Abstract
Biologically active epoxyeicosatrienoic acid regioisomers (EETs) are synthesized from arachidonic acid by cytochrome P450 epoxygenases of endothelial, myocardial and renal tubular cells. EETs relax vascular smooth muscle and decrease inflammatory cell adhesion and cytokine release. Renal EETs promote sodium excretion and vasodilation to decrease hypertension. Cardiac EETs reduce infarct size following ischemia-reperfusion injury and decrease fibrosis and inflammation in heart failure. In diabetes, EETs improve insulin sensitivity, increase glucose tolerance and reduce the renal injury. These actions of EETs emphasize their therapeutic potential. To minimize metabolic inactivation, 14,15-EET agonist analogs with stable epoxide bioisosteres and carboxyl surrogates were developed. In pre-clinical rat models, a subset of agonist analogs, termed EET-A, EET-B and EET-C22, are orally active with good pharmacokinetic properties. These orally active EET agonists lower blood pressure and reduce cardiac and renal injury in spontaneous and angiotensin hypertension. Other beneficial cardiovascular actions include improved endothelial function and cardiac anti-remodeling actions. In rats, EET analogs effectively combat acute and chronic kidney disease including drug- and radiation-induced kidney damage, hypertension and cardiorenal syndrome kidney damage, and metabolic syndrome and diabetes nephropathy. The compelling pre-clinical efficacy supports the prospect of advancing EET analogs to human clinical trials for kidney and cardiovascular diseases.
Keywords: epoxyeicosatrienoic acid, hypertension, diabetes, epoxygenase, inflammation, heart failure, kidney injury
Introduction
Epoxyeicosatrienoic acids (EETs) are cytochrome P450 (CYP) metabolites of arachidonic acid (1–3). Four regioisomeric EETs are biosynthesized. They differ in the placement of the epoxide at the cis-double bonds of arachidonic acid resulting in 14,15-, 11,12-, 8,9- and 5,6-EET. A limited number of isozymes of the CYP gene family catalyze EET biosynthesis, and these isozymes vary in their metabolic repertoire, i.e., relative abundance and stereochemistry, of the EET regioisomers produced (4–7). Additionally, CYP isozyme expression varies with gender, cell type, metabolic state, starvation, exposure to xenobiotics and other factors. The biological significance of these differences in EET synthesis is not completely understood. As with other eicosanoids, activation of phospholipases and the release of arachidonic acid from membrane phospholipids is the step regulating cellular EET synthesis and release (8)(Figure 1). Cell injury, shear stress and hormones such as bradykinin, acetylcholine, angiotensin II and adenosine are among the phospholipase activators that initiate EET synthesis (9–14). CYP epoxygenases are expressed and EETs are synthesized in cells and tissues of cardiovascular importance. Some of these cells include endothelial cells of arteries and veins, cardiac myocytes, renal tubular epithelial cells, macrophages and adrenal zona glomerulosa cells (15–19).
Figure 1.

EET synthesis and metabolism pathways
EETs function as local, autocrine or paracrine mediators (13, 20). They are synthesized by one cell type and exert their biological activity on the same or a different adjacent cell type. For example, in arteries, endothelial cells synthesize EETs that may stimulate endothelial cell growth and migration or relax adjacent smooth muscle cells (9, 21, 22). In some cases, EET regioisomers and stereoisomers have differing biological activities (21–24). For many activities, only a single racemic EET regioisomer has been tested. Thus, the pharmacology of the EETs is incomplete.
Several pathways terminate the biological activity of the EETs (8) (Figure 1). The EET regioisomers and stereoisomers vary in their inactivation by these pathways (25–28). EETs are fatty acids with a free carboxyl group at carbon-1 so are substrates for β-oxidation and esterification. β-Oxidation results in shortening of the 20 carbon backbone to 18-, 16 or 14-carbon metabolites and loss of biological activity (8, 23, 25). This process is relatively slow requiring hours for complete metabolism in cultured cells. EETs are rapidly esterified into cellular phospholipids (8, 25, 26). Free EETs are biologically active, so esterification diminishes the concentration of free EETs and terminates activity. Esterified EETs may make subtle changes in the lipid environment of cell membranes and affect cell function. Activation of phospholipases will release the esterified EETs and restore the active form. Soluble epoxide hydrolase (sEH) hydrates the epoxide group of the EETs to produce the corresponding vicinal diols or dihydroxyeicosatrienoic acids (DHETs) (27, 29–31). This metabolic pathway is also rapid. In most instances, the cis-epoxide group is required for full biological activity, and the DHETs are inactive or less active. There are activities described for the EETs where the corresponding DHETs have not been tested so caution should be used in generalizing about EETs being active and DHETs not. ω-Oxidation of the EETs and EET conjugation to glutathione are documented but have not been extensively studied (32, 33).
Summary of Cardiovascular Actions of the EETs
The cardiovascular actions of the EETs are briefly summarized. However, more detailed reviews are available (8, 20, 31, 34–37).
Regulation of vascular tone
EETs function as endothelium-derived hyperpolarizing factors (EDHFs) in arteries of humans and experimental animals including cows and pigs (9, 13, 18, 38). EETs are synthetized by endothelial cells, but not smooth muscle cells, and are released in response to bradykinin, acetylcholine and shear stress (9, 10, 13, 18, 39). Endothelial cells contain CYP 2C8, 2C9 and 2J2 and release predominately 14,15- and 11,12-EET. These EETs are transferred to the smooth muscle where they activate the large conductance, calcium-activated potassium (BKCa) channel through a guanine nucleotide binding (G) protein-mediated mechanism (14, 40, 41). This increases K efflux and hyperpolarizes the smooth muscle cell membrane resulting in relaxation. The four EET regioisomers are equally active in vasorelaxation (9). The relaxations to bradykinin and acetylcholine are endothelium-dependent whereas relaxations to the EETs are endothelium-independent (42, 43). These relaxation are inhibited by the BKCa blockers iberiotoxin and charybdotoxin, the CYP inhibitors miconazole, MS-PPOH and CYP2C antisense oligonucleotides and the EET antagonist, 14,15-epoxyeicosa-5Z-enoic acid (14,15-EE5ZE) (9, 18, 38, 43, 44). In normal and hypertensive subjects treated with aspirin and NG-monomethyl-L-arginine to block vascular prostaglandin and nitric oxide synthesis, bradykinin increases forearm blood flow that is inhibited by the CYP inhibitor sulfaphenazole (45). Other studies document the dilator role of EETs in humans and human arteries (38, 43, 46).
Renal blood flow and sodium excretion
In rats, 11,12-EET dilates renal arteries whereas 14,15-EET is less active (47). CYP is expressed in the thick ascending limb and collecting duct and converts arachidonic acid to EETs (48–51). Endogenous and exogenous EETs inhibit the epithelial sodium channel (ENaC) resulting in sodium excretion (48, 50, 52, 53). Deletion of CYP2c44 in mice increases ENaC activity, sodium retention and blood pressure (50, 52, 54). The blood pressure is lowered by the ENaC inhibitor amiloride. Thus, renal tubular EETs may also regulate blood pressure.
Hypertension
Since EETs cause vasodilation and increase renal sodium excretion, an anti-hypertension action is predictable. Inhibition of sEH lowers blood pressure in spontaneously hypertensive (SH) rats as well as rats with angiotensin II- and deoxycorticosterone acetate/salt-hypertension (55–59). Similarly, overexpression of CYPs reduces blood pressure in SH rats and hypertensive mice (56, 60).
Cardioprotection
EETs reduce infarct size and improve functional recovery of the myocardium after ischemia in mice, rats, rabbits and dogs (61–65). In perfused mouse hearts, global ischemia followed by reperfusion produces contractile dysfunction (62, 63). Increasing endogenous EETs by increased CYP expression or reducing the activity of sEH diminishes the contractile dysfunction. This is blocked by CYP inhibition or the EET antagonist 14,15-EE5ZE. Without treatment, thirty minutes of ischemia and 2 hours of reperfusion produce a 55% infarct (65). Both 11,12- and 14,15-EET are equipotent in reducing the size of the infarct to 31% in anesthetized rats. Similar results were obtained in anesthetized dogs (64). Importantly, exogenous EETs are effective when given before ischemia or, most notably, at reperfusion. Inhibition of ATP-sensitive K (KATP) channels with glibenclamide blocks the beneficial effects of EETs in mice, rats and dogs (63–65). Other studies implicate the glycogen synthase kinase-3β, phosphatidylinositol-3 (PI-3)-kinase, mitogen-activated protein (MAP) kinase and reactive oxygen species in the cardioprotection afforded by EETs (61, 63, 65).
Heart failure
sEH was identified as a heart failure susceptibility gene in SH heart failure rats (66). In mice, inhibition of sEH increases endogenous EETs, improves pressure overload heart failure and reduces the incidence of arrhythmias (67). Similarly, cardiac overexpression of CYP2J2 reduces mortality and arrhythmias in transverse aortic constriction and isoproterenol infusion models of heart failure compared to wild type mice (68). Angiotensin-induced cardiac fibrosis, inflammation and hypertrophy are also reduced in CYP2J2 mice (69, 70). Thus, endogenous EETs are beneficial in models of both systolic and diastolic heart failure.
Platelet adhesion and fibrinolysis
EETs inhibit expression of the adhesion molecule P-selectin on human platelets under basal conditions and with ADP-stimulation (71). Platelet adhesion to human endothelial cells is decreased. This is mediated by KCa channel activation and hyperpolarization of the platelet membrane. 11,12-EET is more active than 8,9- or 14,15-EET. The concentrations required for activity are μM. In human and bovine endothelial cells, the four EET regioisomers in nM concentrations increase the expression and release of tissue plasminogen activator (tPA); however, 11,12-EET is the most active (72). EET release of tPA requires a guanine nucleotide binding protein, Gαs, and an increase in cyclic AMP. Thus, EETs inhibit platelet adhesion and stimulate thrombolysis.
Inhibition of inflammation
EETs decrease inflammation by a variety of mechanisms (73, 74). Tumor necrosis factor-α (TNFα) increases the expression of vascular cell adhesion molecule-1 (VCAM-1), intercellular adhesion molecule-1 (ICAM-1) and E-selectin in human and bovine endothelial cells (74, 75). Interleukin-1α and lipopolysaccharide act similarly. 11,12-EET inhibits the expression of these adhesion molecules by all three inflammatory stimuli in nM concentrations. 8,9-EET and 5,6-EET are less active than 11,12-EET. 14,15-EET is without activity. The structural features of 11,12-EET required for this activity were defined using a series of analogs (24). TNFα also increases the adhesion of mononuclear cells to the endothelium in mice in vivo (75). It is also blocked by 11,12-EET, but not 14,15-EET. In endothelial cells, TNFα increases the nuclear accumulation of NFκB by increasing the phosphorylation IκB by IκB kinase (IKK) and its degradation. 11,12-EET inhibits IKK preventing NFκB translocation to the nucleus and transcription of adhesion molecules. Participation of peroxisome proliferator-activated receptor-γ (PPARγ) in EET inhibition of NFκB was suggested; however, several discrepancies question this mechanism (76). EETs bind and activate PPARγ in μM concentrations but inhibit NFκB in nM concentrations (75, 76). Also, 14,15-EET activates PPARγ but does not inhibit NFκB in endothelial cells. NFκB regulates numerous pro-inflammatory pathways including adhesion molecule expression, cytokine release and cyclooxygenase expression (77). Thus, it represents a key mechanism in the anti-inflammatory action of the EETs. Increasing endogenous EET production by increased expression of CYPs or inhibition of sEH increases IκB accumulation and inhibits the cytokine release and accumulation of inflammatory cells associated with renal, lung, hepatic and cardiac injury (57, 70, 78, 79).
