potassium-competitive acid blockade: a new therapeutic...

15
Associate editor: G.J. Sanger Potassium-competitive acid blockade: a new therapeutic strategy in acid-related diseases Kjell Andersson * , Enar Carlsson AstraZeneca R&D, 431 83 Mo ¨lndal, Sweden Abstract Current therapies to treat gastroesophageal reflux disease (GERD), peptic ulcer disease (PUD), and other acid-related diseases either prevent stimulation of the parietal cell (H 2 receptor antagonists, H 2 RAs) or inhibit gastric H + ,K + -ATPase (e.g., proton pump inhibitors, PPIs). Of the 2 approaches, the inhibition of the final step in acid production by PPIs provides more effective relief of symptoms and healing. Despite the documented efficacy of the PPIs, therapeutic doses have a gradual onset of effect and do not provide complete symptom relief in all patients. There is scope for further improvements in acid suppressive therapy to maximize healing and offer more complete symptom relief. It is unlikely that cholecystokinin 2 (CCK 2 , gastrin) receptor antagonists, a class in clinical trials, will be superior to H 2 RAs or PPIs. However, a new class of acid suppressant, the potassium-competitive acid blockers (P-CABs), is undergoing clinical trials in GERD and other acid-related diseases. These drugs block gastric H + ,K + -ATPase by reversible and K + -competitive ionic binding. After oral doses, P-CABs rapidly achieve high plasma concentrations and have linear, dose-dependent pharmacokinetics. The pharmacodynamic properties reflect the pharmacokinetics of this group (i.e., the effect on acid secretion is correlated with plasma concentrations). These agents dose dependently inhibit gastric acid secretion with a fast onset of action and have similar effects after single and repeated doses (i.e., full effect from the first dose). Animal studies comparing P-CABs with PPIs suggest some important pharmacodynamic differences (e.g., faster and better control of 24-hr intragastric acidity). Studies in humans comparing PPIs with P-CABs will help to define the place of this new class in the management of acid-related diseases. D 2005 Elsevier Inc. All rights reserved. Keywords: Gastroesophageal reflux; Peptic ulcer disease; Potassium-competitive acid blocker; Proton pump inhibitor Abbreviations: CCK, cholecystokinin; cyclic AMP, cyclic adenosine monophosphate; ECL, enterochromaffin-like cell; GERD, gastroesophageal reflux disease; H 2 RA, histamine H 2 receptor antagonist; P-CAB, potassium-competitive acid blocker; PPI, proton pump inhibitor. Contents 1. Introduction ..................................... 295 2. Physiology of acid secretion ............................. 295 3. Structure and properties of gastric H + ,K + -ATPase .................. 296 4. Targeting gastric acid secretion ........................... 296 4.1. H 2 receptor antagonists and H 3 receptor agonists .............. 297 4.2. Muscarinics and cholecystokinin 2 receptor antagonists ............ 297 4.3. Proton pump inhibitors ............................ 297 5. Potassium-competitive acid blockers ......................... 299 5.1. Development of the potassium-competitive acid blocker class ........ 299 5.2. Mechanism of potassium-competitive acid blocker inhibition of gastric H + ,K + -ATPase ................................ 300 5.3. Selectivity of potassium-competitive acid blockers for gastric H + ,K + -ATPase . 301 0163-7258/$ - see front matter D 2005 Elsevier Inc. All rights reserved. doi:10.1016/j.pharmthera.2005.05.005 * Corresponding author. Integrative Pharmacology, AstraZeneca R&D, Mo ¨ lndal, Sweden. E-mail address: [email protected] (K. Andersson). Pharmacology & Therapeutics 108 (2005) 294 – 307 www.elsevier.com/locate/pharmthera

Upload: others

Post on 22-Oct-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

  • www.elsevier.com/locate/pharmthera

    Pharmacology & Therapeutic

    Associate editor: G.J. Sanger

    Potassium-competitive acid blockade: a new therapeutic

    strategy in acid-related diseases

    Kjell Andersson *, Enar Carlsson

    AstraZeneca R&D, 431 83 Mölndal, Sweden

    Abstract

    Current therapies to treat gastroesophageal reflux disease (GERD), peptic ulcer disease (PUD), and other acid-related diseases either

    prevent stimulation of the parietal cell (H2 receptor antagonists, H2RAs) or inhibit gastric H+,K+-ATPase (e.g., proton pump inhibitors, PPIs).

    Of the 2 approaches, the inhibition of the final step in acid production by PPIs provides more effective relief of symptoms and healing.

    Despite the documented efficacy of the PPIs, therapeutic doses have a gradual onset of effect and do not provide complete symptom relief in

    all patients. There is scope for further improvements in acid suppressive therapy to maximize healing and offer more complete symptom

    relief. It is unlikely that cholecystokinin2 (CCK2, gastrin) receptor antagonists, a class in clinical trials, will be superior to H2RAs or PPIs.

    However, a new class of acid suppressant, the potassium-competitive acid blockers (P-CABs), is undergoing clinical trials in GERD and

    other acid-related diseases. These drugs block gastric H+,K+-ATPase by reversible and K+-competitive ionic binding. After oral doses,

    P-CABs rapidly achieve high plasma concentrations and have linear, dose-dependent pharmacokinetics. The pharmacodynamic properties

    reflect the pharmacokinetics of this group (i.e., the effect on acid secretion is correlated with plasma concentrations). These agents dose

    dependently inhibit gastric acid secretion with a fast onset of action and have similar effects after single and repeated doses (i.e., full effect

    from the first dose). Animal studies comparing P-CABs with PPIs suggest some important pharmacodynamic differences (e.g., faster and

    better control of 24-hr intragastric acidity). Studies in humans comparing PPIs with P-CABs will help to define the place of this new class in

    the management of acid-related diseases.

    D 2005 Elsevier Inc. All rights reserved.

    Keywords: Gastroesophageal reflux; Peptic ulcer disease; Potassium-competitive acid blocker; Proton pump inhibitor

    Abbreviations: CCK, cholecystokinin; cyclic AMP, cyclic adenosine monophosphate; ECL, enterochromaffin-like cell; GERD, gastroesophageal reflux

    disease; H2RA, histamine H2 receptor antagonist; P-CAB, potassium-competitive acid blocker; PPI, proton pump inhibitor.

    Contents

    0163-7258/$ - see fro

    doi:10.1016/j.pharmth

    * Corresponding aut

    E-mail address: k

    1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295

    2. Physiology of acid secretion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295

    3. Structure and properties of gastric H+,K+-ATPase . . . . . . . . . . . . . . . . . . 296

    4. Targeting gastric acid secretion . . . . . . . . . . . . . . . . . . . . . . . . . . . 296

    4.1. H2 receptor antagonists and H3 receptor agonists . . . . . . . . . . . . . . 297

    4.2. Muscarinics and cholecystokinin2 receptor antagonists . . . . . . . . . . . . 297

    4.3. Proton pump inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297

    5. Potassium-competitive acid blockers. . . . . . . . . . . . . . . . . . . . . . . . . 299

    5.1. Development of the potassium-competitive acid blocker class . . . . . . . . 299

    5.2. Mechanism of potassium-competitive acid blocker inhibition of gastric

    H+,K+-ATPase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300

    5.3. Selectivity of potassium-competitive acid blockers for gastric H+,K+-ATPase . 301

    nt matter D 2005 Elsevier Inc. All rights reserved.

    era.2005.05.005

    hor. Integrative Pharmacology, AstraZeneca R&D, Mölndal, Sweden.

    [email protected] (K. Andersson).

    s 108 (2005) 294 – 307

  • K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307 295

    5.4. Pharmacokinetics of potassium-competitive acid blockers . . . . . . . . . . 301

    5.5. Pharmacodynamics of potassium-competitive acid blockers . . . . . . . . . 302

    5.6. Pharmacodynamic comparisons of potassium-competitive acid blockers with

    other agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302

    6. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303

    References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303

    1. Introduction

    By the 19th century, gastric acid was known to be

    important in many physiological processes, such as protein

    digestion and the absorption of calcium and iron. Alongside

    these functions, sterilization of food is now also recognized

    as a vital function of gastric acid. Over time, awareness of

    its central role in the etiology of peptic ulcer disease (PUD),

    and more recently, gastroesophageal reflux disease (GERD),

    has also grown. This knowledge, together with an increased

    understanding of the physiology of acid production, directed

    the search for therapies in these diseases through inhibition

    of gastric acid secretion. However, until just under 30 years

    ago, patients with acid-related diseases had to rely on

    antacids, atropine, and other interventions (e.g., behavioral)

    for the relief of symptoms. The treatment of PUD was

    challenging and, apart from surgery, employed largely

    ineffective approaches that included a bland diet, antacids,

    and anticholinergics. Although antacids, and other

    approaches such as dietary changes and sucralfate, provide

    a degree of relief from the symptoms of GERD, healing of

    erosions was very difficult to achieve except with high and

    frequent doses of antacids (Grove et al., 1985).

    The emergence of H2 receptor antagonists (H2RAs) in

    the 1970s represented the first major advance in the

    treatment of GERD and PUD, as they provided better

    symptom control and allowed higher healing rates

    compared with antacids. However, these agents have a

    relatively short duration of action, their effect is dimin-

    ished by meal-stimulated secretion, and tolerance to their

    antisecretory effect can develop. Other approaches used to

    treat GERD (e.g., prokinetics, sucralfate) have generally

    not met expectations.

    The first proton pump inhibitor (PPI), omeprazole, was

    approved for use in 1988 (Lindberg et al., 1990) and was the

    forerunner of a more effective class of agents than H2RAs.

    PPIs provide superior symptom relief and achieve higher

    healing rates in GERD and PUD than do H2RAs (Chiba et

    al., 1997; van Pinxteren et al., 2001; Salas et al., 2002).

    Despite the undoubted efficacy of PPIs, there are still areas

    in which they could be improved upon (e.g., faster and

    better symptom control and more rapid healing). The quest

    for better therapy has driven research into other acid-

    suppressive treatments. This review examines the develop-

    ment and properties of current and potential treatments to

    suppress acid production, with particular emphasis on the

    potassium-competitive acid blocker (P-CAB) class.

    2. Physiology of acid secretion

    Hydrochloric acid (HCl) is secreted into the lumen of the

    stomach by parietal cells in the glands of the oxyntic

    mucosa. Gastric H+,K+-ATPase is fundamental to this

    process of acid secretion. This enzyme, located in the apical

    membrane of the parietal cell, transports H+ into the parietal

    cell canaliculus in exchange for K+. The secretion of H+ is

    accompanied by the passage of Cl� across the apical

    membrane into the canaliculus, which ensures that acid

    secretion is electroneutral. For each H+ ion that is moved

    into the canaliculus by the action of the H+,K+-ATPase, a

    HCO3� ion is moved out of the parietal cell cytoplasm by a

    basolateral Cl�/HCO3� exchanger that also delivers a Cl�

    ion into the cytosol. Cl� is secreted into the canaliculus via

    Cl� channels in the apical membrane of the parietal cell. It is

    possible that more than 1 type of Cl� channel is involved in

    the movement of Cl� across the membrane, but currently,

    there is evidence only for the ClC-2 channel (Malinowska et

    al., 1995).

