Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Jan 1.
Published in final edited form as: Biochimie. 2010 Sep 22;93(1):46–51. doi: 10.1016/j.biochi.2010.09.009

Targeting cyclooxygenases-1 and -2 in neuroinflammation: therapeutic implications

Saba Aïd 1, Francesca Bosetti 1
PMCID: PMC3008299  NIHMSID: NIHMS245557  PMID: 20868723

Abstract

Neuroinflammation has been implicated in the pathogenesis or the progression of a variety of acute and chronic neurological and neurodegenerative disorders, including Alzheimer’s disease. Prostaglandin H synthases or cyclooxygenases (COX-1 and COX-2) play a central role in the inflammatory cascade by converting arachidonic acid into bioactive prostanoids. In this review, we highlighted recent experimental data that challenge the classical view that the inducible isoform COX-2 is the most appropriate target to treat neuroinflammation. First, we discussed data showing that COX-2 activity is linked to anti-inflammatory and neuroprotective actions and is involved in the generation of novel lipid mediators with pro-resolution properties. Then, we reviewed recent data demonstrating that COX-1, classically viewed as the homeostatic isoform, is actively involved in brain injury induced by pro-inflammatory stimuli including Aβ, lipopolysaccharide, IL-1β, and TNF-α. Overall, we suggest revisiting the traditional views on the roles of each COX during neuroinflammation and we propose COX-1 inhibition as a viable therapeutic approach to treat CNS diseases with a marked inflammatory component.

Cyclooxygenases -1 and -2 in the central nervous system

Prostaglandin H synthases, or cyclooxygenases (COX), exist in two isoforms COX-1 and COX-2, encoded by different genes. A third isoform (COX-3 or COX-1b) has been first described in canine as a splice variant of COX-1 gene but its physiological role at this point remains unknown. Indeed, in rodents and humans, COX-1b encodes proteins with completely different amino acid sequences than COX-1 or COX-2 and without COX activity [1]. COX play a central role in the inflammatory cascade by converting arachidonic acid (AA), released from membrane phospholipids by a phospholipase A2 (PLA2), into bioactive prostanoids. Both COX isoforms catalyze the same reactions: dioxygenation of arachidonic acid (AA) to yield prostaglandin G2 (PGG2), and a peroxidase reaction, which converts PGG2 to prostaglandin H2 (PGH2). PGH2 is then transformed into PGE2, PGF, PGD2, PGI2 and TXB2 through specific terminal synthases (Fig. 1). The eicosanoids synthesized by COX are powerful lipid mediators that exert a variety of biological effects by acting on multiple G coupled protein receptors. The two COX isoforms share 60% homology in their amino acids sequence and are both integral membrane homodimers proteins of the endoplasmic reticulum and nucleus, with roughly comparable kinetics. However, they differ in their regulatory mechanisms, cell localization, and function. COX-2 was first identified as a key element of the acute inflammatory response because its expression is rapidly induced by various inflammogens [2]. Indeed, COX-2 gene has several transcriptional regulatory elements, including NF-κB, Sp1, a TATA box, CAAT Enhancer Binding Protein Beta (C/EBP β), and cAMP response element-binding (CREB) consensus sequences, interacting with trans-acting factors generated by multiple signaling pathways [3, 4]. In contrast, COX-1 promoter lacks a TATA or CAAT box, has a high GC content, and contains several SP1 elements. Moreover, COX-2, but not COX-1, contains a unique 27 amino acid sequence near its C-terminus, an instability element involved in COX-2 protein degradation [4]. Because COX-2 is typically induced by inflammatory stimuli in the majority of tissues, it was presumed to be the only isoform responsible for propagating the inflammatory response and thus traditionally considered as the best target for anti-inflammatory drugs. COX-2 selective inhibition, however, has been associated with an increased risk of severe cardiovascular adverse events, which led to the voluntary withdrawal from the market of some COX-2 selective inhibitors [5]. In the brain, both COX-1 and COX-2 are constitutively expressed. In physiological conditions, COX-1 is mainly expressed in microglia and perivascular cells (a macrophage-derived vascular cell type) [68] and COX-2 is found in post synaptic dendrites and excitatory terminals, particularly in the cortex, hippocampus and amygdala, with both neuronal and vascular associations [9]. Thus, it is not surprising that in the central nervous system (CNS) COX-2 has been implicated in important physiological functions such as synaptic activity, long-term potentiation, long-term depression, memory consolidation, and neurovascular coupling during functional hyperemia [10, 11]. Furthermore, COX-2, but not COX-1, can oxygenate endocannabinoids, which represent an important metabolic pathway in neurons to regulate excitatory synaptic transmission [12]. Even in physiological conditions, COX-1 or COX-2-derived prostanoids seem to have distinct functions. For instance, PGE2 resulting from COX-2, but not COX-1 activity, is necessary for the induction of long-term potentiation and spatial learning in vivo, whereas COX-1 inhibition facilitates baseline synaptic transmission [13].

Fig. 1. The metabolic pathway of arachidonic acid (AA).

