 313. Nishizuka, Y. (1986) Studies and perspectives of protein kinase C. Science 233:305. 314. Nizankowska-Blaz, T., Abramowicz, T., Korczowski, R. and Rusin, J. (1988) Triiodothyronine and thyroxine in human, cow's and formula milk. Exp. Clin. Endocrinol. 91:116. 315. Nogami, M., Suko, M. and Miyamato, T. (1990) The effect of platelet-activating factor on [~4C] aminopyrine uptake by isolated guinea pig parietal cells. Biochem. and Biophys. Res. Comm. 168: 1047. 316. Okada, M., Ohmura, E., Kamiya, I., Murakami, H., Tsushima, T. and Shizume, K. (1991) Transforming growth factor (TGF)-cr in human milk. Life Sci. 48:1151. 317. Okada, Y. and Ueda, S. (1984) Electrical membrane responses to secretagogues in parietal cells of the rat gastric mucosa in culture. J. Physiol. 354:109. 318. Okumura, T., Uehara, A., Kitamori, S., Takasugi, Y. and Namiki, M. (1991) Central basic fibroblast growth factor inhibits gastric ulcer formation in rats. Biochem. Biophys. Res. Commun. 177: 809. 319. Okumura, T., Uehara, A., Tsuji, K., Taniguchi, Y., Kitamori, S., Shibata, Y., Okamura, K., Takasugi, Y. and Namiki, M. (1991) Central nervous system action of basic fibroblast growth factor: Inhibition of gastric acid and pepsin secretion. Biochem. Biophys. Res. Commun. 175: 527. 320. Olsen, P. S., Kirkegaard, P., Poulsen, S. S. and Nexo, E. (1986) Vasoactive intestinal polypeptide and acetylcholine stimulate exocrine secretion of epidermal growth factor from the rat submandibular gland. Regul. Peptides 15:37. 321. Orloff, S. L., Bunnett, N. W., Walsh, J. H. and Debas, H. T. (1992) Intestinal acid inhibits gastric acid secretion by neural and hormonal mechanism in rats. Am. J. Physiol. 262:G165. 322. Osband, M. and McCaffrey, R. (1979) Solubilization, separation and partial characterization of histamine HI and H2 receptors from calf thymocyte membranes. J. Biol. Chem. 254:9970. 323. Oshiman, K. I., Motojima, K., Mahmood, S., Shimada, A., Tamura, S., Maeda, M. and Fntai, M. (1991) Control region and gastric specific transcription of the rat H +, K+-ATPase o~subunit gene. FEBS Letters 281:250. 324. Ostrowski, J. and Bomsztyk, K. (1989) Interaction of signal transdnction pathways in mediating acid secretion by rat parietal cells. Am. J. Physiol. 256:C873. 325. Parker, J., Daniel, L. W. and Waite, M. (1987) Evidence of protein kinase C involvement in phorbol diester-stimulated arachidonic acid release and prostaglandin synthesis. J. Biol. Chem. 262: 5385. 326. Pearce, P. and Funder, J. W. (1980) Histamine H2 receptor antagonist: radioreceptor assay for antiandrogen side-effects. Clin. Exp. Pharmacol. Physiol. 7:442.

Pharmacological Control of Gastric Acid Secretion

363

327. Peden, N. R., Richards, D. A., Saunders, J. H. B. and Wormsley, K. G. (1979) Pharmacologically effective plasma concentrations of ranitidine. Lancet ii, 199. 328. Pendleton, R. G., Torchiana, M. L., Chung, C., Cook, P., Wiese, S. and Clineschmidt, B. V. (1983) Studies on MK-208 (YM-II170) a new, slowly dissociabIe H2-receptor antagonist. Arch. Int. Pharmac. Ther. 266 : 4.