    The exchange of H+ for K+ requires a relatively high

    level of K+ in the parietal cell canaliculus. There is evidence

    that K+ transported into the cytoplasm by the enzyme is

    recycled and returns to the canaliculus via specific K+

    channels in the apical membrane of the cell. To date, 3

    different types of K+ channels (KCNQ1, Kir2.1, and Kir4.1)

    with properties consistent with a role in acid secretion have

    been identified in animals (Dedek & Waldegger, 2001;

    Grahammer et al., 2001; Fujita et al., 2002; Lambrecht et al.,

    2004; Malinowska et al., 2004). It is uncertain what the

    counterpart(s) of these channels may be in humans.

    Production of gastric acid by the parietal cell is a tightly

    regulated process and is triggered by physiological stim-

    ulation of receptors located on the basolateral membrane of

    the cell. The stimuli include histamine, gastrin, and

    acetylcholine, each of which binds to a specific receptor.

    This ligand–receptor interaction involves an elevation of

    intracellular calcium and/or cyclic adenosine monophos-

    phate (cyclic AMP) as part of the signal transduction

    pathways. Activation of these pathways result in morpho-

    logical and ultrastructural changes that lead to the inclusion

    of the gastric H+,K+-ATPase into the apical membrane of the

    cell and, ultimately, acid secretion.

    In the cytoplasmic space of the unstimulated parietal cell

    are tubulovesicles, which contain the gastric H+,K+-ATPase.

    The apical membrane of the parietal cell has canaliculi that

    invaginate from its surface and in the resting cell that are

  • K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307296

    lined with short microvilli. When the cell is stimulated, the

    microvilli on the apical membrane elongate, the canalicular

    spaces enlarge, and the number of tubulovesicles decreases.

    It is generally agreed that the tubulovesicular membrane

    fuses with the apical plasma membrane (Forte & Yao, 1996;

    Duman et al., 2002). Prior to stimulation of the cell, the

    H+,K+-ATPase is inactive due to the low permeability of the

    tubulovesicular membranes to K+ (Reenstra & Forte, 1990).

    The enzyme does not appear to undergo any chemical

    modification on membrane fusion (Dunbar & Caplan,

    2001), but its consequent placement near to K+ and Cl�

    channels, together with the availability of K+ in the

    canaliculus, provides the necessary environment to allow

    ATP-driven secretion of HCl (Forte & Yao, 1996).

    The concentration of H+ in the parietal cell canaliculi

    results in a pH of 1.0 or less, compared with a pH of 7.4 in

    the blood and parietal cell cytoplasm. To be able to create

    such a pH gradient requires a lot of energy (e.g., ATP),

    hence the parietal cell is the most mitochondria-rich cell in

    the body. The pH in the canaliculi is much lower than that in

    other acidic compartments such as lysosomes, endosomes,

    and chromaffin granules. In these locations, the pH ranges

    from 4.5 to 6.5 (Futai et al., 2000). Gastric H+,K+-ATPase

    has been suggested to be present at sites other than the

    stomach (e.g., the cortical collecting duct of the kidney;

    [Kraut et al., 2001], rat vascular smooth muscle cells

    [McCabe & Young, 1992], human leucocytes [Ritter et al.,

    1998], and rat cardiac myocytes [Beisvag et al., 2003]). The

    enzymes at these locations do not generate a very low pH,

    but probably contribute to acid–base and potassium homeo-

    stasis (van Driel & Callaghan, 1995; Sangan et al., 1997;

    Kraut et al., 2001). Other H+,K+-ATPase isoforms have been

    reported to occur in various cells (e.g., in the colon;

    Rajendran et al., 1998). For a discussion of the selectivity

    of P-CABs for gastric H+,K+-ATPase, see Section 5.3.

    3. Structure and properties of gastric H+,K+-ATPase

    The gastric H+,K+-ATPase is an a/h heterodimer, witheach subunit having distinct functions. The a subunit, whichhas 10 helical transmembrane segments (M1–M10), is

    responsible for the catalytic activity of the enzyme and

    contains sites for ATP binding, as well as the cation (K+ and

    H+) binding site. The latter binding site is located near the

    middle of the membrane domains in the a subunit (Munsonet al., 2000). It appears to be formed by amino acid residues

    from M4, M5, and M6, with the ion being held in place by 6

    oxygen atoms provided by these residues (Koenderink et al.,

    2004). The h subunit has a single transmembrane domain.The subunit is required for the functional expression of the

    enzyme.

    The exchange of H+ for K+ involves conformational

    changes in the tertiary structure of gastric H+,K+-ATPase

    (Vander Stricht et al., 2001). According to the Post-Albers

    model, there are 2 important conformational states. In the E1

    form, the ion-binding site faces the parietal cell cytoplasm

    and has high affinity for H+ but low affinity for K+. In the E2form, the ion-binding site faces the extracellular lumen with

    low affinity for H+ and high affinity for K+. The relative

    affinity of the 2 forms for K+ may be determined by

    differences between the E1 and E2 forms in the topography

    of the K+ ion binding site or in the path through which the

    ion accesses its binding site (Vagin et al., 2003). The K+

    affinity is also influenced by a salt bridge from M5 to M6

    that exists only when the enzyme is in the E2 form

    (Koenderink et al., 2004).

    H+ binds to the cytoplasmic face of the enzyme when it

    is in the E1 form. The E1 form also binds ATP from the

    cytoplasm to form a phosphoenzyme (E1P), which

    provides the energy for the change to the (phosphorylated)

    E2 form of the enzyme. The conformational shift causes

    the translocation of H+ from the parietal cell cytoplasm

    into the canaliculus. The phosphorylated E2 form binds K+

    from the canaliculus. This ion is necessary for the

    subsequent dephosphorylation of the H+,K+-ATPase

    (Rabon et al., 1993). In particular, K+ appears to affect

    the conformation of a large loop in which there is a

    phosphorylation domain, and it is involved in stabilizing a

    hairpin between M5 and M6 that appears to be involved in

    linking ATP hydrolysis to cation transport (Swarts et al.,

    1998; Gatto et al., 1999). When K+ occupies the cation

    binding site, it may activate the enzyme by neutralizing a

    negative charge (or charges) that inhibit the dephosphor-

    ylation reaction (Swarts et al., 1998; Hermsen et al., 2000).

    Upon dephosphorylation, the conformation of the enzyme

    returns to the E1 form in which the binding site is again

    exposed to the cytoplasm. The enzyme then releases K+

    into the parietal cell cytoplasm. Thus, K+ is not only the

    counterion for H+, but is also essential for the catalytic

    cycle of gastric H+,K+-ATPase.

    4. Targeting gastric acid secretion

    The treatment of peptic ulcer disease (PUD) has been

    based on Karl Schwartz’s dictum of Fno acid, no ulcer_.Although recent advances in our understanding have high-

    lighted the multifactorial pathogenesis of both PUD and

    GERD, gastric acid is still recognized as a central

    component in both diseases. There is a correlation between

    healing of GERD lesions and the proportion of time (over

    24 hr) when intragastric pH is greater than 4 (Bell et al.,

    1992). Furthermore, the period during which esophageal pH

    is less than 4 increases on moving from endoscopy-negative

    GERD patients to patients with worsening grades of

    esophagitis (Fiorucci et al., 1992). Therapeutically, the

    degree of acid suppression and its relationship with a

    positive outcome (e.g., symptom relief, healing) has been

    documented both for GERD and PUD (Howden et al., 1994;

    Chiba et al., 1997; Selby et al., 2000; van Pinxteren et al.,

    2003).

  • K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307 297

    The groundwork for the current therapeutic approaches

    to controlling gastric acid production was laid in the 19th

    century. Through the work of many scientists (e.g., Prout,

    Gunzberg, Beaumont, Heidenhain, and Cooper), the phys-

    iology of gastric secretion and the pathophysiology of PUD

    have been increasingly elucidated. There has also been an

    increasing understanding of the role of gastric acid in

    GERD. By the mid-20th century, this growing knowledge

    led to several pharmaceutical companies (e.g., SmithKline

    & French, Servier, Searle, and Astra) developing drugs to

    inhibit the production of gastric acid.

    4.1. H2 receptor antagonists and H3 receptor agonists

    The development of H2RAs can be traced back to the

    early 1960s when Sir James Black and his colleagues at

    SmithKline & French embarked on a research program to

    identify histamine receptors. This work resulted in the

    development of burimamide, the first selective antagonist of

    the H2 receptor. Burimamide was the lead compound in a

    search for more potent antagonists. It was tested in humans

    but was not sufficiently active by the oral route to be

    developed as a medicine. Chemical modifications and

    subsequent screening led to the development of cimetidine.

    This H2RAwas approved in 1976 in the UK and in 1977 in

    the USA and was the progenitor of a group of compounds

    that has been recognized as a breakthrough in the treatment

    of PUD and GERD.

    H2RAs are more effective than antacids at providing

    symptom relief and lesion healing in both PUD and GERD.

    However, these compounds have a relatively short duration

    of effect and their effects are mitigated by food ingestion. A

    further limitation is that tolerance can develop after a few

    days of continuous dosing, a clinically relevant phenom-

    enon in some patients (Lachman & Howden, 2000;

    Komazawa et al., 2003).

    The work of Black and colleagues caused a resurgence of

    interest in histamine, and a third histamine receptor, H3, has

    since been discovered. This receptor is detected in the

    peripheral and central nervous system and in several

    nonneural tissues (Barocelli & Ballabeni, 2003). Centrally

    located H3 receptors probably contribute to the regulation of

    acid secretion and inhibit cholinergic-stimulated acid

    production. The H3 receptor agonist R(a)-methylhistaminehas had variable and contradictory effects on gastric acid

    secretion in vitro and in vivo (Barocelli & Ballabeni, 2003).

    Thus, it is difficult to predict whether this or other agents

    will be effective treatments for acid-related diseases.

    Currently, there does not appear to be an H3 receptor

    agonist in clinical development for the treatment of gastric

    acid-related diseases.

    4.2. Muscarinics and cholecystokinin2 receptor antagonists

    Other stimulatory pathways in gastric acid secretion have

    been the focus of approaches to block acid production; they

    include muscarinic (M3) and cholecystokinin2 (CCK2,

    gastrin) receptor antagonists. There are currently no M3-

    specific antagonists available in the clinical setting for the

    treatment of acid-related diseases. However, anticholinergic

    agents (e.g., pirenzepine, telenzepine, and atropine) have a

    long history in the treatment of PUD (Eltze et al., 1985;

    Fiorucci et al., 1988), but as they inhibit M1 receptors at

    locations outside the stomach, they are associated with

    anticholinergic side effects that limit their use.

    The development of the CCK2 antagonists has been

    reviewed by Black and Kalindjian (2002). Acting via CCK2receptors, gastrin mediates acid secretion primarily by

    causing release of histamine from enterochromaffin-like

    (ECL) cells and also, at least in vitro, by directly stimulating

    parietal cells (Kopin et al., 1992; Cabero et al., 1993; Prinz

    et al., 1993, 1994). A number of CCK2 receptor antagonists

    have been evaluated and shown to reduce acid secretion

    (YF476 [Steel, 2002], YM022 [Nishida et al., 1994],

    RP73870 [Pendley et al., 1995], spiroglumide [Beltinger

    et al., 1999], and S0509 [Takeuchi et al., 1999]), but their

    development has been discontinued.