Fig. 1

AA is released from membrane phospholipids by a phospholipase A2 (PLA2) enzyme and metabolized by COX-1 or COX-2 into PGH2 which is further converted to prostaglandins and thromboxanes by specific terminal synthases. AA can also be metabolized to leukotrienes through lipooxygenases, and to epoxyeicosatrienoic acid (EET) and dihydroxyacids through cytochrome p450 epoxidase. The cyclopentenones PGJ2 and 15d-PGJ2 are non-enzymatic degradation products of PGD2.

Recent data have challenged the classical view describing COX-1 as the isoform merely responsible for physiological production of prostanoids and COX-2 as the major pro-inflammatory isoform. The use of genetic mouse models in combination with selective pharmacological inhibitors helped to identify the specific roles of COX-1 and COX-2 during neuroinflammation [14].

COX-2 activity is necessary to switch off neuroinflammation

Several studies have now demonstrated that COX-2 genetic deletion or pharmacological inhibition can worsen the response to neuroinflammatory stimuli. Specifically, Gilroy and collaborators provided the first evidence of COX-2 anti-inflammatory properties in a carrageenan-induced pleurisy model [15]. In the brain, our and other groups demonstrated that inhibition or genetic deletion of COX-2 exacerbated the neuroinflammatory response to an endotoxin challenge [1620]. Lipopolysaccharide (LPS), a gram-negative bacterial cell surface proteoglycan, known also as bacterial endotoxin, has been widely used to activate the innate immune response in both the periphery and the brain [21]. The LPS model is particularly relevant to examine activation of brain innate immunity, since it specifically and directly targets microglia, the immune resident cells in the brain [22]. Indeed, LPS binds to CD14 protein and potently activates toll-like receptor 4, which expression has been demonstrated in vivo in microglia but not in neurons [23]. LPS causes massive resident microglial activation and peripheral leukocyte infiltration into the CNS, accompanied by a robust and transient transcriptional activation of genes encoding pro-inflammatory cytokines, chemokines, prostaglandins (PGs), thromboxanes (Tx) and free radical-generating enzymes (Fig. 2) [24]. Using the LPS model of neuroinflammation, we showed that COX-2 deletion or selective pharmacological inhibition with celecoxib increase neuronal damage, glial activation, and the expression of brain cytokines and ROS-expressing enzymes, such as the pro-inflammatoy cytokine IL-1β and the p67phox subunit of NADPH oxidase, a major source of superoxide during neuroinflammation [17, 20]. COX-2 gene deletion also increases blood-brain barrier (BBB) permeability and leukocyte infiltration [16, 17]. Activated microglia and infiltrated leukocytes further amplify the neuroinflammatory response and neuronal damage by releasing TNF-α, IL-1β, CCL3/macrophage inflammatory protein1α (MIP-1α), CXCL2/MIP-2α and the matrix metalloproteinase-3 and -9, all of which had increased expression in the brain of LPS-challenged COX-2 deficient mice compared with their respective wild type controls [16, 17, 20]. Consistent with these reports, COX-2 deletion exacerbated endotoxin-induced ocular inflammation [19], and selective pharmacological inhibition of COX-2 with NS-398 increased the transcription of inflammatory genes (mPGES-1, TLR2, CD14, MCP-1) in vascular associated brain cells and parenchymal microglia after systemic injection of LPS [18]. These findings in models of primary neuroinflammation describe a different picture from the described role of COX-2 in mediating the progression of ischemic and or/excitotoxic brain injury [25]. This apparent discrepancy can be explained depending on the cell types involved in the specific injury model. Thus, COX-2, because of its constitutive expression in pyramidal neurons, could mediate injury in models that directly challenge neurons. Consistent with this hypothesis, COX-2 deletion did not affect markers of inflammation and oxidative stress in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) model, which selectively injures neurons [26]. Furthermore, transgenic mice overexpressing human COX-2 via neuron-specific Thy-1 promoter, causing elevated brain PG levels, showed similar neuroinflammatory response and neuronal damage after direct activation of glial innate immunity [27]. Therefore, protective or toxic effect of COX-2 on neuronal viability depends on whether the primary stimulus is inflammatory or excitotoxic and on the cell type targeted (glia/neurons).

Fig. 2. COX-1 and COX-2 distinct roles during the progression of neuroinflammation.

Fig. 2

Inflammatory stimuli target primarily microglia, the immune resident cells of the brain. They activate pro-inflammatory signaling pathways (NFκB, STAT/JAK) as well as the AA cascade, leading to the release of cytokines, chemokines, reactive oxygen species (ROS), prostaglandins (PG) and thromboxanes (TX). Being predominantly expressed by microglia, COX-1 contributes within seconds to minutes to the local eicosanoid pool at the site of inflammation, before the induction of COX-2. Chemokines and eicosanoids attract leukocytes and later monocytes to the site of inflammation. During the normal progression of neuroinflammation, resolution occurs with a shift from pro-inflammatory eicosanoids to anti-inflammatory cyclopentanone PGs (PGD2 and 15-dPGJ2) synthesis. Cyclopentanone PGs attenuate the induction of inflammatory genes. With time, a class switching occurs towards pro-resolving lipid mediators generated by aspirin-acetylated COX-2 and lipooxygenases (LOX) such as (AA)-derived lipoxins and DHA-derived resolvins and neuroprotectins. DHA can also be directly metabolized by COX-2 into anti-inflammatory oxo-derivatives (13-EFOX) in activated macrophages.

