当前位置: 首页 > 医学版 > 期刊论文 > 内科学 > 循环研究杂志 > 2005年 > 第10期 > 正文
编号:11255988
Modulation of the Ca2 Permeable Cation Channel TRPV4 by Cytochrome P450 Epoxygenases in Vascular Endothelium
     The Department of Physiology (J.V., G.O., A.J., T.V., B.N.), Campus Gasthuisberg, KU Leuven, Belgium

    Institut fe Kardiovaskulre Physiologie (B.F., I.F., R.B.), Johann Wolfgang Goethe-Universitt Theodor-Stern Kai 7, Frankfurt am Main, Germany

    Department of Entomology and UC Davis Cancer Center (C.M., B.D.H.), University of California, Davis

    Jichi Medical School (M.S.) Yakushiji, Kawachi, Tochigi, Japan.

    Abstract

    TRPV4 is a broadly expressed Ca2+-permeable cation channel in the vanilloid subfamily of transient receptor potential channels. TRPV4 gates in response to a large variety of stimuli, including cell swelling, warm temperatures, the synthetic phorbol ester 4-phorbol 12,13-didecanoate (4-PDD), and the endogenous lipid arachidonic acid (AA). Activation by cell swelling and AA requires cytochrome P450 (CYP) epoxygenase activity to convert AA to epoxyeicosatrienoic acids (EETs) such as 5,6-EET, 8,9-EET, which both act as direct TRPV4 agonists. To evaluate the role of TRPV4 and its modulation by the CYP pathway in vascular endothelial cells, we performed Ca2+ imaging and patch-clamp measurements on mouse aortic endothelial cells (MAECs) isolated from wild-type and TRPV4eC/eC mice. All TRPV4-activating stimuli induced robust Ca2+ responses in wild-type MAECs but not in MAECs isolated from TRPV4eC/eC mice. Upregulation of CYP2C expression by preincubation with nifedipine enhanced the responses to AA and cell swelling in wild-type MAECs, whereas responses to other stimuli remained unaffected. Conversely, inhibition of CYP2C9 activity with sulfaphenazole abolished the responses to AA and hypotonic solution (HTS). Moreover, suppression of EET hydrolysis using 1-adamantyl-3-cyclo-hexylurea or indomethacin, inhibitors of soluble epoxide hydrolases (sEHs), and cyclooxygenases, respectively, enhanced the TRPV4-dependent responses to AA, HTS, and EETs but not those to 4-PDD or heat. Together, our data establish that CYP-derived EETs modulate the activity of TRPV4 channels in endothelial cells and shows the unraveling of novel modulatory pathways via CYP2C modulation and sEH inhibition.

    Key Words: TRP channels endothelium epoxygenases epoxide hydrolases

    Introduction

    TRPV4 is a Ca2+ entry channel belonging to the vanilloid subfamily of the transient receptor potential (TRP) channels.1,2 It is expressed in a broad range of tissues, including lung, spleen, kidney, testis, fat, brain, cochlea, skin, smooth muscle, liver, and vascular endothelium.3eC5 The physiological role of TRPV4 in these tissues may be highly divergent because TRPV4 is able to respond to a wide variety of physical, thermal, and chemical stimuli.

    Initially, TRPV4 was put forward as a mechanosensor or osmosensor. This was based on the finding that the channel opens in response to hypotonic cell swelling3eC8 as well as shear stress.9 Indeed, mice lacking the TRPV4 gene show a disturbance of osmoregulation and an increased mechanical nociceptive threshold.10,11 Furthermore, TRPV4 functions as a transducer of hypo-osmotic stimuli in primary afferent nociceptors7 and plays an essential role in taxol-induced nociceptive behavioral responses to mechanical and hypotonic stimulation of the hind paw.12

    More recently, it was found that TRPV4 can also be activated by heating,13,14 and that it is required for normal thermal responsiveness in vivo. The TRPV4eC/eC mice "select" a warmer floor temperature in a gradient apparatus than their wild-type littermates and exhibit prolonged withdrawal latencies during acute tail heating.15 Moreover, TRPV4 can be activated by a large variety of chemical stimuli, including low pH and citrate11 and the synthetic phorbol-derivative 4-phorbol 12,13-didecanoate (4-PDD), which can be considered the most robust and direct channel activator.16 Recently, we have shown that arachidonic acid (AA) and the endocannabinoid anandamide (arachidonoylethanolamide) can act as endogenous activators of TRPV4. AA and anandamide act via the cytochrome P450 (CYP) epoxygenase-dependent formation of epoxyeicosatrienoic acids (EETs; 5,6-EET and 8,9-EET), which act as direct channel activators.17