329~ Penston, J. G. and Wormsley, K. G. (1989) Histamine H2-receptor antagonists versus prostaglandins in the treatment of peptic ulcer disease. Drugs 37:391. 330. Peskar, B. M., Respondek, M., MOiler, K. M. and Peskar, B. A. (1991) A role for nitric oxide in capsaicin-induced gastroprotection. Eur. J. Pharmacol. 198:133. 331. Pfeiffer, A., Rochlitz, H., Herz, A. and Paumgartner, G. (1988) Stimulation of acid secretion and phosphoinositol production by rat parietal cell muscarinic M2 receptors. Am. J. Physiol. 254: G622. 332. Postius, S., Ruoff, H. J. and Szelenyi, I. (1985) Prostaglandin formation by isolated gastric parietal and nonparietal cells of the rat. Br. J. Pharmac. 84:871. 333. Poynter, D., Pick, C. R., Harcourt, R. A., Selway, S. A. M., Ainge, G., Harman, I. W., Spurling, N. W., Fluck, P. A. and Cook, J. L. (1985) Association of long lasting unsurmountable histamine H2 blockade and gastric carcinoid turnouts in the rat. Gut 26: 1284. 334. Premont, R. T. and Lyengar, R~ (1990) Adenylyl cyclase and its regulation by Gs. In: lyengar, R. and Birnbuamer, L. (eds): G proteins. Academic Press, New York, p. I47. 335. Prost, A., Emami, S. and Gespach, C. (1984) Desensitization by histamine of H2 receptormediated adenylate cyclase activation in the human gastric cancer cell line HGT-1. FEBS Lett. 177: 227. 336. Rabon, E., Gunther, R. D., Soumarmon, A., Bassilian, S., Lewin, M. and Sachs, G. (1985) Solubilization and reconstitution of the gastric H, K-ATPase. J. Biol. Chem. 260:10200. 337. Ramano, M., Razandi, M. and Ivey, K..1. (1989) Protection of gastric epithelial cell monoJayers from a human cell line by omeprazole in vitro. Scand. J. GastroenteroL 24:5t3. 338. Rangachari, P. K. (1992) Histamine: mercurial messenger in the gut. Am. J. Physiol. 262:G1. 339. Ray, T. K., Nandy, J., Pidhorodeckyj, N. and Meng-Ai, Z. (1982) Polyamines are inhibitors of gastric acid secretion. Biochemistry 79: 1448. 340. Read, M. A., Chick, P., Hardy, K. J. and Shulkes, A. (1992) Ontogeny of gastrin, somatostatin, and the H+/K+-ATPase in the ovine fetus. Endocrinology 130: 1688. 341. Reenstra, W. W., Shortridge, B. and Forte, .1, G. (1985) Inhibition of the gastric (H + + K+) ATPase by fenoctimine. Biochem. Pharmacol. 34:2331. 342. Regardh, C. G., Gabrielsson, M., Hoffman, K. J., LOfberg, I. and Skanberg, I. (1985) Pharmacokinetics and metabolism of omeprazole in animals and man--an overview. Stand. J. Gastroenterol. 20: 79. 343. Rehfeldt, W,, Hass, R. and Goppelt-Struebe, M. (1991) Characterization of phosphotipase A2 in monocytic cell lines. Biochem. J. 276:631. 344. ReyI-Desmars, F., Le Roux, S., Linard, C., Benkouka, F. and Lewin, M. J. M. (1989) Purification jusqu'8 apparente homog6n6it6 du r6cepteur de la somatostatine de ta lign6e cellulaire humaine HGT-1 d'origine gastrique. C. R. Acad. Sci. Paris 308:251. 345. Reyl, F. J. and Lewin, M. J. M. (1981) Somatostatin is a potent activator of phosphoprotein phosphatases in the digestive tract. Biochim. Biophys Acta 675:297. 346. Reyl, F. J. and Lewin, M. J. M. (1982) Intracellular receptor for somatostatin in gastric mucosa~ cells: Decomposition and reconstitution of somatostatin-stimulated phosphoprotein phosphatases. Proc. Natl. Acad. Sci, USA 79:978. 347. Rhee, S. G., Suh, P.-G., Ryu, S. H. and Lee, S. Y. (1989) Studies of inositol phospholipidspecific phospholipase C. Science 244:546. 348. Rhee, S. G. (1991) Inositol phospholipid-specific phospholipase C: Interaction of the ~,1 isoform with tyrosine kinase. TIBS 16:297. 349. Richards, M. H. (1991) Pharmacology and second messenger interactions of cloned muscarinic receptors. Biochem. Pharmacol. 42:1645. 350. Ries, R. K., Gilbert, D. A. and Katon, W. (1984) Tricyclic antidepressant therapy for peptic ulcer disease. Arch. Intern. Med. 144:556. 35L Roberts, L. J., Bloomgarden, Z. T., Marney, S. R,, Rabin, D. and Oates, J. A. (1983) Histamine release from a gastric carcinoid: provocation by pentagastrin and inhibition by somatostatin. Gastroenterot. 84:272. 352. Roche, S., Bali, J.-P., Gatleyrand, J.-C. and Magous, R. (1991) Characterization of a gastrin type receptor on rabbit gastric parietal cells using L365,260 and L364,7t8. Am. J, Physiol. 2(~:G182. 353. Roche, S., Bali, J.-P. and Magous, R. (1990) Involvement of a pertussis toxin-sensitive G protein in the action of gastrin on gastric parietal cells. Biochim. Biophys. Acta 1055:287.