These recent findings suggest that COX-2-derived products can mediate a protective effect in the progression and/or the resolution of inflammation in the brain after endotoxin activation of brain innate immunity. In this regard, the recent discovery of novel lipid mediators may underline the mechanical basis for COX-2 anti inflammatory or pro-resolving properties [28, 29].

COX-2 derived anti-inflammatory and resolving mediators

In the resolution phase that follows an inflammatory stimulus, inflammation is cleared allowing the tissue to return to a non-inflamed, homeostatic state. Recent evidence indicates that the resolution phase is an active process, mediated by local-acting and specialized lipid mediators with immunoregulatory properties named lipoxins, cyclopentenone PGs, resolvins, and protectins [30]. The synthesis of some of these bioactive lipid mediators that contribute to the resolution of inflammation requires COX-2 activity [31, 32]. Cyclopentanone PGs such as PGJ2 and 15-deoxy-Δ12,14-PGJ2 (15dPGJ2) are non-enzymatic breakdown products of COX-2-derived PGD2 (Fig. 1 & 2) [29]. Gilroy and colleagues demonstrated that selective COX-2 inhibitors, by blocking the production of COX-2 derived PGD2 and its metabolite 15dPGJ2, disturb the resolution phase of inflammation, and delay the return to homeostasis [15]. 15dPGJ2 has been shown to suppress iNOS activity and the expression of pro-inflammatory cytokines (TNF-α, IL-1β, IL-6), decrease monocyte migration, and cause polymorphonuclear monocyte and macrophage apoptosis [29, 33]. 15dPGJ2 is also a potent endogenous ligand of peroxisome proliferator–activated receptor γ, an important immunoregulator. Supporting the hypothesis that COX-2 activity is important in the resolution of inflammation, COX-2 deficient mice show increased infiltration of leukocytes, particularly neutrophils, in the brain after central endotoxin challenge [16].

Recently, additional bioactive lipids with anti-inflammatory and pro-resolution properties, generated via COX-2 metabolism of the ω-3 polyunsaturated fatty acids (PUFAs) docosahexaenoic acid (DHA) and eicosapentaenoic acid (EPA), have been characterized [28, 34]. The identification of these ω-3 PUFAs-derived anti-inflammatory compounds is especially relevant for CNS inflammation because in brain phospholipids, DHA and AA, are the most abundant PUFAs, particularly in retina and synaptic terminals [35]. Groeger and collaborators discovered new oxo-derivative of ω-3 PUFAs (EFOX), including DHA, that are generated by COX-2 in macrophages activated by phorbol myristic acetate, LPS and interferon-γ (IFN-γ). The study reported that these EFOX dose-dependently reduced the expression of iNOS and COX-2 and also suppressed the production of cytokines (IL-6, MCP-1) and nitric oxide metabolites in the media of monocyte/macrophage cell lines and in primary murine bone marrow-derived macrophages [34]. Interestingly, COX-2 acetylation by aspirin further increased the generation of these mediators. Aspirin irreversibly acetylates COX-1 (at serine residue 530), blocking its activity and the synthesis of PGs and TX, which have pro-inflammatory and pro-thrombotic properties. However, aspirin also acetylates COX-2 (at serine residue 516), switching the synthesis of classical PGs and TX to the synthesis of anti-inflammatory and resolving mediators lipoxins, resolvins and protectins [30].

Lipoxins are generated from AA, resolvins D and protectins from DHA, resolvins E from EPA via lipooxygenases (LOX) or concerted action of COX and LOX via cell-cell interaction (Fig. 2&3) [29]. However, after aspirin administration, acetylated COX-2 can generate epimers of lipoxins, resolvins or protectins and increase their formation [36], effects that are blocked by selective COX-2 inhibitors [37]. These aspirin-triggered protective mediators share the same biological activity with their analogs selectively formed through the LOX biosynthetic pathway. They regulate the magnitude and duration of inflammation by attenuating neutrophil infiltration, and initiating the clearance of apoptotic leukocytes [29]. Aspirin-triggered resolvins and neuroprotectins have been found in both mouse and human brain; and exert potent biological actions in vivo on leukocyte trafficking and on the regulation of cytokine expression by glial cells [38]. Neuroprotectins have demonstrated neuroprotective effects in experimental brain damage, oxidatively stressed retinal pigment epithelial cells, and β-amyloid-induced neurotoxicity in human brain cells [39]. Further studies are needed to examine whether these COX-2-derived protective mediators can be synthesized by inflamed nervous tissues and exert anti-inflammatory effects in vivo.

Fig. 3. COX-2 derived novel protective lipid mediators.

Fig. 3

By acetylating COX-2, aspirin inhibits the synthesis of prostaglandins and thromboxanes from AA or EPA, and triggers the generation of epimeric forms of lipoxins (from AA), resolvins E (from EPA), resolvins D and neuroprotectins (from DHA), via lipooxygenases. COX-2 can also directly metabolize DHA to anti-inflammatory oxo-derivatives (13-EFOX), with 17-EFOX being the aspirin-triggered form. 15R-HETE: 15R-hydroxy-eicosatetraenoic acid; 18R-HEPE: 18R-hydroxy-eicosapentaenoic acid; 17R-HDHA: 17R-hydroxy-docosahexaenoic acid.