    EETs are known to have antihypertensive and anti-inflammatory properties and play a role in the maintenance of renal vascular function.18 They have been identified as endothelium-derived hyperpolarizing factors19,20 and elicit vasodilatation attributable to hyperpolarization by activating smooth muscle Ca2+-activated potassium channels.21,22 Although Ca2+-activated K+ channels play a central role in mediating many of the effects of EETs in vascular tissue, other molecular targets for EETs may exist. Therefore, the main objective of our work was to evaluate TRPV4 as a possible EET-regulated Ca2+ entry channel in mouse aorta endothelium. We report here that pharmacological agents that affect CYP epoxygenase activity and EET metabolism can efficiently modulate TRPV4, a target for EETs in endothelial cells.

    Materials and Methods

    Cell Culture, Transfection, and TRPV4eC/eC Mouse

    Mouse aortic endothelial cells (MAECs) were isolated using the "primary explant technique" described in detail previously.23 For all manipulations and measurements, only MAECs from the first and the second passage were used. The TRPV4+/+ and TRPV4eC/eC mice11 were maintained on ad labium diet chow and tap water for drinking and used at 16 weeks of age for isolation of endothelial cells. For the CYP2C expression experiments, mice aortas were isolated, cleaned of adherent connective tissue, and cut into rings, which were incubated in culture medium containing either solvent or nifedipine (0.1 eol/L) for 18 hours. Thereafter, the rings were recovered and endothelial cell proteins harvested by shaking in Triton X-100 lysis buffer for 30 minutes at 4°C. EA.hyb926 cells,24 which are derived from human umbilical vein endothelial cells (HUVECs), were transiently transfected with the murine TRPV4 (mouse mTRP12; accession number CAC20703) vector, and cloned as a BamHI fragment into the BcII site of the pCAGGS/IRES-GFP vector, which allows detection of transfected cells based on their green fluorescence when illuminated at 475 nm. For transfection, we used GeneJuice Transfection Reagent (Novagen). Green fluorescent negative cells from the same batch were used as controls.

    RNA Extraction, cDNA Synthesis, and Analysis

    Detection of TRPV4 mRNA expression in MAECs was performed as described.25 Briefly, total RNA from MAECs was isolated with the RNeasy kit (Qiagen) and cDNA was synthesized using Ready-to-Go You-Prime First-Strand Beads (Amersham Biosciences), according to the manufacturer recommendations. Specific TRPV4 and -actin fragments were amplified in a 50-e蘈 polymerase chain reaction (PCR) with 25 pmol/L of each of the primers 5'-GAT GGA GGA GAA AGG TCG TG and 5'-GAG AAC TGT CTC CAG GTT GC (for amplification of a 733-bp TRPV4 fragment) or 25 pmol/L of each of the primers 5'-GTC GTC GAC AAC GGC TCC GGC and 5'-CAC AGC CTG GAT GGC TAC GTA C (for amplification of a 350-bp -actin fragment), one tenth of cDNA synthesis reactions, and Taq DNA polymerase (Amersham Biosciences), according to manufacturer protocol.