364

Mirossay, Di Gioia, Chastre, Emami and Gespach

354. Roche, S., Gusdinar, T., Bali, J.-P. and Magous, R. (1991) "Gastrin" and "CCK" receptors on histamine- and somatostatin-containing cells from rabbit fundic mucosa-I. Biochem. Pharmacol. 42:765. 355. Roche, S., Gusdinar, T., Bali, J.-P. and Magous, R. (1991) Biphasic kinetics of inositol 1,4,5-trisphosphate accumulation in gastrin-stimulated parietal cells. FEBS Letters 282:147. 356. Roche, S., Gusdinar, T., Bali, J.-P. and Magous, R. (1991) Relationship between inositol 1,4,5-trisphosphate mass level and (14C)aminopyrine uptake in gastrin-stimulated parietal cells. Mol. Cell. Endocrinol. 77:109. 357. Roche, S. and Magous, R. (1989) Gastrin and CCK-8 induce inositol 1,4,5-trisphosphate formation in rabbit gastric parietal cells. Biochim. Biophys. Acta 1014: 313. 358. Rodbell, M. (1980) The role of hormone receptors and GTP-regulatory proteins in membrane transduction. Nature 284:17. 359. Rosselin, G., Boige, N., Dupont, C. and Gespach, C. (1983) Receptors coupled to the cyclic AMP system in human gastric epithelium. In Gut Peptides and Ulcer, Miyoshi, A. ed., Biomedical Research Foundation, p 38. 360. Rosselin, G. (1989) Les r6cepteurs de surface des cellules 6pith61iales digestives. Ann. Endoerinol. 50: 447. 361. Ruat, M., K6rner, M., Garbarg, M., Gros, C., Schwartz, J. C., Tertiuk, W. and Ganellin, C. R. (1988) Characterization of histamine HI-receptor binding peptides in guinea pig brain using [125]iodoazidophenpyramine, an irreversible specific photoaffinity probe. Proc. Natl. Acad. Sci. USA 85:2743. 362. Ruat, M., Traiffort, E., Arrang, J.-M., Leurs, R. and Schwartz, J.-C. (1991) Cloning and tissue expression of a rat histamine H2-receptor gene. Biochem. Biophys. Res. Commun. 179:1470. 363. Rubin, W. A. (1973) A fine structural characterization of the proliferated endocrine cells in atrophic gastric mucosa. Am. J. Pathol. 70:109. 364. Ruiz, M. C. and Michelangeli, F. (1986) Stimulation of oxyntic and histaminergic cells in gastric mucosa by gastrin C-terminal tetrapeptide. Am. J. Physiol. 251: G529. 365. Ruoff, H. J., Reutter, K. and Schepp, W. (1985) Histamine content and release of isolated rat gastric mucosal cells. Agents and Actions 16: 202. 366. Rutten, M. J. and Machen, T. E. (1981) Histamine, cyclic AMP and activation events in piglet gastric mucosa in vitro. Gastroenterol. 80: 928. 367. Ryberg, B., Bishop, A. E., Bloom, S. R., Carlsson, E., Hakanson, R., Larsson, H., Mattsson, H., Polak, J. M. and Sundler F. (1989) Omeprazole and ranitidine, antisecretagogues with different modes of action, are equally effective in causing hyperplasia of enterochromaffin-like cells in rat stomach. Regul. peptides 25:235. 368. Ryu, S. H., Suh, P.-G., Cho, K. S., Lee, K.-Y. and Rhee, S. G. (1987) Bovine brain cytosol contains three immunologically distinct forms of inositolphospholipid-specific phospholipase C. Proc. Natl. Acad. Sci. USA 84:6649. 369. Saccomani, G., Psarras, C. G., Smith, P. R., Kirk, K. L. and Shoemaker, R. L. (1991) Histamine-induced chloride channels in apical membrane of isolated rabbit parietal cells. Am. J. Physiol. 260: C1000. 370. Saigenji, K., Fukutomi, H. and Nakazawa, S. (1987) Famotidine; postmarketing clinical experience. Scand. J. Gastroenter. 22:34. 371. Saltiel, A. R. and Cuatrecasas, P. (1988) In search of a sound messenger for insulin. Am. J. Physiol. 255:C1. 372. Sanders, M. J., Amirian, D. A., Ayalon, A. and Soll, A. H. (1983) Regulation of pepsinogen release from canine chief cells in primary monolayer culture. Am. J. Physiol. 245:G641. 373. Sandvik, A. K., Waldum, H. L. (1991) CCK-B (gastrin) receptor regulates gastric histamine release and acid secretion. Am. J. Physiol. 260:G925. 374. Seperas, E., Kauffman, G. and Tach6, Y. (1991) Role of central prostaglandin E2 in the regulation of gastric acid secretion in the rat. Eur. J. Pharm. 209: 1. 375. Savola, M., Savola, J.-M. and Puurunen, J. (1989) e:2-adrenoceptor-mediated inhibition of gastric acid secretion by medetomidine is efficiently antagonized by atipamezole in rats. Arch. int. Pharmacodyn. 301: 267. 376. Sax, M. J. (1987) Clincially important adverse effects and drug interactions with H2-receptor antagonists: An update. Pharmacotherapy 7:110S. 377. Saxena, S. P., Brandes, L. J., Becker, A. B., Simons, J. K., LaBella, F. S. and Gerrard, J. M. (1989) Histamine is an intracellular messenger mediating platelet aggregation. Science 243:1596. 378. Scarpignato, C., Tramacere, R. and Zappia, L. (1987) Antisecretory and antiulcer effect of the H2-receptor antagonist famotidine in the rat: comparison with ranitidine. Br. J. Pharmac. 92:153.