    Two CCK2 receptor antagonists in clinical studies

    include Z360 and itriglumide (Vakil, 2004). Z360 is in

    early clinical development for reflux esophagitis and gastric

    ulcer. In preclinical studies, Z360 was more potent than

    famotidine at inhibiting pentagastrin-stimulated acid secre-

    tion and also reduced meal-induced acid production in rats

    and dogs (Miura et al., 2001; Morita et al., 2001). The status

    of itriglumide is uncertain: Phase I clinical studies on

    gastrin- and meal-induced gastric secretion were reported to

    be in progress, and Phase II clinical studies in peptic ulcer

    were scheduled to be complete by the end of 2003

    (Rottapharm, 2004). In pentagastrin-stimulated rats, itriglu-

    mide was less potent than ranitidine or omeprazole when

    administered intravenously, but intraduodenally, it was 3

    times more potent than ranitidine and twice as potent as

    omeprazole (Makovec et al., 1999).

    It is unlikely that there will be a role for CCK2antagonists alone as alternative antisecretory therapy to

    PPIs or H2RAs, especially given the development of

    tolerance with at least 1 member of this class (i.e., YF476;

    Black & Kalindjian, 2002; Steel, 2002).

    4.3. Proton pump inhibitors

    One of several companies that started up research

    programs with the intention of developing a drug that

    would reduce gastric acid secretion was Hässle (a

    research unit within AB Astra, Sweden). The initial idea

    was to find a drug to inhibit the release of gastrin but the

    approach was not successful. However in 1971 the French

    company Servier reported (Malen & Danree, 1971) that

    the compound CMN 131 (a gastrin receptor antagonist)

    inhibited acid secretion but resulted in acute toxicity in

    animals. The thioamide group of this compound was

    assumed to be responsible for its toxicity, and therefore

  • Fig. 1. Mode of action of a PPI. PPIs are weak bases that concentrate in the

    parietal cell canaliculus, where they undergo a proton-catalyzed, 3-step

    process to generate the active sulfenamide. This moiety interacts covalently

    with sulfhydryl groups on cysteine residues in the transmembrane domains

    of the gastric H+,K+-ATPase and thereby inhibits the enzyme.

    K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307298

    this group was eliminated by incorporating it into or

    between heterocyclic ring systems. Based on this strategy,

    substituted benzimidazoles with potent antisecretory prop-

    erties but without acute toxicity were identified. A

    number of such compounds (H124/26, timoprazole, and

    picoprazole) were selected as candidate drugs during the

    1970s, but all were associated with safety problems in

    long-term toxicity studies. The breakthrough came with

    the synthesis of omeprazole in 1979. This compound was

    the most potent of a series of candidate drugs synthesized

    and tested, and it did not show any significant effects in

    the initial repeat-dose toxicity studies in animals. Thus,

    the project started with a lead compound CMN 131, a

    gastrin receptor antagonist, and ended with the discovery

    of a complete new mechanism of action: inhibition of

    gastric H+,K+-ATPase.

    Omeprazole was taken into clinical trials, and the first

    publication on the efficacy in man was in 1983 (Gustavsson

    et al., 1983). This and subsequent studies led to omeprazole

    becoming available in Europe in 1988. Omeprazole is

    recognized as a major breakthrough in the treatment of

    GERD and PUD. Moreover, omeprazole is effective in the

    treatment of Zollinger–Ellison syndrome (Lambers et al.,

    1984), which was previously difficult to treat and often

    required surgical intervention.

    Other available PPIs include lansoprazole, pantoprazole,

    rabeprazole, and esomeprazole, all of which are substituted

    benzimadazoles. Tenatoprazole, which is in clinical devel-

    opment, has an imidazopyridine ring instead of a benzima-

    dazole ring but has the same mechanism of action as the

    substituted benzimadazole PPIs. As illustrated by omepra-

    zole, PPIs are more effective than H2RAs at reducing

    intragastric pH and maintaining pH >4 for a longer period

    of time (Bell et al., 1992). They are generally accepted to be

    superior to H2RAs in the treatment of GERD and PUD, as

    they relieve symptoms and heal lesions more effectively

    than H2RAs (Chiba et al., 1997; van Pinxteren et al., 2001;

    Salas et al., 2002). In addition, tolerance to PPIs has not

    been documented (Tefera et al., 2001). The superior efficacy

    of PPIs over H2RAs is attributed to the fact that they inhibit

    gastric H+,K+-ATPase, independently of the nature of the

    stimulus.

    All PPIs are lipophilic compounds with weak base

    properties and pKa values ranging from 3.8 to 5.0. Thus,

    they easily penetrate cell membranes and are accumulated in

    the highly acidic (pH ¨1.0) parietal cell canaliculi. For

    example, omeprazole with a pKa of 4.0 would theoretically

    concentrate 1000-fold in the parietal cell versus blood.

    However, this equilibrium concentration will probably never

    be achieved in vivo as the chemical stability of the

    protonated form of omeprazole is extremely low. Within

    milliseconds, the compound is degraded to form the active

    inhibitor, the sulfenamide (Lindberg et al., 1987; Fig. 1).

    Again, almost instantaneously, the sulfenamide binds

    covalently to cysteine residues (in particular, cysteine 813

    although other residues are involved) on the luminal side of

    the a-subunit of the enzyme (Besancon et al., 1993, 1997).These cysteine residues do not appear to be necessary for

    enzyme functioning, but their binding of the sulfenamide

    may disrupt or prevent the conformational change of the

    enzyme.

    This mechanism of action results in the unique character-

    istics of the pharmacokinetic–pharmacodynamic effect

    pattern of PPIs. The inhibitory effect on acid secretion is

    related to the amount of sulfenamide formed, which, for a

    given dose, is related to the area under the plasma

    concentration time curve (AUC). Thus, the maximal effect

    of a given dose is correlated with the AUC rather than to the

    Cmax, and there is no direct correlation between blood

    plasma concentration and effect at any given time. The

    duration of effect is determined by the half-life of the

    sulfenamide–enzyme complex (at least 24 hr but may be as

    long as 48 hr; Metz et al., 2002), rather than to the half-life

    of the PPI in blood plasma (e.g., 1–2 hr for omeprazole).

    Therapeutic oral doses of PPIs reach steady state and thus

    achieve their maximal effect levels after 4–5 days of daily

    dosing (Fig. 2). Therefore, PPIs in therapeutic doses have a

    slow, cumulative onset of effect (e.g., 24–43% inhibition of

    acid secretion on the first day of treatment and ¨80%

    inhibition at steady state (Cederberg et al., 1992; Dammann

    & Burkhardt, 1999).

    After PPI administration, there is a return of acid

    secretion that is partly due to de novo synthesis of the

    enzyme (Im et al., 1985a; Gedda et al., 1995) and partly to

    the dissolution of the enzyme–sulfenamide complex, owing

    to the effect of endogenous glutathione (Im et al., 1985b;

    Fujisaki et al., 1991). This dissociation may lead to the

    reactivation of the enzyme, together with the release of a

    sulfide corresponding to the PPI.

    In addition to a cumulative onset of effect, PPIs have a

    relatively slow onset of acute effect. PPIs are chemically

    unstable at low pH, such as in the stomach (e.g., the half-

    lives at pH 1.2 range from 1.3 to 4.7 min; Kromer et al.,

    1998) and must be given in an acid-protected form, such as

    enteric-coated granules. While such preparations allow PPIs

  • Fig. 2. Pharmacodynamic profile of a PPI demonstrating that there is a cumulative onset of effect at therapeutic oral doses.

    K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307 299

    to reach the intestine intact from where they can be

    absorbed, they delay absorption.

    Although PPIs have markedly advanced the treatment

    of acid-related diseases, areas of medical need remain

    (e.g., faster and more complete symptom relief, more

    rapid healing). The degree and speed of onset of symptom

    relief are important to patients (Kleinman et al., 2002).

    However, not all patients with GERD experience complete

    symptom relief after initial PPI therapy, which may

    explain why many patients are unsatisfied with PPI

    treatment (Crawley & Schmitt, 2000; Robinson & Shaw,

    2002; Bytzer, 2003). Improved control of 24-hr intra-

    gastric acidity would be likely to enhance symptom

    resolution and healing, but as the limitations of PPIs

    mainly relate to their shared mechanism of action, it is

    likely that only incremental developments can be

    achieved, making it attractive to consider other potential

    approaches. The clinical utility of many of these strategies

    (e.g., H3 receptor agonists, CCK2 receptor antagonists) is

    likely to be limited, but 1 class of developmental agents

    that holds promise for the treatment of acid-related

    diseases is the P-CABs.

    5. Potassium-competitive acid blockers

    5.1. Development of the

    potassium-competitive acid blocker class

    During acid secretion, the surface of gastric

    H+,K+-ATPase faces the extremely acidic parietal cell

    canaliculus with a high affinity for K+. Given the

    importance of the cation for enzyme function, agents that

    compete with the binding of K+ have the potential to

    block acid secretion. It is this principle that underlies the

    mode of action of P-CABs (previously known as acid

    pump antagonists or APAs). These agents can be classified

    as imidazopyridines (e.g., SCH28080, AZD0865, and

    BY841), pyrimidines (e.g., revaprazan), imidazonaphthyr-

    idine (e.g., soraprazan), or quinolines (e.g., SK&F96067

    and SK&F97574; Wallmark et al., 1987; Keeling et al.,

    1991; Wurst & Hartmann, 1996; Tsukimi et al., 2000;

    Park et al., 2003b; Briving et al., 2004).

    The prototype P-CAB, SCH28080, was developed by

    Schering-Plough (Ene et al., 1982; Long et al., 1983). In a

    study published in 1982, this compound was shown to

    inhibit gastric acid secretion in humans (Ene et al., 1982),

    although its mechanism of action was not fully known at the

    time. Subsequently, SCH28080 was shown to block gastric

    H+,K+-ATPase by competing with K+ (Beil et al., 1986).

    The development of SCH28080 ceased because of hepatic

    toxicity. Despite this, the SCH28080 molecule has been

    used extensively to investigate the functions of gastric

    H+,K+-ATPase and provide insights into the mechanism of

    action of the P-CAB class. Other compounds based on

    SCH28080 were found to have improved bioavailability and

    safety profiles (Kaminski et al., 1987, 1989; Scott et al.,

    1987), including SCH33405. This compound was also

    tested in humans (Sachs et al., 1995), but no publications

    on these clinical studies are available.

    Other P-CABs have provided further proof of principle

    for this class of agents. SK&F96067 (BY067; Keeling et al.,

    1988, 1991) was one of a series of quinolones that inhibited

    gastric acid secretion and was investigated in early clinical

    studies. It inhibited pentagastrin-simulated acid secretion

    and was more effective than ranitidine in raising intragastric

    pH (Pope & Parsons, 1993). SK&F97574 (BY574), an

    analogue of SK&F96067 but with a longer duration of

    action, inhibited H+,K+-ATPase in a potassium-competitive

    fashion, blocked H+ transport in isolated gastric vesicles

    (Pope et al., 1995), and inhibited histamine-stimulated

    gastric acid secretion in the Heidenhain pouch dog (Parsons

    et al., 1995). Preliminary toxicological studies did not reveal

    any untoward findings (Leach et al., 1995), but although

    taken into clinical studies (Pope & Parsons, 1993), the drug

    is no longer under development. BY841 was found to be

    more potent and more selective for the H+,K+-ATPase than

    both SK&F96067 and SK&F97574 (Wurst & Hartmann,

    1996). Based on positive preclinical and clinical Phase I

    data (Wurst & Hartmann, 1996), BY841 entered Phase II

    studies (e.g., comparing healing of reflux esophagitis

    following BY841 and omeprazole treatment), but its

    development was subsequently discontinued without pub-

    lication of trial results.