Nuclear erythroid-related factor 2 (Nrf2) contributes to the resolution of inflammation by increasing the expression of antioxidant proteins and enzymes (80). 14,15-EET activates Nrf2 accumulation and nuclear translocation in lung epithelial cells (81). Also, 11,12-EET, 14,15-EET and CYP2C8 overexpression increases Nrf2 expression and reduces TNFα-induced endothelial cell reactive oxygen species production and apoptosis (82). These effects are blocked by transfection with a Nrf2 small interfering RNA. While less studied than NFκB, Nrf2 may represent an important pathway for the anti-inflammatory actions of EETs.
Angiogenesis and wound healing
Endogenous and exogenous EETs promote angiogenesis, thus increasing endothelial cells growth, migration and tube formation (21, 22, 83–85). These in vitro activities require μM concentrations of the EETs. The regioisomer(s) causing these activities vary with the vascular source and species of endothelial cells. 11,12-EET is the most extensively studied. Angiogenesis is an important mechanism contributing to EET accelerating wound healing and organ regeneration (86, 87).
Diabetes
Endogenous EETs contribute to glucose tolerance, insulin sensitivity and glucose uptake as well as reduce the kidney and liver damage and improve the wound healing in diabetes. Increasing endogenous EETs by increasing CYP or decreasing sEH reduces blood pressure, reverses insulin resistance and increases glucose tolerance in fructose-treated rats and db/db mice (88–90). Renal injury and inflammation are reduced in streptozotocin-induced diabetic mice and FVB mice on a high fat diet (91, 92). In contrast, glucose tolerance, insulin sensitivity and glucose uptake are decreased in CYP2c44 deleted mice compared to wild type mice (93). Plasma EET concentrations positively correlates with insulin sensitivity in human subjects (93). 11,12-EET also improves wound healing in ob/ob mice by decreasing accumulation of inflammatory cells, reducing cytokine production and increasing angiogenesis (87). Similarly, increases in CYP expression in db/db mice decrease hepatic inflammation by the same mechanisms (89).
EET Receptors and Signaling Mechanisms
As indicated above, a number of cell signaling pathways are implicated in the actions of the EETs (8, 37). For example, KCa channel activation is involved in relaxation of vascular smooth muscle and inhibition of platelet adhesion to the endothelium (41, 71). EETs activate BKCa channels and adenylyl cyclase/cyclic AMP accumulation through Gαs (21, 41, 72). Endothelial tPA release, Trp channels and KCa channel activation as well as inhibition of smooth muscle cell migration use cyclic AMP as a second messenger (21, 72, 94, 95). NFκB inhibition, possibly Nrf2 activation, mediates the anti-inflammatory activity of the EETs (75, 81, 82). EETs also increase the ERK1/2 and Akt kinase pathways to promote cell growth (83). Other signaling pathways have been described (37, 61, 63, 65).
Characterization of EET receptors has lagged behind the discovery of signaling mechanisms. Clearly, defining the EET receptors is important to the therapeutic development of EET agonist analogs. A high affinity binding site for 3H-14,15-EET was described in mononuclear cells and membranes (96–98). The binding is specific, saturable and reversible. We designed 20-125I-14,15-EE8ZE and –EE5ZE as chemically stable, high specific activity agonist and antagonist radioligands, respectively (99–101). In cell membranes, specific, saturable, reversible and high affinity (Kd=8.5 nM) binding occurs with 20-125I-14,15-EE8ZE (101). 14,15- and 11,12-EET and 14,15-EE5ZE displaced the radioligand (IC50=12–40 nM) whereas 14,15-DHET and other inactive 14,15-EET analogs do not. Similar results were obtained for antagonist binding with 20-125I-14,15-EE5ZE (Kd=1.1 nM) (99).
The role of G-proteins in EET action has been investigated. The specific binding of 20-125I-14,15-EE8ZE is inhibited by GTPγS implicating G protein coupling to the binding (101). Using cell-attached patch clamp measurements of BKCa channels in smooth muscle cells, 14,15- and 11,12-EET increase K channel activity at 1–100 nM concentrations (9, 41). In contrast, the EETs are without activity in inside-out pulled off patches; however, GTP, but not ATP, restores EET-induced K channel activity. The EET activity was inhibited by an anti-Gαs, but not anti-Gαi or anti-Gβγ, specific antibodies. Similarly, in endothelial cells, an anti-Gαs antibody and small interfering RNAs inhibit EET activities (21, 72). Others have also confirmed that some EET actions are GTP-dependent and Gαs mediated (102, 103). These studies indicate that a high affinity 14,15-EET receptor exists. It is a GPCR coupled to Gαs in smooth muscle and endothelial cells. The identity of this 14,15-EET receptor is not known. There are no studies documenting high affinity EET receptors for 11,12- or 8,9-EET. However, PPARγ, GPR40, E prostanoid type 2 (EP2) and thromboxane prostanoid (TP) receptors have been described as possible receptors for EETs (76, 104–107). Their order of EET potency and requirement for μM EET concentrations for activation exclude them as the high affinity EET receptors.
Initially, differences in the rank order of potency of prostaglandins provided compelling evidence for multiple prostaglandin receptors (108). This may be the case for EETs as well, but the data comparing EET regioisomers are incomplete. For some biological activities, one regioisomer is more active than others; however, the regioisomers are equally active in other cases (9, 21, 22, 72, 95). This suggests that multiple EET receptors exist. Additionally, regioisomer-specific EET antagonists have been developed and implicate multiple EET receptors (44, 109–111). In coronary arteries, 14,15-EE5ZE inhibits the relaxations to all four EET regioisomers. In contrast, 14,15-dihydroxy-E5ZE inhibits relaxations to 14,15-EET without affecting 11,12- or 8,9-EET relaxations (109). 11,12,20-Trihydroxy-E8ZE inhibits relaxations to 11,12-EET, but not 14,15- or 8,9-EET (110). These data suggest there are separate EET receptors for 14,15-, 11,12- and 8,9-EET and indicate the need to investigate and characterize possible 11,12- and 8,9-EET receptors.
Therapeutic Considerations
The EET pathway may be enhanced by direct and indirect strategies that include increasing EET synthesis by induction of cellular CYP epoxygenase(s), decreasing EET degradation to DHETs by sEH inhibition and EET agonists. Each of these strategies has been used (31, 55, 112, 113). Increasing epoxygenases and inhibition of sEH increase the accumulation of all of the EETs. EET regioisomers differ in their actions so untoward or off-target effects may accompany the therapeutic benefit. Furthermore, inhibitors of sEH depend on EET biosynthesis to increase endogenous levels of EETs. Endogenous EET synthesis may vary with disease or concomitant drug ingestion and blunt or enhance the therapeutic effect of the sEH inhibitor. Orally active, metabolically stable EET agonist analogs represent an analog of a single regioisomer and are therefore expected to be more selective in their action. Also, the dose of the agonist analog is directly related to the therapeutic effect.
EET Agonist Analogs and Structure Activity Relationships (SARs)
We initiated a long-term, inter-laboratory program of EET analog development. Our goals included: (i) elucidation of SARs, (ii) improve potency, (iii) obviate autooxidation, (iv) extend in vivo half-life, (v) increase water solubility, and (vi) streamline chemical syntheses/reduce production costs. The earliest effort to develop more robust analogs involved simple modification of the carboxylic acid of 14,15-EET (1) (114). (The numbers in bold refer to the specific structures provided in the text.) The N-methylsulfonylamides of 14,15-EET (2) and 11,12-EET (3) were introduced to reduce β-oxidation and esterification of the parent EETs during in vitro studies of mitogenesis and tyrosine phosphorylation stimulation in renal epithelial cells (115). Analog 3 was subsequently utilized by Imig et al to demonstrate the activation of protein kinase A during afferent arteriolar vasodilation and again to evaluate renal microvascular reactivity during angiotensin II-induced hypertension (115, 116). However, 2 and 3 are rapidly hydrolyzed in plasma and other biological milieu and should be considered as primarily pro-drugs when used for in vivo applications.

Following the proposal by Campbell et al. that EETs were EDHFs in some species, the first generation of structural analogs was evaluated using a bovine coronary artery isometric tension assay and the results were compared with equimolar 1 (9, 23). These studies identified the carboxylic acid (cf., analog 4) and an intact epoxide (cf., analog 5) as essential to vascular relaxation. Systematic removal of the olefins beginning one at a time, then two at a time led to analog 6 which retained full vasorelaxation activity with respect to 1 whereas the Δ5,6-regioisomer 7 and fully saturated analog 8 proved to be antagonists. Analog 6, with its simplified structure and stability towards autoxidation, was selected as the lead scaffold for the next generation of analogs. Analog 7, due to its greater potency as a competitive antagonist compared with 8, has found wide utility as a pharmacological tool (44).

The impact of 14,15-EET stereochemistry upon vasodilation was also tested (23, 117). The 14(S),15(R)-enantiomer 9 was more potent than its antipode 10 which in turn was stronger than trans-14,15-EET 11. Other studies have also reported enantioselective-dependent EET responses and metabolism (27, 97).

A parallel study of 11,12-EET looking at the structural determinants for inhibition of TNF-α-induced VCAM-1 expression allowed us to propose a recognition or binding domain map (Figure 2) (24). Again, a cis-Δ8,9-olefin was critical for maximum activity. An ionizable (acidic) group at C(1) that participates in ionic bonding and hydrophobic regions flanking the epoxide/olefinic moieties were likewise important. Introduction of heteroatoms or bulky functional groups in these regions as well as at the ω-end proved detrimental.
Figure 2.

Proposed recognition/binding domain map of 11,12-EET
Substitution of the epoxide with an acyclic ether led to a series of analogs, e.g., 12 (NUDSA) that essentially maintained the atom sequence without the rigidity and susceptibility of the three-membered epoxide to sEH cleavage (118). This was counterbalanced by a partial loss of potency compared with 11,12-EET.

The second generation of analogs focused primarily on the search for more efficacious, sEH-resistant epoxide bioisosteres or surrogates (119). Candidates were evaluated for both vascular relaxation and inhibition of sEH (Table 1). The simple 1,3-disubstituted urea 13 was about equal to the acyclic ether 12; however, truncated versions such as 14 performed poorly so this approach was terminated. Mono-N-alkylation of the urea gave mixed results depending upon the location of introduction (15 vs 16); vasorelaxation of the di-N-alkylation version 17 seemed to share the negative and positive contributions in equal proportions. Notably, as the level of substitution increased (13 vs 17), the ability to inhibit sEH declined progressively and was essentially independent of the vasorelaxant effects. Regrettably, the impressive activity of thiourea analog 18 was dampened by its rapid metabolism in microsomal fractions. Since sulfur is considered a weaker hydrogen bond acceptor than oxygen, the improvement in ED50 for 18 compared with 13 would be consistent with coordination to a metal center in the epoxide binding region. Various urethanes were prepared as represented by 19, but none rose above the activity of urea 13. A series of amides were instructive: progressing from secondary amide 20 to the tertiary amides 21 or 22 saw a loss of more than two orders of magnitude in sEH inhibition. Importantly, increasing the steric size of the nitrogen substituent from N-methyl 21 to N-isopropyl 22 significantly boosted the vasorelaxation efficacy. On the other hand, reversing the placement order of the amide, 20 vs 23, almost abolished the analogs ability to relax the vessels (121). Oxamide 24 stands out as a good choice for a reagent that displays EET-like EDHF effects, yet has little sEH inhibitory activity. Unlike urea 17, N,N′-dimethyloxamide 25 was not as effective in either assay as its unsubstituted parent. A variety of oxygen-containing heterocycles, e.g., 26, were also evaluated, but with no promising results. The reader will appreciate that all of the epoxide bioisosteres in Table 1 are achiral, i.e., do not contain an asymmetric center. This architectural simplification, we anticipate, will lead to agonists capable of binding universally to receptors or recognition sites irrespective of the sites’ enantiomeric preferences.