    Protein Extraction and Immunodetection

    Isolated MAECs grown in standard medium on Matrigel (Becton Dickinson) were collected by centrifugation at 500g for 5 minutes at 4°C (all subsequent steps are performed at 4°C), resuspended in 3 mL ice-cold lysis buffer (50 mmol/L Tris, pH 7.5, 10 mmol/L KCl, 1.5 mmol/L MgCl2, 1 mmol/L EDTA, and 1 mmol/L phenylmethylsulfonyl fluoride [PMSF]), and a protease inhibitor cocktail (10 e/mL leupeptin and antipain and 2 e/mL chymostatin and pepstatin). After rotation on an end-over-end stirrer for 30 minutes, cells were subjected for 3 freezeeCthaw cycles and then were passed 10 times through a 26-gauge needle. Normal osmolarity of resulted lysates were restored by supplementation with 150 mmol/L NaCl and then they were centrifuged at 4000g for 15 minutes to remove nuclei, mitochondria, and any remaining large cellular fragments. Obtained supernatants were centrifuged at 20 000g for 20 minutes and then 100 000g for 1 hour, yielding low-speed (P20/20) and high-speed (P100/60) pellets of membrane fractions, which were solubilized in a cold PBS (10 mmol/L phosphate buffer, pH 7.4, 137 mmol/L NaCl, and 2.7 mmol/L KCl,) containing 1% Triton X-100, 0.2% sodium dodecyl sulfate (SDS), 1 mmol/L PMSF, and a protease inhibitor cocktail. Protein concentrations were determined by the bicinchoninic acid assay method, using BSA as a standard. Samples (2 e) were solubilized in 5 e蘈 of 3-fold concentrated sample buffer (240 mmol/L Tris, pH 6.8, 30% glycerol, 6% SDS, 3% dithiothreitol, and 0.015% bromophenol blue) by heating to 56°C for 15 minutes and then subjected to 8% SDS-PAGE. Separated proteins were transferred to a nitrocellulose membrane (Schleicher & Schuell) and probed with purified polyclonal rabbit anti-TRPV4 (1:1000 dilution)13 or antieCsecretory pathway Ca2+-ATPase 1 (SPCA1; 1:3000 dilution)26 antibodies for 1 hour at room temperature. Immunoreactive complexes were visualized by chemiluminescence using anti-rabbit IgG antibodies conjugated to horseradish peroxidase (1:4000 dilution; Amersham Biosciences).

    For immunodetection of CYP2C, MAECs were lysed in a buffer containing 50 mmol/L Tris/HCl, pH 7.5, 150 mmol/L NaCl, 100 mmol/L NaF, 15 mmol/L Na4P2O7, 2 mmol/L Na3VO4, 1% (v/v) Triton X-100, 44 e/mL PMSF and the protease inhibitor cocktail, left on ice for 10 minutes and centrifuged at 10 000g for 10 minutes. Proteins (50 e) were subjected to SDS-PAGE as described above, and the CYP2C specific bands were visualized using a polyclonal antibody raised in rabbit against CYP2C9 (Abcam). To ensure equal protein loading, the blots were reprobed with polyclonal anti-platelet/endothelial cell adhesion molecule-1 (PECAM-1) antibody raised in goat (Santa Cruz Biotechnology). The detection of the specifically bound antibodies was performed as described above.

    Solutions

    For electrophysiological measurements, the standard extracellular solution contained (in mmol/L): 150 NaCl, 6 CsCl, 1 MgCl2, 5 CaCl2, 10 glucose, and 10 HEPES, buffered at pH 7.4 with NaOH. The osmolarity of this solution, as measured with a vapor pressure osmometer (5 Wescor 5500; Schlag), was 320±5 mosmol/L. When measuring swelling-activated responses, we used an isotonic solution containing (in mmol/L): 105 NaCl, 6 CsCl, 5 CaCl2, 1 MgCl2, 10 HEPES, 90 D-mannitol, 10 glucose, and buffered pH 7.4 with NaOH (320±5 mosmol/L). Cell swelling was induced by omitting mannitol from this solution (240 mosmol/L; 25% reduction of osmolarity).

    The pipette solution was composed of (in mmol/L): 20 CsCl, 100 Asp, 1 MgCl2, 10 HEPES, 4 Na2ATP, 10 BAPTA, and 5.56 CaCl2. The free Ca2+ concentration of this solution is &200 nmol/L. The non-protein kinase CeCactivating phorbol ester (Sigma Aldrich) was applied at a 1-eol/L concentration from a 10-mmol/L stock solution in ethanol. AA (Sigma) was used at a concentration of 3 eol/L from a 10-mmol/L stock solution in ethanol, and 5,6-EET and 8,9-EET (Biomol) were applied at a concentration of 500 nmol/L and 1 eol/L, respectively. Nifedipine (2 eol/L; Sigma Aldrich), sulfaphenazole (1 eol/L; Sigma Aldrich), 1-adamantyl-3-cyclo-hexylurea (ACU; 2 eol/L), and indomethacin (2 eol/L, Sigma Aldrich) were applied to the extracellular medium for 18 hours before the application of the different activation stimuli. ACU was synthesized by standard procedures, and high purity was supported by proton and carbon nuclear magnetic resonance, high-performance liquid chromatography (HPLC)eCmass spectrometry, HPLC-time of flight, and other means.27 Indomethacin was dissolved in 1% Na bicarbonate to an end concentration of 10 eol/L. In the case of activation by heat, the perfusate was warmed using a water jacket device. We changed the temperature in the bath from 22°C to 43°C in 50 seconds. Other experiments were performed at room temperature (22°C to 25°C).