Another ω-3 PUFA, EPA, present at low concentrations in membrane phospholipids, acts as a substrate for COX-2 for the synthesis of an alternative family of eicosanoids called 3 series-PGs and TX, which are typically less potent than their AA-derived counterparts, because they act as weak agonists at their respective eicosanoid receptors (Fig. 3) [40].

COX-1 is an important player during neuroinflammation

COX-1 recently emerged as a prominent player in CNS neuroinflammation [7, 16, 17, 24, 4144]. In aged rats, brain levels of COX-1-derived TXB2 and COX-1 expression in the hippocampus are increased and may contribute to the increased brain susceptibility to inflammation and neurodegenerative diseases. Indeed, COX-1 has been shown to support inflammatory processes and facilitate pro-inflammatory PG upregulation in several models of neuroinflammation [16, 17, 24, 42, 43, 45]. Owing to its predominant localization in microglia, brain immune resident cells, COX-1-derived products seem to be particularly important in modulating the initial phase of the inflammatory response. A recent study also identified COX-1 expression in perivascular cells (a subset of brain resident macrophages), where its expression was increased in response to a systemic LPS challenge, suggesting a role for COX-1 in mediating the immune-to-brain signaling [7]. Our group showed that COX-1 deficient mice have a decreased inflammatory response, leukocyte infiltration, oxidative stress and neuronal damage after central injection of LPS or Aβ1–42 [16, 24, 42], and these effects were associated with altered PG levels. LPS or Aβ1–42 –induced upregulation of brain PGE2, PGF and TXB2 was reduced by genetic deletion of COX-1 and by SC-560, a selective COX-1 inhibitor [24, 42]. In agreement with these results, Matousek and collaborators recently demonstrated that COX-1 expression and PGE2 were upregulated up to 2 months following a sustained IL-1β overexpression, a model of chronic neuroinflammation, and COX-1 genetic deletion or pharmacological inhibition completely abrogated IL-1β-mediated PGE2 increase [43]. Thus, COX-1 seems to play a role not only in acute but also in chronic neuroinflammatory conditions. In this regard, in a mouse animal model of Parkinson’s disease, associated with astrogliosis, microglia activation, inflammation and neurodegeneration, COX-1 expression was increased and COX-1 inhibition ameliorated mice survival [44]. Moreover, COX-1 inhibition or genetic deletion attenuated BBB disruption during TNF-α- or LPS-induced neuroinflammation [16, 45]. The exact mechanisms by which COX-1 is involved in neuroinflammation remain to be elucidated. Because COX-1 is predominantly present in microglia and perivascular cells, it can be hypothesized that the enzyme can be mobilized within seconds to minutes after an acute challenge, thereby contributing to the local eicosanoid pool at the site of inflammation, even before the induction of COX-2 expression (Fig. 2).

Overall, recent data imply that COX-1 plays a critical role in the process of neuroinflammation and neurodegeneration and that COX-2 may mediate either a neurotoxic or an anti-inflammatory role depending on the stimulus, the cell type targeted by the insult and the substrate used for eicosanoid synthesis.

Therapeutical implications for managing neuroinflammation in neurological and neurodegenerative disorders

It is becoming clear that neuroinflammation, mediated predominantly by microglia, represents an important component of several neurodegenerative diseases and may contribute to cognitive impairment in the elderly. Supporting this concept, several observational studies have shown that long term use of non-steroidal anti-inflammatory drugs (NSAIDs) reduces the incidence of AD and age-related dementia [46, 47]. However, treatment trials or secondary prevention trials, mostly using selective COX-2 inhibitors, in AD patients with mild to moderate cognitive impairment failed to show any benefit, except for a small trial with indomethacin, a COX-1 preferential inhibitor [46, 48]. The discrepancy between the epidemiological studies and the clinical trials may be ascribed to several reasons, including timing of administration, off-target effects, and selectivity of the NSAIDs used. One possible explanation for the failure of clinical trials is that targeting COX-2 after the onset of dementia, when the inflammatory cascade has already started, is not effective, and treatment should begin before the onset of AD symptoms. Indeed, the first and only primary prevention trial (ADAPT) to test the effect of celecoxib (a selective COX-2 inhibitor) and naproxen (a mixed COX-2/COX-1 inhibitor) on AD incidence did not show any reduction in AD risks, and even suggested detrimental effects on cognitive function in the treated groups compared to placebo [49, 50]. Because of cardiovascular safety concerns, the drugs were stopped when the average treatment duration was two years. However, subjects were then followed for two more years. After exclusion of individuals with baseline cognitive syndromes, the later hazard ratio for the naproxen-treated group was significantly reduced (0.33; 95% CI 0.11 – 0.98). A neuroprotective effect of naproxen was also evident in CSF, where the tau/Aβ ratio was reduced by more than 40% in the group originally assigned to naproxen. This effect was unabated 41 months following treatment termination, and it remained apparent in a smaller sample of CSF collected an average 13 months after the first. None of these results were seen in those assigned to celecoxib (John Breitner, personal communication). While these findings deserve further investigation, they suggest that NSAIDs can accelerate AD pathogenesis in patients with relatively advanced disease, but naproxen appears to protect against neurodegeneration and incidence of dementia in the subset of patients with healthier brains. Thus, a single agent produced contrasting effects at different stages of AD pathogenesis.