Table 1.
Selected examples of second generation 14,15-EET analogs.a
| Compd | Analog | Vascular | Relax. | sEHi |
|---|---|---|---|---|
| % (10 μM) | ED50 (μM) | IC50 (nM) | ||
| 13 |
|
63 | 7.5 | 46 |
| 14 |
|
16 | >10 | 1355 |
| 15 |
|
59 | 7.6 | 71.5 |
| 16 |
|
88 | 3.2 | 1451 |
| 17 |
|
83 | 3.2 | 8484 |
| 18 |
|
99 | 1.5 | 770 |
| 19 |
|
53 | 9.3 | 3480 |
| 20 |
|
64.5 | 6.1 | 79 |
| 21 |
|
64.5 | 4.3 | 11194 |
| 22 |
|
100 | 1.7 | 10712 |
| 23 |
|
12 | >10 | 13877 |
| 24 |
|
89 | 1.7 | 58712 |
| 25 |
|
70 | 4.4 | 17622 |
| 26 |
|
11 | >10 | 7147 |
At 10 μM, 14,15-EET induces 85% of the maximum vasorelaxation and its ED50 is 2.2 μM. Recombinant human sEH was utilized.
The remaining major modification addressed replacement of the carboxylic acid and drew extensively from the large pool of established carboxylic acid bioisosteres (120). Collectively, they comprise the third generation of 14,15-EET analogs and were mainly targeted at minimizing or preventing β-oxidation and esterification while simultaneously promoting better water solubility, improving oral bioavailability, and extending in vivo half-life (121). Some typical examples are shown in Table 2. Small molecular weight amides such as asparate 27 were easily prepared from the parent carboxylic acid and reasonably functional while conversely phosphonate 28 was less effective despite its inherent greater acidity versus a carboxylic acid. Many small heterocyclic bioisosteres satisfied the principle goals of improved potency, water solubility, and oral bioavailability, e.g., 5-sulfonyl-1H,1,2,3-triazole 29, tetrazole 30, and oxathiadiazole-2-oxides 32 and 33. The latter two had virtually identical vasorelaxation effects, but starkly different behavior on sEH since the former contained a urea and the latter an N-isopropyl amide as their respective epoxide surrogates. Notwithstanding the wide use 2,4-thiazolidinedione as a bioisostere in the glitazone class of antidiabetic drugs and elsewhere, 31 was disappointing (122). In a significant departure from the basic carbon backbone utilized in the majority of examples, it is clear from the fused bicyclic example 34 than there is wide latitude in the design of EET agonists. With only a few exceptions, the reiterative approach ending in the third generation of analogs has delivered a cornucopia of drug candidates that are comparable or surpass the activity of 14,15-EET either with or without significant sEH inhibition. From these examples, a potential therapeutic should emerge as they are further vetted through advanced absorption, distribution, metabolism, excretion screens.
Table 2.
Representative third generation 14,15-EET analogs.a
| Compd | Analog | Vascular | Relax. | sEHi |
|---|---|---|---|---|
| % (10 μM) | ED50 (μM) | IC50 (nM) | ||
| 27 |
|
91 | 1.6 | 392 |
| 28 |
|
71 | >10 | >500 |
| 29 |
|
92 | 3.1 | 23 |
| 30 |
|
110 | 1.1 | 32 |
| 31 |
|
47 | >10 | 57 |
| 32 |
|
109 | 0.34 | 10 |
| 33 |
|
109 | 0.32 | >500 |
| 34 |
|
96 | 1.3 | >500 |
At 10 μM, 14,15-EET induces 85% of the maximum vasorelaxation and its ED50 is 2.2 μM. Recombinant human sEH was utilized.

Therapeutic Application of Orally Active EET Agonists in Pre-Clinical Rat Models
The therapeutic potential for manipulating EETs has been well recognized for the past decade with the development of sEH inhibitors that increase the EET to DHET ratio (31, 123). The promising biological actions of EETs in a number of pre-clinical disease models has led to interest in developing EET-based therapeutic strategies and inspired the development of sEH inhibitors (Figure 3). As a result, there have been human clinical trials for hypertension, diabetes, and chronic obstructive pulmonary disease with two sEH inhibitors (31, 124). Specifically, sEH inhibition improves endothelial function in obese smokers with chronic obstructive pulmonary disease (124). Despite this positive cardiovascular outcome, sEH inhibitors have not yet garnered approval for use in humans. Drawbacks to this approach include: (i) indirect increases in endogenous EETs and epoxides of other fatty acids, (ii) requirement for a functioning CYP epoxygenase for EET generation and (iii) chemical and metabolic lability of endogenous EETs. Our group developed EET agonist analogs with key features important for stability and bioavailability. Third generation of orally active EET agonist analogs demonstrate great potential as therapy for cardiovascular and kidney diseases in pre-clinical rat models (121). There have been no clinical studies with these EET agonist analogs.
Figure 3.

Therapeutic uses of orally active EET agonist analogs
The precise molecular and protein target for the EETs has yet to be identified which does not allow for target-based screening drug design approaches. As indicated in the EET receptor section above, identification of EET receptors is important and active research is ongoing in this area. Our group has taken a structure-activity approach to define the pharmacophore for EET agonist analogs. Cell-signaling mechanisms are extensively described for EETs. We have determined that EET analogs utilize these cell-signaling mechanisms and that EET antagonists block these actions (37, 121). The pharmacophore for EET agonists was used to develop second and third generation orally active EET agonist drug candidates through a phenotypic screening process (121). Interestingly, when comparing FDA drugs approved between 1999–2008, phenotypic screening was the most successful approach for first-in-class drugs whereas target-based screening was most successful for follower drugs (125). The phenotypic screening data for EET analogs in cardiovascular and kidney diseases by our group and others provide clear and convincing evidence that EET analogs are viable drug candidates.
Cardiovascular Diseases
Numerous drugs on the market exist for the treatment of cardiovascular diseases like hypertension and heart failure and include renin-angiotensin system blockers, diuretics, and beta-adrenergic blockers. Even with this being the case, there is significant need for better cardiovascular therapeutic agents and none of the drugs used for heart failure with reduced systolic function are effective in diastolic dysfunction. With an aging population and increased prevalence of metabolic syndrome, therapies for diastolic heart failure are needed. A major shortfall with current therapeutics is that co-morbid events, end organ damage, and cardiovascular-related mortality remain problematic (126–128). Endothelial dysfunction is predictive for cardiovascular events in pre-clinical animal models and humans (128). EET agonist analogs provide a novel therapeutic approach that improves endothelial function and decreases end organ damage in cardiovascular disease (37).
Initial studies with EET agonist analogs focused on hypertension. A second generation EET analog, NUDSA 12, was successfully used in vivo in rodents (129, 130). NUDSA was developed by evaluating the second generation of EET analogs for dilation of the afferent arteriole and mesenteric artery via activation of vascular smooth muscle cell KCa channels (118, 131). Modification of EET analogs was based on the SAR for afferent arteriolar dilation and pharmacological properties necessary for in vivo administration (129). A single intraperitoneal (i.p.) injection identified EET analogs that dramatically decrease blood pressure to normal values over 4 to 12 hours in rats with established angiotensin hypertension (129). Importantly, EET analogs with no or weak afferent arteriolar dilatory actions fail to lower blood pressure in angiotensin hypertensive rats (129). EET agonist analogs that lower blood pressure in angiotensin hypertension were then tested in SH rats. EET analogs demonstrate variability in lowering blood pressure in SH rats and require at least three days for a maximum blood pressure lowering effect (129). After discontinuing the EET analog, blood pressure rises to hypertensive levels over 4 to 5 days. These findings raise the possibility that mechanisms other than vasodilation are necessary for EET analogs to be antihypertensive. EETs and EET analogs inhibit ENaC and promote natriuresis and may serve as a longer-term mechanism to decrease blood pressure (113). The discovery that the injection of NUDSA achieves adequate therapeutic levels to lower blood pressure to normal in two hypertensive animal models was a major breakthrough (129). Additionally, in heme oxygenase 2 null mice that develop metabolic syndrome, NUDSA dramatically lowers blood pressure, lowers blood glucose, decreases body weight, and improves vascular function in these metabolic syndrome mice (130). Although NUDSA demonstrated great therapeutic potential, the major drawback is that NUDSA requires i.p. administration to be effective.
The design, synthesis, and screening of a series of approximately 50 EET analogs resulted in three novel orally active EET analogs (EET-A 27, EET-B 34, and EET-C22 30)(112, 121). These EET analogs were evaluated for their ability to relax coronary arteries and sEH inhibition. Potential EET analogs with therapeutic anti-hypertensive activity were determined in a similar manner in angiotensin hypertensive and SH rats (113). EET-A and EET-B demonstrate anti-hypertensive activity with i.p. injection (112, 113). In addition to vasodilator activity, EET-A inhibits sEH with an IC50 of 392 nM and EET-B has no sEH inhibitory activity below 30 μM (121). Oral administration of the EET analogs, EET-A or EET-B to SH and angiotensin II hypertensive rats for two weeks lowers blood pressure to normal levels and decreases urinary albumin excretion, an indicator for decreased renal injury (112, 113). Importantly, LC-MS-MS methods to measure plasma and tissue EET analog levels show that orally administered EET analogs achieve plasma levels that are sufficient for biologically activity (112, 132, 133). EET-A administered orally for two weeks prevents the development of angiotensin-mediated malignant hypertension in Cyp1a1-Ren-2 transgenic rats (134). However, when EET-A is administered for one week during the late stage of established malignant hypertension where significant renal damage is present, EET-A fails to lower blood pressure (134). A longer duration of EET analog treatment may be necessary to lower blood pressure at a late stage of malignant hypertension. In a thorough evaluation of EET-A, it vasodilates arterioles, inhibits ENaC activity in cultured collecting ducts, reduces ENaC expression in nephrons, and lowers blood pressure in Cyp2c44−/− mice with salt-sensitive hypertension (112, 113, 134). Thus, the blood pressure lowering actions for EET analogs depends on vasodilator and natriuretic actions. Additionally, these findings are consistent with EET analog antihypertensive actions being multi-factorial and not purely due to vasodilation and as such, requires days to weeks for maximal blood pressure lowering to occur in hypertensive pre-clinical animal models (112, 113). Another consistent finding in hypertension models treated with EET analogs is a marked decrease in renal macrophage infiltration and diminished monocyte chemoattractant protein-1 (MCP-1) levels (112, 113, 133). Although renal inflammation contributes to hypertension, the contribution of the renal anti-inflammatory actions of EET analogs to lowering blood pressure remains to be determined. More importantly, the biological actions of EET analogs have the potential to lower co-morbid events by decreasing end organ damage, improving endothelial function, and significantly reducing cardiovascular related mortality.