    Electrophysiological Recordings and Ca2+ Measurements

    Whole-cell membrane currents and Ca2+ measurements were measured as described in detail previously.8 For every condition, at least 20 cells from at least 3 independent experiments were assayed. The calibration procedure is described previously.8

    Data Analysis

    Electrophysiological data were analyzed using WinASCD software (Guy Droogmans). The Origin 6.1 software package (OriginLab Corporation) was used for statistical analysis and data display. Data are expressed as the mean±SEM. Statistics on Ca2+ increases are done on peak responses. Significance between individual groups was tested using Student’s t test for "normal" data and the nonparametric and distribution-free KolmogoroveCSmirnoff Test for "non-normal" data.

    Results

    TRPV4 Is Expressed in MAECs

    To confirm the expression of TRPV4, we first performed RT-PCR analysis on mRNA extracted from MAECs isolated from wild-type (TRPV4+/+) and TRPV4-deficient (TRPV4eC/eC) mice. In the TRPV4-deficient mouse, a PKG-neo cassette has been inserted into exon 4 of the TRPV4 gene, resulting in the premature termination and the loss of TRPV4 expression.11 As shown in Figure 1A, the TRPV4 fragment was amplified from TRPV4+/+ MAECs but not from TRPV4eC/eC MAECs. Subsequently, we examined TRPV4 expression in MAECs at the protein level. Using a differential centrifugation protocol, we prepared low-speed (P20/20) and high-speed (P100/60) membrane fractions from MAECs isolated from TRPV4+/+ and TRPV4eC/eC mice. The P20/20 fraction comprised the plasma membrane together with a part of large intracellular membrane fragments (the plasma membraneeCenriched fraction), whereas other intracellular membranes were present in the P100/60 fraction (data not shown). Specific anti-TRPV4 antibodies13 recognized 2 bands of &110 and 120 kDa in both fractions isolated from TRPV4+/+ MAECs (Figure 1B). The TRPV4 band intensity was higher in the P20/20 fraction, suggesting that the majority of TRPV4 is most likely present in the plasma membrane. In contrast, neither of these 2 bands were present in isolates from TRPV4eC/eC mice.

    Next, we performed intracellular Ca2+ measurements to evaluate functional expression of TRPV4 channels in MAECs. We compared the response to 6 different TRPV4-activating stimuli in wild-type and TRPV4eC/eC MAECs: 4-PDD (1 eol/L), a 25% hypotonic solution (HTS), AA (3 eol/L), 5,6-EET (500 nmol/L), 8,9-EET (1 eol/L), and heating to 42°C (Figure 2A through 2F). All 6 stimuli evoked robust increases in [Ca2+]i in TRPV4+/+ MAECs. In TRPV4eC/eC MAECs, responses to 4-PDD, 5,6-EET, and 8,9-EET were fully abolished (Figure 2A, 2D, and 2E), whereas the responses to HTS, AA, and heat were reduced but still detectable (Figure 2B, 2C, and 2F). As a control stimulus, we used ATP (100 nmol/L), which causes an increase in [Ca2+]i in endothelial cells via activation of purinergic receptors. As shown in Figure 2G, ATP induced a similar Ca2+ signal in TRPV4+/+ and TRPV4eC/eC MAECs, indicating that other aspects of Ca2+ handling are not drastically changed in the knockout cells. Together, RT-PCR, Western blot analysis, and [Ca2+]i measurements demonstrate that TRPV4 is functionally expressed in the plasma membrane of MAECs.

    CYP Epoxygenase Modulators Selectively Affect TRPV4 Responses to AA and Cell Swelling

    The above results indicated that the increase in [Ca2+]i in MAECs induced by 5,6-EET and 8,9-EET is fully dependent on TRPV4. These EETs, which are known to have vasorelaxant properties, are formed from AA through the action of CYP epoxygenases (Figure 3A). We therefore tested whether pharmacological modulation of CYP epoxygenase activity affected TRPV4 activity in MAECs.