The different distribution patterns of COX-1 and COX-2 expression in the AD brain could implicate that these enzymes are involved in different cellular processes in the pathogenesis of AD. Indeed, COX-1 is predominantly expressed by microglia in rodent and human brain [6] and may be involved in the neuroinflammatory process associated with AD. COX-1 expressing microglia surround amyloid plaques and are increased during AD [6, 8]. In contrast, COX-2 expression has not been detected in microglia or astrocytes in the AD brain. Thus, we propose that the beneficial effects of NSAIDs in preventing AD could be due to the inhibition of COX-1 rather than COX-2. An increase in COX-1 expressing microglia and/or COX-1 expression has been observed also in other neurological diseases with a marked inflammatory component, such as traumatic brain injury, Creutzfeld-Jacob disease, HIV-related dementia, and Parkinson’s disease, where COX-1 expression often correlated with microglia activation, disease severity and tissue damage (reviewed in [41]).

On the other hand, the discovery of novel COX-2-derived lipid mediators with pro-resolving properties suggests that inhibiting COX-2 could be detrimental in certain neuroinflammatory conditions. Selective COX-2 inhibitors (coxibs) were developed and preferred to classical, mixed NSAIDs because of the lower gastrointestinal toxicity, attributed to the inhibition of COX-1-derived PGs. However, animal studies have shown that specific inhibition of COX-1 alone does not alter gastric or intestinal mucosal integrity, and gastrotoxicity was observed when both COX-1 and COX-2 were inhibited [5153]. Unfortunately, the failure of the clinical trials with COX-2 selective inhibitors has hindered appropriate follow up on the potential use of COX-1 preferential inhibitors. The only clinical trial with low dose aspirin treatment over 2 years (AD2000) did not demonstrate any benefit in patients with mild to moderate probable AD, and reported hemorrhagic adverse events due to aspirin intake [54]. Low dose aspirin, within the range used for cardiovascular protection, is indeed associated with gastroduodenal toxicity, but several trials have shown that gastric side effects can be countered by concomitant use of proton pump inhibitors [55].