    Four representatives of the P-CAB class are currently in

    clinical development to improve the treatment of acid-

  • Fig. 3. Mode of action of a P-CAB. As a P-CAB concentrates in the parietal

    cell canaliculi, it is instantaneously protonated. It then binds ionically to the

    gastric H+,K+-ATPase and inhibits acid secretion.

    K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307300

    related diseases (Table 1), with most information being

    available on AZD0865 and revaprazan. Papers have been

    published recently on other molecules (e.g., SPI447 [Ushiro

    et al., 1997; Tsukimi et al., 2000], YJA20379-8 [Sohn et al.,

    1999; Kim et al., 2000], and YJA20379-6 [Kim et al.,

    1998]), but it is not clear if these agents are being evaluated

    in humans.

    5.2. Mechanism of potassium-competitive

    acid blocker inhibition of gastric H+,K+-ATPase

    P-CABs are lipophilic, weak bases that have high pKavalues and are stable at low pH. This combination of

    properties allows them to concentrate in acidic environ-

    ments. For example, the concentration of a P-CAB with a

    pKa of 6.0 would theoretically be expected to be 100,000-

    fold higher in the parietal cell canaliculus (pH 1) than in the

    plasma (pH 7.4). The concentration of P-CABs in the gastric

    mucosa is demonstrated by in vitro and in vivo studies with

    AZD0865 and revaprazan (Park et al., 2003a; Briving et al.,

    2004; Holstein et al., 2004c). In a radiolabeling study in rats,

    the concentration of revaprazan reached peak levels in most

    tissues at 4 hr after oral dosing. At 8 hr after oral dosing, the

    radioactivity was highly localized in the gastric wall (Park et

    al., 2003a). AZD0865 was tested in 2 different preparations

    with isolated H+,K+-ATPase: in ion-tight and in ion-leaky

    vesicles. In the ion-tight vesicle preparation, the high-

    affinity K+-binding site faces intravesically and the vesicles

    have an acidic internal environment. In the ion-leaky

    vesicle, no acidic environment is formed. AZD0865 was

    found to be more potent in ion-tight versus ion-leaky

    vesicles, which suggests that it concentrates in regions of

    low pH (Briving et al., 2004). Further evidence that

    AZD0865 concentrates in acidic environments is provided

    by studies in Heidenhain pouch dogs. For example, the

    concentration of AZD0865 in gastric juice exceeded the

    plasma concentration at ¨2 hr after dose (Holstein et al.,

    2004b), and while the P-CAB was detectable in gastric juice

    at 24 hr postdose, it was undetectable in plasma in most

    animals (Holstein et al., 2004c).

    On entering an acidic environment, P-CABs are instantly

    protonated and it is in this form that it is thought to bind to

    and inhibit the enzyme (Fig. 3). In keeping with this

    hypothesis, the potency of P-CABs increases as pH falls

    Table 1

    P-CABS currently in clinical development

    P-CAB Description Clinical

    development

    phase

    Company

    AZD0865 Imidazopyridine Phase II AstraZeneca

    CS526 (R105266) Pyrrolopyridazine Phase I Sankyo and

    Ube/Novartis

    Revaprazan

    (YH1885, Revanex)

    Pyrimidine Phase III Yuhan

    Soraprazan (BY359) Imidazonaphthyridine Phase II Altana

    (Briving et al., 1988, 2004; Keeling et al., 1991; Pope et al.,

    1995; Tsukimi et al., 2000). In ion-leaky vesicles, the IC50value for AZD0865 at pH 6.4 was ¨7 times lower

    compared with the value at pH 7.4 (0.13 and 1.00 AM,respectively). This corresponds very well to the availability

    of 6- to 7-fold more protonated drug at pH 6.4 (Briving et

    al., 2004) The protonated form of a P-CAB inhibits the

    H+,K+,-ATPase by binding ionically to it (Wallmark et al.,

    1987; Keeling et al., 1991; Tsukimi et al., 2000; Park et al.,

    2003b; Briving et al., 2004), as illustrated by the recovery of

    enzyme activity after washout of AZD0865 and revaprazan

    (Park et al., 2003a; Briving et al., 2004).

    There appear to be separate binding sites on the gastric

    H+,K+-ATPase for P-CABs and K+. For example, K+

    affinity is not affected by mutations in the membrane

    domains that reduce affinity for SCH28080 (and vice versa;

    Asano et al., 1999; Lambrecht et al., 2000; Munson et al.,

    2000; Vagin et al., 2001, 2002). Nevertheless, the K+-

    competitive nature of P-CAB binding indicates that some of

    the residues at or near the cation binding sites are involved

    in P-CAB binding. Research indicates that P-CABs have a

    luminal site of action (Wallmark et al., 1987; Keeling et al.,

    1989; Pope & Parsons, 1993; Briving et al., 2004).

    Investigations with SCH28080 indicate the likely binding

    site of P-CABs. An early photoaffinity labeling study

    suggested that the first extracellular loop (between M1

    and M2) was the direct binding site of SCH28080 (Munson

    et al., 1991), although subsequent mutational studies

    indicate that this is not the case (Asano et al., 1997,

    1999). Several areas of research (e.g., mutational studies,

    NMR studies) indicate that the SCH28080 binding site is

    located in the M1 to M6 and possibly M8 domains (Watts et

    al., 2001; Vagin et al., 2002, 2003; Asano et al., 2004; Yan

    et al., 2004). SCH28080 and SPI447 both bind in a cavity

    formed by the M1, M4, M5, M6, and M8 transmembrane

    segments, and by loops formed by M5/M6, M7/M8, and

    M9/M10 (Asano et al., 2004). This cavity appears to be

    separate from the cation-binding site.

    P-CABs are believed to bind to gastric H+,K+-ATPase

    when the enzyme is in its phosphorylated E2 form and/or

    in its (nonphosphorylated) E2 form (Keeling et al., 1989;

    Mendlein & Sachs, 1990). Support is provided by the

    observed 10-fold increase in the binding affinity of

    SCH28080 for the enzyme in the presence of ATP

  • K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307 301

    (Wallmark et al., 1987; Keeling et al., 1989). Conceptually,

    the binding of a P-CAB to the phosphorylated E2 form is

    consistent with the fact that K+ binds to the same form of

    the enzyme and that P-CABs compete with this ion. The

    docking of a P-CAB in its binding site on the gastric

    H+,K+-ATPase appears to stabilize the enzyme in the E2conformation (Asano et al., 2004) and prevents the

    translocation of H+ ions. A P-CAB molecule may be

    unable to bind to the E1 form of the enzyme due to re-

    arrangement of transmembrane segments and loops that

    form the P-CAB binding site. It appears that the

    orientation of the loop between the M3 and M4 domains

    in the E1 form of the enzyme prevents a P-CAB molecule

    from occupying its binding site (Asano et al., 2004).

    5.3. Selectivity of potassium-competitive

    acid blockers for gastric H+,K+-ATPase

    There are 3 main classes of enzymes that translocate

    H+ in biological systems: P-type, F-type, and V-type of

    H+-ATPases. V-type H+-ATPases are found in endomem-

    branes and plasma membranes, and the F-type is located

    in mitochondria. The P-type H+-ATPases are part of a

    general class of ion-translocating ATPase that are charac-

    terized by the formation of a covalently phosphorylated

    enzyme intermediate as part of their catalytic cycle

    (Rabon & Reuben, 1990). Members of this class include

    gastric H+,K+-ATPase, Na+,K+-ATPase, and Ca+-ATPases

    of sarcoplasmic reticulum and plasma membrane

    (Sachs, 1994). The amino acid sequence of the human

    H+,K+-ATPase is highly homologous (¨60%) to that of

    the human Na+,K+-ATPase (Maeda et al., 1990). The

    importance of K+ for physiological functioning dictates

    that P-CABs must not affect other processes which

    involve this cation, such as the ion exchange by

    Na+,K+-ATPase. Data from studies of P-CABs no

    longer in development show that they have a much

    higher selectivity for gastric H+,K+-ATPase than for

    Na+,K+-ATPase (Keeling et al., 1991; Kromer et al., 2000;

    Tsukimi et al., 2000): SCH28080 has little effect on

    Na+,K+-ATPase activity even at a concentration of 100

    AM compared with an IC50 value of 1.3 AM for theinhibition of the H+,K+-ATPase (Beil et al., 1986);

    SK&F96067 was 32-fold more selective for H+,K+-ATPase

    than Na+,K+-ATPase; and SK&F97574 was 60-fold more

    selective for H+,K+-ATPase than Na+,K+-ATPase (Pope et

    al., 1995). The data available on compounds that are

    currently in clinical development also demonstrate high

    selectivity for gastric H+,K+-ATPase. AZD0865 is more

    than 100-fold selective for gastric H+,K+-ATPase over

    Na+,K+-ATPase. At a concentration of 100 AM, itreduced Na+,K+-ATPase activity by only 9% compared

    with an IC50 value of 1.0 AM for the inhibition of theH+,K+-ATPase (Andersson et al., 2004). Revaprazan is

    also more than 100-fold more selective for H+,K+-ATPase

    over Na+,K+-ATPase (Park et al., 2003b).

    H+,K+-ATPase isoforms found in rat and guinea pig

    distal colon are not inhibited significantly by SCH28080

    (Cougnon et al., 1996). In support of the targeting of

    P-CABs for gastric H+,K+-ATPase in the parietal cell

    canaliculus, AZD0865 does not appear to affect kidney

    function despite the occurrence of the enzyme in the

    cortical collecting duct of the kidney (Andersson et al., in

    press). The weak base properties and, thus, the super-

    concentration of P-CABs at areas of low pH probably

    contribute significantly to the targeting, as P-CABs are

    unlikely to concentrate to any great extent at sites with

    relatively high pH.

    P-CABs are also more selective for gastric H+,K+-ATPase

    than for other H+-translocating ATPases, such as

    vacuolar-ATPases, that are responsible for generating an

    acidic pH in intracellular compartments. This selectivity

    is illustrated by the selectivity of SK&F 96067 for

    vacuolar-ATPase derived from avian osteoclasts and by the

    insensitivity to SCH28080 in a similar preparation (Pope

    & Sachs, 1992; Mattsson et al., 1993). As this type of

    ATPase is highly conserved, it is unlikely that mammalian

    vacuolar-ATPase will be inhibited by P-CABs.

    5.4. Pharmacokinetics of

    potassium-competitive acid blockers

    After oral doses, P-CABs rapidly achieve peak plasma

    concentrations in both animals and humans. This is partly

    because the compounds are stable at low pH and so can be

    administered as immediate-release formulations. A clinical

    study of BY841 reported peak serum concentrations

    between 0.5 and 1.5 hr after administration (Wurst &

    Hartmann, 1996). In a volunteer study, the plasma concen-

    tration of revaprazan reached peak levels within 1.3 to 2.5 hr

    after a single dose (Yu et al., 2004). The Cmax of AZD0865

    after oral administration to Heidenhain pouch dogs was

    generally achieved 0.5–1 hr postdose (Andersson et al.,

    2004). In humans, AZD0865 was rapidly absorbed and

    maximum plasma concentrations occurred within 1 hr in

    most subjects (Nilsson et al., 2005).