Cardioprotective actions for EETs and EET analogs depend on their multi-functional actions (135–137). Endogenous and exogenous EETs and sEH inhibitors are beneficial in cardiac ischemia reperfusion injury. Pathophysiological responses to myocardial ischemia reperfusion injury such as myocardial infarction and reduced left ventricular function are improved by EET treatment (136, 137). Importantly, EETs given after reperfusion improve left ventricular heart function (64, 65). Likewise, NUDSA improves cardiac function when given to mice after myocardial infarction (138). Mice were subjected to left anterior descending artery ligation, and five days later NUDSA treatment was started and continued for four weeks. By echocardiography and cardiac histological analysis, NUDSA treatment returns left ventricular end diastolic area and fractional area to normal. More importantly, cardiac fibrosis and measures of diastolic dysfunction significantly decreased at four weeks following myocardial infarction with NUDSA (138). EET agonist analog treatment attenuates the progression of congestive heart failure and cardiac fibrosis following myocardial infarction induced by a 30 minute left coronary artery occlusion in SH rats (139). EET-B was administered for seven weeks following a myocardial infarction. It diminishes congestive heart failure-induced lung edema and dramatically reduces myocardial fibrosis at the ischemic zone (139). Cardioprotection by EETs and EET analogs depends on activation of pro-survival mechanisms, opposing apoptosis, and mitochondrial protection. EET analogs increase OPA1 oligomers and mitochondrial cristae density, activate mitochondrial KATP channels, and activate MAP kinase pro-survival signaling mechanisms to prevent cardiac cell apoptosis following an ischemic stress (136). EET mediated ischemia reperfusion protection of the heart also involves sarcKATP channel activation and PI-3-kinase signaling mechanisms. EET analogs decrease left ventricular hypertrophy and cardiac fibrosis in hypertension (137). Importantly, EET-based therapies consistently prevent cardiac fibrosis in hearts after myocardial infarction, in hearts subjected to pressure overload, and in animals with insulin resistance (37, 137). Taken together, these findings strongly support the idea that EET analog treatment will provide beneficial anti-remodeling in the injured myocardium and may provide a novel approach to treating both systolic and diastolic cardiac dysfunction.
Kidney Diseases
EET agonist analogs combat kidney damage and improve renal function in acute and chronic kidney disease models (132, 133). EET analogs have actions that are ideally suited to treat kidney diseases including vasodilation, inhibition of platelet adhesion, blood pressure lowering action that takes days to weeks, and long-term anti-inflammatory, anti-apoptotic, and anti-fibrotic actions over the course of weeks to months. More importantly, the combined renal and cardiovascular actions make EET analogs an outstanding therapeutic candidate for kidney diseases associated with drug-induced organ toxicity, hypertension, cardiorenal syndrome, and diabetes.
Orally active EET agonist analogs were initially evaluated in a drug-induced acute kidney injury animal model. EET analogs prevent cisplatin-induced increases in blood urea nitrogen (BUN), plasma creatinine, albuminuria, renal tubular cast formation as well as urinary N-acetyl-(D)-glucosaminidase activity (NAG) and kidney injury molecule-1 (KIM-1) excretion (132). Renal injury markers are diminished 40–80% by EET analogs in cisplatin-induced nephrotoxicity. EET analogs also reduce renal oxidative stress, inflammation, apoptosis, and endoplasmic reticulum stress. In fact, these findings are consistent with other reports that EETs reduce critical apoptotic events including Bcl2 activated proapoptotic signaling and caspase-3 activity (140, 141). Likewise, EET analogs protect against the progression of chronic kidney injury in hypertension (112, 113, 133, 134). EET analog treatment of Dahl salt sensitive hypertensive, SH, or angiotensin hypertensive rats markedly reduces urinary albumin along with significant reductions in glomerular injury, intra-tubular cast formation, and kidney fibrosis (112, 113, 133, 134). Renal inflammation, oxidative stress, and endoplasmic reticulum stress dramatically decreases with EET analog treatment in hypertensive rats. Inflammatory MCP-1 levels and macrophage infiltration are consistently lowered by 60–80% with EET analogs in hypertensive animal models. EET analogs also protect the glomerular barrier by preserving nephrin expression and decrease renal fibrosis associated with hypertensive kidney injury (133). In Cyp1a1-Ren-2 transgenic malignant hypertensive rats administration of EET analogs for two weeks decreased oxidative stress, improved renal hemodynamics, and diminished renal injury (134). These data indicate that EET analogs prevent and treat kidney disease if given during the progression of acute or chronic kidney disease. The effectiveness of EET analogs in late stages of kidney diseases needs further study.
Orally active EET analogs protect the kidney from chronic injury through combined anti-inflammatory, anti-apoptotic, and anti-fibrotic mechanisms (142–144). The ability for EET-A to mitigate radiation nephropathy caused by total body irradiation (TBI) indicates that EET analog intervention prevents progression of chronic kidney disease (142). EET-A treatment started two days after TBI and continuing for three months results in reduced BUN, albuminuria, and improved renal histopathological changes. Renal inflammation and oxidative stress do not contribute to TBI mediated renal injury. Therefore, TBI evaluates the ability of EET analog to directly combat renal endothelial dysfunction and epithelial apoptosis. Renal parenchymal apoptosis in TBI rats is reduced by EET-A actions on the p53/Fas/FasL (Fas ligand) apoptotic pathway (142). Renal afferent arteriolar endothelial dysfunction following TBI is dramatically decreased by EET analog treatment (142). These findings clearly show that EET analogs would be an effective treatment in kidney diseases where renal endothelial dysfunction and epithelial apoptosis are prominent.
EET analogs have significant therapeutic potential for organ transplantation. Not only is acute and chronic rejection of implanted organs a problem, but the major immunosuppressive treatment to prevent organ rejection, calcineurin inhibitors, causes hypertension and renal injury. Late rejection is the principal cause of graft failure. Consequently, the ability of an EET analog to oppose cyclosporine-mediated hypertension and renal injury was determined in rats. Oral treatment with an EET agonist analog decreases blood pressure and renal injury induced by four weeks of cyclosporine exposure (143). EET analog treatment also decreases cyclosporine-induced renal oxidative stress, inflammation, fibrosis, and apoptosis by 50–80%. Additionally, the EET analog eliminates cyclosporine-induced apoptosis in cultured NRK-52E renal epithelial cells (143). EET analog treatment decreases endoplasmic reticulum stress and activation of apoptotic pathways by decreasing GRP78 levels in renal epithelial cells (132, 133, 143). More recently, experimental studies in the unilateral ureter obstruction (UUO) mouse kidney fibrosis model have determined the ability of EET analogs to decrease fibrosis, independent of other confounding pathologies causing kidney damage. The anti-fibrotic mechanism by which EET-A diminishes chronic kidney disease was evaluated. EET-A treatment for 10 days produces clear anti-fibrotic actions by decreasing renal collagen positive area, kidney hydroxyproline content, and renal α-smooth muscle actin levels by 50–70% (144). A critical factor that contributes to kidney fibrosis is renal epithelial to mesenchymal transition (EMT) and involves renal epithelial cells that undergo phenotypic and functional changes to myofibroblasts (145). In UUO mice, the anti-fibrotic actions of EET analog treatment are due to a marked decline in E-cadherin transcription factors and renal EMT signaling pathways (144). Taken together, these findings demonstrate that EET analogs prevent and treat kidney diseases through direct renal epithelial cellular protective actions.
Summary of EET Agonist Analog Therapeutic Considerations
EET agonist analogs have exceptional therapeutic potential to combat cardiovascular and kidney diseases (Figure 3). They are effective for short-term or long-term therapy depending on the cardiovascular or renal disease being treated. These EET analogs possess biological activities ideally suited to treat these diseases. Novel, orally active EET analogs have been tested in a range of pre-clinical disease models and demonstrated clear benefit. Importantly, EET analogs not only prevent disease progression but also effectively intervene after the start of the disease. Cardiovascular diseases amenable to treatment by EET analogs include hypertension, drug-induced cardiotoxicity, myocardial infarction, left ventricular fibrosis, chronic left heart disease, and acute coronary artery syndrome. EET analogs also have tremendous potential to treat drug-induced nephrotoxicity, acute kidney injury, hypertensive chronic kidney disease, nephritis, and diabetic nephropathy and ischemia reperfusion injury. Organ transplantation is an intriguing potential therapeutic use for EET analogs. EET analogs could protect the organ being implanted as well as protect the kidney from the deleterious actions of calcineurin inhibitors like cyclosporine. Defining and identifying the EET receptors will be an important step in developing EET agonists for human therapeutics. Next steps will be to determine and prioritize the clinical circumstances that are best suited for EET agonist treatment and make significant advances towards human clinical trials.
Acknowledgments
The authors thank Ms. Kimberly Howard for her secretarial assistance. Supported by grants from the National Institutes of Health (HL-83297, HL-111392 and DK-103616), the Dr. Ralph and Marian Falk Medical Research Trust Bank of America, N.A., Trustee grant and Robert A. Welch Foundation (I-0011).
References