In summary, identifying the specific roles of COX-1 and COX-2 in CNS inflammation and their coupling to downstream receptors could lead to revisit current therapeutic strategies and open new approaches for the treatment of neuroinflammatory conditions. COX-1, traditionally considered merely as the homeostatic isoform, is instead actively involved in brain injury induced by pro-inflammatory stimuli such as Aβ, LPS, IL-1β and TNF-α. On the other hand, a major breakthrough in the lipid field has uncovered novel COX-2 derived lipid mediators with anti-inflammatory and pro-resolution properties, which may underlie recent reports demonstrating COX-2 neuroprotective function during neuroinflammatory conditions. Therefore, NSAIDs with higher selectivity for COX-1 rather than COX-2 are more likely to reduce neuroinflammation and should be further investigated as a potential therapeutic approach in neurodegenerative diseases with a marked inflammatory component. The major drawback for the use of selective COX-1 inhibitor was the belief that they will cause more gastrointestinal toxicity than selective COX-2 inhibitors. It is now generally admitted that both COX-1 and COX-2 inhibition are required to elicit NSAID-induced gastric injury. As for many other drugs, the ratio benefit/risk should be weighed. Moreover, the novel pro-resolving functions of COX-2-derived mediators in the termination of inflammation should be considered in the design of innovative anti-inflammatory therapies.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Kis B, Snipes JA, Busija DW. Acetaminophen and the cyclooxygenase-3 puzzle: sorting out facts, fictions, and uncertainties. J Pharmacol Exp Ther. 2005;315:1–7. doi: 10.1124/jpet.105.085431. [DOI] [PubMed] [Google Scholar]
  • 2.O’Banion MK. Cyclooxygenase-2: molecular biology, pharmacology, and neurobiology. Crit Rev Neurobiol. 1999;13:45–82. doi: 10.1615/critrevneurobiol.v13.i1.30. [DOI] [PubMed] [Google Scholar]
  • 3.Morita I. Distinct functions of COX-1 and COX-2. Prostaglandins Other Lipid Mediat. 2002;68–69:165–175. doi: 10.1016/s0090-6980(02)00029-1. [DOI] [PubMed] [Google Scholar]
  • 4.Kang YJ, Mbonye UR, DeLong CJ, Wada M, Smith WL. Regulation of intracellular cyclooxygenase levels by gene transcription and protein degradation. Prog Lipid Res. 2007;46:108–125. doi: 10.1016/j.plipres.2007.01.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Funk CD, FitzGerald GA. COX-2 inhibitors and cardiovascular risk. J Cardiovasc Pharmacol. 2007;50:470–479. doi: 10.1097/FJC.0b013e318157f72d. [DOI] [PubMed] [Google Scholar]
  • 6.Hoozemans JJ, Rozemuller AJ, Janssen I, De Groot CJ, Veerhuis R, Eikelenboom P. Cyclooxygenase expression in microglia and neurons in Alzheimer’s disease and control brain. Acta Neuropathol (Berl) 2001;101:2–8. doi: 10.1007/s004010000251. [DOI] [PubMed] [Google Scholar]
  • 7.Garcia-Bueno B, Serrats J, Sawchenko PE. Cerebrovascular cyclooxygenase-1 expression, regulation, and role in hypothalamic-pituitary-adrenal axis activation by inflammatory stimuli. J Neurosci. 2009;29:12970–12981. doi: 10.1523/JNEUROSCI.2373-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Yermakova AV, Rollins J, Callahan LM, Rogers J, O’Banion MK. Cyclooxygenase-1 in human Alzheimer and control brain: quantitative analysis of expression by microglia and CA3 hippocampal neurons. J Neuropathol Exp Neurol. 1999;58:1135–1146. doi: 10.1097/00005072-199911000-00003. [DOI] [PubMed] [Google Scholar]
  • 9.Wang H, Hitron IM, Iadecola C, Pickel VM. Synaptic and vascular associations of neurons containing cyclooxygenase-2 and nitric oxide synthase in rat somatosensory cortex. Cereb Cortex. 2005;15:1250–1260. doi: 10.1093/cercor/bhi008. [DOI] [PubMed] [Google Scholar]
  • 10.Stefanovic B, Bosetti F, Silva AC. Modulatory role of cyclooxygenase-2 in cerebrovascular coupling. Neuroimage. 2006;32:23–32. doi: 10.1016/j.neuroimage.2006.03.014. [DOI] [PubMed] [Google Scholar]
  • 11.Hewett SJ, Bell SC, Hewett JA. Contributions of cyclooxygenase-2 to neuroplasticity and neuropathology of the central nervous system. Pharmacol Ther. 2006;112:335–357. doi: 10.1016/j.pharmthera.2005.04.011. [DOI] [PubMed] [Google Scholar]
  • 12.Slanina KA, Schweitzer P. Inhibition of cyclooxygenase-2 elicits a CB1-mediated decrease of excitatory transmission in rat CA1 hippocampus. Neuropharmacology. 2005;49:653–659. doi: 10.1016/j.neuropharm.2005.04.019. [DOI] [PubMed] [Google Scholar]
  • 13.Cowley TR, Fahey B, O’Mara SM. COX-2, but not COX-1, activity is necessary for the induction of perforant path long-term potentiation and spatial learning in vivo. The European journal of neuroscience. 2008;27:2999–3008. doi: 10.1111/j.1460-9568.2008.06251.x. [DOI] [PubMed] [Google Scholar]
  • 14.Bosetti F. Arachidonic acid metabolism in brain physiology and pathology: lessons from genetically altered mouse models. J Neurochem. 2007;102:577–586. doi: 10.1111/j.1471-4159.2007.04558.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Gilroy DW, Colville-Nash PR, Willis D, Chivers J, Paul-Clark MJ, Willoughby DA. Inducible cyclooxygenase may have anti-inflammatory properties. Nat Med. 1999;5:698–701. doi: 10.1038/9550. [DOI] [PubMed] [Google Scholar]