    The P-CABs investigated to date exhibit linear pharma-

    cokinetics. The Cmax and AUC of BY841 increased in

    proportion with increasing dose (20–400 mg; Wurst &

    Hartmann, 1996). The serum t1/2 of BY841 was unchanged

    on repeated dosing, although a slight increase in AUC was

    observed (Wurst & Hartmann, 1996). The oral bioavail-

    ability of AZD0865 in Heidenhain pouch dogs was ¨50%

    and the compound had linear pharmacokinetics over the

    dose range 0.125–1 Amol/kg (Holstein et al., 2004b,2004d). Consistent with the findings from animal studies,

    there was a proportional increase in AUC and Cmax with

    dose in human subjects receiving single oral doses of

    AZD0865 (0.08–4.0 mg/kg; Nilsson et al., 2005). The

    plasma concentration–time profiles of AZD0865 are inde-

    pendent of the number of doses given (Holstein et al.,

    2004b, 2004d). For example, there was no increase in the

  • K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307302

    plasma concentration of AZD0865 as a result of repeated

    administration to chronic fistula rats or Heidenhain pouch

    dogs (Holstein et al., 2004a, 2004c).

    In rats and dogs, revaprazan had linear pharmacoki-

    netics for oral doses of 2–30 mg/kg. In this dose range,

    oral bioavailability was 41–47% in rats and 43–59% in

    dogs (Park et al., 2003a). However, at doses above

    200 mg/kg, the AUC of revaprazan did not show dose

    proportionality, which was attributed to poor water

    solubility of the drug (Han et al., 1998). On repeated

    administration to rats and dogs, the daily pharmacokinetic

    profiles of revaprazan were similar during 7 days of

    dosing (Park et al., 2003a). A clinical study confirmed

    that revaprazan has linear pharmacokinetic characteristics

    and demonstrated little accumulation after multiple admin-

    istrations (Yu et al., 2004).

    The lipophilicity and chemical stability of the P-CAB,

    together with the ionic binding between the agent and the

    H+,K+-ATPase, facilitates an equilibrium between blood and

    the secretory canaliculi, albeit with the concentration in the

    latter being much higher.

    5.5. Pharmacodynamics of

    potassium-competitive acid blockers

    The rapid absorption of P-CABs is mirrored by a fast

    onset of acid inhibition. In animal and clinical studies,

    there was a rapid inhibition of acid secretion with BY841

    (Wurst & Hartmann, 1996; Kromer et al., 2000). A single

    dose of BY841 quickly raised intragastric pH to ¨6 in

    the pentagastrin-stimulated fistula dog and did so faster

    than omeprazole (Wurst & Hartmann, 1996). In healthy

    volunteers, a single 50 mg oral dose of BY841 increased

    intragastric pH to about 6 within 30–60 min (Wurst &

    Hartmann, 1996). For AZD0865, peak antisecretory effect

    in the rat was achieved within 2 hr after an oral dose of

    1 Amol/kg (Holstein et al., 2004d). The onset of effectwas also fast in dogs (e.g., peak antisecretory effect was

    achieved within 3 hr postdose; Holstein et al., 2004d). In

    humans, high doses of AZD0865 resulted in over 95%

    inhibition of acid secretion within 1 hr after oral dosing

    (Nilsson et al., 2005). In humans, revaprazan at single

    doses of 150 mg and above rapidly increased mean

    intragastric pH (Yu et al., 2004).

    Fig. 4. Theoretical pharmacodynamic profile of a P-CAB demonstrating that these

    of acid inhibition with subsequent, repeated doses.

    The peak levels of acid secretion inhibition with

    AZD0865 did not change over 5- and 14-day dosing

    periods in Heidenhain pouch dogs and full effect was

    achieved with the first dose (Andersson et al., 2004; Hol-

    stein et al., 2004b). Similarly, intragastric pH was similar at

    days 1 and 7 in a clinical study of revaprazan (Park et al.,

    2002; Fig. 4).

    P-CABs exhibit a classical (sigmoid) dose–response

    profile, with the magnitude and duration of effect being

    determined by dose, pKa, and plasma half-life. BY841 had a

    dose-dependent duration of action in animal studies

    (Kromer et al., 2000). In chronic fistula rats, AZD0865

    provided dose-dependent inhibition of acid secretion (ED50value for inhibition of acid output of 0.3 Amol/kg), withdoses above 10 Amol/kg almost completely inhibitinggastric acid production 24 hr after administration (Holstein

    et al., 2004a). Dose-dependent effects were also seen in

    Heidenhain pouch dogs (ED50=0.28 Amol/kg; Holstein etal., 2004b). Confirming the relationship between pharma-

    cokinetics and effect, there was a close correlation between

    maximum inhibition of acid output and the logarithm of

    Cmax for AZD0865 (EC50=130 nmol/L; Andersson et al.,

    2004). In humans, AZD0865 demonstrated a dose–effect

    relationship with a dose-dependent duration of inhibition of

    acid secretion; more than 95% inhibition was sustained for

    up to 15 hr for 0.8 and 1 mg/kg doses (EC50=100 nmol/L;

    Nilsson et al., 2005). Dose-related pharmacodynamics were

    noted in 46 healthy volunteers after single doses of

    revaprazan (Yu et al., 2004). In a multiple-dose, crossover

    study in which patients received 100, 150, and 200 mg once

    daily for 7 days, mean intragastric pH on day 7 was 1.9, 3.5,

    and 4.2, respectively (Park et al., 2002).

    5.6. Pharmacodynamic comparisons of

    potassium-competitive acid blockers with other agents

    To date, there have been only a limited number of

    published studies comparing P-CABs and other inhibitors of

    acid secretion. In the pentagastrin-stimulated fistula dog, a

    single dose of BY841 raised intragastric pH more rapidly

    than omeprazole or ranitidine and elevated it almost to

    neutrality, whereas omeprazole and ranitidine produced only

    a moderate increase in pH (Kromer et al., 2000). When

    histamine was used as a stimulus in the Ghosh-Schild rat,

    agents achieve their full effect with the first dose and provide similar levels

  • Fig. 5. Comparison of acid blockade by the P-CAB, BY841, and omeprazole, a PPI. Note acid inhibition occurring within 30 min and the higher pH with

    BY841. Reproduced from Wurst and Hartmann (1996) with permission from Yale J Biol Med.

    K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307 303

    BY841 and ranitidine displayed similar efficacy (Kromer et

    al., 2000). Revaprazan was 3 times more potent than

    omeprazole at inhibiting basal acid secretion in pylorus-

    ligated rats and chronic fistula rats (Park et al., 2003b), and

    was also more potent than omeprazole in inhibiting

    histamine-stimulated gastric acid secretion in Heidenhain

    pouch dogs. However, there was no significant difference in

    inhibitory potency between revaprazan and omeprazole on

    pentagastrin-stimulated gastric acid secretion in the lumen-

    perfused rat (Park et al., 2003b). Soraprazan was found to be

    more effective than ranitidine in raising pH within the

    mucous layer of the fistula dog (Kromer et al., 2000).

    There is only 1 published study in humans that allows

    direct comparison to be made between P-CABs and other

    therapies, but it does indicate pharmacodynamic differences

    between the compounds. BY841 100 mg twice daily

    markedly elevated intragastric pH and prolonged the

    percentage of time at which pH was �4 on the first day ofdosing (Wurst & Hartmann, 1996; Fig. 5). In contrast,

    omeprazole 20 mg once daily had only minor effects. After 7

    days of dosing, BY841 was judged to be at least comparable

    to omeprazole. The authors of the study expected the

    differences between the agents to translate into a shortening

    of healing time and a more rapid relief of symptoms. Taken

    together, this clinical study and the animal studies indicate

    that P-CABs offer more rapid and more profound elevation

    of pH than a PPI or an H2RA, although the relative effects

    appear to differ according to the model employed.

    6. Summary

    The evolution of our understanding of the biochemistry

    and physiology of gastric acid secretion has led to the

    development of therapies to inhibit gastric acid secretion.

    PPIs are currently recognized to be the most effective

    available agents for the treatment of acid-related diseases.

    They offer superior symptom control and healing rates

    compared with H2RAs in both PUD and GERD. However,

    PPIs exhibit a delayed onset of acute effect and achieve full

    effect only slowly and incrementally over several dose cycles.

    A number of alternative therapeutic strategies have been

    pursued with the objective of further improving the manage-

    ment of acid-related diseases. Of these, the agents that have

    progressed the furthest are CCK2 receptor antagonists and

    P-CABs, with representatives of both classes in clinical trials.

    CCK2 receptor antagonists are unlikely to become

    alternative therapies to H2RAs or PPIs, especially given the

    phenomenon of tolerance observed with an example of the

    class. Thus, P-CABs appear to offer the most promise of the

    newer agents, with initial studies comparing P-CABs with

    PPIs suggesting pharmacodynamic differences between the 2

    classes. P-CABs rapidly achieve therapeutic plasma levels

    and concentrate in the acidic environment of the parietal cell

    canaliculus. Once there, these compounds block gastric

    H+,K+-ATPase by a K+ competitive binding at or near the

    K+ binding site. They achieve their full effect quickly and

    provide similar acid inhibition with the first dose and

    subsequent, repeated doses. The results of comparative

    clinical studies with PPIs will help to define the place of P-

    CABs in the management of acid-related diseases.

    References

    Andersson, K., Aurell Holmberg, A., Briving, C., & Holstein, B. (2004).

    Preclinical profile of AZD0865, a novel, selective, potassium-compet-

    itive acid blocker. Gastroenterology 126, A-56.

    Andersson, K., Hallberg, H., & Holstein, B. (in press). AZD0865, a

    potassium-competitive acid blocker (P-CAB), does not affect kidney

    gastric H+,K+-ATPase in the rat. Gut.

  • K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307304

    Asano, S., Matsuda, S., Tega, Y., Shimizu, K., Sakamoto, S., & Takeguchi,

    N. (1997). Mutational analysis of putative SCH 28080 binding sites of

    the gastric H+, K+-ATPase. J Biol Chem 272, 17668–17674.

    Asano, S., Matsuda, S., Hoshina, S., Sakamoto, S., & Takeguchi, N. (1999).

    A chimeric gastric H+,K+-ATPase inhibitable with both ouabain and

    SCH 28080. J Biol Chem 274, 6848–6854.

    Asano, S., Yoshida, A., Yashiro, H., Kobayashi, Y., Morisato, A., & Ogawa,

    H., et al. (2004). The cavity structure for docking the K(+)-competitive

    inhibitors in the gastric proton pump. J Biol Chem 279, 13968–13975.

    Barocelli, E., & Ballabeni, V. (2003). Histamine in the control of gastric

    acid secretion: a topic review. Pharmacol Res 47, 299–304.

    Beil, W., Hackbarth, I., & Sewing, K. F. (1986). Mechanism of gastric

    antisecretory effect of SCH 28080. Br J Pharmacol 88, 19–23.

    Beisvag, V., Falck, G., Loennechen, J. P., Qvigstad, G., Jynge, P.,

    Skomedal, T., et al. (2003). Identification and regulation of the gastric

    H+/K+-ATPase in the rat heart. Acta Physiol Scand 179, 251–262.

