You are on page 1of 12

ARTICLE IN PRESS

Journal of Thermal Biology 31 (2006) 208–219


www.elsevier.com/locate/jtherbio

Review

Cyclooxygenase: Past, present and future. A tribute to John R. Vane


(1927–2004)
Regina M. Botting
The William Harvey Research Institute, The John Vane Science Centre, St. Bartholomew’s and the London School of Medicine and Dentistry,
Queen Mary, University of London, Charterhouse Square, London EC1M 6BQ, UK
Accepted 22 November 2005

Abstract

Study of the prostaglandins led directly to the elucidation of the mode of action of aspirin which inhibits their synthesis. John Vane
elegantly demonstrated in 1971 that non-steroid anti-inflammatory drugs (NSAIDs) blocked cyclooxygenase (COX), the enzyme which
makes prostaglandins. In 1991, Daniel Simmons described the gene which expresses a second cyclooxygenase, COX-2. This discovery
explained the anti-inflammatory actions of NSAIDs, inhibition of COX-2, and their side actions, inhibition of COX-1. Within 8 years
selective COX-2 inhibitors became available for the treatment of inflammation without the disadvantage of gastric toxicity. Recently, a
COX-1 variant protein, named COX-3, sensitive to inhibition with acetaminophen, was characterised, cloned and expressed.
r 2005 Elsevier Ltd. All rights reserved.

Keywords: Prostaglandin; Cyclooxygenase-1; Cyclooxygenase-2; Cyclooxygenase-3; Non-steroid anti-inflammatory drugs; Rofecoxib; Celecoxib;
Lumiracoxib; Valdecoxib; Etoricoxib; Acetaminophen

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
2. Cyclooxygenase-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
3. Cyclooxygenase-2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
4. Functions of COX-1 and COX-2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
4.1. Cardiovascular system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
4.2. Gastrointestinal tract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
4.3. Kidney . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
4.4. Central nervous system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
4.5. Reproduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
5. Inhibitors of COX-1 and COX-2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
5.1. The non-steroid anti-inflammatory drugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
5.2. Selective COX-2 inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
6. Other cyclooxygenases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
7. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215

Fax: +44 208 677 9671.


E-mail address: r.m.botting@qmul.ac.uk.

0306-4565/$ - see front matter r 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jtherbio.2005.11.008
ARTICLE IN PRESS
R.M. Botting / Journal of Thermal Biology 31 (2006) 208–219 209

1. Introduction

Cyclooxygenase (COX) or prostaglandin endoperoxide


synthase (PGHS) is the key enzyme in the synthesis of
prostaglandins from their precursor, arachidonic acid.
Prostaglandins are lipid mediators made by most cells in
the body except for red blood cells and released by almost
any type of chemical or mechanical stimulus. They produce
a remarkably broad spectrum of effects that embraces
practically every biological function as well as being of
prime importance as mediators of pain, fever and swelling
in inflammation. Inhibition of their biosynthesis by the
non-steroid anti-inflammatory drugs (NSAIDs) causes
major changes to the pathophysiological functions of the
organism.
More than 70 years ago, two American gynaecologists,
Kurzrok and Lieb (1930) observed that strips of human
uterus either relaxed or contracted when exposed to human
semen. Some years later, von Euler in Sweden reported the
presence of a substance in seminal fluid that contracted
smooth muscle and lowered blood pressure in experimental
animals. He identified the active material as a lipid-soluble
acid and named it ‘‘prostaglandin’’ as it appeared to
originate from the prostate gland (von Euler, 1936). He
was wrong in this assumption since most of the prosta-
glandins are made in the seminal vesicles. Technical
advances in the 1960s allowed the demonstration that
‘‘prostaglandins’’ were a family of lipid compounds of Fig. 1. The arachidonic acid cascade. The precursor of the prostaglandins,
unique structure. More than 100 related substances derived arachidonic acid, is cleaved from cell membrane phospholipids by
from arachidonic acid and synthesised by COX have now phospholipase A2. It is then converted by COX-1 or COX-2 into unstable
been identified. endoperoxides, PGG2 and PGH2. The endoperoxide, PGH2 is metabolised
Prostaglandin E1 (PGE1) and PGF1a were isolated and by synthases to primary prostaglandins PGD2, PGE2 and PGF2a as well as
TXA2 and PGI2 (prostacyclin). TXA2 is found mostly in platelets and
their structures elucidated in 1962 (Bergström et al., 1962) PGI2 mainly in endothelial cells and in the stomach mucosa. PGE2 is made
and in 1964, PGE2 was synthesised from arachidonic acid in the kidney and at sites of inflammation. TXA2 is metabolised to the
using an enzyme preparation from sheep seminal vesicles inactive TXB2 and PGI2 is converted into the inactive metabolite, 6-keto-
(Bergström et al., 1964). Prostaglandins of the D, E and Fa PGF1a. (Reprinted with permission from ‘‘Principles of Immunopharma-
series are referred to as the ‘‘primary prostaglandins’’ and cology’’, Nijkamp, F.P. and Parnham, M.J. eds. 2005. Birkhäuser Verlag,
Basel, Switzerland.)
they are synthesised in a stepwise manner by a complex
system of microsomal enzymes beginning with COX. Thus,
the precursor, arachidonic acid, is first cyclised to form the
endoperoxide derivative, prostaglandin G2 and then the enzyme, prostacyclin synthase. PGI2 has a double-ring
reduced by peroxidase to form prostaglandin H2. These structure, closed by an oxygen bridge between carbons 6
two endoperoxides, which are chemically unstable, are then and 9 (Fig. 1). It is hydrolysed non-enzymatically to a
isomerised enzymatically or non-enzymatically into differ- much less active, stable compound, 6-keto-PGF1a.
ent products, such as PGD2, PGE2 or PGF2a. The presence of the different prostaglandin synthases
The endoperoxide, PGH2, is also metabolised into two varies from tissue to tissue. For example, lung and spleen
unstable and highly biologically active compounds with tissues are able to synthesise the whole range of products
structures that differ from those of the primary prosta- but other tissues cannot; hence, platelets synthesise mainly
glandins. One of these is thromboxane A2 (TXA2), formed TXA2, whereas the blood vessel walls primarily produce
by an enzyme, thromboxane synthase, first isolated from PGI2 (Bunting et al., 1983). PGE2 is an important mediator
platelets (Needleman et al., 1976). TXA2 has a very short in the kidney and in inflammation, whereas PGD2 is made
chemical half-life of about 30 s; it breaks down non- in mast cells and in the brain.
enzymatically into the stable and relatively inactive
thromboxane B2 (TXB2). 2. Cyclooxygenase-1
The other route of metabolism of PGH2 is to prostacy-
clin or prostaglandin I2 (PGI2), yet another unstable In 1976, COX-1 with a molecular mass of 71 kDa was
compound with a half-life of around 3 min, formed by isolated from sheep seminal vesicles (Hemler and Lands,
ARTICLE IN PRESS
210 R.M. Botting / Journal of Thermal Biology 31 (2006) 208–219