    To increase CYP2C expression, MAECs were preincubated with nifedipine. In line with previous work,28 we found that this preincubation led to an increase in protein expression (Figure 3B). We investigated the effects of nifedipine pretreatment on the [Ca2+]i response to the different TRPV4-activating stimuli in MAECs derived from TRPV4+/+ and TRPV4eC/eC mice. The basal [Ca2+]i was not affected by preincubation with nifedipine (data not shown), indicating that basal activity of TRPV4 is not affected by nifedipine. Direct application of nifedipine did not change [Ca2+]i in MAECs ([Ca2+]i=2±1 nmol/L; n=53). Importantly, pretreatment of TRPV4+/+ MAECs with nifedipine significantly enhanced the response to AA (Figure 4B) and HTS (Figure 4C). This effect of nifedipine was not observed in TRPV4eC/eC MAECs (Figure 4B and 4C). Moreover, nifedipine pretreatment did not affect the responses to 4-PDD (Figure 4A), exogenously applied 5,6-EET (Figure 4D), 8,9-EET (Figure 4E), or heat (Figure 4F) in MAECs from TRPV4+/+ or TRPV4eC/eC mice. In addition, the TRPV4-independent response to ATP (100 nmol/L) was not affected by nifedipine pretreatment (data not shown).

    Next, we investigated the effect of sulfaphenazole,29eC31 an inhibitor of CYP2C enzymes, on the [Ca2+]i response to the different TRPV4-activating stimuli in MAECs derived from TRPV4+/+ and TRPV4eC/eC mice. Sulfaphenazole did not affect the basal [Ca2+]i level in MAECs from either TRPV4 +/+ or TRPV4eC/eC mice (data not shown), indicating that the compound does not directly affect TRPV4 activity. However, sulfaphenazole strongly reduced the response of TRPV4+/+ MAECs to AA (Figure 4H) or HTS (Figure 4I). This inhibitory effect of sulfaphenazole was not observed in TRPV4eC/eC MAECs. Sulfaphenazole had no effect on the responses to 4-PDD (Figure 4G), exogenously applied 5,6-EET (Figure 4J), 8,9-EET, heating (Figure 4L), or ATP (data not shown) in cells from TRPV4+/+ and TRPV4eC/eC mice.

    Together, the results obtained with nifedipine and sulfaphenazole indicate that modulation of CYP2C activity selectively affects the TRPV4-dependent responses to AA and HTS. These data demonstrate that activation of TRPV4 in MAECs by HTS and AA requires the CYP2C-dependent formation of EETs, whereas 4-PDD and heat activate the channel via a different, CYP-independent pathway.

    Inhibition of EET Degradation Increases the TRPV4-Dependent Ca2+ Responses to AA, Cell Swelling, and EETs

    Most EETs are metabolized by soluble epoxide hydrolase (sEH)eCdependent epoxide hydration into the biologically less active dihydroxyeicosatrienoic acids (DHETs)32 (Figure 3A). We used a selective inhibitor of sEH, ACU,27,33 to investigate the effect of reduced EET breakdown on the TRPV4 activity in MAECs. In TRPV4+/+ MAECs, ACU (2 eol/L) significantly enhanced the responses to AA (Figure 5B), HTS (Figure 5C), and 8,9-EET (Figure 5E) but had no effect on the responses to 4-PDD (Figure 5A), 5,6-EET (Figure 5D), or heat (Figure 5F). In TRPV4-deficient cells, ACU only caused a modest enhancement of the response to AA (Figure 5B). ACU had no effect on the response to ATP stimulation in either TRPV4eC/eC or TRPV4+/+ MAECs (data not shown).

    The metabolism of 5,6-EET differs from that of the other EETs in that it is a much weaker substrate for sEH. In contrast, 5,6-EET is rapidly metabolized by cyclooxygenase (COX) enzymes to generate 5,6-epoxy-prostaglandin34 (Figure 3A). To test the consequence of inhibition of 5,6-EET metabolism, we used the COX inhibitor indomethacin. Similar to the effects of ACU, the indomethacin (2 eol/L) enhanced the [Ca2+]i response to AA (Figure 5H) and HTS (Figure 5I) without affecting the responses to 4-PDD (Figure 5G) or heat (Figure 5I). Indomethacin potentiated the response to 5,6-EET (Figure 5J) but not that to 8,9-EET (Figure 5K), consistent with 5,6-EET but not 8,9-EET being a COX substrate. Again, effects of indomethacin were only observed in endothelial cells that expressed TRPV4 (Figure 5G through 5L), and indomethacin had no effects on basal [Ca2+]i or on the response to ATP (data not shown).