Pharmacological Control of Gastric Acid Secretion

365

379. Schenker, S., Dicke, J., Johnson, R. F., Mor, L. L. and Henderson, G. I. (1987) Human placental transport of cimetidine..I. Clin. Invest. 80:1428. 380. Schepp, W., Kath, D., Zimmerhackl, B., Schusdziarra, V. and Classen, M. (1989) Leukotrienes C4 and D4 potentiate acid production by isolated rat parietal cells. Gastroenterol. 97: 1420. 381. Schepp, W., Miederer, S. E. and Ruoff, H. J. (1985) Effects of hormones (calcitonin, GIP) and pharmacological antagonists (ranitidine and famotidine) on isolated rat parietal cells. Regulatory Peptides 12: 297. 382. Schepp, W. and Ruoff, H. J. (1984) Adenylate cyclase and H + production of isolated rat parietal cells in response to glucagon and histamine. Eur. J. Pharmac. 98:9. 383. Schjoldager, B. T. G., Mortensen, P. E., Christiansen, J., Orskov, C. and Hoist, J. J. (1989) GLP-I (glucagon-like peptide 1) and truncated GLP-1, fragments of human proglucagon, inhibit gastric acid secretion in humans. Dig. Dis. Sci. 34:703. 384. Schmidtler, J., Schepp, W., Janczewska, I., Weigert, N., Ftirlinger, C., Schusdziarra, V. and Classen, M. (1991) GLP-l-(7-36)amide, -(1-37), and -(1-36) amide: potent cAMP stimuli of rat parietal cell function. Am. J. Physiol. 260:G940. 385. Schnefel, S., Pr6frock, A., Hinsch, K.-D. and Schulz, I. (1990) Cholecystokinin activates Gil-, Gi2-, Gi3- and several Gs-proteins in rat pancreatic acinar cells. Biochem. J. 269:483. 386. Scholes, P., Cooper, A., Major, J. J., Walters, M. and Wilde, C. (1976) Characterization of an adenylate cyclase system sensitive to histamine H2-receptor excitation in cells from dog gastric mucosa. Agents and Actions 6: 677. 387. Schubert, M. L., Hightower, J. and Makhlouf, G. M. (1989) Linkage between somatostatin and acid secretion: evidence from use of pertussis toxin. Am. J. Physiol. 256:G418. 388. Schubert, M. L. (1991) The effect of vasoactive intestinal polypeptide on gastric acid secretion is predominantly mediated by somatostatin. Gastroenterol. 100:1195. 389. Schubert, M. L,, Jong, M. J. and Makhlouf, G. M. (1991) Bombesin/GRP-stimulated somatostatin secretion is mediated by gastrin in the antrum and intrinsic neurons in the funduns. Am. J. Physiol. 261: G885. 390. Schulz, S., Chinkers, M. and Garbers, D. L. (1989) The guanylate cyclase/receptor family of proteins. FASEB J. 3:2026. 391. Schulz, S., Ynen, P. S. T. and Garbers, D. L. (1991) The expanding family of guanylyl cyclases. TIPS 12:116. 392. Schunack, W. (1987) What are the differences between the H2-receptor antagonists? Aliment. Pharmac. Therapy 1: 493S. 393. Scott, J. B., Glaccum, M. B., Zoller, M. J., Uhler, M. D,, Helfman, D. M., McKnight, G. S. and Krebs, E. G. (1987) The molecular cloning of a type II regulatory subunit of the cAMP-dependent protein kinase from rat skeletal muscle and mouse brain. Proc. Natl. Acad. Sci. USA 84:5192. 394. Seamon, K. B. and Daly, J. W. (1981) Activation of adenylate cyclase by the diterpene forskolin does not require the guanine nucleotide regulatory protein. J. Biol. Chem. 256:9799. 395. Seidler, U., Beinborn, M, and Sewing, K. F. (1989) Inhibition of acid formation in rabbit parietal cells by prostaglandins is mediated by the prostaglandin E2 receptor. Gastroenterol. 96:314. 396. Sekiguchi, M., Sakakibara, K. and Fujii, G. (1978) Establishment of cultured cell lines derived from a human gastric carcinoma. Jap. J. Exp. Med. 48:61. 397. Sekiguchi, T., Nishioka, T., Kogure, M., Kusano, M., Horikoshi, T., Sugiyama, T. and Kobayashi, S. (1987) Once-daily administration of famotidine for reflux esophagitis. Scand. J. Gastroenter. 22: 51. 398. Shah, G. V., Kacsoh, B., Seshardri, R., Grosvenor, C. E. and Crowley, W. R. (1989) Presence of calcitonin-like peptide in rat milk: Possible physiological role in regulation of neonatal prolactin secretion. Endocrinology 125:6l. 399. Sharp, J. D., White, D. L., Chiou, G. X., Goodson, T., Gamboa, G. C., McClure, D., Burgett, S., Hoskins, J., Skatrud, P. L., Sportsman, J. R., Becker, G. W., Kang, L. H., Roberts, E. F. and Kramer, R. M. (1991) Molecular cloning and expression of human Ca2+-sensitive cytosolic phospholipase A2. J. Biol. Chem. 266:14850. 400. Shaw, G. P., Hatt, J. F., Anderson, N. G. and Hanson, P. J. (1987) Action of epidermal growth factor on acid secretion by rat isolated parietal cells. Biochem. J. 244:699. 401. Shepherd-Rose, A. J. and Pendleton, R. G. (1985) Studies on the H2-receptor antagonism of MK-208 in isolated rabbit gastric glands. Eur. J. Pharmac. 106:423. 402. Shull, G. E. (1990) CDNA cloning of the fl-subunit of the rat gastric H, K-ATPase. J. Biol. Chem. 265:12123. 403. Sigrist-Nelson, K., Krasso, A., Miiller, R. K. M. and Fischli, A. E. (1987) Ro 18-5364, a potent new inhibitor of the gastric (H + + K+)-ATPase. Eur. J. Biochem. 166:453.