1976) and it was cloned by three separate groups in 1988


(DeWitt and Smith, 1988; Merlie et al., 1988; Yokoyama
et al., 1988). This enzyme exhibits both COX and
hydroperoxidase activities and is derived from an mRNA
of 2.8 kb. Garavito and his colleagues (Picot et al., 1994)
determined the three-dimensional structure of COX-1,
which consists of three independent folding units: an
epidermal growth factor-like domain, a membrane-binding
section and an enzymatic domain (Fig. 2). The sites for
COX and peroxidase activity are adjacent but spatially
distinct. The enzyme integrates into only a single leaflet of
the membrane lipid bilayer and thus the position of the
COX channel allows arachidonic acid to gain access to the
active site from the interior of the bilayer. The COX active
site is a long, hydrophobic channel and Garavito et al.
Fig. 3. The side effects of NSAIDs are caused by inhibition of the
(Malkowski et al., 2000) provide evidence to suggest that constitutive enzyme, COX-1, which synthesises prostaglandins that serve
some of the NSAIDs such as flurbiprofen inhibit COX-1 by essential physiological functions such as causing appropriate platelet
excluding arachidonate from the upper portion of the aggregation, protection of the gastric mucosa, inhibition of thrombogen-
channel. Binding sites for arachidonic acid and NSAID, esis and maintenance of renal function. The therapeutic effects of NSAIDs
are due to inhibition of COX-2, an enzyme induced by inflammatory
tyrosine 385 and serine 530, are positioned at the apex of
factors released by bacteria, the vascular endothelium or other cells
the long active site whereas arginine 120 lies close to the involved in the inflammatory response. (Reprinted with permission from
opening of the COX channel. There may be other sub-sites ‘‘Principles of Immunopharmacology’’, Nijkamp, F.P. and Parnham, M.J.
for binding of the precursor in this narrow channel. eds. 2005. Birkhäuser Verlag, Basel, Switzerland.)
COX-1 is the constitutive isoform of COX which has
clear physiological functions. It performs a ‘‘housekeep-
ing’’ function to synthesise prostaglandins which regulate largely remains stable, but small (2- to 4-fold) increases in
normal cell activity (Fig. 3). Its activation leads, for expression can occur in response to stimulation with
instance, to the production of prostacyclin which, when hormones or growth factors (DeWitt, 1991; Wu et al.,
released by the endothelium, is anti-thrombogenic (Mon- 1991).
cada et al., 1976) and when released by the gastric mucosa,
it is cytoprotective (Whittle et al., 1978). It is also COX-1 in 3. Cyclooxygenase-2
platelets that leads to TXA2 production, causing aggrega-
tion of the platelets to prevent inappropriate bleeding There were various clues in the literature suggesting that
(Funk et al., 1991). The concentration of the enzyme there may be a second COX enzyme. As early as 1972,
Smith and Lands (1972) and Flower and Vane (1972)
speculated on the existence of isoenzymes. Others (Lysz
and Needleman, 1982; Lysz et al., 1988) also suggested two
distinct forms of COX.
Rosen et al. (1989) were studying the regulation of COX
in cultures of epithelial cells from sheep trachea and found
an increase in activity of COX during prolonged cell
culture. The increase in activity was not accounted for by a
coordinate increase in 70 kDa COX protein, and a
corresponding increase in the mRNA of 2.8 kb. However,
they found an increase in a second mRNA of 4.0 kb and
suggested that this was derived from a distinct COX-
related gene which encoded for a protein with COX
activity.
Needleman and his group (Raz et al., 1989; Fu et al.,
1990; Masferrer et al., 1990) reported that bacterial
lipopolysaccharide (LPS) increased the synthesis of pros-
Fig. 2. COX-1 and COX-2 catalytic channels. The hatched areas in COX-1 taglandins in human monocytes in vitro and in mouse
are those that are more accessible in COX-2 due to amino acid peritoneal macrophages in vivo. This increase, but not the
substitutions. Tyr 385 and Arg 120 function as binding sites for basal level of enzyme, was inhibited by dexamethasone and
arachidonic acid and NSAIDs, while Ser 530 is acetylated irreversibly
by aspirin. SP ¼ side pocket; ES ¼ extra space. (Reprinted with permis-
associated with de novo synthesis of new COX protein.
sion from ‘‘Principles of Immunopharmacology’’, Nijkamp, F.P. and This further gave rise to the concept of ‘‘constitutive’’ and
Parnham, M.J. eds. 2005. Birkhäuser Verlag, Basel, Switzerland.) ‘‘inducible’’ forms of COX.
ARTICLE IN PRESS
R.M. Botting / Journal of Thermal Biology 31 (2006) 208–219 211

The breakthrough came from biologists outside the field relaxes vascular smooth muscle resulting in vasodilatation.
of prostaglandins. Simmons and his colleagues were Both effects are brought about through activation of a Gs
studying immediate early genes and discovered a gene protein, stimulation of adenylate cyclase and accumulation
encoding a second form of COX, induced by v-src, serum of cyclic AMP (Tateson et al., 1977; Oliva and Nicosia,
or phorbol esters in chicken embryo fibroblasts (Simmons 1987). Prostacyclin inhibits platelet aggregation (platelet–
et al., 1989; Xie et al., 1991). It was encoded by a 4.1 kb platelet interaction) at much lower concentrations than
mRNA, similar in size to that reported by Rosen et al. those needed to inhibit adhesion (platelet–collagen inter-
(1989). They cloned the gene, deduced the structure of the action) (Higgs et al., 1978). Thus, prostacyclin may permit
enzyme protein and found it homologous to COX but to platelets to stick to and interact with damaged vascular
no other known protein. Some months later, Herschmann tissue, so allowing platelets to participate in the repair of a
and his colleagues (Kujubu et al., 1991) found a similar vessel wall while at the same time preventing or limiting
gene in the mouse, as did Simmons et al. (1991). Others thrombus formation.
confirmed a distinct isoform of COX, COX-2 (O’Banion Immunocytochemical studies suggest that expression of
et al., 1991; Sirois and Richards, 1992). COX-2 in normal blood vessels is very low (Crofford et al.,
Both COX-1 and COX-2 have a molecular weight of 1994; Schonbeck et al., 1999). However, shear stress
71 kDa and the amino acid sequence of COX-2 shows a induced by laminar but not by turbulent flow upregulates
60% homology with the sequence of the non-inducible COX-2 in vascular endothelial cell cultures (Topper et al.,
enzyme. The mRNA for the inducible enzyme approx- 1996). In fact, studies in human volunteers given selective
imates 4.5 kb and that of the constitutive enzyme, 2.8 kb. COX-2 inhibitors show, by estimating the prostacyclin
The inhibition by the glucocorticoids of the expression of metabolite (2,3-dinor-6-keto-PGF1a) in urine, that most of
COX-2 is an additional aspect of the anti-inflammatory the prostacyclin produced by the endothelium derives from
action of the corticosteroids. The levels of COX-2, induced COX-2 (Catella–Lawson et al., 1999; McAdam
normally very low in cells, are tightly controlled by a et al., 1999). However, high levels of COX-2 can be
number of factors including cytokines, intracellular mes- detected in inflamed tissues, angiogenic microvessels and
sengers and by the availability of substrate. Thus, increased atherosclerotic plaques (Sano et al., 1992).
synthesis of prostaglandins is achieved in inflammation
due to up-regulation of COX-2 by inflammatory stimuli 4.2. Gastrointestinal tract
(Fig. 3).
The three-dimensional structure of COX-2 has also been In most species, including humans, the bulk of the
published (Luong et al., 1996). It closely resembles the gastroprotective prostaglandins is synthesised by COX-1
structure of COX-1, except that the COX-2 active site is although small amounts of COX-2 can also be found
slightly larger and can accommodate bigger structures than expressed constitutively in the normal rat stomach (Karg-
those which are able to fit into the active site of COX-1. A man et al., 1996). The so-called ‘‘cytoprotective’’ action of
secondary internal side pocket of COX-2 contributes the prostaglandins in preventing gastric erosions and
significantly to the larger volume of the active site in this ulceration is mainly brought about by endogenously
enzyme (Fig. 2), although the central channel is also wider produced prostacyclin and PGE2. Their synthesis can be
by approximately 17%. demonstrated to occur in every part of the gastrointestinal
tract with the highest concentrations being produced in
4. Functions of COX-1 and COX-2 gastric smooth muscle and gastric mucosa (Whittle and
Salmon, 1983). This ‘‘cytoprotective’’ action in preventing
4.1. Cardiovascular system gastric damage is complex but depends on the reduction in
secretion of gastric acid (Robert et al., 1967), increase in
Blood platelets contain only COX-1, which converts gastric mucosal blood flow through vasodilatation of
arachidonic acid to the potent aggregatory and vasocon- mucosal blood vessels (Whittle et al., 1978) and increased
strictor eicosanoid TXA2, the major COX product formed secretion of viscous mucus (Johansson and Kollberg, 1979)
by platelets. TXA2 has a half-life at body pH and and of duodenal bicarbonate thus effectively neutralising
temperature of 30 s, degrading to inactive TXB2 (Needle- stomach acid production (Allen and Garner, 1980).
man et al., 1976). Prostacyclin generated in the vessel wall Surprisingly, transgenic mice with a deleted COX-1 gene
may be the physiological antagonist of this system in did not spontaneously develop stomach ulcers (Langen-
platelets. Normal vascular endothelial and smooth muscle bach et al., 1995). The reason for this unexpected finding
cells express COX-1 which produces mainly prostacyclin. became clear when rats were treated with a combination of
According to this concept, prostacyclin and TXA2 repre- a selective COX-1 and a selective COX-2 inhibitor and
sent biologically opposite poles of a mechanism for rapidly developed gastric damage (Wallace et al., 2000).
regulating platelet–vessel wall interaction and the forma- Thus, inactivation of both COX-1 and COX-2 results in
tion of haemostatic plugs and intra-arterial thrombi the development of gastrointestinal damage but not
(Moncada and Vane, 1982). By stimulating IP receptors, inhibition of activity of either enzyme alone (Gretzer
prostacyclin potently inhibits aggregation of platelets and et al., 2001). This result was confirmed in transgenic
ARTICLE IN PRESS
212 R.M. Botting / Journal of Thermal Biology 31 (2006) 208–219