    Thus, inhibition of the metabolism of 5,6-EET and 8,9-EET increases the TRPV4 responses to AA and cell swelling. This indicates that the activity of sEH and COX enzymes reduces cellular EET levels in MAECs, leading to lowered activity of TRPV4.

    Effects of Modulating EET Generation and Degradation on Currents Through TRPV4 Channels

    To investigate the consequences of modulating CYP activity on EET degradation on TRPV4 currents, we measured TRPV4 channel activation directly using the whole-cell patch-clamp technique. For this purpose, we used HUVEC-derived EA.Hyb926 cells (EA cells) transiently transfected with TRPV4, which yield more robust TRPV4 current responses than native MAECs. In [Ca2+]i measurements on TRPV4-transfected EA cells (data not shown), we obtained qualitatively identical results as in TRPV4+/+ MAECs. In whole patch-clamp experiments using TRPV4-transfected EA cells, AA (3 eol/L) activated large currents showing the characteristic features of the TRPV4 channel: moderate outward rectification, a reversal potential around +10 mV, and S-shaped currenteCvoltage curves.2 In agreement with its effects in MAECs, AA-induced current amplitudes were larger in cells pretreated with nifedipine to increase CYP2C expression (Figure 6C, 6D, and 6I), smaller in cells treated with the CYP2C inhibitor sulfaphenazole (Figure 6E, 6F, and 6I) and larger in cells treated with the sEH inhibitor ACU (Figure 6G through 6I). In contrast, current responses to application of 4-PDD (1 eol/L) were unaffected by nifedipine, sulfaphenazole, or ACU (Figure 6J).

    Discussion

    In this study, we evaluated whether the Ca2+ entry channel TRPV4 functions as a molecular target for EETs in MAECs and assessed the effects of pharmacological modulation of EET metabolism on TRPV4 activity. The results of the present investigation demonstrate that the response of TRPV4 to AA and cell swelling is modulated by formation and breakdown of EETs. Indeed, stimulation of the formation of EETs from AA, by increasing CYP2C expression, enhanced the TRPV4 response to AA or cell swelling, whereas inhibition of CYP2C expression blocked the response to these stimuli. Inhibition of COX and sEH enzymes to reduce EET breakdown enhanced the effect of AA, cell swelling, and EETs. By using MAECs from TRPV4eC/eC mice, we confirmed that these effects were mediated by TRPV4. Moreover, responses to 4-PDD and heat were independent of EET metabolism, confirming that they activate TRPV4 via separate pathways.

    We initially tested the effects of different well-described TRPV4-activating stimuli on Ca2+ responses in these cells and observed that stimulation with 4PDD, AA, cell swelling, heat, 5,6-EET, and 8,9-EET all resulted in robust responses in TRPV4+/+ cells. In TRPV4eC/eC MAECs, we did not detect any response to 4PDD, 5,6-EET, and 8,9-EET, whereas clear but reduced responses were obtained with AA, cell swelling, and heat. A possible explanation for this residual effect in the knockout cells could be that AA or downstream metabolites of AA activate Ca2+ permeable channels other than TRPV4 or a possible upregulation of other channel types. Likewise, other temperature-sensitive TRP channels such as TRPV3, which also demonstrate an activation threshold between 25°C and 42°C,35 may account for the residual response to heat. The presence of other swelling-activated Ca2+ channels like TRPV2 or TRPM336,37 could explain the Ca2+ increase caused by cell swelling in TRPV4eC/eC cells.

    Our data clearly illustrate the existence of several ways in which TRPV4 channels can be activated. Activation by AA and HTS is highly sensitive to pharmacological agents that affect EET metabolism. AA is the substrate for CYP epoxygenases expressed in endothelial cells, which mediate the production of the TRPV4-activating EETs 5,6-EET and 8,9-EET. Activation of TRPV4 by cell swelling can most probably be attributed to the liberation of endogenous AA from phospholipids attributable to activation of phospholipase A2.2,8 The mode of TRPV4 activation by -phorbols and heat is clearly distinct from that by AA or swelling; CYP induction, CYP inhibition, and inhibition of sEH or COX failed to affect TRPV4 activation by these stimuli. Moreover, activation of TRPV4 by -phorbols and heat depends on a specific tyrosine residue, Tyr555, which likely forms part of the -phorbol binding site.8