  • 16.Choi SH, Aid S, Choi U, Bosetti F. Cyclooxygenases-1 and -2 differentially modulate leukocyte recruitment into the inflamed brain. Pharmacogenomics J. 2009 doi: 10.1038/tpj.2009.68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Aid S, Silva AC, Candelario-Jalil E, Choi SH, Rosenberg GA, Bosetti F. Cyclooxygenase-1 and -2 differentially modulate lipopolysaccharide-induced blood-brain barrier disruption through matrix metalloproteinase activity. J Cereb Blood Flow Metab. 2010;30:370–380. doi: 10.1038/jcbfm.2009.223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Blais V, Turrin NP, Rivest S. Cyclooxygenase 2 (COX-2) inhibition increases the inflammatory response in the brain during systemic immune stimuli. J Neurochem. 2005;95:1563–1574. doi: 10.1111/j.1471-4159.2005.03480.x. [DOI] [PubMed] [Google Scholar]
  • 19.Tuo J, Tuaillon N, Shen D, Chan CC. Endotoxin-induced uveitis in cyclooxygenase-2-deficient mice. Investigative ophthalmology & visual science. 2004;45:2306–2313. doi: 10.1167/iovs.03-0756. [DOI] [PubMed] [Google Scholar]
  • 20.Aid S, Langenbach R, Bosetti F. Neuroinflammatory response to lipopolysaccharide is exacerbated in mice genetically deficient in cyclooxygenase-2. Journal of neuroinflammation [electronic resource] 2008;5:17. doi: 10.1186/1742-2094-5-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Bonow RH, Aid S, Zhang Y, Becker KG, Bosetti F. The brain expression of genes involved in inflammatory response, the ribosome, and learning and memory is altered by centrally injected lipopolysaccharide in mice. Pharmacogenomics J. 2009;9:116–126. doi: 10.1038/tpj.2008.15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Keene CD, Chang R, Stephen C, Nivison M, Nutt SE, Look A, Breyer RM, Horner PJ, Hevner R, Montine TJ. Protection of hippocampal neurogenesis from toll-like receptor 4-dependent innate immune activation by ablation of prostaglandin E2 receptor subtype EP1 or EP2. Am J Pathol. 2009;174:2300–2309. doi: 10.2353/ajpath.2009.081153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Lehnardt S, Massillon L, Follett P, Jensen FE, Ratan R, Rosenberg PA, Volpe JJ, Vartanian T. Activation of innate immunity in the CNS triggers neurodegeneration through a Toll-like receptor 4-dependent pathway. Proc Natl Acad Sci U S A. 2003;100:8514–8519. doi: 10.1073/pnas.1432609100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Choi SH, Langenbach R, Bosetti F. Genetic deletion or pharmacological inhibition of cyclooxygenase-1 attenuate lipopolysaccharide-induced inflammatory response and brain injury. FASEB J. 2008;22:1491–1501. doi: 10.1096/fj.07-9411com. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Candelario-Jalil E, Fiebich BL. Cyclooxygenase inhibition in ischemic brain injury. Curr Pharm Des. 2008;14:1401–1418. doi: 10.2174/138161208784480216. [DOI] [PubMed] [Google Scholar]
  • 26.Teismann P, Tieu K, Choi DK, Wu DC, Naini A, Hunot S, Vila M, Jackson-Lewis V, Przedborski S. Cyclooxygenase-2 is instrumental in Parkinson’s disease neurodegeneration. Proc Natl Acad Sci U S A. 2003;100:5473–5478. doi: 10.1073/pnas.0837397100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Aid S, Parikh N, Palumbo S, Bosetti F. Neuronal overexpression of cyclooxygenase-2 does not alter the neuroinflammatory response during brain innate immune activation. Neurosci Lett. 478:113–118. doi: 10.1016/j.neulet.2010.04.076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Serhan CN, Chiang N, Van Dyke TE. Resolving inflammation: dual anti-inflammatory and pro-resolution lipid mediators. Nat Rev Immunol. 2008;8:349–361. doi: 10.1038/nri2294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Lawrence T, Willoughby DA, Gilroy DW. Anti-inflammatory lipid mediators and insights into the resolution of inflammation. Nat Rev Immunol. 2002;2:787–795. doi: 10.1038/nri915. [DOI] [PubMed] [Google Scholar]
  • 30.Serhan CN, Yacoubian S, Yang R. Anti-inflammatory and proresolving lipid mediators. Annu Rev Pathol. 2008;3:279–312. doi: 10.1146/annurev.pathmechdis.3.121806.151409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Schwab JM, Chiang N, Arita M, Serhan CN. Resolvin E1 and protectin D1 activate inflammation-resolution programmes. Nature. 2007;447:869–874. doi: 10.1038/nature05877. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Chan MM, Moore AR. Resolution of inflammation in murine autoimmune arthritis is disrupted by cyclooxygenase-2 inhibition and restored by prostaglandin E2-mediated lipoxin A4 production. J Immunol. 184:6418–6426. doi: 10.4049/jimmunol.0903816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Gilroy DW, Colville-Nash PR, McMaster S, Sawatzky DA, Willoughby DA, Lawrence T. Inducible cyclooxygenase-derived 15-deoxy (Delta)12-14PGJ2 brings about acute inflammatory resolution in rat pleurisy by inducing neutrophil and macrophage apoptosis. Faseb J. 2003;17:2269–2271. doi: 10.1096/fj.02-1162fje. [DOI] [PubMed] [Google Scholar]
  • 34.Groeger AL, Cipollina C, Cole MP, Woodcock SR, Bonacci G, Rudolph TK, Rudolph V, Freeman BA, Schopfer FJ. Cyclooxygenase-2 generates anti-inflammatory mediators from omega-3 fatty acids. Nat Chem Biol. 6:433–441. doi: 10.1038/nchembio.367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Bazan NG. Cell survival matters: docosahexaenoic acid signaling, neuroprotection and photoreceptors. Trends Neurosci. 2006;29:263–271. doi: 10.1016/j.tins.2006.03.005. [DOI] [PubMed] [Google Scholar]