COX-1 gene-deficient mice, in which selective inhibition of COX-2 mRNA in neurones of the hippocampus (March-
COX-2 caused small bowel ulcers (Sigthorsson et al., 2002). eselli and Bazan, 1996) and acute stress raises levels in the
Thus, the lack of spontaneous intestinal damage in COX-1 cerebral cortex (Yamagata et al., 1993). Constitutive
knockout mice may be explained by the upregulation of expression of COX-2 mRNA which may be involved with
COX-2 protein when COX-1 is absent (Tanaka et al., nociceptive processing is found in rat spinal cord (Beiche
2002). et al., 1996; Yaksh and Svensson, 2001).
Although very little COX-2 is found in healthy intestinal There is good evidence that COX-2 induced in endothe-
tissues, COX-2 is highly expressed in human and animal lial cells of hypothalamic blood vessels (Cao et al., 1995,
colon cancer cells as well as in human colorectal 1996, 1998; Matsumura et al., 1998) by circulating LPS or
adenocarcinomas (Thun et al., 1991; Luk, 1996). A clinical IL-1 (Ek et al., 2001) is involved in fever generation. Fever
trial of celecoxib in patients with familial adenomatous is produced by PGE2 which enters the organum vasculo-
polyposis showed a 30% reduction in intestinal polyps sum laminae terminalis (OVLT) and activates the thermo-
(Steinbach et al., 2000) and this indication for the drug has regulatory centre (Blatteis and Sehic, 1997). The febrigenic
been allowed by the FDA. Human gastric and breast response to LPS or IL-1 is abolished in COX-2 knockout
tumours also express higher levels of COX-2 protein than mice (Li et al., 1999, 2001; Steiner et al., 2005) suggesting
surrounding normal tissues (Ristimäki et al., 1997; Parrett that the PGE2 which causes fever originates from the COX-
et al., 1997). These tumours may also be susceptible to 2 gene. Microsomal prostaglandin E synthase-1, expressed
treatment with selective COX-2 inhibitors. in endothelial cells, also plays an important role in the
development of fever. Mice deficient in this enzyme showed
4.3. Kidney no fever or PGE2 synthesis in the central nervous system in
response to injection of LPS. However, these mice
The normal kidney cortex produces mainly PGE2 and responded with a fever to an intracerebral injection of
prostacyclin as well as small amounts of TXA2 (Farman et PGE2 (Engblom et al., 2003).
al., 1987) while the renal medulla produces mostly PGE2
for which it has a synthetic capacity approximately 20 4.5. Reproduction
times that of the cortex (Zusman and Keiser, 1977). Renal
PGE2 is mostly synthesised by COX-1, however mesangial Both COX-1 and COX-2 are expressed in the pregnant
cells in the macula densa contain constitutive COX-2 which uterus, fetal membranes and umbilical cord. COX-2
synthesises prostacyclin that may directly stimulate renin mRNA increases in the human amnion and placenta
secretion (Harris et al., 1994; Harris, 1996). immediately before and after the start of labour (Gibb and
Production of prostaglandins is not essential for the Sun, 1996) and preterm labour can be delayed by
maintenance of normal kidney function. However, vasodi- administration of selective COX-2 inhibitors (Sawdy
lator prostaglandins are important in patients with kidneys et al., 1997). Female COX-2 knockout mice are mostly
compromised by disease, for example congestive heart infertile, producing very few offspring due to a reduction in
failure, liver cirrhosis or renal insufficiency and in animal ovulation (Lim et al., 1997). In addition, the kidneys of
models of disease states. Inhibition of COX by NSAIDs these mice do not develop fully after birth which reduces
may result in renal ischaemia when prostaglandin synthesis their life span to between 8 and 16 weeks (Dinchuk et al.,
is reduced. 1995; Morham et al., 1995). Kidney maturation also ceased
prematurely when selective COX-2 inhibitors were fed
4.4. Central nervous system chronically to normal mice during pregnancy (Kömhoff
et al., 2000). However, mice of a mixed strain did not
Neurones throughout the brain express COX-1, but it is exhibit this severe kidney pathology (Laulederkind et al.,
most abundant in the forebrain (Yamagata et al., 1993; 2002) and lived a full life span.
Breder et al., 1995) where it may be involved in sensory The mouse uterine epithelium is a major source of
processing and control of the autonomic nervous system. It prostaglandins during labour and female COX-1 knockout
is the most prolific isoenzyme in the mouse brain since mice experience parturition failure (Reese et al., 2000)
deletion of the COX-1 gene reduces PGE2 levels by 70% reversed by administration of prostaglandins. Prostaglan-
(Ayoub et al., 2004). However, ‘‘basal’’ levels of COX-2 dins synthesised by COX-1 are also essential for the
mRNA and COX-2 protein are expressed in the forebrain survival of fetuses during parturition, since the majority of
without stimulation by pro-inflammatory stimuli (Yama- offspring born to homozygous COX-1 knockout mice do
gata et al., 1993; Breder et al., 1995; Cao et al., 1995), but not survive (Langenbach et al., 1995). This high mortality
their expression can be increased in neuronal, glial and of the pups may be due to the failure of the ductus
endothelial cells by treatment with LPS, IL-1 or TNF arteriosus to close after birth. Patency of the ductus
(Breder and Saper, 1996; Cao et al., 1996, 1998). The so- arteriosus as well as its closure is dependent on prosta-
called constitutive COX-2 is particularly high in neonate glandins made by COX-1 and COX-2 as offspring with a
brains where it is probably induced by intense nerve double COX-1 and -2 knockout die shortly after birth with
activity in the developing brain. Electrical seizures increase an open ductus (Reese et al., 2000; Loftin et al., 2001).
ARTICLE IN PRESS
R.M. Botting / Journal of Thermal Biology 31 (2006) 208–219 213

Thus, the patency of the ductus arteriosus in wild type mediators and that aspirin might well be blocking their
animals depends on the vasodilator action of PGE2 acting synthesis. Using the supernatant of a homogenate from
on EP4 receptors in the ductus (Segi et al., 1998). However, guinea pig lung as a source of prostaglandin synthase,
the compensatory dilator action of nitric oxide has been Vane found a dose-dependent inhibition of prostaglandin
suggested as the mechanism which maintains a patent formation by aspirin, salicylate and indomethacin but not
ductus in COX-1 and -2 null mice (Baragatti et al., 2003; by morphine (Vane, 1971). Two other reports from the
Coceani et al., 2005). The use of indomethacin as a same laboratory supported and extended his finding. Smith
tocolytic has the disadvantage of closing the ductus in and Willis found that aspirin prevented the release of
utero (Heymann et al., 1976). prostaglandins from aggregating human platelets (Smith
and Willis, 1971) and Ferreira, Moncada and Vane
demonstrated that aspirin-like drugs blocked prostaglandin
5. Inhibitors of COX-1 and COX-2 release from the perfused, isolated spleen of the dog
(Ferreira et al., 1971).
5.1. The non-steroid anti-inflammatory drugs Following the elucidation of the mechanism of action of
aspirin, many more NSAIDs were developed by pharma-
The NSAIDs were so named to distinguish them from ceutical companies. These include naproxen, ibuprofen,
the anti-inflammatory steroids also used for the treatment piroxicam, diclofenac and many others. The inhibition of
of inflammation. The oldest of these is aspirin or prostaglandin synthesis explains all the actions of the
acetylsalicylic acid, a derivative of salicylate originally NSAIDs. They prevent the pathological over-production
obtained from plant sources. Its main therapeutic actions of prostaglandins by COX-2 which contribute to the
are analgesic, antipyretic and anti-inflammatory and its inflammatory process (therapeutic effects) and prevent
major side actions are gastrotoxic, anti-thrombotic and the physiological formation of prostanoids by COX-1 (side
delay of the birth process. Aspirin was synthesised by Felix effects). For example, the ulcerogenic activity of aspirin
Hoffman at the Bayer Company in Germany in 1897, but arises from the inhibition of prostaglandin production in
its mechanism of action remained unknown until 1971. the stomach which has an important cytoprotective
Before 1971, inhibition of several enzymes by aspirin had function in the gastric mucosa. The inhibition by aspirin
been observed and there were different hypotheses about is due to the irreversible acetylation of serine 530 in the
its mode of action. Against this background of knowledge, COX-1 active site or the serine 516 in the active site of
Piper and Vane took over the investigation of the mode of COX-2 (Van der Ouderaa et al., 1980). In contrast to this
action of aspirin. They employed the technique of irreversible action of aspirin, other NSAIDs such as
continuous bioassay using the cascade bioassay system indomethacin or ibuprofen produce reversible COX
(Vane, 1964) developed by Vane in the mid-1960s for use inhibition by competing with the substrate, arachidonic
with blood or an artificial salt solution (Fig. 4). Piper and acid, for binding to the active site of COX-1 or COX-2
Vane found that the anaphylactic release of prostaglandins (Vane et al., 1990). This inhibition is responsible for the
from isolated perfused guinea pig lungs was blocked by analgesic, antipyretic and anti-inflammatory actions of the
aspirin and similar drugs (Palmer and Piper, 1970). Vane NSAIDs as well as their gastrotoxicity, anti-thrombotic
postulated that the various stimuli which released pros- action (Fig. 3) and delay of the birth process. The anti-
taglandins were in fact ‘‘turning on’’ the synthesis of these thrombotic action is reversed when the NSAID is
eliminated but the anti-thrombotic action of aspirin is
irreversible until new platelets are formed after 8–10 days
(Vane et al., 1998).