    Bell, N. J., Burget, D., Howden, C. W., Wilkinson, J., & Hunt, R. H. (1992).

    Appropriate acid suppression for the management of gastro-oesopha-

    geal reflux disease. Digestion 51(Suppl 1), 59–67.

    Beltinger, J., Hildebrand, P., Drewe, J., Christ, A., Hlobil, K., Ritz, M., et al.

    (1999). Effects of spiroglumide, a gastrin receptor antagonist, on acid

    secretion in humans. Eur J Clin Invest 29, 153–159.

    Besancon, M., Shin, J. M., Mercier, F., Munson, K., Miller, M., Hersey,

    S., et al. (1993). Membrane topology and omeprazole labeling of

    the gastric H+, K(+)-adenosinetriphosphatase. Biochemistry 32,

    2345–2355.

    Besancon, M., Simon, A., Sachs, G., & Shin, J. M. (1997). Sites of reaction

    of the gastric H,K-ATPase with extracytoplasmic thiol reagents. J Biol

    Chem 272, 22438–22446.

    Black, J., & Kalindjian, S. (2002). Gastrin agonists and antagonists.

    Pharmacol Toxicol 91, 275–281.

    Briving, C., Andersson, B. M., Nordberg, P., & Wallmark, B. (1988).

    Inhibition of gastric H+/K+-ATPase by substituted imidazo[1,2-a]pyr-

    idines. Biochim Biophys Acta 946, 185–192.

    Briving, C., Svensson, K., Maxvall, I., & Andersson, K. (2004). Mechanism

    of action of AZD0865, an H+,K+-ATPase selective, potassium-

    competitive acid blocker. Gastroenterology 126, A-333 (M1436).

    Bytzer, P. (2003). Goals of therapy and guidelines for treatment success in

    symptomatic gastroesophageal reflux disease patients. Am J Gastro-

    enterol 98, S31–S39.

    Cabero, J. L., Li, Z. Q., & Mardh, S. (1993). Gastrin action on aminopyrine

    accumulation in isolated pig parietal cells requires cAMP. Biochim

    Biophys Acta 1177, 245–252.

    Cederberg, C., Lind, T., Rohss, K., & Olbe, L. (1992). Comparison of once-

    daily intravenous and oral omeprazole on pentagastrin-stimulated acid

    secretion in duodenal ulcer patients. Digestion 53, 171–178.

    Chiba, N., De Gara, C. J., Wilkinson, J. M., & Hunt, R. H. (1997). Speed of

    healing and symptom relief in grade II to IV gastroesophageal reflux

    disease: a meta-analysis. Gastroenterology 112, 1798–1810.

    Cougnon, M., Planelles, G., Crowson, M. S., Shull, G. E., Rossier, B. C., &

    Jaisser, F. (1996). The rat distal colon P-ATPase alpha subunit encodes a

    ouabain-sensitive H+, K+-ATPase. J Biol Chem 271, 7277–7280.

    Crawley, J. A., & Schmitt, C. M. (2000). How satisfied are chronic

    heartburn suffers with their prescription medications? Results of the

    Patient Unmet Needs Survey. J Clin Outcomes Manag 7, 29–34.

    Dammann, H. G., & Burkhardt, F. (1999). Pantoprazole versus omeprazole:

    influence on meal-stimulated gastric acid secretion. Eur J Gastroenterol

    Hepatol 11, 1277–1282.

    Dedek, K., & Waldegger, S. (2001). Colocalization of KCNQ1/KCNE

    channel subunits in the mouse gastrointestinal tract. Pflugers Arch 442,

    896–902.

    Duman, J. G., Pathak, N. J., Ladinsky, M. S., McDonald, K. L., & Forte, J.

    G. (2002). Three-dimensional reconstruction of cytoplasmic membrane

    networks in parietal cells. J Cell Sci 115, 1251–1258.

    Dunbar, L. A., & Caplan, M. J. (2001). Ion pumps in polarized cells: sorting

    and regulation of the Na+, K+- and H+, K+-ATPases. J Biol Chem 276,

    29617–29620.

    Eltze, M., Gonne, S., Riedel, R., Schlotke, B., Schudt, C., & Simon, W. A.

    (1985). Pharmacological evidence for selective inhibition of gastric acid

    secretion by telenzepine, a new antimuscarinic drug. Eur J Pharmacol

    112, 211–224.

    Ene, M. D., Khan-Daneshmend, T., & Roberts, C. J. (1982). A study of the

    inhibitory effects of SCH 28080 on gastric secretion in man. Br J

    Pharmacol 76, 389–391.

    Fiorucci, S., Clausi, G. G., Farinelli, M., Santucci, L., Farroni, F., Pelli,

    M. A., et al. (1988). Do anticholinergics interact with histamine H2

    receptor antagonists on night intragastric acidity in active duodenal

    ulcer patients? Am J Gastroenterol 83, 1371–1375.

    Fiorucci, S., Santucci, L., Chiucchiu, S., & Morelli, A. (1992). Gastric

    acidity and gastroesophageal reflux patterns in patients with esoph-

    agitis. Gastroenterology 103, 855–861.

    Forte, J. G., & Yao, X. (1996). The membrane-recruitment-and-recycling

    hypothesis of gastric HCl secretion. Trends Cell Biol 6, 45–48.

    Fujisaki, H., Shibata, H., Oketani, K., Murakami, M., Fujimoto, M., &

    Wakabayashi, T., et al. (1991). Inhibitions of acid secretion by E3810

    and omeprazole, and their reversal by glutathione. Biochem Pharmacol

    42, 321–328.

    Fujita, A., Horio, Y., Higashi, K., Mouri, T., Hata, F., Takeguchi, N., et al.

    (2002). Specific localization of an inwardly rectifying K(+) channel,

    Kir4.1, at the apical membrane of rat gastric parietal cells; its possible

    involvement in K(+) recycling for the H(+)-K(+)-pump. J Physiol 540,

    85–92.

    Futai, M., Oka, T., Sun-Wada, G., Moriyama, Y., Kanazawa, H., & Wada,

    Y. (2000). Luminal acidification of diverse organelles by V-ATPase in

    animal cells. J Exp Biol 203(Pt 1), 107–116.

    Gatto, C., Lutsenko, S., Shin, J. M., Sachs, G., & Kaplan, J. H. (1999).

    Stabilization of the H,K-ATPase M5M6 membrane hairpin by K+ ions.

    Mechanistic significance for p2-type ATPases. J Biol Chem 274,

    13737–13740.

    Gedda, K., Scott, D., Besancon, M., Lorentzon, P., & Sachs, G. (1995).

    Turnover of the gastric H+,K(+)-adenosine triphosphatase alpha subunit

    and its effect on inhibition of rat gastric acid secretion. Gastro-

    enterology 109, 1134–1141.

    Grahammer, F., Herling, A. W., Lang, H. J., Schmitt-Graff, A., Wittekindt,

    O. H., Nitschke, R., et al. (2001). The cardiac K+ channel KCNQ1 is

    essential for gastric acid secretion. Gastroenterology 120, 1363–1371.

    Grove, O., Bekker, C., Jeppe-Hansen, M. G., Karstoft, E., Sanchez, G.,

    Axelsson, C. K., et al. (1985). Ranitidine and high-dose antacid in

    reflux oesophagitis. A randomized, placebo-controlled trial. Scand J

    Gastroenterol 20, 457–461.

    Gustavsson, S., Adami, H. O., Loof, L., Nyberg, A., & Nyren, O. (1983).

    Rapid healing of duodenal ulcers with omeprazole: double-blind dose-

    comparative trial. Lancet 2, 124–125.

    Han, K. S., Kim, Y. G., Yoo, J. K., Lee, J. W., & Lee, M. G. (1998).

    Pharmacokinetics of a new reversible proton pump inhibitor, YH1885,

    after intravenous and oral administrations to rats and dogs: hepatic first-

    pass effect in rats. Biopharm Drug Dispos 19, 493–500.

    Hermsen, H. P., Koenderink, J. B., Swarts, H. G., & de Pont, J. J. (2000).

    The carbonyl group of glutamic acid-795 is essential for gastric H+, K+-

    ATPase activity. Biochemistry 39, 1330–1337.

    Holstein, B., Florentzson, M., Andersson, M., & Andersson, K. (2004a).

    AZD0865, a potassium-competitive acid blocker (P-CAB), has a long

    duration of effect in the rat. Am J Gastroenterol 99(Suppl 10), S8.

    Holstein, B., Holmberg, A., Florentzson, M., Aurell Holmberg, A.,

    Andersson, M., & Andersson, K. (2004b). AZD0865, a potassium-

    competitive acid blocker (P-CAB), provides predictable inhibition of

    acid secretion with repeated dosing in the dog. Eur J Pharm Sci

    23(Suppl 1), S8.

    Holstein, B., Holmberg, A., Florentzson, M., Aurell Holmberg, A.,

    Andersson, M., & Andersson, K. (2004c). AZD0865–a new potas-

    sium-competitive acid blocker–exhibits maximal gastric antisecretory

    effects from first dose. Gastroenterology 126, A-334 (M1437).

    Holstein, B., Holmberg, A., Florentzson, M., Aurell Holmberg, A.,

    Andersson, M., & Andersson, K. (2004d). Pharmacokinetic and

  • K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307 305

    pharmacodynamic profiles of AZD0865, a novel, potassium-compet-

    itive acid blocker. Gastroenterology 126, A-334 (M1440).

    Howden, C. W., Burget, D. W., & Hunt, R. H. (1994). Appropriate acid

    suppression for optimal healing of duodenal ulcer and gastro-oesopha-

    geal reflux disease. Scand J Gastroenterol Suppl 201, 79–82.

    Im, W. B., Blakeman, D. P., & Davis, J. P. (1985a). Irreversible inactivation

    of rat gastric (H+-K+)-ATPase in vivo by omeprazole. Biochem Biophys

    Res Commun 126, 78–82.

    Im, W. B., Blakeman, D. P., & Sachs, G. (1985b). Reversal of antisecretory

    activity of omeprazole by sulfhydryl compounds in isolated rabbit

    gastric glands. Biochim Biophys Acta 845, 54–59.

    Kaminski, J. J., Perkins, D. G., Frantz, J. D., Solomon, D. M., Elliott, A. J.,

    Chiu, P. J., et al. (1987). Antiulcer agents: 3. Structure–activity–

    toxicity relationships of substituted imidazo[1,2-a]pyridines and a

    related imidazo[1,2-a]pyrazine. J Med Chem 30, 2047–2051.

    Kaminski, J. J., Puchalski, C., Solomon, D. M., Rizvi, R. K., Conn, D. J.,

    Elliott, A. J., et al. (1989). Antiulcer agents: 4. Conformational

    considerations and the antiulcer activity of substituted imidazo[1,2-

    a]pyridines and related analogues. J Med Chem 32, 1686–1700.

    Keeling, D. J., Laing, S. M., & Senn-Bilfinger, J. (1988). SCH 28080 is a

    lumenally acting, K+-site inhibitor of the gastric (H++K+)-ATPase.

    Biochem Pharmacol 37, 2231–2236.