366

Mirossay, Di Gioia, Chastre, Emami and Gespach

404. Sinha, S. K. and Yunis, A. A. (1983) Isolation of colony stimulating factor from human milk. Biochem. Biophys. Res. Commun. 114:797. 405. Skoglund, M. L., Gerber, J. G., Murphy, R. C. and Nies, A. S. (1980) Prostaglandin production by intact isolated gastric parietal cells. Eur. J. Pharmac. 66:145. 406. Smith, C. D., Uhing, R. J. and Snyderman, R. (1987) Nucleotide regulatory protein-mediated activation of phospholipase C in human polymorphonuclear leukocytes is disrupted by phorbol esters. J. Biol. Chem. 262:6121. 407. Smith, J. L., Gamal, M. A., Chremos, A. N. and Graham, D. Y. (1985) Famotidine, a new H2-receptor antagonist. Effect on parietal, nonparietal, and pepsin secretion in man. Digestive Diseases and Sciences 30:308. 408. Solcia, E., Capella, C., Buffa, R., Frigerio, B. and Fiocca, R. (1980) Pathology of the Zollinger-Ellison syndrome. Prog. Surg. Pathol. 1: 112. 409. Soil, A. H., Amirian, D. A., Thomas, L. P., Reedy, T. J. and Elashoff, J. D. (1984) Gastrin receptors on isolated canine parietal cells. J. Clin. Invest. 73:1434. 410. Soil, A. H., Toomey, M., Culp, D., Shanahan, F. and Beaven, M. A. (1988) Modulation of histamine release from canine fundic mucosal mast cells. Am. J. Physiol. 254:G40. 411. Soil, A. H. and Toomey, M. (1989) /3-adrenergic and prostanoid inhibition of canine fundic mucosal mast cells. Am. J. Physiol. 256:G727. 412. Soll, A. H. (1978) The actions of secretagogues on oxygen uptake by isolated mammalian epithelial cells. J. Clin. Invest. 61:370. 413. Soil, A. H. (1978) The interaction of histamine with gastrin and carbamylcholine on oxygen uptake by isolated mammalian parietal cells. J. Clin. Invest. 61: 381. 414. Soumarmon, A., Abastado, M., Bonfils, S. and Lewin, M. J. M. (1980) C1- Transport in gastric microsomes. J. Biol. Chem. 255:11682. 415. Speir, G., Takeuchi, K., Peitsch, W. and Johnson, L. (1982) Mucosal gastrin receptor VII. Upand down-regulation. Am. J. Physiol. 242: G243. 416. Stachura, J., Krause, W. J. and Ivey, K. J. (1981) Ultrastructure of endocrine-like cells in lamina propria of human gastric mucosa. Gut. 22:534. 417. Stein, R., Pinkas-Kramarski, R. and Sokolovsky, M. (1988) Cloned M1 muscarinic receptors mediate both adenylate cyclase inhibition and phosphoinositide turnover. The E M B O J. 7:3031. 