5.2. Selective COX-2 inhibitors

In 1998, an estimated 5.8 billion dollars was spent world


wide on anti-inflammatory drugs and 1.8 billion dollars of
this accounted for sales of NSAIDs in the USA. However,
all NSAIDs at anti-inflammatory doses cause side effects,
the most serious being irritation, bleeding and ulcers of the
gastrointestinal tract. Indeed, several epidemiological
studies have estimated the degree of gastric damage caused
by different NSAIDs (Henry et al., 1996) since 34–46% of
patients on NSAID therapy will have some form of
gastrointestinal gastric event (Coles et al., 1983). It has
Fig. 4. Diagram of the blood-bathed organ technique. Blood or salt
been estimated that in the USA alone, some 100,000
solution superfuses a series of isolated organs, the movements of which, in patients are hospitalised each year because of perforations,
response to drugs, are recorded. ulcers or bleeding (PUBs) in the stomach (Fries, 1999) and
ARTICLE IN PRESS
214 R.M. Botting / Journal of Thermal Biology 31 (2006) 208–219

10,000–20,000 of these patients die in intensive care. of COX-1 (Kurumbail et al., 1996). Thus, they are potent
Gastric side effects range from mild dyspepsia to PUBs inhibitors of COX-2 and weak inhibitors of COX-1. When
and even ibuprofen, which is recognised as a mild gastric the selectivity is estimated by the human whole blood
irritant, causes problems in a significant proportion of assay, celecoxib has a selectivity of 7.6 and rofecoxib a
patients. For many years, there was a clear need for anti- selectivity of 35 in favour of COX-2 (Riendeau et al., 2001).
inflammatory drugs which did not damage the stomach. Second-generation selective COX-2 inhibitors have now
The discovery of the inducible COX-2 highlighted the been developed with higher selectivities for COX-2. The
differences in pharmacology of the two COX enzymes successor to celecoxib, valdecoxib (Bextra) has a selectivity
(Mitchell et al., 1993). Aspirin, indomethacin and ibupro- ratio of 30 and etoricoxib (Arcoxia), the successor to
fen are much less active against COX-2 than against COX- rofecoxib, has a selectivity of 106 (Riendeau et al., 2001).
1 (Meade et al., 1993). Indeed, the strongest inhibitors of Parecoxib, an injectable pro-drug of valdecoxib, is also
COX-1 such as aspirin, indomethacin and piroxicam are available for the treatment of acute pain (Cheer and Goa,
the NSAIDs which cause the most damage to the stomach 2001) and lumiracoxib is currently in development (Ding
(Lanza, 1989). The spectrum of activities of some 10 and Jones, 2002) and has undergone clinical trials
standard NSAIDs against the two enzymes ranges from a (Schnitzer et al., 2004). Lumiracoxib differs in structure
high selectivity towards COX-1 (166-fold for aspirin; from other coxibs. It is a phenyl acetic acid derivative
Mitchell et al., 1993; Akarasereenont et al., 1994) through instead of a sulfonamide or sulfone, and compared to non-
to equiactivity on both (0.5-fold for diclofenac; Warner et selective NSAIDs causes no more cardiovascular side
al., 1999). effects and less gastrointestinal complications (Schnitzer
The range of activities of NSAIDs against COX-1 et al., 2004).
compared with COX-2 explains the variations in the side The Celecoxib Long-Term Arthritis Safety Study
effects of NSAIDs at their anti-inflammatory doses. Drugs (CLASS) in 8000 patients for 6 months demonstrated a
which have a high potency against COX-2 and a low COX- lower incidence of ulcer complications in patients receiving
2/COX-1 activity ratio will have potent anti-inflammatory celecoxib than those on ibuprofen or diclofenac. However,
activity with few side effects on the stomach and kidney. in patients taking aspirin as well as celecoxib, the incidence
Garcia Rodriguez and Jick (1994) have published a of ulcers was no better than in the comparator group
comparison of epidemiological data on the side effects of (Silverstein et al., 2000). Moreover, when the trial was
NSAIDs. Piroxicam and indomethacin in anti-inflamma- continued for 12 months, there was no difference in the
tory doses showed high gastrointestinal toxicity. These number of gastric adverse events between the celebrex and
drugs have a much higher potency against COX-1 than comparator NSAIDs treated patients (FDA Celecoxib
against COX-2 (Vane and Botting, 1995). Thus, when Hearing July 2, 2001).
epidemiological results are compared with COX-2/COX-1 The ‘‘Vioxx’’ Gastrointestinal Outcomes Research (VIG-
ratios, there is a parallel relationship between gastrointest- OR) Trial (Bombardier et al., 2000) comparing rofecoxib
inal side effects and COX-2/COX-1 ratios. with naproxen in 8000 patients for 9 months reported fewer
Nimesulide, etodolac and meloxicam were identified in serious gastrointestinal adverse events with rofecoxib.
the 1980s as potent anti-inflammatory drugs with low Patients who were taking aspirin were excluded from the
ulcerogenic activity in the rat stomach. In some instances, trial. However, the incidence of myocardial infarction was
this was also shown to parallel low activity against higher among patients in the rofecoxib group than among
prostaglandin synthesis in the rat stomach. After the those in the naproxen group (0.4% versus 0.1%). The
characterisation of the COX-2 gene, these three drugs were higher incidence of heart attacks may be a general risk
each found preferentially to inhibit COX-2 rather than factor for all selective COX-2 inhibitors. Recent trials
COX-1, with a variation in their COX-2/COX-1 ratios of investigating the efficacy of celecoxib (Solomon et al., 2005)
between 0.1 and 0.01 depending on the test system used. and rofecoxib (Bresalier et al., 2005) in preventing color-
Ratios obtained by the human whole blood assay, which ectal adenomas demonstrated a higher incidence of
measures inhibition of COX-1 in platelets and COX-2 in cardiovascular events, including myocardial infarction, in
mononuclear cells stimulated with LPS, are now generally the drug treated group compared to the placebo. It has
accepted as the best reflection of the inhibitory activity of been suggested that selective COX-2 inhibitors prevent the
the drugs in humans (Patrignani et al., 1994). synthesis of the anti-thrombotic prostaglandin, prostacy-
Selective inhibitors of COX-2 were introduced in 1999. clin, by endothelial cells while leaving unopposed the action
The first NSAIDs to be introduced as selective COX-2 of the pro-thrombotic thromboxane in platelets (McAdam
inhibitors were celecoxib (Celebrex) and rofecoxib (Vioxx). et al., 1999). Recent epidemiological studies show that all
In place of the carboxyl group of the non-steroid anti- NSAIDs in anti-inflammatory doses increase the risk of
inflammatory acids, the structure of celecoxib contains a myocardial infarction (Hippisley-Cox and Coupland,
sulfonamide group and that of rofecoxib contains a 2005). This may be difficult to explain since the majority
methylsulfone. The sulphur-containing phenyl rings of of NSAIDs have a greater selectivity for COX-1 than for
these drugs bind into the side pocket of the COX catalytic COX-2, which is the basis for their gastrotoxicity and for
channel of COX-2 but interact weakly with the active site their anti-thrombotic action.
ARTICLE IN PRESS
R.M. Botting / Journal of Thermal Biology 31 (2006) 208–219 215