- 1.Capdevila J, Chacos N, Werringloer J, Prough RA, Estabrook RW. Liver microsomal cytochrome P450 and the oxidative metabolism of archidonic acid. Proc Natl Acad Sci USA. 1981;78:5362–6. doi: 10.1073/pnas.78.9.5362. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Chacos N, Falck JR, Wixtrom FC, Capdevila J. Novel epoxides formed during the liver cytochrome P450 oxidation of arachidonic acid. Biochem Biophys Res Comm. 1982;104:916–22. doi: 10.1016/0006-291x(82)91336-5. [DOI] [PubMed] [Google Scholar]
- 3.Falck JR, Schueler VJ, Jacobson HR, Siddhanta AK, Pramanik B, Capdevila J. Arachidonate epoxygenase: identification of epoxyeicosatrienoic acids in rabbit kidney. J Lipid Res. 1987;28:840–6. [PubMed] [Google Scholar]
- 4.Capdevila JH, Dishman E, Karara A, Falck JR. Cytochrome P450 arachidonic acid epoxygenase: Stereochemical characterization of epoxyeicosatrienoic acids. Methods Enzymol. 1991;206:441–53. doi: 10.1016/0076-6879(91)06113-h. [DOI] [PubMed] [Google Scholar]
- 5.Zeldin DC. Epoxygenase pathways of arachidonic acid metabolism. J Biol Chem. 2001;276:36059–62. doi: 10.1074/jbc.R100030200. [DOI] [PubMed] [Google Scholar]
- 6.Capdevila JH, Falck JR. Biochemical and molecular characteristics of the cytochrome P450 arachidonic acid monooxygenase. Prostaglandins Lipid Mediators. 2000;62:271–92. doi: 10.1016/s0090-6980(00)00085-x. [DOI] [PubMed] [Google Scholar]
- 7.Zeldin DC, Dubois RN, Falck JR, Capdevila JH. Molecular cloning, expression and characterization of an endogenous human cytochrome P450 arachidonic acid epoxygenase isoform. Arch Biochem Biophys. 1995;322:7686. doi: 10.1006/abbi.1995.1438. [DOI] [PubMed] [Google Scholar]
- 8.Spector AA, Norris AW. Action of epoxyeicosatrienoic acids on cellular function. Am J Physiol. 2007;292:C996–C1012. doi: 10.1152/ajpcell.00402.2006. [DOI] [PubMed] [Google Scholar]
- 9.Campbell WB, Gebremedhin D, Pratt PF, Harder DR. Identification of epoxyeicosatrienoic acids as endothelium-derived hyperpolarizing factors. Circ Res. 1996;78:415–23. doi: 10.1161/01.res.78.3.415. [DOI] [PubMed] [Google Scholar]
- 10.Huang A, Sun D, Jacobson A, Carroll MA, Falck JR, Kaley G. Epoxyeicosatrienoic acids are released to mediate shear stress-dependent hyperpolarization of arteriolar smooth muscle. Circ Res. 2005;96:376–83. doi: 10.1161/01.RES.0000155332.17783.26. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Kopf PG, Gauthier KM, Zhang DX, Falck JR, Campbell WB. Angiotensin II regulates adrenal vascular tone through zona glomerulosa cell-derived EETs and DHETs. Hypertension. 2011;57:323–9. doi: 10.1161/HYPERTENSIONAHA.110.158311. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Cheng MK, Dourmand AB, Jiang H, Falck JR, McGiff JC, Carroll MA. Epoxyeicosatrienoic acids mediate adenosine-induced vasodilation in rat preglomerular microvessels via A2A receptors. Brit J Pharmacol. 2004;141:441–8. doi: 10.1038/sj.bjp.0705640. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Campbell WB, Falck JR. Arachidonic acid metabolites as endothelium-derived hyperpolarizing factors. Hypertension. 2007;49:590–6. doi: 10.1161/01.HYP.0000255173.50317.fc. [DOI] [PubMed] [Google Scholar]
- 14.Gauthier KM, Edwards EM, Falck JR, Reddy DS, Campbell WB. 14,15-Epoxyeicosatrienoic acid represents a transferable endothelium-dependent relaxing factor in bovine coronary arteries. Hypertension. 2005;45:666–71. doi: 10.1161/01.HYP.0000153462.06604.5d. [DOI] [PubMed] [Google Scholar]
- 15.Wu S, Moomaw CR, Tomer KB, Falck JR, Zeldin DC. Molecular cloning and expression of CYP2J2, a human cytochrome P450 arachidonic acid epoxygenase highly expressed in heart. J Biol Chem. 1996;271:3460–8. doi: 10.1074/jbc.271.7.3460. [DOI] [PubMed] [Google Scholar]
- 16.Wei Y, Lin D-H, Kemp R, Yaddanapudi GSS, Nasjletti A, Falck JR, Wang W-H. Arachidonic acid inhibits epithelial Na channel via cytochrome P450 (CYP) epoxygenase-dependent metabolic pathway. J Gen Physiol. 2004;124:719–27. doi: 10.1085/jgp.200409140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Campbell WB, Brady MT, Rosolowsky LJ, Falck JR. Metabolism of arachidonic acid by rat adrenal glomerulosa cells: Synthesis of HETEs and EETs. Endocrinology. 1991;128:2183–94. doi: 10.1210/endo-128-4-2183. [DOI] [PubMed] [Google Scholar]
- 18.Fisslthaler B, Popp R, Kiss L, Potente M, Harder DR, Fleming I, Busse R. Cytochrome P450 2C is an EDHF synthase in coronary arteries. Nature. 1999;401(6752):493–7. doi: 10.1038/46816. [DOI] [PubMed] [Google Scholar]
- 19.Rosolowsky M, Campbell WB. Synthesis of hydroxyeicosatetraenoic acids (HETEs) and epoxyeicosatrienoic acids (EETs) by cultured bovine coronary artery endothelial cells. Biochim Biophys Acta. 1996;1299:267–77. doi: 10.1016/0005-2760(95)00216-2. [DOI] [PubMed] [Google Scholar]
- 20.Campbell WB, Fleming I. Epoxyeicosatrienoic acids and endothelium-dependent responses. Pflugers Arch. 2010;459:881–95. doi: 10.1007/s00424-010-0804-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Ding Y, Fromel T, Popp R, Falck JR, Schunck W-H, Fleming I. The biological actions of 11,12-epoxyeicosatrienoic acid in endothelial cells are specific to the R/S-enantiomer and require the Gs protein. J Phamacol Exp Ther. 2014;350:14–21. doi: 10.1124/jpet.114.214254. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Pozzi A, Macias-Perez I, Abair T, Wei S, Su Y, Zent R, Falck JR, Capdevila J. Characterization of 5,6- and 8,9-epoxyeicosatrienoic acids as potent in vivo angiogenic lipids. J Biol Chem. 2005;280:27138–46. doi: 10.1074/jbc.M501730200. [DOI] [PubMed] [Google Scholar]
- 23.Falck JR, Krishna UM, Reddy YK, Kumar PS, Reddy KM, Hittner SB, Deeter C, Sharma KK, Gauthier KM, Campbell WB. Comparison of the vasodilatory properties of 14,15-EET analogs: Structural requirements for dilation. Am J Physiol. 2003;284:H337–H49. doi: 10.1152/ajpheart.00831.2001. [DOI] [PubMed] [Google Scholar]
- 24.Falck JR, Reddy LM, Reddy YK, Bondlela M, Krishna UM, Ji Y, Sun J, Liao JK. 11,12-Epoxyeicosatrienoic acid (11,12-EET): Structural determinants for inhibition of TNF-alpha-induced VCAM-1 expression. Bioorganic Med Chem Letters. 2003;13:4011–4. doi: 10.1016/j.bmcl.2003.08.060. [DOI] [PubMed] [Google Scholar]
- 25.Fang X, Kaduce TL, VanRollins M, Weintraub NL, Spector AA. Conversion of epoxyeicosatrienoic acids (EETs) to chain-shortened epoxy fatty acids by human skin fibroblasts. J Lipid Res. 2000;41:66–74. [PubMed] [Google Scholar]
- 26.Karara A, Dishman E, Falck JR, Capdevila JH. Endogenous epoxyeicosatrienoyl-phospholipids. J Biol Chem. 1991;266:7561–9. [PubMed] [Google Scholar]
- 27.Zeldin DC, Kobayashi J, Falck JR, Winder BS, Hammock BD, Snapper JR, Capdevila JH. Regio- and enantiofacial selectivity of epoxyeicosatrienoic acid hydration by cytosolic epoxide hydrolase. J Biol Chem. 1993;268:6402–7. [PubMed] [Google Scholar]
- 28.Zeldin DC, Shouzou W, Falck JR, Hammock BD, Snapper JR, Capdevila JH. Metabolism of epoxyeicosatrienoic acids by cytosolic epoxide hydrolase: Substrate structural determinants of asymmetric catalysis. Arch Biochem Biophys. 1995;316:443–51. doi: 10.1006/abbi.1995.1059. [DOI] [PubMed] [Google Scholar]
- 29.Newman JW, Morisseau C, Harris TR, Hammock BD. The soluble epoxide hydrolase encoded by EPXH2 is a bifunctional enzyme with novel lipid phosphate phosphatase activity. Proc Nat Acad Sci USA. 2003;100:1558–63. doi: 10.1073/pnas.0437724100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Chacos N, Capdevila J, Falck JR, Manna S, Martin-Wixtram C, Gill SS, Hammock BD, Estabrook RW. The reaction of arachidonic acid epoxides (epoxyeicosatrienoic acids) with a cytosolic epoxide hydrolase. Arch Biochem Biophy. 1983;223:639–48. doi: 10.1016/0003-9861(83)90628-8. [DOI] [PubMed] [Google Scholar]
- 31.Imig JD, Hammock BD. Soluble epoxide hydrolase as a therapeutic target for cardiovascular diseases. Nature Rev Drug Discov. 2009;8:794–805. doi: 10.1038/nrd2875. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Cowart LA, Wei S, Hsu MH, Johnson EF, Krishna UM, Falck JR, Capdevila J. The CYP4A isoforms hydroxylate epoxyeicosatrienoic acids to form high affinity peroxisome proliferator-activated receptor ligands. J Biol Chem. 2002;20:35105–12. doi: 10.1074/jbc.M201575200. [DOI] [PubMed] [Google Scholar]
- 33.Spearman ME, Prough RA, Estabrook RW, Falck JR, Manna S, Leibman KC, Murphy RC, Capdevila J. Novel glutathione conjugates formed from epoxyeicosatrienoic acids (EETs) Arch Biochem Biophys. 1985;242:225–30. doi: 10.1016/0003-9861(85)90496-5. [DOI] [PubMed] [Google Scholar]
- 34.Larsen BT, Campbell WB, Gutterman DD. Beyond vasodilation: Non-vasomotor roles of epoxyeicosatrienoic acids in the cardiovascular system. TIPS. 2007;28:32–8. doi: 10.1016/j.tips.2006.11.002. [DOI] [PubMed] [Google Scholar]
- 35.Fleming I. Epoxyeicosatrienoic acids, cell signaling and angiogenesis. Prostag and Other Lipid Mediat. 2007;82:60–7. doi: 10.1016/j.prostaglandins.2006.05.003. [DOI] [PubMed] [Google Scholar]
- 36.Fleming I. Vascular cytochrome P450 enzymes: Physiology and pathophysiology. TCM. 2008;18:20–5. doi: 10.1016/j.tcm.2007.11.002. [DOI] [PubMed] [Google Scholar]
- 37.Imig JD. Epoxides and soluble epoxide hydrolase in cardiovascular physiology. Physiol Rev. 2012;92:101–30. doi: 10.1152/physrev.00021.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Archer SL, Gragasin FS, Wu X, Wang S, McMurtry S, Kim DH, Platonov M, Koshal A, Hasimoto K, Campbell WB, Falck JR, Michelakis ED. Endothelium-derived hyperpolarizing factor in human internal mammary artery is 11,12-epoxyeicosatrienoic acid and causes relaxation by activating smooth muscle BKca channels. Circulation. 2003;107:769–76. doi: 10.1161/01.cir.0000047278.28407.c2. [DOI] [PubMed] [Google Scholar]
- 39.Gauthier KM, Baewer DV, Hittner SB, Hillard CJ, Nithipatikom K, Reddy DS, Falck JR, Campbell WB. Endothelium-derived 2-arachidonylglycerol: An intermediate in vasodilatory eicosanoid release in bovine coronary arteries. Am J Physiol. 2005;288:H1344–H51. doi: 10.1152/ajpheart.00537.2004. [DOI] [PubMed] [Google Scholar]
- 40.Gebremedhin D, Harder DR, Pratt PF, Campbell WB. Bioassay of an endothelium-derived hyperpolarizing factor from bovine coronary arteries: role of a cytochrome P450 metabolite. J Vasc Res. 1998;35:274–84. doi: 10.1159/000025594. [DOI] [PubMed] [Google Scholar]