    Keeling, D. J., Taylor, A. G., & Schudt, C. (1989). The binding of a K+

    competitive ligand, 2-methyl,8-(phenylmethoxy)imidazo(1,2-a)pyridine

    3-acetonitrile, to the gastric (H++K+)-ATPase. J Biol Chem 264,

    5545–5551.

    Keeling, D. J., Malcolm, R. C., Laing, S. M., Ife, R. J., & Leach, C. A.

    (1991). SK&F 96067 is a reversible, lumenally acting inhibitor of the

    gastric (H++K+)-ATPase. Biochem Pharmacol 42, 123–130.

    Kim, K. B., Chang, M. S., Chung, Y. K., Sohn, S. K., Kim, S. G., & Choi,

    W. S. (1998). Biochemical and pharmacological characteristics of 3-

    butyryl-8-methoxy-4-[(2-thiophenyl)amino]quinoline, a new proton-

    pump inhibitor, in rabbit gastric microsomes and in rats. J Pharm

    Pharmacol 50, 521–529.

    Kim, H. J., Han, K. S., Kim, Y. G., Chung, Y. K., Chang, M. S., & Lee, M.

    G. (2000). Pharmacokinetics of a new proton pump inhibitor, YJA-

    20379-8, after intravenous and oral administration to spontaneously

    hypertensive rats and DOCA-salt-induced hypertensive rats. Biopharm

    Drug Dispos 21, 293–302.

    Kleinman, L., McIntosh, E., Ryan, M., Schmier, J., Crawley, J.,

    Locke, G. R., et al. (2002). Willingness to pay for complete

    symptom relief of gastroesophageal reflux disease. Arch Intern Med

    162, 1361–1366.

    Koenderink, J. B., Swarts, H. G., Willems, P. H., Krieger, E., & de Pont, J.

    J. (2004). A conformation-specific interhelical salt bridge in the K+

    binding site of gastric H,K-ATPase. J Biol Chem 279, 16417–16424.

    Komazawa, Y., Adachi, K., Mihara, T., Ono, M., Kawamura, A., Fujishiro,

    H., et al. (2003). Tolerance to famotidine and ranitidine treatment after

    14 days of administration in healthy subjects without Helicobacter

    pylori infection. J Gastroenterol Hepatol 18, 678–682.

    Kopin, A. S., Lee, Y. M., McBride, E. W., Miller, L. J., Lu, M., Lin, H. Y.,

    et al. (1992). Expression cloning and characterization of the canine

    parietal cell gastrin receptor. Proc Natl Acad Sci U S A 89, 3605–3609.

    Kraut, J. A., Helander, K. G., Helander, H. F., Iroezi, N. D., Marcus, E.

    A., & Sachs, G. (2001). Detection and localization of H+-K+-

    ATPase isoforms in human kidney. Am J Physiol Renal Physiol 281,

    F763–F768.

    Kromer, W., Kruger, U., Huber, R., Hartmann, M., & Steinijans, V. W.

    (1998). Differences in pH-dependent activation rates of substituted

    benzimidazoles and biological in vitro correlates. Pharmacology 56,

    57–70.

    Kromer, W., Postius, S., & Riedel, R. (2000). Animal pharmacology of

    reversible antagonism of the gastric acid pump, compared to standard

    antisecretory principles. Pharmacology 60, 179–187.

    Lachman, L., & Howden, C. W. (2000). Twenty-four-hour intragastric pH:

    tolerance within 5 days of continuous ranitidine administration. Am J

    Gastroenterol 95, 57–61.

    Lambers, C. B., Lind, T., Moberg, S., Jansen, J. B., & Olbe, L. (1984).

    Omeprazole in Zollinger–Ellison syndrome. Effects of a single dose

    and of long-term treatment in patients resistant to histamine H2-receptor

    antagonists. N Engl J Med 310, 758–761.

    Lambrecht, N., Munson, K., Vagin, O., & Sachs, G. (2000). Comparison of

    covalent with reversible inhibitor binding sites of the gastric H, K-

    ATPase by site-directed mutagenesis. J Biol Chem 275, 4041–4048.

    Lambrecht, N. W., Yakubov, I., Scott, D., & Sachs, G. (2004, Dec 21).

    Identification of the K efflux channel coupled to the gastric H,K ATPase

    during acid secretion. Physiol Genomics (electronic publication ahead of

    print: http://physiolgenomics.physiology.org/cgi/reprint/00212.2004v1

    (accessed 16 February 2005).

    Leach, C. A., Brown, T. H., Ife, R. J., Keeling, D. J., Parsons, M. E.,

    Theobald, C. J., et al. (1995). Reversible inhibitors of the gastric

    (H+/K+)-ATPase: 4. Identification of an inhibitor with an intermediate

    duration of action. J Med Chem 38, 2748–2762.

    Lindberg, P., Brändström, A., & Wallmark, B. (1987). Structure–activity

    relationships of omeprazole and their mechanism of action. Trends

    Pharmacol Sci 8, 402.

    Lindberg, P., Brandstrom, A., Wallmark, B., Mattsson, H., Rikner, L., &

    Hoffmann, K. J. (1990). Omeprazole: the first proton pump inhibitor.

    Med Res Rev 10, 1–54.

    Long, J. F., Chiu, P. J., Derelanko, M. J., & Steinberg, M. (1983). Gastric

    antisecretory and cytoprotective activities of SCH 28080. J Pharmacol

    Exp Ther 226, 114–120.

    Maeda, M., Oshiman, K., Tamura, S., & Futai, M. (1990). Human gastric

    (H++K+)-ATPase gene. Similarity to (Na++K+)-ATPase genes in

    exon/intron organization but difference in control region. J Biol Chem

    265, 9027–9032.

    Makovec, F., Revel, L., Letari, O., Mennuni, L., & Impicciatore, M. (1999).

    Characterization of antisecretory and antiulcer activity of CR 2945, a

    new potent and selective gastrin/CCK(B) receptor antagonist. Eur J

    Pharmacol 369, 81–90.

    Malen, C. E., & Danree, B. H. (1971). New thiocarboxamides

    derivatives with specific gastric antisecretory properties. J Med Chem

    14(3), 244–246.

    Malinowska, D. H., Kupert, E. Y., Bahinski, A., Sherry, A. M., &

    Cuppoletti, J. (1995). Cloning, functional expression, and character-

    ization of a PKA-activated gastric Cl� channel. Am J Physiol 268,C191–C200.

    Malinowska, D. H., Sherry, A. M., Tewari, K. P., & Cuppoletti, J.

    (2004). Gastric parietal cell secretory membrane contains PKA- and

    acid-activated Kir2.1 K+ channels. Am J Physiol Cell Physiol 286,

    C495–C506.

    Mattsson, J. P., Lorentzon, P., Wallmark, B., & Keeling, D. J. (1993).

    Characterization of proton transport in bone-derived membrane

    vesicles. Biochim Biophys Acta 1146, 106–112.

    McCabe, R. D., & Young, D. B. (1992). Evidence of a K(+)-H(+)-ATPase

    in vascular smooth muscle cells. Am J Physiol 262, H1955–H1958.

    Mendlein, J., & Sachs, G. (1990). Interaction of a K(+)-competitive

    inhibitor, a substituted imidazo[1,2a] pyridine, with the phospho- and

    dephosphoenzyme forms of H+,K(+)-ATPase. J Biol Chem 265,

    5030–5036.

    Metz, D. C., Ferron, G. M., Paul, J., Turner, M. B., Soffer, E., Pisegna,

    J. R., et al. (2002). Proton pump activation in stimulated parietal

    cells is regulated by gastric acid secretory capacity: a human study.

    J Clin Pharmacol 42, 512–519.

    Miura, N., Yoneta, T., Ukawa, H., Fukuda, Y., Eta, R., Mera, Y., et al.

    (2001). Pharmacological profiles of Z-360, a novel CCK/gastrin

    (CCK2) receptor antagonist with excellent oral potency. Gastroenterol-

    ogy 120, A311.

    Morita, H., Miura, N., Hori, Y., Matsunaga, Y., Ukawa, H., Suda, H., et al.

    (2001). Effects of Z-360, a novel CCKB/gastrin (CCK2) receptor

    antagonist, on meal-induced acid secretion and experimental ulcer

    models in dogs and rats. Gastroenterology 120, A311.

    Munson, K. B., Gutierrez, C., Balaji, V. N., Ramnarayan, K., & Sachs, G.

    (1991). Identification of an extracytoplasmic region of H+,K(+)-ATPase

    http://physiolgenomics.physiology.org/cgi/reprint/00212.2004v1http://physiolgenomics.physiology.org/cgi/reprint/00212.2004v1

  • K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307306

    labeled by a K(+)-competitive photoaffinity inhibitor. J Biol Chem 266,

    18976–18988.

    Munson, K., Lambrecht, N., Shin, J. M., & Sachs, G. (2000). Analysis of

    the membrane domain of the gastric H(+)/K(+)-ATPase. J Exp Biol

    203(Pt 1), 161–170.

    Nilsson, C., Albrektson, E., Rydholm, H., Rohss, K., Hassan-Alin, M., &

    Hasselgren, G. (2005). Tolerability, pharmacokinetics and effects on

    gastric acid secretion after single oral doses of the potassium-

    competitive acid blocker AZD0865 in healthy male subjects. Gastro-

    enterology 128(4 Suppl. 2), A528.

    Nishida, A., Takinami, Y., Yuki, H., Kobayashi, A., Akuzawa, S., Kamato,

    T., et al. (1994). YM022 [(R)-1-[2,3-dihydro-1-(2V-methylphenacyl)-2-oxo-5-phenyl-1H-1,4-benzodiazepin-3-yl]-3-(3-methylphenyl)urea], a

    potent and selective gastrin/cholecystokinin-B receptor antagonist,

    prevents gastric and duodenal lesions in rats. J Pharmacol Exp Ther

    270, 1256–1261.

    Park, S., Song, K., Moon, B. S., Joo, S., Choi, M., & Chung, I., et al.

    (2002). The effect of 7 days of dosing with YH1885 on 24-hour

    intragastric acidity and plasma gastrin concentration in healthy male

    subjects. Gut 51(Suppl 3) (Abstr. TUE-G-312).

    Park, S., Ahn, B., Lee, B., Kang, H., & Song, K. (2003a). Pharmacokinetics

    of YH1885. Gut 52(Suppl VI).

    Park, S., Lee, S., Song, K., Lee, B., Kang, H., & Kang, J. (2003b). The

    pharmacological properties of a novel acid pump antagonist, YH1885.

    Gut 52(Suppl VI).

    Parsons, M. E., Rushant, B., Rasmussen, T. C., Leach, C., Ife, R. J., Postius,

    S., et al. (1995). Properties of the reversible, K(+)-competitive inhibitor

    of the gastric (H+/K+)-ATPase, SK&F 97574: II. Pharmacological

    properties. Biochem Pharmacol 50, 1551–1556.

    Pendley, C. E., Fitzpatrick, L. R., Capolino, A. J., Davis, M. A., Esterline,

    N. J., Jakubowska, A., et al. (1995). RP 73870, a gastrin/cholecysto-

    kinin-B receptor antagonist with potent anti-ulcer activity in the rat.

    J Pharmacol Exp Ther 273, 1015–1022.

    Pope, A. J., & Parsons, M. E. (1993). Reversible inhibitors of the gastric

    H+/K(+)-transporting ATPase: a new class of anti-secretory agent.