In 2004, Merck Frosst withdrew rofecoxib from the macrophages when incubated with diclofenac for 48 h
market and is currently involved in litigation with victims expressed a COX-2-like activity which was more sensitive
of cardiovascular adverse events who were treated with to inhibition by acetaminophen than LPS-induced COX-2.
rofecoxib. Selective COX-2 inhibitors are now required to In addition, LPS-induced COX-2 was a membrane-bound
carry a warning of the risk of cardiovascular side effects. enzyme whereas COX-2 induced with diclofenac existed in
Valdecoxib causes severe toxic skin rashes as well as the cytoplasm of the cells. If such a COX-2 variant is
possible myocardial infarctions and the FDA has recom- identified in vivo, it may explain some therapeutic actions
mended that its sales should be suspended (Scrip, 2005). of acetaminophen which cannot be attributed to inhibition
Valdecoxib and parecoxib are associated with an increased of a variant of COX-1.
incidence of cardiac events if used for pain after coronary
artery bypass grafting (Nussmeier et al., 2005).
7. Conclusions
6. Other cyclooxygenases
The prostaglandin field has been one of the most exciting
areas of research in recent times. More than 70 years after
Acetaminophen is generally considered an NSAID but
the marketing of aspirin, John Vane established its
its mechanism of action has not been fully resolved. It is a
mechanism of action. Just 20 years later, with the increase
weak inhibitor of isolated COX-1 and COX-2 but
in research into the properties of COX, a second
demonstrates antipyretic and analgesic properties when
isoenzyme, COX-2, was revealed by Daniel Simmons in
taken internally. It has no anti-inflammatory actions. In
1991. In the eight years that followed, the eagerly awaited
1972, Flower and Vane postulated the existence of a COX
in dog brain more sensitive to inhibition with acetamino- selective inhibitors of COX-2 reached the market, which
alleviated the symptoms of inflammatory disease without
phen than the COX in rabbit spleen. More recently,
damaging the stomach. The discovery of COX-2 inhibitors
Simmons and his colleagues characterised and cloned a
generated great enthusiasm and competition to design
COX enzyme in dog brain which, unlike COX-1 and COX-
more effective drugs. However, increasing selectivity for
2, was sensitive to inhibition with acetaminophen (Chan-
COX-2 also increased toxicity, since the anti-thrombotic
drasekharan et al., 2002). This was a splice variant of
prostacyclin is formed by COX-2 and inhibiting its
COX-1 termed by the authors, COX-3, which consists of a
synthesis precipitated heart attacks. The problem of this
COX-1 mRNA that retains intron-1. In dogs, intron-1 is 90
nucleotides in length and represents an in frame insertion side action has not yet been resolved. Is it possible to
obtain the benefit of low gastrotoxicity and avoid the
into the portion of the COX-1 open reading frame
danger of a heart attack? Perhaps lumiracoxib has
encoding the N-terminal hydrophobic signal peptide. This
provided the solution (see above). It now appears that by
variant produces enzyme protein containing the encoded
taking any NSAID, patients risk experiencing a heart
intron-1 sequence when expressed in insect cells. The
attack.
activity of the protein is preferentially inhibited by
A third COX, COX-3, has also been postulated, based
analgesic antipyretic drugs such as acetaminophen and
on the sensitivity of a COX-1 variant from dog brain to
may explain the therapeutic actions of this class of
compounds. However, although intron-1 is of similar size acetaminophen. However, when the COX-3 mRNA of
mouse or human is translated into an enzyme protein, the
in all species, it is out of frame in humans and rodents.
intron-1 insertion is out of frame. Thus, the mechanism by
Thus, it would require additional mechanisms such as the
which conversion of mRNA occurs into active enzyme
use of alternative splice sites, ribosomal frameshifting or
protein in mouse and human requires further investigation.
RNA editing to make a functional protein (Chandrase-
When this question is resolved, the mechanism of action of
kharan et al., 2002; Dinchuk et al., 2003; Simmons, 2003;
a drug even older than aspirin will be clarified.
Simmons et al., 2004). However, COX-1 variant protein
has been identified in both mouse and human tissues
(Chandrasekharan et al., 2002; Dou et al., 2004; Qin et al., References
2005), although the candidate protein identified by Qin
et al. (2005) is not selectively inhibited by acetaminophen. Akarasereenont, P., Mitchell, J.A., Thiemermann, C., Vane, J.R., 1994.
Since high levels of peroxides (such as those that exist in Relative potency of nonsteroidal anti-inflammatory drugs as inhibitors
inflamed tissues) abolish the action of acetaminophen, an of cyclooxygenase-1 or cyclooxygenase-2. Br. J. Pharmacol. 112
(Suppl.), 183.
alternative mechanism has been proposed to explain its Allen, A., Garner, A., 1980. Mucus and bicarbonate secretion in the
action. This hypothesis proposes that the inhibitory stomach and their possible role in mucosal protection. Gut 21,
activity of acetaminophen on COX depends on the level 249–262.
of peroxides in the target cells, (Ouellet and Percival, 2001; Ayoub, S.S., Botting, R.M., Goorha, S., Colville-Nash, P.R., Willoughby,
Boutaud et al., 2002). D.A., 2004. Acetaminophen-induced hypothermia in mice is mediated
by a prostaglandin endoperoxide synthase 1 gene-derived protein.
An inducible COX-2 enzyme sensitive to inhibition with Proc. Natl. Acad. Sci. USA 101, 11165–11169.
low concentrations of acetaminophen was described in Baragatti, B., Brizzi, F., Ackerley, C., Barogi, S., Ballou, L.R., Coceani,
1999 (Simmons et al., 1999). Cultured J774.2 mouse F., 2003. Cyclooxygenase-1 and cyclooxygenase-2 in the mouse ductus
ARTICLE IN PRESS
216 R.M. Botting / Journal of Thermal Biology 31 (2006) 208–219

arteriosus: individual activity and functional coupling with nitric oxide rheumatoid synovial tissues. Effects of interleukin-1 beta, phorbol
synthase. Br. J. Pharmacol. 139, 1505–1515. ester and corticosteroids. J. Clin. Invest. 93, 1095–1101.
Beiche, F., Scheuerer, S., Brune, K., Geisslinger, G., Goppelt-Struebe, M., DeWitt, D.L., 1991. Prostaglandin endoperoxide synthase: regulation of
1996. Up-regulation of cyclooxygenase-2 mRNA in the rat spinal cord enzyme expression. Biochim. Biophys. Acta 1083, 121–134.
following peripheral inflammation. FEBS. Lett. 390, 165–169. DeWitt, D.L., Smith, W.L., 1988. Primary structure of prostaglandin G/H
Bergström, S., Ryhage, R., Samuelsson, B., Sjövall, J., 1962. The structure synthase from sheep vesicular gland determined from the complemen-
of prostaglandin E, F1 and F2. Acta Chem. Scand. 16, 501–502. tary DNA sequence. Proc. Natl. Acad. Sci. USA 85, 1412–1416.
Bergström, S., Danielsson, H., Samuelsson, B., 1964. The enzymatic Dinchuk, J.E., Car, B.D., Focht, R.J., Johnston, J.J., Jaffee, B.D.,
formation of prostaglandin E2 from arachidonic acid. Prostaglandins Covington, M.B., et al., 1995. Renal abnormalities and an altered
and related factors 32. Biochim. Biophys. Acta 90, 207–210. inflammatory response in mice lacking cyclooxygenase II. Nature 378,
Blatteis, C.M., Sehic, E., 1997. Fever: how may circulating pyrogens signal 406–409.
the brain? News Physiol. Sci. 12, 1–9. Dinchuk, J.E., Liu, R.Q., Trzaskos, J.M., 2003. COX-3: in the wrong
Bombardier, C., Laine, L., Reicin, A., Shapiro, D., Burgos-Vargas, R., frame in mind. Immunol. Lett. 86, 121.
Davis, B., Day, R., Ferraz, M.B., Hawkey, C.J., Hochberg, M.C., et Ding, C., Jones, G., 2002. Lumiracoxib (Novartis). Drugs 5, 1168–1172.
al., 2000. Comparison of upper gastrointestinal toxicity of rofecoxib Dou, W., Jiao, Y., Goorha, S., Raghow, R., Ballou, L.R., 2004.
and naproxen in patients with rheumatoid arthritis. N. Engl. J. Med. Nociception and the differential expression of cyclooxygenase-1
343, 1520–1528. (COX-1), the COX-1 variant retaining intron-1 (COX-1v), and
Boutaud, O., Aronoff, D.M., Richardson, J.H., Marnett, L.J., Oates, J.A., COX-2 in mouse dorsal root ganglia (DRG). Prostaglandins Other
2002. Determinants of the cellular specificity of acetaminophen as an Lipid Mediat. 74, 29–43.
inhibitor of prostaglandin H2 synthases. Proc. Natl. Acad. Sci. USA Ek, M., Engblom, D., Saha, S., Blomqvist, A., Jakobsson, P.J., Ericsson-
99, 7130–7135. Dahlstrand, A., 2001. Pathway across the blood-brain barrier. Nature
Breder, C.D., Saper, C.B., 1996. Expression of inducible cyclooxygenase (Lond.) 410, 430–431.
mRNA in the mouse brain after systemic administration of bacterial Engblom, D., Saha, S., Engström, L., Westman, M., Audoly, L.P.,
lipopolysaccharide. Brain Res. 713, 64–69. Jakobsson, P.-J., Blomqvist, A., 2003. Microsomal prostaglandin E
Breder, C.D., Dewitt, D., Kraig, R.P., 1995. Characterization of inducible synthase-1 is the central switch during immune-induced pyresis. Nat.
cyclooxygenase in rat brain. J. Comp. Neurol. 355, 296–315. Neurosci. 6, 1137–1138.
Bresalier, R.S., Sandler, R.S., Quan, H., Bolognese, J.A., Oxenius, B., Farman, N., Pradelles, P., Bonvalet, J.P., 1987. PGE2, PGF2a, 6-keto-
Horgan, K., Lines, C., Riddell, R., Morton, D., Lanas, A., Konstam, PGF1a, and TXB2 synthesis along the rabbit nephron. Am. J. Physiol.
M.A., Baron, J.A. for the Adenomatous Polyp Prevention on Vioxx 252, F53–F59.
(APPROVe) Trial Investigators, 2005. Cardiovascular events asso- FDA Celecoxib Website 2001: http://www.fda.gov/ohrms/dockets/ac/
ciated with rofecoxib in a colorectal adenoma chemoprevention trial. cder01.htm#Arthritis.
N. Engl. J. Med. 352, 1092–1102. Ferreira, S.H., Moncada, S., Vane, J.R., 1971. Indomethacin and aspirin
Bunting, S., Moncada, S., Vane, J.R., 1983. The prostacyclin-thrombox- abolish prostaglandin release from spleen. Nature 231, 237–239.
ane A2 balance: pathophysiological and therapeutic implications. Br. Flower, R.J., Vane, J.R., 1972. Inhibition of prostaglandin synthetase in
Med. Bull. 39, 271–276. brain explains the antipyretic activity of paracetamol (4-Acetamido-
Cao, C., Matsumura, K., Yamagata, K., Watanabe, Y., 1995. Induction phenol). Nature 240, 410–411.
by lipopolysaccharide of cyclooxygenase-2 mRNA in rat brain; its Fries, J.F., 1999. Gastrointestinal toxicity of nonsteroidal anti-inflamma-
possible role in the febrile response. Brain Res. 697, 187–196. tory drugs. N. Engl. J. Med. 341, 1397–1398.
Cao, C., Matsumura, K., Yamagata, K., Watanabe, Y., 1996. Endothelial Fu, J.-Y., Masferrer, J.L., Seibert, K., Raz, A., Needleman, P., 1990. The
cells of the brain vasculature express cyclooxygenase-2 mRNA in induction and suppression of prostaglandin H2 synthase (cycloox-
response to systemic interleukin-1b: a possible site of prostaglandin ygenase) in human monocytes. J. Biol. Chem. 265, 16737–16740.
synthesis responsible for fever. Brain Res. 733, 263–272. Funk, C.D., Funk, L.B., Kennedy, M.E., Pong, A.E., FitzGerald, G.A.,
Cao, C., Matsumura, K., Yamagata, K., Watanabe, Y., 1998. Cycloox- 1991. Human platelet/erythroleukemia cell prostaglandin G/H
ygenase-2 is induced in brain blood vessels during fever evoked by synthase: cDNA cloning, expression and gene chromosomal assign-
peripheral or central administration of tumor necrosis factor. Mol. ment. FASEB J. 5, 2304–2312.
Brain Res. 56, 45–56. Garcia Rodriguez, L.A., Jick, H., 1994. Risk of upper gastrointestinal
Catella-Lawson, F., McAdam, B., Morrison, B.W., Kapoor, S., Kujubu, bleeding and perforation associated with individual non-steroidal anti-
D., Antes, L., Lasseter, K.C., Quan, H., Gertz, B.J., FitzGerald, G.A., inflammatory drugs. Lancet 343, 769–772.
1999. Effects of specific inhibition of cyclooxygenase-2 on sodium Gibb, W., Sun, M., 1996. Localization of prostaglandin H synthase type 2
balance, hemodynamics, and vasoactive eicosanoids. J. Pharmacol. protein and mRNA in term human fetal membranes and decidua.
Exp. Ther. 289, 735–741. J. Endocrinol. 150, 497–503.
Chandrasekharan, N.V., Dai, H., Roos, K.L., Evanson, N.K., Tomsik, J., Gretzer, B., Maricic, N., Respondek, M., Schuligoi, R., Peskar, B.M.,
Elton, T.S., Simmons, D.L., 2002. COX-3, a cyclooxygenase-1 variant 2001. Effects of specific inhibition of cyclooxygenase-1 and cycloox-
inhibited by acetaminophen and other analgesic/antipyretic drugs: ygenase-2 in the rat stomach with normal mucosa and after acid
cloning, structure, and expression. Proc. Natl. Acad. Sci. USA 99, challenge. Br. J. Pharmacol. 132, 1565–1573.
13926–13931. Harris, R.C., 1996. The macula densa: recent developments. J. Hypertens.
Cheer, S.M., Goa, K.L., 2001. Parecoxib (parecoxib sodium). Drugs 61, 14, 815–822.
1133–1141. Harris, R.C., McKanna, J.A., Akai, Y., Jacobson, H.R., Dubois, R.N.,
Coceani, F., Barogi, S., Brizzi, F., Ackerley, C., Seidlitz, E., Kelsey, L., Breyer, M.D., 1994. Cyclooxygenase-2 is associated with the macula
Ballou, L.R., Baragatti, B., 2005. Cyclooxygenase isoenzymes and densa of rat kidney and increases with salt restriction. J. Clin. Invest.
patency of ductus arteriosus. Prostaglandins Leukot. Essent. Fatty 94, 2504–2510.
Acids 72, 71–77. Hemler, M., Lands, W.E., 1976. Purification of the cyclooxygenase that
Coles, L.S., Fries, J.F., Kraines, R.G., Roth, S.H., 1983. From experiment forms prostaglandins: demonstration of two forms of iron in the
to experience: side effects of nonsteroidal anti-inflammatory drugs. holoenzyme. J. Biol. Chem. 251, 5575–5579.
Am. J. Med. 74, 820–828. Henry, D., Lim, L.L.-Y., Garcia Rodriguez, L.A., Perez Gutthann, S.,
Crofford, L.J., Wilder, R.L., Ristimäki, A.P., Sano, H., Remmers, E.F., Carson, J.L., Griffin, M., et al., 1996. Variability in risk
Epps, H.R., Hla, T., 1994. Cyclooxygenase-1 and -2 expression in of gastrointestinal complications with individual non-steroidal
ARTICLE IN PRESS
R.M. Botting / Journal of Thermal Biology 31 (2006) 208–219 217