- 41.Li P-L, Campbell WB. Epoxyeicosatrienoic acids activate potassium channels in coronary smooth muscle through guanine nucleotide binding protein. Circ Res. 1997;80:877–84. doi: 10.1161/01.res.80.6.877. [DOI] [PubMed] [Google Scholar]
- 42.Campbell WB, Falck JR, Gauthier K. Role of epoxyeicosatrienoic acids as endothelium-derived hyperpolarizing factor in bovine coronary arteries. Med Sci Monitor. 2001;7:578–84. [PubMed] [Google Scholar]
- 43.Larsen BT, Miura H, Hatoum OA, Campbell WB, Hammock BD, Zeldin DC, Falck JR, Gutterman DD. Epoxyeicosatrienoic and dihydroxyeicosatrienoic acids dilate human coronary arterioles via BKca channels: Implications for soluble epoxide hydrolase inhibition. Am J Physiol. 2005;290:H491–H9. doi: 10.1152/ajpheart.00927.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Gauthier KM, Deeter C, Krishna UM, Reddy YK, Bondlela M, Falck JR, Campbell WB. 14,15-Epoxyeicosa-5(Z)-enoic acid: A selective epoxyeicosatrienoic acid antagonist that inhibits endothelium-dependent hyperpolarization and relaxation in coronary arteries. Circ Res. 2002;90:1028–36. doi: 10.1161/01.res.0000018162.87285.f8. [DOI] [PubMed] [Google Scholar]
- 45.Taddei S, Varsari D, Cipriano A, Ghiadoni L, Glaetta F, Franzoni F, Magagna A, Virdis A, Salvetti A. Identification of a cytochrome P450 2C9-derived endothelium-derived hyperpolarizing factor in human essential hypertensive patients. J Am Coll Cardiol. 2006;48:508–15. doi: 10.1016/j.jacc.2006.04.074. [DOI] [PubMed] [Google Scholar]
- 46.Bellein J, Remy-Jouet I, Jacob M, Blot E, Mercier A, Lucas D, Dreano Y, Gutlerrez L, Donnadieu N, Thuillez C, Joannides R. Impaired role of epoxyeicosatrienoic acids in the regulation of basal conduit artery diameter during essential hypertension. Hypertension. 2012;60:1415–21. doi: 10.1161/HYPERTENSIONAHA.112.201087. [DOI] [PubMed] [Google Scholar]
- 47.Imig JD, Navar LG, Roman RJ, Reddy KK, Falck JR. Actions of epoxygenase metabolites on the preglomerular vasculature. J Am Soc Nephrol. 1996;7:2364–70. doi: 10.1681/ASN.V7112364. [DOI] [PubMed] [Google Scholar]
- 48.Sun P, Lin D-H, Wang T, Babilonia E, Wang Z, Jin Y, Kemp R, Nasjletti A, Wang W-H. Low Na intake suppresses expression of CYP2C23 and arachidonic acid-induced inhibition of ENaC. Am J Physiol. 2006;291:F1192–F200. doi: 10.1152/ajprenal.00112.2006. [DOI] [PubMed] [Google Scholar]
- 49.Sun P, Antoun J, Lin D-H, Gotlinger KH, Capdevila J, Wang W-H. Cyp2C44 epoxygenase is essential for preventing the renal sodium absorption during icnreasing dietary potassium intake. Hypertension. 2012;59:339–47. doi: 10.1161/HYPERTENSIONAHA.111.178475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Wang W-H, Zhang C, Lin D-H, Wang L, Graves JP, Zeldin DC, Capdevila J. Cyp2C44 epoxygenase in th collecting duct is essential for the high K intake-induced antihypertensive effect. Am J Physiol. 2014;307:F453–F60. doi: 10.1152/ajprenal.00123.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Zhao X, Pollock DM, Inscho EW, Zeldin DC, Imig JD. Decreased renal cytochrome P450 2C enzymes and impaired vasodilation are associated with angiotensin salt-sensitive hypertension. Hypertension. 2003;41:709–14. doi: 10.1161/01.HYP.0000047877.36743.FA. [DOI] [PubMed] [Google Scholar]
- 52.Capdevila J, Wang W-H. Role of cytochrome P450 epoxygenase in regulating renal membrane transport and hypertension. Curr Opin Nephrol Hypertens. 2013;22:163–9. doi: 10.1097/MNH.0b013e32835d911e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Pidkovka N, Rao R, Mei S, Gong Y, Harris RC, Wang WH, Capdevila J. Epoxyeiccosatrienoic acids regulate epithelial sodium channel activity by extracellular signal-regulated kinase 1/2 (ERK 1/2)-mediated phosphorylation. J Biol Chem. 2013;288:5223–31. doi: 10.1074/jbc.M112.407981. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Capdevila J, Pidkovka N, Mei S, Gong Y, Falck JR, Imig JD, Harris RC, Wang W-H. The Cyp2C44 epoxygenase regulates epithelial sodium channel activity and blood pressure responses to increased dietary salt. J Biol Chem. 2014;289:4377–86. doi: 10.1074/jbc.M113.508416. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Huang H, Morisseau C, Wang J, Yang T, Falck JR, Hammock BD, Wang W-H. Increasing or stabilizing renal epoxyeicosatrienoic acid production attenuates abnormal renal function and hypertension. Am J Physiol. 2007;293:F342–F349. doi: 10.1152/ajprenal.00004.2007. [DOI] [PubMed] [Google Scholar]
- 56.Xiao B, Li X, Yan J, Yu X, Yang G, Xiao X, Voltz JW, Zeldin DC, Wang D-W. Overexpression of cytochrome P450 epoxygenase prevents development of hypertension in spontanenously hypertensive rats by enhancing atrial natriuretic peptide. J Phamacol Exp Ther. 2010;334:784–94. doi: 10.1124/jpet.110.167510. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Muller DN, Theuer J, Shagdarsuren E, Kaergel E, Honeck H, Park JK, Maarkovic M, Barbosa-Sicard E, Deshend R, Wellner M, Kirsch T, Fiebeler A, Rothe M, Haller H, Luft FC, Schunck W-H. A peroxidome proliferator-activated receptor-alpha activator induces renal CYP2C23 activity and protects from angiotensin II-induced renal injury. Am J Pathol. 2004;164:521–32. doi: 10.1016/s0002-9440(10)63142-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Imig JD, Zhao X, Capdevila J, Morisseau C, Hammock BD. Soluble epoxide hydrolase inhibition lowers blood pressure in angiotensin II hypertension. Hypertension. 2002;39:690–4. doi: 10.1161/hy0202.103788. [DOI] [PubMed] [Google Scholar]
- 59.Jung O, Brandes RP, Kim IH, Schweda F, Schmidt R, Hammock BD, Fleming I. Soluble epoxide hydrolase is a main effector of angiotensin II-induced hypertension. Hypertension. 2005;45:759–65. doi: 10.1161/01.HYP.0000153792.29478.1d. [DOI] [PubMed] [Google Scholar]
- 60.Lee CR, Imig JD, Edin ML, Foley J, DeGraff LM, Bradbury JA, Graves JP, Lih FB, Clark J, Myer P, Perrow AL, Lepp AN, Kannon MA, Ronnekleiv OK, Alkayed NJ, Falck JR, Tomer KB, Zeldin DC. Endothelial expression of human cytochrome P450 epoxygenases lower blood pressure and attenuates hypertension-induced renal injury in mice. FASEB J. 2010;24:3770–81. doi: 10.1096/fj.10-160119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Gauthier KM, Yang W, Gross GJ, Campbell WB. Roles of epoxyeicosatrienoic acids in vascular regulation and cardiac preconditioning. J Cardiovas Pharmacol. 2007;50:601–8. doi: 10.1097/FJC.0b013e318159cbe3. [DOI] [PubMed] [Google Scholar]
- 62.Seubert JM, Siani CJ, Graves J, DeGraff LM, Bradbury JA, Lee CR, Goralski K, Carey MA, Luria A, Newman JW, Hammock BD, Falck JR, Roberts H, Rockman HA, Murphy E, Zeldin DC. Role of soluble epoxide hydrolase in postischemic recovery of heart contractile function. Circ Res. 2006;99:442–50. doi: 10.1161/01.RES.0000237390.92932.37. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Seubert JM, Yang B, Bradbury A, Graves J, DeGraff LM, Gabel S, Gooch R, Foley J, Newman JW, Mao L, Rockman HA, Hammock BD, Murphy E, Zeldin DC. Enhanced postischemic functional recovery in CYP2J2 transgenic hearts involves mitochondrial ATP-sensitive K channels and p42/p44 MAPK pathway. Circ Res. 2004;95:506–14. doi: 10.1161/01.RES.0000139436.89654.c8. [DOI] [PubMed] [Google Scholar]
- 64.Nithipatikom K, Moore JM, Isbell MA, Falck JR, Gross GJ. Epoxyeicosatrienoic acids in cardioprotection: ishemic versus reperfusion injury. Am J Physiol. 2006;291:H537–H42. doi: 10.1152/ajpheart.00071.2006. [DOI] [PubMed] [Google Scholar]
- 65.Gross GJ, Hsu AK, Falck JR, Nithipatikom K. Mechanisms by which epoxyeicosatrienoic acids (EETs) elicit cardioprotection in rat hearts. J Mol Cell Cardiol. 2007;42:687–91. doi: 10.1016/j.yjmcc.2006.11.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Monti J, Fisher J, Paskas S, Heinig M, Schultz H, Gosele C, Heuser A, Fisher R, Schmidt C, Schirdewan A, Gross V, Hummel O, Maatz H, Patone G, Saar K, Vingron M, Weldon SM, Lindpaintner K, Hammock BD, Rohde K, Dietz R, Cook SA, Schunck W-H, Luft FC, Nubner N. Soluble epoxide hydrolase is a susceptibility factor for heart failure in a rat model of human disease. Nature Genetics. 2008;40:529–37. doi: 10.1038/ng.129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Xu D, Li N, He Y, Timofeyev V, Lu L, Tsai H-J, Kim I-H, Tuteja D, Mateo RKP, Singapuri A, Davis BB, Low R, Hammock BD, Chiamvimonvat N. Prevention and reversal of cardiac hypertrophy by soluble epoxide hydrolase inhibitors. Proc Nat Acad Sci USA. 2006;103:18733–8. doi: 10.1073/pnas.0609158103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Westphal C, Spallek B, Konkel A, Marko L, Qadri F, DeGraff LM, Schubert C, Bradbury A, Regitz-Zagrosek V, Falck JR, Zeldin DC, Muller DN, Schunck W-H, Fischer R. CYP2J2 overexpression protects against arrhythmia susceptibility in cardiac hypertrophy. PLoS One. 2013;8:e73490. doi: 10.1371/journal.pone.0073490. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Wang X, Ni L, Yang L, Duan Q, Chen C, Edin ML, Zeldin DC, Wang D-W. CYP2J2-derived epoxyeicosatrienoic acids suppress endoplasmic reticulum stress in heart failure. Mol Pharmacol. 2014;85:105–15. doi: 10.1124/mol.113.087122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.He Z, Zhang X, Chen C, Wen Z, Hoopes SL, Zeldin DC, Wang D-W. Cardiomyocyte-specific expression of CYP2J2 presents development of cardiac remodelling induced by angiotensin II. Cardiovas Res. 2015;105:304–17. doi: 10.1093/cvr/cvv018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Krotz F, Riexinger T, Buerkle MA, Nithipatikom K, Gloe T, Sohn HY, Campbell WB, Pohl U. Membrane potential-dependent inhibition of platelet adhesion to endothelial cells by epoxyeicosatrienoic acids. Arterioscler Thromb Vasc Biol. 2003;24:595–600. doi: 10.1161/01.ATV.0000116219.09040.8c. [DOI] [PubMed] [Google Scholar]
- 72.Node K, Ruan X-L, Dai J, Yang S-X, Graham L, Zeldin DC, Liao JK. Activation of Gas mediates induction of tissue-type plasminogen activator gene transcription by epoxyeicosatrienoic acids. J Biol Chem. 2001;276:15983–9. doi: 10.1074/jbc.M100439200. [DOI] [PubMed] [Google Scholar]
- 73.Thomson SJ, Askari A, Bishop-Bailey D. Anti-inflammatory effects of epoxyeicosatrienoic acids. Int J Vasc Med. 2012 doi: 10.1155/2012/605101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Campbell WB. New role for epoxyeicosatrienoic acids as anti-inflammatory mediators. TiPS. 2000;21:125–7. doi: 10.1016/s0165-6147(00)01472-3. [DOI] [PubMed] [Google Scholar]