    Trends Pharmacol Sci 14, 323–325.

    Pope, A. J., & Sachs, G. (1992). Reversible inhibitors of the gastric

    (H+/K+)-ATPase as both potential therapeutic agents and probes of

    pump function. Biochem Soc Trans 20, 566–572.

    Pope, A. J., Boehm, M. K., Leach, C., Ife, R. J., Keeling, D., & Parsons, M.

    E. (1995). Properties of the reversible K(+)-competitive inhibitor of the

    gastric (H+/K+)-ATPase, SK&F 97574: I. In vitro activity. Biochem

    Pharmacol 50, 1543–1549.

    Prinz, C., Kajimura, M., Scott, D. R., Mercier, F., Helander, H. F., & Sachs,

    G. (1993). Histamine secretion from rat enterochromaffinlike cells.

    Gastroenterology 105, 449–461.

    Prinz, C., Scott, D. R., Hurwitz, D., Helander, H. F., & Sachs, G. (1994).

    Gastrin effects on isolated rat enterochromaffin-like cells in primary

    culture. Am J Physiol 267, G663–G675.

    Rabon, E. C., & Reuben, M. A. (1990). The mechanism and structure of the

    gastric H,K-ATPase. Annu Rev Physiol 52, 321–344.

    Rabon, E. C., Smillie, K., Seru, V., & Rabon, R. (1993). Rubidium

    occlusion within tryptic peptides of the H,K-ATPase. J Biol Chem 268,

    8012–8018.

    Rajendran, V. M., Singh, S. K., Geibel, J., & Binder, H. J. (1998).

    Differential localization of colonic H(+)-K(+)-ATPase isoforms in

    surface and crypt cells. Am J Physiol 274, G424–G429.

    Reenstra, W. W., & Forte, J. G. (1990). Characterization of K+ and Cl�conductances in apical membrane vesicles from stimulated rabbit

    oxyntic cells. Am J Physiol 259, G850–G858.

    Ritter, M., Schratzberger, P., Rossmann, H., Woll, E., Seiler, K., Seidler,

    U., et al. (1998). Effect of inhibitors of Na+/H+-exchange and

    gastric H+/K+ ATPase on cell volume, intracellular pH and

    migration of human polymorphonuclear leucocytes. Br J Pharmacol

    124, 627–638.

    Robinson, M., & Shaw, K. (2002). Proton pump inhibitor attitudes and

    usage: a patient survey. Pharm Ther J 27, 202–206.

    Rottapharm. (2004). CR2945 (Itriglumide) Product profile. Available at:

    http://www.rotta.com/cr2945.htm. Accessed December 17, 2004.

    Sachs, G. (1994). The gastric H+,K+-ATPase. In L. R. Johnson (Ed.),

    Physiology of the Gastrointestinal Tract (pp. 1119–1138). New York’

    Raven Press.

    Sachs, G., Shin, J. M., Briving, C., Wallmark, B., & Hersey, S. (1995). The

    pharmacology of the gastric acid pump: the H+,K+ ATPase. Annu Rev

    Pharmacol Toxicol 35, 277–305.

    Salas, M., Ward, A., & Caro, J. (2002). Are proton pump inhibitors the first

    choice for acute treatment of gastric ulcers? A meta analysis of

    randomized clinical trials. BMC Gastroenterol 2, 17.

    Sangan, P., Rajendran, V. M., Mann, A. S., Kashgarian, M., & Binder, H. J.

    (1997). Regulation of colonic H-K-ATPase in large intestine and kidney

    by dietary Na depletion and dietary K depletion. Am J Physiol 272,

    C685–C696.

    Scott, C. K., Sundell, E., & Castrovilly, L. (1987). Studies on the mechanism

    of action of the gastric microsomal (H++K+)-ATPase inhibitors SCH

    32651 and SCH 28080. Biochem Pharmacol 36, 97–104.

    Selby, N. M., Kubba, A. K., & Hawkey, C. J. (2000). Acid suppression in

    peptic ulcer haemorrhage: a Fmeta-analysis_. Aliment Pharmacol Ther

    14, 1119–1126.

    Sohn, S. K., Chang, M. S., Choi, W. s., Chung, Y. K., Kim, K. B., Woo, T.

    W., et al. (1999). Biochemical and pharmacological characteristics of a

    newly synthesized H+/K+-ATPase inhibitor, YJA20379-8, 3-butyryl-4-

    [R-1-methylbenzylamino]-8-ethoxy-1,7-naphthyridine, in pigs and rats.

    J Pharm Pharmacol 51, 1359–1365.

    Steel, K. (2002). Gastrin and gastrin receptor ligands—a review of recent

    patent literature. IDrugs 5, 689–695.

    Swarts, H. G., Hermsen, H. P., Koenderink, J. B., Schuurmans Stekhoven,

    F. M., & de Pont, J. J. (1998). Constitutive activation of gastric H+,K+-

    ATPase by a single mutation. EMBO J 17, 3029–3035.

    Takeuchi, K., Hirata, T., Yamamoto, H., Kunikata, T., Ishikawa, M., &

    Ishihara, Y. (1999). Effects of S-0509, a novel CCKB/gastrin receptor

    antagonist, on acid secretion and experimental duodenal ulcers in rats.

    Aliment Pharmacol Ther 13, 87–96.

    Tefera, S., Hatlebakk, J. G., & Berstad, A. (2001). Stability of gastric

    secretory inhibition during 6-month treatment with omeprazole in

    patients with gastroesophageal reflux disease. Am J Gastroenterol 96,

    969–974.

    Tsukimi, Y., Ushiro, T., Yamazaki, T., Ishikawa, H., Hirase, J.,

    Narita, M., et al. (2000). Studies on the mechanism of action of

    the gastric H+,K(+)-ATPase inhibitor SPI-447. Jpn J Pharmacol 82,

    21–28.

    Ushiro, T., Tsukimi, Y., & Tanaka, H. (1997). Inhibition of gastric

    H+, K+-ATPase by 3-amino-5-methyl-2-(2-methyl-3-thienyl)imi-

    dazo[1,2-a]thieno[3,2-c]pyridine, SPI-447. Jpn J Pharmacol 75,

    303–306.

    Vagin, O., Munson, K., Lambrecht, N., Karlish, S. J., & Sachs, G. (2001).

    Mutational analysis of the K+-competitive inhibitor site of gastric H,K-

    ATPase. Biochemistry 40, 7480–7490.

    Vagin, O., Denevich, S., Munson, K., & Sachs, G. (2002). SCH28080, a

    K+-competitive inhibitor of the gastric H,K-ATPase, binds near the M5-

    6 luminal loop, preventing K+ access to the ion binding domain.

    Biochemistry 41, 12755–12762.

    Vagin, O., Munson, K., Denevich, S., & Sachs, G. (2003). Inhibition

    kinetics of the gastric H,K-ATPase by K-competitive inhibitor

    SCH28080 as a tool for investigating the luminal ion pathway. Ann N

    Y Acad Sci 986, 111–115.

    Vakil, N. (2004). Review article: new pharmacological agents for the

    treatment of gastro-oesophageal reflux disease. Aliment Pharmacol

    Ther 19, 1–9.

    Vander Stricht, D. V., Raussens, V., Oberg, K. A., Ruysschaert, J. M.,

    & Goormaghtigh, E. (2001). Difference between the E1 and E2

    conformations of gastric H+/K+-ATPase in a multilamellar lipid

    film system. Characterization by fluorescence and ATR-FTIR

    spectroscopy under a continuous buffer flow. Eur J Biochem 268,

    2873–2880.

    http://www.rotta.com/cr2945.htm

  • K. Andersson, E. Carlsson / Pharmacology & Therapeutics 108 (2005) 294–307 307

    van Driel, I. R., & Callaghan, J. M. (1995). Proton and potassium transport

    by H+/K(+)-ATPases. Clin Exp Pharmacol Physiol 22, 952–960.

    van Pinxteren, B., Numans, M. E., Bonis, P. A., & Lau, J. (2001). Short-

    term treatment with proton pump inhibitors, H2-receptor antagonists

    and prokinetics for gastro-oesophageal reflux disease-like symptoms

    and endoscopy negative reflux disease. Cochrane Database Syst Rev

    (CD002095).

    van Pinxteren, B., Numans, M. E., Lau, J., de Wit, N. J., Hungin, A. P., &

    Bonis, P. A. (2003). Short-term treatment of gastroesophageal reflux

    disease. J Gen Intern Med 18, 755–763.

    Wallmark, B., Briving, C., Fryklund, J., Munson, K., Jackson, R.,

    Mendlein, J., et al. (1987). Inhibition of gastric H+,K+-ATPase and

    acid secretion by SCH 28080, a substituted pyridyl(1,2a)imidazole.

    J Biol Chem 262, 2077–20784.

    Watts, J. A., Watts, A., & Middleton, D. A. (2001). A model of

    reversible inhibitors in the gastric H+/K+-ATPase binding site

    determined by rotational echo double resonance NMR. J Biol Chem

    276, 43197–43204.

    Wurst, W., & Hartmann, M. (1996). Current status of acid pump antagonists

    (reversible PPIs). Yale J Biol Med 69, 233–243.

    Yan, D., Hu, Y. D., Li, S., & Cheng, M. S. (2004). A model of 3D-structure

    of H+, K+-ATPase catalytic subunit derived by homology modeling.

    Acta Pharmacol Sin 25, 474–479.

    Yu, K. S., Bae, K. S., Shon, J. H., Cho, J. Y., Yi, S. Y., Chung, J. Y., et al.

    (2004). Pharmacokinetic and pharmacodynamic evaluation of a novel

    proton pump inhibitor, YH1885, in healthy volunteers. J Clin

    Pharmacol 44, 73–82.

  • 本文献由“学霸图书馆-文献云下载”收集自网络,仅供学习交流使用。

    学霸图书馆(www.xuebalib.com)是一个“整合众多图书馆数据库资源,

    提供一站式文献检索和下载服务”的24 小时在线不限IP

    图书馆。

    图书馆致力于便利、促进学习与科研,提供最强文献下载服务。

    图书馆导航:

    图书馆首页 文献云下载 图书馆入口 外文数据库大全 疑难文献辅助工具

    http://www.xuebalib.com/cloud/http://www.xuebalib.com/http://www.xuebalib.com/cloud/http://www.xuebalib.com/http://www.xuebalib.com/vip.htmlhttp://www.xuebalib.com/db.phphttp://www.xuebalib.com/zixun/2014-08-15/44.htmlhttp://www.xuebalib.com/

    Potassium-competitive acid blockade: A new therapeutic strategy in acid-related diseasesIntroductionPhysiology of acid secretionStructure and properties of gastric H+,K+-ATPaseTargeting gastric acid secretionH2 receptor antagonists and H3 receptor agonistsMuscarinics and cholecystokinin2 receptor antagonistsProton pump inhibitors

    Potassium-competitive acid blockersDevelopment of the potassium-competitive acid blocker classMechanism of potassium-competitive acid blocker inhibition of gastric H+,K+-ATPaseSelectivity of potassium-competitive acid blockers for gastric H+,K+-ATPasePharmacokinetics of potassium-competitive acid blockersPharmacodynamics of potassium-competitive acid blockersPharmacodynamic comparisons of potassium-competitive acid blockers with other agents

    SummaryReferences