anti-inflammatory drugs: results of a collaborative meta-analysis. Br. Lysz, T.W., Zweig, A., Keeting, P.E., 1988. Examination of mouse and rat
Med. J. 312, 1563–1566. tissues for evidence of dual forms of the fatty acid cyclooxygenase.
Heymann, M.A., Rudolph, A.M., Silverman, N.H., 1976. Closure of the Biochem. Pharmacol. 37, 921–927.
ductus arteriosus in premature infants by inhibition of prostaglandin Malkowski, M.G., Ginell, S.L., Smith, W.L., Garavito, R.M., 2000. The
synthesis. N. Engl. J. Med. 295, 530–533. productive conformation of arachidonic acid bound to prostaglandin
Higgs, E.A., Moncada, S., Vane, J.R., 1978. Inflammatory effects of synthase. Science 289, 1933–1937.
prostacyclin (PGI2) and 6-oxo-PGF1a in the rat paw. Prostaglandins Marcheselli, V.L., Bazan, N.G., 1996. Sustained induction of prostaglan-
16, 153–162. din endoperoxide synthase-2 by seizures in hippocampus. J. Biol.
Hippisley-Cox, J., Coupland, C., 2005. Risk of myocardial infarction in Chem. 271, 24794–24799.
patients taking cyclo-oxygenase-2 inhibitors or conventional non- Masferrer, J.L., Zweifel, B.S., Seibert, K., Needleman, P., 1990. Selective
steroid anti-inflammatory drugs: population based nested case-control regulation of cellular cyclooxygenase by dexamethasone and endotoxin
analysis. Br. J. Med. 330, 1366–1372. in mice. J. Clin. Invest. 86, 1375–1379.
Johansson, C., Kollberg, B., 1979. Stimulation by intragastrically Matsumura, K., Cao, C., Ozaki, M., Morii, H., Nakadate, K., Watanabe,
administered E2 prostaglandins of human gastric mucus output. Eur. Y., 1998. Brain endothelial cells express cyclooxygenase-2 during
J. Pharmacol. 9, 229–232. lipopolysaccharide-induced fever: light and electron microscopic
Kargman, S., Charlson, S., Cartwright, M., Frank, J., Riendeau, D., et al., immunocytochemical studies. J. Neurosci. 18, 6279–6289.
1996. Characterisation of prostaglandin G/H synthase 1 and 2 in rat, McAdam, B.F., Catella-Lawson, F., Mardini, I.A., Kapoor, S., Lawson,
dog, monkey and human gastrointestinal tracts. Gastroenterology 111, J.A., FitzGerald, G.A., 1999. Systemic biosynthesis of prostacyclin by
445–454. cyclooxygenase (COX)-2: the human pharmacology of a selective
Kömhoff, M., Wang, J.-L., Cheng, H.-F., Langenbach, R., McKanna, inhibitor of COX-2. Proc. Natl. Acad. Sci. USA 96, 272–277.
J.A., Harris, R.C., Breyer, M.D., 2000. Cyclooxygenase-2-selective Meade, E.A., Smith, W.L., DeWitt, D.L., 1993. Differential inhibition of
inhibitors impair glomerulogenesis and renal cortical development. prostaglandin endoperoxide synthase (cyclooxygenase) isozymes by
Kidney Int. 57, 414–422. aspirin and other non-steroidal anti-inflammatory drugs. J. Biol.
Kujubu, D.A., Fletcher, B.S., Varnum, B.C., Lim, R.W., Herschman, Chem. 268, 6610–6614.
H.R., 1991. TIS10, a phorbol ester tumor promoter-inducible mRNA Merlie, J.P., Fagan, D., Mudd, J., Needleman, P., 1988. Isolation and
from Swiss 3T3 cells, encodes a novel prostaglandin synthase/ characterisation of the complementary DNA for sheep seminal vesicle
cyclooxygenase homologue. J. Biol. Chem. 266, 12866–12872. prostaglandin endoperoxide synthase (cyclooxygenase). J. Biol. Chem.
Kurumbail, R.G., Stevens, A.M., Gierse, J.K., McDonald, J.J., Stegeman, 263, 3550–3553.
R.A., Pak, J.Y., Gildehaus, D., Miyashiro, J.M., Penning, T.D., Mitchell, J.A., Akarasereenont, P., Thiemermann, C., Flower, R.J., Vane,
Seibert, K., et al., 1996. Structural basis for selective inhibition of J.R., 1993. Selectivity of nonsteroidal anti-inflammatory drugs as
cyclooxygenase-2 by anti-inflammatory agents. Nature 384, 644–648. inhibitors of constitutive and inducible cyclooxygenase. Proc. Natl.
Kurzrok, R., Lieb, C.C., 1930. Biochemical studies of human semen II Acad. Sci. USA 90, 11693–11697.
action of semen on the human uterus. Proc. Soc. Exp. Biol. Med. 28, Moncada, S., Vane, J.R., 1982. The role of prostaglandins in platelet-
268–272. vessel wall interactions. In: Nossel, H.L., Vogel, H.J. (Eds.),
Langenbach, R., Morham, S.G., Tiano, H.F., Loftin, C.D., Ghanayem, Pathobiology of the Endothelial Cell. Academic Press, New York,
B.I., Chulada, P.C., Mahler, J.F., Lee, C.A., Goulding, E.H., pp. 253–285.
Kluckman, K.D., et al., 1995. Prostaglandin synthase 1 gene Moncada, S., Gryglewski, R., Bunting, S., Vane, J.R., 1976. An enzyme
disruption in mice reduces arachidonic acid-induced inflammation isolated from arteries transforms prostaglandin endoperoxides to an
and indomethacin-induced gastric ulceration. Cell 83, 483–492. unstable substance that inhibits platelet aggregation. Nature 263,
Lanza, F.L., 1989. A review of gastric ulcer and gastroduodenal injury in 663–665.
normal volunteers receiving aspirin and other non-steroidal anti- Morham, S.G., Langenbach, R., Loftin, C.D., Tiano, H.F., Voulouma-
inflammatory drugs. Scand. J. Gastroenterol. 24 (Suppl. 163), 24–31. nos, N., Jennette, J.C., Mahler, J.F., Kluckman, K.D., Ledford, A.,
Laulederkind, S.J., Wall, B.M., Ballou, L.R., Raghow, R., 2002. Renal Lee, C.A., et al., 1995. Prostaglandin synthase 2 gene disruption causes
pathology resulting from PGHS-2 gene ablation in DBA/B6 mice. severe renal pathology in the mouse. Cell 83, 473–482.
Prostaglandins Other Lipid Mediat. 70, 161–168. Needleman, P., Moncada, S., Bunting, S., Vane, J.R., Hamberg, M.,
Li, S., Wang, Y., Matsumura, K., Ballou, L.R., Moreham, S.G., Blatteis, Sammuelsson, B., 1976. Identification of an enzyme in platelet
C.M., 1999. The febrile response to lipopolysaccharide is blocked in microsomes which generates thromboxane A2 from prostaglandin
cyclooxygenase-2/ mice. Brain Res. 825, 86–94. endoperoxides. Nature 261, 558–560.
Li, S., Ballou, L.R., Morham, S.G., Blatteis, C.M., 2001. Cyclooxygenase- Nussmeier, N.A., Whelton, A.A., Brown, M.T., Langford, R.M., Hoeft,
2 mediates the febrile response of mice to interleukin-1b. Brain Res. A., Parlow, J.L., Boyce, S.W., Verburg, K.M., 2005. Complications of
910, 163–173. the COX-2 inhibitors parecoxib and valdecoxib after cardiac surgery.
Lim, H., Paria, B.C., Das, S.K., Dinchuk, J.E., Langenbach, R., Trzaskos, N. Engl. J. Med. 352, 1081–1091.
J.M., Dey, S.K., 1997. Multiple female reproductive failures in O’Banion, M.K., Sadowski, H.B., Winn, V., Young, D.A., 1991. A serum-
cyclooxygenase 2-deficient mice. Cell 91, 197–208. and glucocorticoid-regulated 4-kilobase mRNA encodes a cycloox-
Loftin, C.D., Trivedi, D.B., Tiano, H.F., Clark, J.A., Lee, C.A., Epstein, ygenase-related protein. J. Biol. Chem. 266, 23261–23267.
J.A., Morham, S.G., Breyer, M.D., Nguyen, M., Hawkins, B.M., et Oliva, D., Nicosia, S., 1987. PGI2 receptors and molecular mechanisms in
al., 2001. Failure of ductus arteriosus closure and remodelling in platelets and vasculature: state of the art. Pharmacol. Res. Commun.
neonatal mice deficient in cyclooxygenase-1 and cyclooxygenase-2. 19, 735–765.
Proc. Natl. Acad. Sci. USA 98, 1059–1064. Ouellet, M., Percival, M.D., 2001. Mechanism of acetaminophen
Luk, G.D., 1996. Prevention of gastrointestinal cancer—the potential role inhibition of cyclooxygenase isoforms. Arch. Biochem. Biophys. 387,
of NSAIDs in colorectal cancer. Schweiz. Med. Wochenschr. 126, 273–280.
801–812. Palmer, M.A., Piper, P.J., 1970. The release of RCS from chopped lung
Luong, C., Miller, A., Barnett, J., Chow, J., Ramesha, C., Browner, M.F., and its antagonism by anti-inflammatory drugs. Br. J. Pharmacol. 40,
1996. Flexibility of the NSAID binding site in the structure of human 581.
cyclooxygenase-2. Nat. Struct. Biol. 3, 927–933. Parrett, M.L., Harris, R.E., Joarder, F.S., Ross, M.S., Clausen, K.P.,
Lysz, T.W., Needleman, P., 1982. Evidence for two distinct forms of fatty Robertson, F.M., 1997. Cyclooxygenase-2 gene expression in human
acid cyclooxygenase in brain. J. Neurochem. 38, 1111–1117. breast cancer. Int. J. Oncol. 10, 503–507.
ARTICLE IN PRESS
218 R.M. Botting / Journal of Thermal Biology 31 (2006) 208–219