  • 36.Serhan CN, Hong S, Gronert K, Colgan SP, Devchand PR, Mirick G, Moussignac RL. Resolvins: a family of bioactive products of omega-3 fatty acid transformation circuits initiated by aspirin treatment that counter proinflammation signals. J Exp Med. 2002;196:1025–1037. doi: 10.1084/jem.20020760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Serhan CN, Clish CB, Brannon J, Colgan SP, Chiang N, Gronert K. Novel functional sets of lipid-derived mediators with antiinflammatory actions generated from omega-3 fatty acids via cyclooxygenase 2-nonsteroidal antiinflammatory drugs and transcellular processing. J Exp Med. 2000;192:1197–1204. doi: 10.1084/jem.192.8.1197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Hong S, Gronert K, Devchand PR, Moussignac RL, Serhan CN. Novel docosatrienes and 17S-resolvins generated from docosahexaenoic acid in murine brain, human blood, and glial cells. Autacoids in anti-inflammation. J Biol Chem. 2003;278:14677–14687. doi: 10.1074/jbc.M300218200. [DOI] [PubMed] [Google Scholar]
  • 39.Niemoller TD, Stark DT, Bazan NG. Omega-3 fatty acid docosahexaenoic acid is the precursor of neuroprotectin D1 in the nervous system. World Rev Nutr Diet. 2009;99:46–54. doi: 10.1159/000192994. [DOI] [PubMed] [Google Scholar]
  • 40.Calder PC. Polyunsaturated fatty acids and inflammatory processes: New twists in an old tale. Biochimie. 2009;91:791–795. doi: 10.1016/j.biochi.2009.01.008. [DOI] [PubMed] [Google Scholar]
  • 41.Choi SH, Aid S, Bosetti F. The distinct roles of cyclooxygenase-1 and -2 in neuroinflammation: implications for translational research. Trends Pharmacol Sci. 2009;30:174–181. doi: 10.1016/j.tips.2009.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Choi SH, Bosetti F. Cyclooxygenase-1 null mice show reduced neuroinflammation in response to beta-amyloid. Aging (Albany NY) 2009;1:234–244. doi: 10.18632/aging.100021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Matousek SB, Hein AM, Shaftel SS, Olschowka JA, Kyrkanides S, O’Banion MK. Cyclooxygenase-1 mediates prostaglandin E(2) elevation and contextual memory impairment in a model of sustained hippocampal interleukin-1beta expression. J Neurochem. 114:247–258. doi: 10.1111/j.1471-4159.2010.06759.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Gu XL, Long CX, Sun L, Xie C, Lin X, Cai H. Astrocytic expression of Parkinson’s disease-related A53T alpha-synuclein causes neurodegeneration in mice. Mol Brain. :3–12. doi: 10.1186/1756-6606-3-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Candelario-Jalil E, Taheri S, Yang Y, Sood R, Grossetete M, Estrada EY, Fiebich BL, Rosenberg GA. Cyclooxygenase inhibition limits blood-brain barrier disruption following intracerebral injection of tumor necrosis factor-alpha in the rat. J Pharmacol Exp Ther. 2007;323:488–498. doi: 10.1124/jpet.107.127035. [DOI] [PubMed] [Google Scholar]
  • 46.McGeer PL, McGeer EG. NSAIDs and Alzheimer disease: epidemiological, animal model and clinical studies. Neurobiol Aging. 2007;28:639–647. doi: 10.1016/j.neurobiolaging.2006.03.013. [DOI] [PubMed] [Google Scholar]
  • 47.Vlad SC, Miller DR, Kowall NW, Felson DT. Protective effects of NSAIDs on the development of Alzheimer disease. Neurology. 2008;70:1672–1677. doi: 10.1212/01.wnl.0000311269.57716.63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Rogers J, Kirby LC, Hempelman SR, Berry DL, McGeer PL, Kaszniak AW, Zalinski J, Cofield M, Mansukhani L, Willson P, et al. Clinical trial of indomethacin in Alzheimer’s disease. Neurology. 1993;43:1609–1611. doi: 10.1212/wnl.43.8.1609. [DOI] [PubMed] [Google Scholar]
  • 49.Group AR, Lyketsos CG, Breitner JC, Green RC, Martin BK, Meinert C, Piantadosi S, Sabbagh M. Naproxen and celecoxib do not prevent AD in early results from a randomized controlled trial. Neurology. 2007;68:1800–1808. doi: 10.1212/01.wnl.0000260269.93245.d2. [DOI] [PubMed] [Google Scholar]
  • 50.Martin BK, Szekely C, Brandt J, Piantadosi S, Breitner JC, Craft S, Evans D, Green R, Mullan M. Cognitive function over time in the Alzheimer’s Disease Anti-inflammatory Prevention Trial (ADAPT): results of a randomized, controlled trial of naproxen and celecoxib. Archives of neurology. 2008;65:896–905. doi: 10.1001/archneur.2008.65.7.nct70006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Wallace JL, McKnight W, Reuter BK, Vergnolle N. NSAID-induced gastric damage in rats: requirement for inhibition of both cyclooxygenase 1 and 2. Gastroenterology. 2000;119:706–714. doi: 10.1053/gast.2000.16510. [DOI] [PubMed] [Google Scholar]
  • 52.Tanaka A, Hase S, Miyazawa T, Ohno R, Takeuchi K. Role of cyclooxygenase (COX)- 1 and COX-2 inhibition in nonsteroidal anti-inflammatory drug-induced intestinal damage in rats: relation to various pathogenic events. J Pharmacol Exp Ther. 2002;303:1248–1254. doi: 10.1124/jpet.102.041715. [DOI] [PubMed] [Google Scholar]
  • 53.Menozzi A, Pozzoli C, Giovannini E, Solenghi E, Grandi D, Bonardi S, Bertini S, Vasina V, Coruzzi G. Intestinal effects of nonselective and selective cyclooxygenase inhibitors in the rat. Eur J Pharmacol. 2006;552:143–150. doi: 10.1016/j.ejphar.2006.08.089. [DOI] [PubMed] [Google Scholar]
  • 54.Bentham P, Gray R, Sellwood E, Hills R, Crome P, Raftery J. Aspirin in Alzheimer’s disease (AD2000): a randomised open-label trial. Lancet Neurol. 2008;7:41–49. doi: 10.1016/S1474-4422(07)70293-4. [DOI] [PubMed] [Google Scholar]
  • 55.Scheiman JM. NSAID-induced peptic ulcer disease: a critical review of pathogenesis and management. Dig Dis. 1994;12:210–222. doi: 10.1159/000171455. [DOI] [PubMed] [Google Scholar]

RESOURCES