418. Sternweis, P. C. and Robishaw, J. D. (1984) Isolation of two proteins with high affinity for guanine nucleotide from membranes of bovine brain. J. Biol. Chem. 259:13806. 419. Stewart, B., Wallmark, B. and Sachs, G. (1981) The interaction of H + and K + with the partial reactions of gastric (H + + K+)-ATPase. J. Biol. Chem. 256:2682. 420. Strathmann, M. P. and Simon, M. I. (1991) GoL12 and Goll3 subunits define a fourth class of G protein o; subunits. Proc. Natl. Acad. Sci. USA 88:5582. 421. Sugatani, J., Fujimura, K., Miwa, M., Mizuno, T., Sameshima, Y. and Saito, K. (1989) Occurrence of platelet-activating factor (PAF) in normal rat stomach and alteration of PAF level by water immersion stress. The FASEB Journal 3:65. 422. Sugden, B. (1989) An intricate route to immortality. Cell 57:5. 423. Sun, X. M. and Hsueh, W. Bowel (1988) Necrosis Induced by Tumor Necrosis Factor in Rats is Mediated by Platelet-activating Factor. J. Clin. Invest. 81:1328. 424. Tachr, Y. (1987) Central nervous system regulation of gastric acid secretion. In Physiology of the Gastrointestinal Tract, Johnson, L. R. ed., Raven Press New York, pp. 911. 425. Takagi, T., Takeda, M. and Maeno, H. (1982) Effect of a new potent H2-blocker on gastric secretion induced by histamine and food in conscious dogs. Arch. int. Pharmacodyn 256:49. 426. Takeyama, M., Kondo, K., Hayashi, Y. and Yajima, H. (1989) Enzyme immunoassay of gastrin releasing peptide (GRP)-like immunoreactivity in milk. Int. J. Peptide Protein Res. 34:70. 427. Tang, W.-J., Krupinski, J. and Gilman, A. G. (1991) Expression and characterization of calmodulin-aetivated (type II) adenylylcyclase. J. Biol. Chem. 266:8595. 428. Tang, W.-J. and Gilman, A. G. (1991) Type-Specific Regulation of Adenylyl Cyclase by G Protein fly subunits. Science 254:1500. 429. Tanner, L. A. and Arrowsmith, J. B. (1988) Bradycardia and H2 antagonists. Ann. Int. Med., Sept. 434. 430. Taylor, K. M. (1973) Comparison of the antihistamien effect of burimamide and mepyramide and their effect on the activity of histamine methyltransferase. Biochem. Pharmacol. 22: 2775. 431. Thompson, W. J., Chang, L. K. and Rosenfeld, G. C. (1981) Histamine regulation of adenylyl cyclase of enriched rat gastric parietal cells. Am. J. Physiol. 240: G76-G84. 432. Thompson, N. T., Bonser, R. W. and Garland, L. G. (1991) Receptor-coupled phospholipase D and its inhibition. TIPS 12:404. 433. Thurston, A. W., Cole, J. A., Hilmann, L. S., Im, J. H., Thorne, P. K., Krause, W. J., Jones,