Patrignani, P., Panara, M.R., Greco, A., Fusco, O., Natoli, C., Iacobelli, the CLASS study: a randomised controlled trial. Celecoxib Long-term
S., Cipollone, F., Ganci, A., Creminon, C., Maclouf, J., et al., 1994. Arthritis Safety Study. J. Am. Med. Assoc. 284, 1247–1255.
Biochemical and pharmacological characterization of the cyclooxy- Simmons, D.L., 2003. Variants of cyclooxygenase-1 and their roles in
genase activity of human blood prostaglandin endoperoxide synthases. medicine. Thromb. Res. 110, 265–268.
J. Pharmacol. Exp. Ther. 271, 1705–1710. Simmons, D.L., Levy, D.B., Yannoni, Y., Erikson, R.L., 1989.
Picot, D., Loll, P.J., Garavito, R.M., 1994. The X-ray crystal structure of Identification of a phorbol ester-repressible v-src-inducible gene. Proc.
the membrane protein prostaglandin H2 synthase-1. Nature (Lond.) Natl. Acad. Sci. USA 86, 1178–1182.
367, 243–249. Simmons, D.L., Xie, W., Chipman, J., Evett, G., 1991. Multiple
Qin, N., Zhang, S.-P., Reitz, T.L., Mei, J.M., Flores, C.M., 2005. Cloning, cyclooxygenases: cloning of a mitogen-inducible form. In: Bailey, M.
expression and functional characterization of human COX-1 splicing (Ed.), Prostaglandin, Leukotrienes, Lipoxins and PAF. Plenum Press,
variants: evidence for intron 1 retention. J. Pharmacol. Exp. Ther. 315, London, pp. 67–78.
1298–1305. Simmons, D.L., Botting, R.M., Robertson, P.M., Madsen, M.L., Vane,
Raz, A., Wyche, A., Needleman, P., 1989. Temporal and pharmacological J.R., 1999. Induction of an acetaminophen-sensitive cyclooxygenase
division of fibroblast cyclooxygenase expression into transcriptional with reduced sensitivity to nonsteroid anti-inflammatory drugs. Proc.
and translational phases. Proc. Natl. Acad. Sci. USA 86, 1657–1661. Natl. Acad. Sci. USA 96, 3275–3280.
Reese, J., Paria, B.C., Brown, N., Zhao, X., Morrow, J.D., Dey, S.K., Simmons, D.L., Botting, R.M., Hla, T., 2004. Cyclooxygenase isozymes:
2000. Coordinated regulation of fetal and maternal PGs directs the biology of prostaglandin synthesis and inhibition. Pharmacol. Rev.
successful birth and postnatal adaptation in the mouse. Proc. Natl. 56, 387–437.
Acad. Sci. USA 97, 9759–9764. Sirois, J., Richards, J.S., 1992. Purification and characterisation of a
Riendeau, D., Percival, M.D., Brideau, C., Charleson, S., Dubé, D., novel, distinct isoform of prostaglandin endoperoxide synthase
Ethier, D., Falgueyret, J.-P., Friesen, R.W., Gordon, R., Greig, G., induced by human chorionic gonadotropin in granulosa cells of rat
Guay, J., Mancini, J., Ouellet, M., Wong, E., Xu, L., Boyce, S., Visco, preovulatory follicles. J. Biol. Chem. 267, 6382–6388.
D., Girard, Y., Prasit, P., Zamboni, R., Rodger, I.W., Gresser, M., Smith, W.L., Lands, W.E.M., 1972. Oxygenation of polyunsaturated fatty
Ford-Hutchinson, A.W., Young, R.N., Chan, C.-C., 2001. Etoricoxib acids during prostaglandin biosynthesis by sheep vesicular gland.
(MK-0663): preclinical profile and comparison with other agents that Biochemistry 11, 3276–3285.
selectively inhibit cyclooxygenase-2. J. Pharmacol. Exp. Ther. 296, Smith, J.B., Willis, A.L., 1971. Aspirin selectively inhibits prostaglandin
558–566. production in human platelets. Nature 231, 235–237.
Ristimäki, A., Honkanen, N., Jänkälä, H., Sipponen, P., Härkönen, M., Solomon, S.D., McMurray, J.J.V., Pfeffer, M.A., Wittes, J., Fowler, R.,
1997. Expression of cyclooxygenase-2 in human gastric carcinoma. Finn, P., Anderson, W.F., Zauber, A., Hawk, E., Bertagnolli, M., for
Cancer Res. 57, 1276–1280. the Adenoma Prevention with Celecoxib (APC) Study Investigators,
Robert, A., Nezamis, J.E., Phillips, J.P., 1967. Inhibition of gastric acid 2005. Cardiovascular risk associated with celecoxib in a clinical trial
secretion by prostaglandins. Am. J. Dig. Dis. 12, 1073–1076. for colorectal adenoma prevention. N. Engl. J. Med. 352, 1071–1080.
Rosen, G.D., Birkenmeier, T.M., Raz, A., Holtzman, M.J., 1989. Steinbach, G., Lynch, P.M., Phillips, R.K., Wallace, M.H., Hawk, E.,
Identification of a cyclooxygenase-related gene and its potential role Gordon, G.B., et al., 2000. The effect of celecoxib, a cyclooxygenase-2
in prostaglandin formation. Biochem. Biophys. Res. Commun. 164, inhibitor, in familial adenomatous polyposis. N. Engl. J. Med. 342,
1358–1365. 1946–1952.
Sano, H., Hla, T., Maier, J.A., Crofford, L.J., Case, J.P., Maciag, T., Steiner, A.A., Rudaya, A.Y., Robbins, J.R., Dragic, A.S., Langenbach,
Wilder, R.L., 1992. In vivo cyclooxygenase expression in synovial R., Romanovsky, A.A., 2005. Expanding the febrigenic role
tissues of patients with rheumatoid arthritis and osteoarthritis and rats of cyclooxygenase-2 to the previously overlooked responses. Am.
with adjuvant and streptococcal cell wall arthritis. J. Clin. Invest. 89, J. Physiol. Regul. Integr. Comp. Physiol. 4 (Epub ahead of print).
97–108. Tanaka, A., Hase, S., Miyazawa, T., Takeuchi, K., 2002. Up-regulation of
Sawdy, R., Slater, D., Fisk, N., Edmonds, D.K., Bennett, P., 1997. Use of cyclooxygenase-2 by inhibition of cyclooxygenase-1: a key to
a cyclo-oxygenase type-2 selective non-steroidal anti-inflammatory nonsteroidal anti-inflammatory drug-induced intestinal damage.
agent to prevent preterm delivery. Lancet 350, 265–266. J. Pharmacol. Exp. Ther. 300, 754–761.
Schnitzer, T.J., Burmester, G.R., Mysler, E., Hochberg, M.C., Doherty, Tateson, J.E., Moncada, S., Vane, J.R., 1977. Effects of prostacyclin
M., Ehrsam, E., Gitton, X., Krammer, G., Mellein, B., Matchaba, P., (PGX) on cyclic AMP concentrations in human platelets. Prostaglan-
Gimona, A., Hawkey, C., on behalf of the TARGET Study Group, dins 13, 389–397.
2004. Comparison of lumiracoxib with naproxen and ibuprofen in the Thun, M.J., Manboodiri, M.M., Heath, C.W.J., 1991. Aspirin use and
Therapeutic Arthritis Research and Gastrointestinal Event Trial reduced risk of fatal colon cancer. N. Engl. J. Med. 325, 1593–1596.
(TARGET), reduction in ulcer complications: randomized controlled Topper, J.N., Cai, J., Falb, D., Gimbrone Jr., M.A., 1996. Identification
trial. Lancet 364, 665–674. of vascular endothelial genes differentially responsive to fluid
Schonbeck, U., Sukhova, G.K., Graber, P., Coulter, S., Libby, P., 1999. mechanical stimuli: cyclooxygenase-2, manganese superoxide dismu-
Augmented expression of cyclooxygenase-2 in human atherosclerotic tase, and endothelial cell nitric oxide synthase are selectively up-
lesions. Am. J. Pathol. 155, 1281–1291. regulated by steady laminar shear stress. Proc. Natl. Acad. Sci. USA
Scrip, August 17th 2005 No. 3081. PJB Publications, London, UK, p. 15. 93, 10417–10422.
Segi, E., Sugimoto, Y., Yamasaki, A., Aze, Y., Oida, H., Nishimura, T., Van der Ouderaa, F.J., Buytenhek, M., Nugteren, D.H., Van Dorp, D.A.,
Murata, T., Matsuoka, T., Ushikubi, F., Hirose, M., et al., 1998. 1980. Acetylation of prostaglandin endoperoxide synthetase with
Patent ductus arteriosus and neonatal death in prostaglandin receptor acetylsalicylic acid. Eur. J. Biochem. 109, 1–8.
EP4-deficient mice. Biochem. Biophys. Res. Commun. 246, 7–12. Vane, J.R., 1964. The use of isolated organs for detecting active substances
Sigthorsson, G., Simpson, R.J., Walley, M., Anthony, A., Foster, R., in the circulating blood. Br. J. Pharmacol. Chemother. 23, 360–373.
Hotz-Behoftsitz, C., Palizban, A., Pombo, J., Watts, J., Morham, S.G., Vane, J.R., 1971. Inhibition of prostaglandin synthesis as a mechanism of
Bjarnason, I., 2002. COX-1 and 2, intestinal integrity and pathogenesis action for aspirin-like drugs. Nature (New Biol.) 231, 232–235.
of nonsteroidal anti-inflammatory drug enteropathy in mice. Gastro- Vane, J.R., Botting, R.M., 1995. New insights into the mode of action of
enterology 122, 1913–1923. anti-inflammatory drugs. Inflamm. Res. 44, 1–10.
Silverstein, F.E., Faich, G., Goldstein, J.L., Simon, L.S., Pincus, T., Vane, J.R., Flower, R.J., Botting, R.M., 1990. History of aspirin and its
Whelton, A., Makuch, R., Eisen, G., Agrawal, N.M., Stenson, W.F., mechanism of action. Stroke (Suppl.IV), 12–13.
et al., 2000. Gastrointestinal toxicity with celecoxib vs nonsteroidal Vane, J.R., Bakhle, Y.S., Botting, R.M., 1998. Cyclooxygenases 1 and 2.
anti-inflammatory drugs for osteoarthritis and rheumatoid arthritis: Annu. Rev. Pharmacol. Toxicol. 38, 97–120.
ARTICLE IN PRESS
R.M. Botting / Journal of Thermal Biology 31 (2006) 208–219 219