- 75.Node K, Huo Y, Ruan X, Yang B, Spiecker M, Ley K, Zeldin D, Liao J. Anti-inflammatory properties of cytochrome P450 epoxygenase-derived eicosanoids. Science. 1999;285:1276–9. doi: 10.1126/science.285.5431.1276. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Liu Y, Zhang Y, Schmelzer K, Lee T-S, Fang X, Zhu Y, Spector AA, Gill S, Morisseau C, Hammock BD, Shyy JY-J. The antiinflammatory effect of laminar flow: the role of PPARγ. epoxyeicosatrienoic acids and soluble epoxide hydrolase. Proc Nat Acad Sci USA. 2005;102:16747–52. doi: 10.1073/pnas.0508081102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Rothwarf DM, Karin M. The NF-kappa B activation pathway: a paradigm in information transfer from membrane to nucleus. Sci STKE. 1999;5:1–16. doi: 10.1126/stke.1999.5.re1. [DOI] [PubMed] [Google Scholar]
- 78.Liu Y, Webb HK, Fukushima H, Michell J, Markova S, O JL, Kroetz DL. Attenuation of cisplatin-induced renal injury by inhibition of soluble epoxide hydrolase involves nuclear factor kappa B signaling. J Phamacol Exp Ther. 2012;341:725–34. doi: 10.1124/jpet.111.191247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Chen W, Yang S, Ping W, Fu X, Xu Q, Wang J. CYP2J2 and EETs protect against lung ischemia/reperfusion injury via anti-inflammatory effects in vivo and in vitro. Cel Physiol Biochem. 2015;35:2043–54. doi: 10.1159/000374011. [DOI] [PubMed] [Google Scholar]
- 80.Ahmed SMU, Luo L, Namani A, Wang XJ, Tang X. Nrf2 signaling pathway: Pivotal roles in inflammation. Biochim Biophy Acta. 2017;1863:585–97. doi: 10.1016/j.bbadis.2016.11.005. [DOI] [PubMed] [Google Scholar]
- 81.Li Y, Yu G, Yuan S, Tan C, Xie J, Ding Y, Lian P, Fu L, Hou Q, Xu B, Wang H. 14,15-Epoxyeicosatrienoic acid suppresses cigarette smoke condensate-induced inflammation in lung epithelial cells by inhibiting autophagy. Am J Physiol. 2016;311:L970–L80. doi: 10.1152/ajplung.00161.2016. [DOI] [PubMed] [Google Scholar]
- 82.Liu W-J, Wang T, Wang B, Liu X-T, He X-W, Liu Y-J, Li Z-X, Tan R, Zeng H-S. CYP2C8-derived epoxyeicosatrienoic acids decreased oxidative stress-induced endothelial apoptosis in development of atherosclerosis: Role of Nrf2 activation. J Huazhong Univ Sci Technol. 2015;35:640–5. doi: 10.1007/s11596-015-1483-5. [DOI] [PubMed] [Google Scholar]
- 83.Yang S, Wei S, Pozzi A, Capdevila J. The arachidonic acid epoxygenase is a component of the signaling mechanisms responsible for VEGF-stimulated angiogenesis. Arch of Biochem and Biophys. 2009;489:82–91. doi: 10.1016/j.abb.2009.05.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Michaelis UR, Fissllthaler B, Medhora M, Harder D, Fleming I, Busse R. Cytochrome P450 2C9-derived epoxyeicosatrienoic acids induce angiogenesis via cross-talk with the epidermal growth factor receptor. FASEB J. 2003;17:770–2. doi: 10.1096/fj.02-0640fje. [DOI] [PubMed] [Google Scholar]
- 85.Potente M, Michaelis UR, Fisslthaler B, Busse R, Fleming I. Cytochrome P450 2C9-induced endothelial cell proliferation involved induction of mitogen-activated protein (MAP) kinase phosphatase-1, inhibition of the c-Jun N-teminal kinase, and up-regulation of cylcin D1. J Biol Chem. 2002;277:15671–6. doi: 10.1074/jbc.M110806200. [DOI] [PubMed] [Google Scholar]
- 86.Panigraphy D, Kalish BT, Huang S, Bielenberg DR, Le HD, Yang J, Edin ML, Lee CR, Benny O, Mudge DK, Butterfield CE, Mammoto A, Mammoto T, Inceoglu B, Jenkins RL, Simpson MA, Akino T, Lih FB, Tomer KB, Ingber DE, Hammock BD, Falck JR, Manthati VL, Kaipainen A, D’Amore PA, Puder M, Zeldin DC, Kieran MW. Epoxyeicosanoids promote organ and tissue regeneration. Proc Nat Acad Sci USA. 2013;110:13528–33. doi: 10.1073/pnas.1311565110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Zhao H, Chen J, Chai J, Zhang Y, Yu C, Pan Z, Gao P, Zong C, Guan Q, Liu Y. Cytochrome P450 (CYP) epoxygenasses as potential targets in the management of impaired diabetic wound healing. Lab Invest. 2017;00:1–10. doi: 10.1038/labinvest.2017.21. [DOI] [PubMed] [Google Scholar]
- 88.Xu X, Zhao CX, Wang L, Tu L, Fang X, Zheng C, Edin ML, Zeldin DC, Wang D-W. Increased CYP2J3 expression reduces insulin resistance in fructose-treated rats and db/db mice. Diabetes. 2010;59:997–1005. doi: 10.2337/db09-1241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Li R, Xu X, Chen C, Wang Y, Gruzdev A, Zeldin DC, Wang D-W. CYP2J2 attenuates metabolic dysfunction in diabetic mice by reducing hepatic inflammation via the PPAR gamma. Am J Physiol. 2015;308:E270–E82. doi: 10.1152/ajpendo.00118.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Luria A, Bettaieb A, Shieh GJ, Liu HC, Inoue H, Tsai HJ, Imig JD, Haj FG, Hammock BD. Soluble epoxide hydrolase deficiency alters pancreatic islet size and improves glucose homeostatis in the model of insulin resistance. Proc Nat Acad Sci USA. 2011;108:9038–43. doi: 10.1073/pnas.1103482108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Chen G, Wang P, Zhao G, Gruzdev A, Zeldin DC, Wang D-W. Cytochrome P450 epoxygenase CYP2J2 attenuates nephropathy in streptozotocin-induced diabetic mice. Prostag and Other Lipid Mediat. 2011;96:63–71. doi: 10.1016/j.prostaglandins.2011.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Roche C, Guerrot D, Harouki N, Duflot T, Besnier M, Remy-Jouet I, Renet S, Dumesnil A, Lejeune A, Morrisseau C, Richard V, Bellien J. Impace of soluble epoxide hydrolase inhibition on early kidney deamage in hyperglycemic overweight mice. Prostag and Other Lipid Mediat. 2015;120:148–54. doi: 10.1016/j.prostaglandins.2015.04.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Gangadhariah MH, Dieckmann BW, Lantier L, Kang L, Wasserman DH, Chiusa M, Caskey CF, Dickerson J, Lou P, Gamboa JL, Capdevila J, Imig JD, Yu C, Pozzi A, Luther JM. Cytochrome P450 epoxygenase-derived epoxyeicosatrienoic acids contribute to insulin sensitivity in mice and humans. Diabetologia. 2017 doi: 10.1007/s00125-017-4260-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Fleming I, Rueben A, Popp R, Fissllthaler B, Schrodt S, Sander A, Haendeler J, Falck JR, Morisseau C, Hammock BD, Busse R. Epoxyeicosatrienoic acids regulate Trp-channel-dependent Ca signaling and hyperpolariztion in endothelial cells. Arterio Thromb Vasc Biol. 2007;27:2612–8. doi: 10.1161/ATVBAHA.107.152074. [DOI] [PubMed] [Google Scholar]
- 95.Sun J, Sui X-X, Bradbury A, Zeldin DC, Conte MS, Liao JK. Inhibition of vascular smooth muscle cell migration by cytochrome P450 epoxygenase-derived eicosanoids. Circ Res. 2002;90:1020–7. doi: 10.1161/01.res.0000017727.35930.33. [DOI] [PubMed] [Google Scholar]
- 96.Wong PY, Lin KT, Yan YT, Ahern D, Iles J, S SY, Bhatt RK, Falck JR. 14(R), 15(S)-Epoxyeicosatrienoic acid receptor in guinea pig mononuclear cell membranes. J Lipid Mediat Cell Signal. 1993;6:199–208. [PubMed] [Google Scholar]
- 97.Wong PY-K, Lai P-S, Falck JR. Mechanism and signal transduction of 14(R), 15 (S)-epoxyeicosatrienoic acid (14,15-EET binding in guinea pig monocytes. Prostag Other Lipid Med. 2000;62:321–33. doi: 10.1016/s0090-6980(00)00079-4. [DOI] [PubMed] [Google Scholar]
- 98.Wong PY-K, Lai P-S, Shen S-Y, Belosludtsev YY, Falck JR. Post-receptor signal transduction and regulation of 14(R), 15(S)-epoxyeicosatrienoic acid (14,15-EET) binding in U-937 cells. J Lipid Med Cell Signal. 1997;16:155–69. doi: 10.1016/s0929-7855(97)00005-9. [DOI] [PubMed] [Google Scholar]
- 99.Chen Y, Falck JR, Tuniki VR, Campbell WB. 20-125Iodo-14,15-epoxyeicosa-5Z-enoic acid: a high affinity radioligand used to characterize the epoxyeicosatrienoic acid antagonist binding site. J Pharmacol Exp Ther. 2009;331:1137–45. doi: 10.1124/jpet.109.157818. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Yang W, Holmes BB, Gopal VR, Kishore RVK, Sangras B, Yi X-Y, Falck JR, Campbell WB. Characterization of 14,15-epoxyeicosatrienoyl-sulfonamides as 14,15-epoxyeicosatrienoic acid agonists: Use for studies of metabolism and ligand binding. J Pharmacol Exp Ther. 2007;321:1023–31. doi: 10.1124/jpet.107.119651. [DOI] [PubMed] [Google Scholar]
- 101.Yang W, Tuniki VR, Anjaiah S, Falck JR, Hillard CJ, Campbell WB. Characterization of epoxyeicosatrienoic acid binding site in U937 membranes using a novel radiolabeled agonist, 20-125I-14,15-epoxyeicosa-8(Z)-enoic acid. J Pharmacol Exp Ther. 2008;324:1019–27. doi: 10.1124/jpet.107.129577. [DOI] [PubMed] [Google Scholar]
- 102.Fukao M, Mason HS, Kenyon JL, Horowitz B, Keef KD. Regulation of BKCa channels expressed in human embryonic kidney 293 cells by epoxyeicosatrienoic acid. Molec Pharmacol. 2001;59:16–23. doi: 10.1124/mol.59.1.16. [DOI] [PubMed] [Google Scholar]
- 103.Hayabuchi Y, Nakaya Y, Matsuoka S, Kuroda Y. Endothelium-derived hyperpolarizing factor activates Ca2+-activated K+ channels in porcine coronary artery smooth muscle cells. J Cardiovasc Pharmacol. 1998;32:642–9. doi: 10.1097/00005344-199810000-00018. [DOI] [PubMed] [Google Scholar]
- 104.Itoh Y, Kawamata Y, Harada M, Kobayashi M, Fujii R, Fukusumi S, Ogi K, Hosoya M, Tanaka Y, Uejima H, Taknaka H, Maruyama M, S R, Okubo S, Kizawa H, Komatsu H, Matsumura F, Noguchi Y, Shinohara T, Hinuma S, Fumisawa Y, Fujino M. Free fatty acids regulate insulin secretion from pancreatic beta cells through GPR40. Nature. 2003;422:173–6. doi: 10.1038/nature01478. [DOI] [PubMed] [Google Scholar]
- 105.Behm DJ, Ogbonna A, Wu C, Burns-Kurtis CL, Douglas SA. Epoxyeicosatrienoic acids function as selective, endogenous antagonists of native thromboxane receptors: Identification of a novel mechanism of vasodilation. J Pharmacol Exp Ther. 2009;328:231–9. doi: 10.1124/jpet.108.145102. [DOI] [PubMed] [Google Scholar]
- 106.Yang C, Kwan YW, Au AL, Poon CC, Zhang Q, Chan SW, Lee SM, Leung GP. 14,15-Epoxyeicosatrienoic acid induces vasorelaxation through prostaglandin EP(2) receptors in rat mesenteric artery. Prostag and Other Lipid Mediat. 2010;93:44–51. doi: 10.1016/j.prostaglandins.2010.06.004. [DOI] [PubMed] [Google Scholar]
- 107.Yang C, Yang J, Xu X, Yang S, Pan S, Zhang C, Leung GP. Vasodilatory effect of 14,15-epoxyeicosatrienoic acid on mesenteric arteries in hypertensive and aged rats. Prost Other Lipid Mediators. 2014;112:1–8. doi: 10.1016/j.prostaglandins.2014.05.001. [DOI] [PubMed] [Google Scholar]
- 108.Coleman RA, Humphrey PPA, Kennedy I, Lumley P. Prostanoid receptors–the development of a working classification. TIPS. 1984;5:303–6. [Google Scholar]