Pharmacological Control of Gastric Acid Secretion

367

J. R., Eber, S. L. and Forte, L. R. (1990) Purification and properties of parathyroid hormone-related peptide isolated from milk. Endocrinology 126:1183. 434. Tielemans, Y., Hakanson, R., Sundler, F. and WiUems, G. (1989) Protififeration of enterochromaffinlike cells in omeprazole-treated hypergastrinemic rats_ Gastroenterol. 96:723. 435. Toh, B. H., Gleeson, P. A., Simpson, R. J., Moritz, R. L., Catlaghan, J. M., Goldkorn, L, Jones, C. M., Martinelli, T. M., Mu, F. T., Humphris, D. C., Pettitt, J_ M , Mort, Y., Masuda, T., Sobieszczuk, P., Weinstock, J., Mantamadiotis, T. and Baldwin, G. S. (1990) The 60- to 90-kDa parietal cell autoantigen associated with autoimmune gastritis is a/3 subunit of the gastric H+/K+-ATPase (proton pump). Proc. Natl. Acad. Sci. USA 87:6418. 436. Topiol, S. and Sabio, M. (1991) The computational design of test compounds with potentially specific biological activity: Histamine-H 2 agonists derived from 5-HT/Hz antagonists. Journal of Computer-Aided Molecular Design 5: 263. 437. Torchiana, M. L., Pendleton, R, G., Cook, P. G., Hanson, C. A. and Clineschmidt, B. V. (1983) Apparent irreversible H2-receptor blocking and prolonged gastric antisecretory activities of 3-N-{3-[3•(1•piperidin•-methy•)ph•nexy]pr•pyl}amin••4-amin•-•,2,5•thi•diaz••e-•-oxide (1643,441). J. Pharmac. Exp. Therap. 224:514. 438. Trzeciakowski, J. P. (1987) Inhibition of guinea pig ileum contractions mediated by a class of histamine receptor resembling the H3 subtype. J. Pharmac. Exp, Therap. 243: 874. 42~9. Tsai, B. S. and Yellin, T. O. (1984) Differences in the interaction of histamine H2 receptor antagonists and tricyclic antidepressants with adenylate cyclase from guinea pig gastric mucosa. Biochem. Pharmac. 33: 3621. 440. Uhing, J. R., Prpic, V., Jiang, H. and Exton, J. H. (1986) Hormone-stimulated polyphosphoinositide breakdown in rat liver plasma membranes. J. Biol. Chem. 261:2140. 441. Ullrich, A. and Schlessinger, J. (1990) Signal transduction by receptors with tyrosine kinase activity. Celt 61: 203. 442. Urushidani, T., Hanzel, D. K. and Forte, J. G. (1989) Characterization of an 80-kDa phosphoprotein involved in parietal cell stimulation. Am. J. Physiol. 256:G1070. 443. Urushidani, T., Hanzel, D. K. and Forte, J. G. (1987) Protein phosphorylation associated with stimulation of rabbit gastric glands. Biochim. Biophys. Acta 930:209. 444. Vanhoutte, P. M. (1990) Endothelium-derived relaxing and contracting factors. Adv. Nephrol. 19:3. 445. Van Tran~ T. and Snyder, S. H. (1981) Histidine decarboxylase. Purification from fetal rat liver, immunologic properties and histochemicai localization in brain and stomach. J. Biol. Chem. 256: 680. 446. Varticovski, L., Druker, B., Morrison, D., Canttey, L. and Roberts, T. (1989) The colony stimulating factor-1 receptor associates with and activates phosphatidylinositol 3-kinase. Nature 342: 699. 447. Vigna, S. R., Mantyh, C. R., Giraud, A. S., Soll, A. H., Walsh, J. H. and Mantyh, P. W. (1987) Localization of specific binding sites for bombesin in the canine gastrointestinal ~ract. Gastroenterol. 93:1287. 448. Vinayek, R., Amantea, M. A., Maton, P. N., Frucht, H., Gardmer, J. D. and Jensen, R. T. (1991) Pharmacokinetics of oral and intravenous omeprazole in patients with the ZollingerEllison syndrome. Gastroenterol. 101:138. 449. Vogelsanger, B., Godrey, P. D. and Brown, R. D. (1991) Rotational spectra of biomolecutes: Histamine. J. Am. Chem. Soc. 113:7864. 450. Wakelam, M., Murphy, G., Hruby, V. and Houslay, M. (1986) Ac'tivation of two signaltransduction systems by glucagon. Nature 323: 68. 451. Wallace, J. L., Cucala, M., Mugridge, K. and Parente, L. (1991) Secretagogue-specific effects of interleukin-1 on gastric acid secretion. Am. J. Physiol. 261:G559o 452. Wallmark, B., Sachs, G., Mardh, S. and Fellenius, E. (1983) Inhibition of gastric (H* + K+)ATPase by the substituted benzimidazole picoprazole. Biochim. Biophys Acta 728:31. 453. Wallmark, B., Jaresten, B. M., Larsson, H., Ryberg, B., Brgndstr6m, A. and Fellenius, E. (1983) Differentiation among inhibitory actions of omeprazole, cimetidine, and SCN- on gastric acid secretion. Am. J. Physiol. 245:G64. 454. Wallmark, B., Br/indstr6m, A. and Larsson, H. (1984) Evidence for acid-induced transformation of omeprazole into an active inhibitor of (H++ K+)-ATPase within the parietal cell. Biochim. Biophys. Acta. 778:549. 455. Warlow, R. S., White, R. and Bernard, C. C. A. (1986) Solubilization and characterization of a low-affinity histamine-binding site on human blood mononuclear cells. Molecular. lmmunol. 23: 393.