von Euler, U.S., 1936. On specific vasodilating and plain muscle expression in cultured endothelial cells. Proc. Natl. Acad. Sci. USA 88,
stimulating substances from accessory genital glands in man and 2384–2387.
certain animals (prostaglandin and vesiglandin). J. Physiol. 88, Xie, W., Chipman, J.G., Robertson, D.L., Erikson, R.L., Simmons, D.L.,
213–234. 1991. Expression of a mitogen-responsive gene encoding prostaglandin
Wallace, J.L., McKnight, W., Reuter, B.K., Vergnolle, N., 2000. NSAID- synthase is regulated by mRNA splicing. Proc. Natl. Acad. Sci. USA
induced gastric damage in rats: requirements for inhibition of both 88, 2692–2696.
cyclooxygenase 1 and 2. Gastroenterology 119, 706–714. Yaksh, T.L., Svensson, C., 2001. Role of spinal cyclooxygenases in
Warner, T.D., Giuliano, F., Vojnovic, I., Bukasa, A., Mitchell, J.A., Vane, nociceptive processing. In: Vane, J.R., Botting, R.M. (Eds.), Ther-
apeutic Roles of Selective COX-2 Inhibitors. William Harvey Press,
J.R., 1999. Nonsteroid drug selectivities for cyclo-oxygenase-1 rather
London, pp. 168–190.
than cyclo-oxygenase-2 are associated with human gastrointestinal
Yamagata, K., Andreasson, K.I., Kaufman, E.W., Barnes, C.A., Worley,
toxicity: a full in vitro analysis. Proc. Natl. Acad. Sci. USA 96,
P.F., 1993. Expression of a mitogen-inducible cyclooxygenase in brain
7563–7568.
neurons: regulation by synaptic activity and glucocorticoids. Neuron
Whittle, B.J.R., Salmon, J.A., 1983. In: Turnberg, L.A. (Ed.), Intestinal 11, 371–386.
Secretion. Smith Kline and French Publications, Welwyn Garden City, Yokoyama, C., Takai, T., Tanabe, T., 1988. Primary structure of sheep
pp. 69–73. prostaglandin endoperoxide synthase deduced from cDNA sequence.
Whittle, B.J.R., Boughton-Smith, N.K., Moncada, S., Vane, J.R., 1978. FEBS Lett. 231, 247–251.
Actions of prostacyclin (PGI2) and its product 6-oxo-PGF1a on the rat Zusman, R.M., Keiser, H.R., 1977. Prostaglandin biosynthesis by rabbit
gastric mucosa in vivo and in vitro. Prostaglandins 15, 955–968. renomedullary interstitial cells in tissue culture. Stimulation by
Wu, K.K., Sanduja, R., Tsai, A.-L., Ferhanoglu, B., Loose-Mitchell, D.S., angiotensin II, bradykinin and arginine vasopressin. J. Clin. Invest.
1991. Aspirin inhibits interleukin 1-induced prostaglandin H synthase 60, 215–223.

You might also like