- 109.Bukhari IA, Gauthier KM, Jagadeesh SG, Sangras B, Falck JR, Campbell WB. 14,15-Dihydroxy-eicosa-5(S)-enoic acid selectively inhibits 14,15-epoxyeicosatrienoic acid-induced relaxations in bovine coronary arteries. J Phamacol Exp Ther. 2011;336:47–55. doi: 10.1124/jpet.110.169797. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Bukhari IA, Shah AJ, Gauthier KM, Walsh KA, Koduru SR, Imig JD, Falck JR, Campbell WB. 11,12,20-Trihydroxy-eicosa-8(Z)-enoic acid: a selective inhibitor of 11,12-EET-induced relaxations of bovine coronary and rat mesenteric arteries. Am J Physiol. 2012;302:H1574–H83. doi: 10.1152/ajpheart.01122.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Gauthier KM, Jagadeesh SG, Falck JR, Campbell WB. 14,15-Epoxyeicosa-5(Z)-enoic-mSI: A 14,15- and 5,6-EET antagonist in bovine coronary arteries. Hypertension. 2003;42:555–61. doi: 10.1161/01.HYP.0000091265.94045.C7. [DOI] [PubMed] [Google Scholar]
- 112.Khan MAH, Falck JR, Manthati VL, Campbell WB, Imig JD. Epoxyeicosatrienoic acid analog attenuates angiotensin II hypertension and kidney injury. Frontiers Pharmacol. 2014;5:216–23. doi: 10.3389/fphar.2014.00216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Khan MAH, Pavlov TS, Staruschenko A, Capdevila J, Falck JR, Campbell WB, Imig JD. Epoxyeicosatrienoic acid analog lowers blood pressure through epithelial sodium channel inhibition. Clin Sci. 2014;127:463–74. doi: 10.1042/CS20130479. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Chen J-K, Falck JR, Reddy KM, Capdevila J, Harris RC. Epoxyeicosatrienoic acids and their sulfonimide derivatives stimulate tyrosine phosphorylation and induce mitogenesis in renal epithelial cells. J Biol Chem. 1998;273:29254–61. doi: 10.1074/jbc.273.44.29254. [DOI] [PubMed] [Google Scholar]
- 115.Imig JD, Inscho EW, Deichmann PC, Reddy KM, Falck JR. Afferent arteriolar vasodilation to the sulfonimide analog of 11,12-epoxyeicosatrienoic acid involves protein kinase A. Hypertension. 1999;33:408–13. doi: 10.1161/01.hyp.33.1.408. [DOI] [PubMed] [Google Scholar]
- 116.Imig JD, Zhao X, Falck JR, Wei S, Capdevila J. Enhanced renal microvascular reactivity to angiotensin II in hypertension is ameliorated by the sulfonimide analog of 11,12-epoxyeicosatrienoic acid. J Hypertension. 2001;19:983–92. doi: 10.1097/00004872-200105000-00020. [DOI] [PubMed] [Google Scholar]
- 117.Gauthier KM, Falck JR, Reddy LM, Campbell WB. 14,15-EET analogs: Characterization of structural requirements for agonist and antagonist activity in bovine coronary arteries. Pharmacol Res. 2004;49:515–24. doi: 10.1016/j.phrs.2003.09.014. [DOI] [PubMed] [Google Scholar]
- 118.Dimitropoulou C, West L, Field MB, White RE, Reddy LM, Falck JR, Imig JD. Protein phosphatase 2A and Ca-activated K channels contribute to 11,12-epoxyeicosatrienoic acid analog mediated mesenteric arterial dilation. Prostag and Other Lipid Mediat. 2007;83:50–61. doi: 10.1016/j.prostaglandins.2006.09.008. [DOI] [PubMed] [Google Scholar]
- 119.Falck JR, Kodela R, Manne R, Atcha KR, Puli N, Dubasi N, Manthati VL, Capdevila J, Yi X-Y, Goldman DM, Morisseau C, Hammock BD, C WB. 14,15-Epoxyeicosa-5,8,11-trienoic acid (14,15-EET) surrogates containing epoxide bioisosteres: Influence upon vascular relaxation and soluble epoxide hydrolase inhibition. J Med Chem. 2009;52:5069–75. doi: 10.1021/jm900634w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Brown N. Bioisosteres in Medicinal Chemistry. Weinheim: Wiley-VCH; 2012. [Google Scholar]
- 121.Falck JR, Koduru SR, Mohapatra S, Manne R, Atcha KR, Manthati VL, Capdevila JH, Christian S, Imig JD, Campbell WB. 14,15-Epoxyeicosa-5,8,11-trienoic acid (14,15-EET) surrogates: Carboxylate modifications. J Med Chem. 2014;57:6965–72. doi: 10.1021/jm500262m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Yasmin S, Jayaprakash V. Thiazolidinediones and PPAR orcchestra as antidiabetic agents: from past to present. Euro J Med Chem. 2017;126:879–93. doi: 10.1016/j.ejmech.2016.12.020. [DOI] [PubMed] [Google Scholar]
- 123.Imig JD. Epoxyeicosatreinoic acids, hypertension and kidney injury. Hypertension. 2015;65:476–82. doi: 10.1161/HYPERTENSIONAHA.114.03585. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Yang L, Cheriyan J, Gutterman DD, Mayer RJ, Ament Z, Griffin JL, Lazaar AL, Newby DE, Tai-Singer R, Wilkinson IB. Mechanisms of vascular dysfunction in COPD and effect of a novel soluble epoxide hydrolase inhibitor in smokers. Chest. 2017;151:555–63. doi: 10.1016/j.chest.2016.10.058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Swinney DC, Anthony J. How were new medicines discovered? Nature Rev Drug Discov. 2011;10:507–19. doi: 10.1038/nrd3480. [DOI] [PubMed] [Google Scholar]
- 126.Barri YM. Hypertension and kidney disease: a deadly connection. Curr Hypertens Rep. 2008;10:39–45. doi: 10.1007/s11906-008-0009-y. [DOI] [PubMed] [Google Scholar]
- 127.Coresh J, Selvin E, Stevens LA, Manzi J, Kusek JW, Eggers P, Van Lente F. Prevalence of chronic kidney disease in the United States. J Am Med Assoc. 2007;298:2038–47. doi: 10.1001/jama.298.17.2038. [DOI] [PubMed] [Google Scholar]
- 128.Hadi HA, Carr CS, Suwaldi AI. Endothelial dysfunction: Cardiovascular risk factor, therapy and outcome. Vascular Health and Risk Management. 2005;1:183–98. [PMC free article] [PubMed] [Google Scholar]
- 129.Imig JD, Elmarakby AA, Nithipatikom K, Wei S, Capdevila J, Tuniki VR, Sangras B, Anjaiah S, Manthati VL, Reddy DS, Falck JR. Development of epoxyeicosatrienoc acid analogs with in vivo anti-hypertensive actions. Frontiers Vasc Physiol. 2010;1:157. doi: 10.3389/fphys.2010.00157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Sodhi K, Inoue K, Gotlinger KH, Canestraro M, Vanella L, Dim DH, Manthati VL, Konduru SR, Falck JR, Schwartzman M, Abraham NG. Epoxyeicosatrienoic acid agonist rescues the metabolic syndrome phenotype of HO-2-null mice. J Phamacol Exp Ther. 2009;331:906–16. doi: 10.1124/jpet.109.157545. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Imig JD, Dimitropoulou C, Reddy DS, White RE, Falck JR. Afferent arteriolar dilation to 11,12-EET alanogs involves PP2A activity and Ca-activated K channels. Microcirculation. 2008;15:137–50. doi: 10.1080/10739680701456960. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Khan MAH, Liu J, Kumar G, S SX, Falck JR, Imig JD. Novel orally active epoxyeicosatrienoic acid analog attenuates cisplatin nephrotoxicity. FASEB J. 2013;27:2946–56. doi: 10.1096/fj.12-218040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Khan HMA, Neckar J, Manthati VL, Errabelli R, Pavlov TS, Staruschenko A, Falck JR, Imig JD. Orally active epoxyeicosatrienoic acid analog attenuates kidney injury in hypertensive Dahl salt-sensitive rat. Hypertension. 2013;62:905–13. doi: 10.1161/HYPERTENSIONAHA.113.01949. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Jichova S, Kopkan L, Huskova Z, Dolezelova S, Neckar J, Kujal P, Vernerova Z, Kramer HJ, Sadowski J, Kompanoowska-Jezierska E, Reddy NR, Falck JR, Imig JD, Cervenka L. Epoxyeicosatrienoic acid analog attenuates devlopment malignant hypertension but does not reverse it once established: a study in Cyp1a1-Ren-2 transgenic rats. J Hypertension. 2016;34:2008–25. doi: 10.1097/HJH.0000000000001029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.El-Sikhry HE, Alsaieh N, Dakarapu R, Falck JR, Seubert JM. Novel roles of epoxyeicosanoids in regulating cardiac mitochondria. PLoS One. 2016;11:e0160380. doi: 10.1371/journal.pone.0160380. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Batchu SN, Lee SB, Qadhi RS, Chaudhary KR, El-Sikhry H, Kodela R, Falck JR, Seubert JM. Cardioprotective effect of a dual acting epoxyeicosatrienoic acid analogue toward ischemia reperfusion injury. Brit J Pharmacol. 2011;162:897–907. doi: 10.1111/j.1476-5381.2010.01093.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Oni-Orisan A, Alsaleh N, Lee CR, Seubert JM. Epoxyeicosatrienoic acids and cardioprotection: the road to translation. J Mol Cell Cardiol. 2014;74:199–208. doi: 10.1016/j.yjmcc.2014.05.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Cao J, Tsenovoy PL, Thompson EA, Falck JR, Touchon R, Sodhi K, Rezzani R, Shapiro JI, Abraham NG. Agonists of epoxyeicosatrienoic acids reduce infarct size and ameliorate cardiac dysfunction via activation of HO-1 and Wnt1 connical pathway. Prostag and Other Lipid Mediat. 2015;116–117:76–86. doi: 10.1016/j.prostaglandins.2015.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Neckar J, Khan HMA, Gross GJ, Falck JR, Campbell WB, Kolar F, Imig JD. Epoxyeicosatrienoic acid analog EET-B attenuates post-myocardial infaraction remodeling and renal injury in spontanelusly hypertensive rats. Am J Physiol. 2017 doi: 10.1042/CS20180728. in revision. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Yang B, Graham L, Dikolov S, Mason RP, Falck JR, Liao JK, Zeldin DC. Overexpression of cytochrome P450 CYP2J2 protects against hypoxia-reoxygenastion injury in cultured bovine aortic endothelial cells. Mol Pharmacol. 2001;60:310–20. doi: 10.1124/mol.60.2.310. [DOI] [PubMed] [Google Scholar]
- 141.Zhao G, Wang J, Xu X, Jing Y, Tu L, Li X, Chen C, Cianflone K, Wang P, Dackor RT, Zeldin DC, Wang DW. Epoxyeicosatrienoic acid protect rat hearts against tumor necrosis factor-alpha-induced injury. J Lipid Res. 2012;53:456–66. doi: 10.1194/jlr.M017319. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Khan HMA, Fish B, Wahl G, Sharma A, Falck JR, P MP, Moulder JE, Imig JD, Cohen EP. Epoxyeicosatrienoic acid analogue mitigates kidney injury in a rat model of radiation nephropathy. Clin Sci. 2016;130:587–99. doi: 10.1042/CS20150778. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Yeborah MM, Khan HMA, Chesnik MA, Sharma A, Paudyai MP, Falck JR, Imig JD. The epoxyeicosatrienoic accid analog, PVPA, ameliorates cyclosporine-induced hypertension and renal injury. Am J Physiol. 2016;311:F576–F85. doi: 10.1152/ajprenal.00288.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Skibba M, Khan HMA, Kolb LL, Yeboah M, Falck JR, Amaradhi R, Imig JD. Epoxyeicosatrienoic acid analog decreases renal fibrosis by reducing epithelial-to-mesenchymal transition. Frontiers Pharmacol. 2017 doi: 10.3389/fphar.2017.00406. in revision. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Zeisberg M, N EG. Mechanisms of tubulointerstitial fibrosis. J Am Soc Nephrol. 2010;21:1819–34. doi: 10.1681/ASN.2010080793. [DOI] [PubMed] [Google Scholar]