368

Mirossay, Di Gioia, Chastre, Emami and Gespach

456. Warrander, S. E., Norris, D. B., Rising, T. J. and Wood, T. P. (1983) 3H-cimetidine and the H2-receptor. Life Sci. 33:1119. 457. Watson, S. and Abbott, A. (1992) Receptor Nomenclature Supplement. TIPS 12:1. 458. Werner, H., Koch, Y., Fridkin, M., Fahrenkrug, J. and Goes, I. (1985) High levels of vasoactive intestinal peptide in human milk. Biochem. Biophys. Res. Commun. 133:228. 459. Whalin, M. E., Scammell, J. G., Strada, S. J. and Thompson, W. J. (1991) Phosphodiesterase II, the cGMP-activable cyclic nucleotide phosphodiesterase, regulates cyclic AMP metabolism in PC12 cells. Mol. Pharmacol. 39:711. 460. Whitman, M., Downes, C. P., Keeler, M., Keeler, T. and Cantley, L. (1988) Type I phosphatidylinositol kinase makes a novel inositol phospholipid, phosphatidylinositol-3phosphate. Nature 332: 644. 461. Wiedenmann, B., Waldherr, R., Buhr, H., Hille, A., Rosa, P. and Huttner, W. B. (1988) Identification of gastroenteropancreatic neuroendocrine cells in normal and neoplastic human tissue with antibodies against synaptophysin, chromgranin A, secretogranin I (chromogranin B) and secretogranin II. Gastroenterol. 95: 1364. 462. Williams, L. T. (1989) Signal transduction by the platelet-derived growth factor receptor. Science 7,43:1564. 463. Wilson, W. E., Lazarus, L. H. and Tomer, K. B. (1989) Bradykinin and kininogens in bovine milk. J. Biol. Chem. 264" 17777. 464. Wollin, A., Barnes, L. D., Hui, Y. S. and Dousa, T. P. (1975) Activation of protein kinase in the guinea pig fundic gastric mucosa by histamine. Life Sci. 17:1303. 465. Wolosin, J. M. and Forte, J. G. (1984) Stimulation of oxyntic cell triggers K+ and Cl-conductances in apical H+-K+-ATPase membrane. Am. J. Physiol. 246: C537. 466. Wouters, W., Van Dun, J., Leysen, J. E. and Laduron, P. M. (1985) Photoaflinity probes for serotonin and histamine receptors. Synthesis and characterization of two azide analogues of ketanserin. J. Biol. Chem. 260:8423. 467. Wright, M. S., Gautvik, V. T. and Gautvik, K. M. (1992) Cloning strategies for peptide hormone receptors. A cta Endocrinol. 126: 97. 468. Yamada, T. (1987) Local regulatory actions of gastrointestinal peptides. In Physiology o f the Gastrointestinal Tract, Johnson, L. R., ed. Raven Press, New York, pp. 131. 469. Yamada, Y., Post, S. R., Wang, K., Tager, H. S., Bell, G. I. and Seino, S. (1992) Cloning and functional characterization of a family of human and mouse somatostatin receptors expressed in brain, gastrointestinal tract, and kidney. Proc. Natl. Acad. Sci. USA 89:251. 470. Yamashita, M., Fukui, H., Sugama, K., Horio, Y., Ito, S., Mizuguchi, H. and Wada, H. (1991) Expression cloning of a cDNA encoding the bovine histamine H 1 receptor. Proc. Natl. Acad. Sci. USA 88:11515. 471. Yarden, Y. and Ullrich, A. (1988) Growth factor receptor tyrosine kinases. Annu. Rev. Biochem. 57: 443. 472. Yokel, R. A. and Dickey, K. M. (1991) Mucosal injury and y-irradiation produce persistent gastric ulcers in the rabbit. Gastroenterol. 100:1201. 473. Yokotani, K., Muramatsu, I. and Fujiwara, M. (1984) Alpha-1 and alpha-2 type adrenoceptors involved in the inhibitory effect of splanchnic nerves on parasympathetically stimulated gastric acid secretion in rats. J. Pharmacol. Exp. Ther. 299:305. 474. Yoshimasa, T., Sibley, D. R., Bouvier, M., Lefkowitz, R, J. and Caron, M. G. (1987) Cross-talk between cellular signalling pathways suggesting by phorbol-ester-induced adenylate cyclase phosphorylation. Nature 327: 67. 475. Young, D., Waitches, G., Birchmeier, C., Fasano, O. and Wiglet, M. (1986) Isolation and Characterization of a New Cellular Oncogene Encoding a Protein with multiple Potential Transmembrane Domains. Cell. 45:711.

Pharmacological control of gastric acid secretion: molecular and cellular aspects.

Bioscience Reports, VoL !2, No. 5, I992 REVIEW Pharmacological Control of Gastric Acid Secretion" Molecular and Cellular Aspects Ladislav Mirossay,...
4MB Sizes 0 Downloads 0 Views