    The results of the present study indicate that a novel way of modulating TRPV4 activity is to interfere with the breakdown of the EETs into DHETs. We found that preventing the breakdown of EETs using the selective sEH inhibitor ACU increased the TRPV4-dependent responses to AA, 8,9-EET, and cell swelling but not those induced by 4-PDD or heat. These data also imply that DHETs are no longer able to activate TRPV4, in line with a previous report.17 The fact that ACU enhanced the AA-induced Ca2+ response in TRPV4eC/eC MAECs could indicate that EETs also affect the activity/open probability of other MAEC channels. The response of endothelial cells to the exogenous application of 5,6-EET was not affected by ACU but was potentiated by the COX inhibitor indomethacin, which is consistent with the fact that 5,6-EET is the only regioisomer metabolized by COX.34

    EETs are biologically important signal transduction molecules that have been implicated in numerous biological processes including communication between endothelial and vascular smooth muscle cells.38 Although some of the effects attributed to EETs, such as enhanced celleCcell communication39 and angiogenesis,40 are membrane potential independent and most likely related to the activation of protein kinase A, it is tempting to speculate that the activation of TRPV4 channels underlies some of the anti-inflammatory and antihypertensive effects of the EETs. Our data not only highlight the potential role of TRPV4 in the regulation of vascular homeostasis but also underscore the role of sEH in the regulation of endothelial cell signaling. Indeed, our finding that sEH inhibitors increase endothelial cell [Ca2+]i via TRPV4 indicate that these drugs have the potential to affect a number of Ca2+-dependent processes in endothelial cells. Enhanced TRPV4 activation may account, at least partially, for the hypotensive effects of sEH inhibitors in vitro and in vivo.32,41,42

    Acknowledgments

    This work was supported by the Human Frontiers Science Programme (HFSP research grant RGP 32/2004), the Belgian Federal Government, the Flemish Government, the Onderzoeksraad KU Leuven (GOA 2004/07, FWO G.0214.99, FWO G. 0136.00; FWO. G.0172.03, and the Interuniversity Poles of Attraction Program, Prime Ministers Office IUAP), Philip Morris Inc. (research grant to I.F.), and the Deutsche Forschungsgemeinschaft (FI 830/2-1 to B.F. and R.B.). We thank M. Caterina (Johns Hopkins School of Medicine, Baltimore, Md) and Professor F. Wuytack (Katholieke Universiteit Leuven, Belgium) for providing anti-TRPV4 and anti-SPCA1 antibodies, respectively. J.V. is a postdoctoral fellow of the FWO (Fonds voor Wetenschappelijk Onderzoek, Flanders).

    This manuscript was sent to Hans Michael Piper, Consulting Editor, for review by expert referees, editorial decision, and final disposition.

    References

    Nilius B, Droogmans G, Wondergem R. Transient receptor potential channels in endothelium:solving the calcium entry puzzle Endothelium. 2003; 10: 5eC15.

    Nilius B, Vriens J, Prenen J, Droogmans G, Voets T. TRPV4 calcium entry channel: a paradigm for gating diversity. Am J Physiol Cell Physiol. 2004; 286: C195eCC205.

    Liedtke W, Choe Y, Marti-Renom MA, Bell AM, Denis CS, Sali A, Hudspeth AJ, Friedman JM, Heller S. Vanilloid receptor-related osmotically activated channel (VR-OAC), a candidate vertebrate osmoreceptor. Cell. 2000; 103: 525eC535.

    Strotmann R, Harteneck C, Nunnenmacher K, Schultz G, Plant TD. OTRPC4, a nonselective cation channel that confers sensitivity to extracellular osmolarity. Nat Cell Biol. 2000; 2: 695eC702.

    Wissenbach U, Bodding M, Freichel M, Flockerzi V. Trp12, a novel Trp related protein from kidney. FEBS Lett. 2000; 485: 127eC134.

    Nilius B, Prenen J, Wissenbach U, Bodding M, Droogmans G. Differential activation of the volume-sensitive cation channel TRP12 (OTRPC4) and volume-regulated anion currents in HEK-293 cells. Pfle筭ers Arch Eur J Physiol. 2001; 443: 227eC233.

    Alessandri-Haber N, Yeh JJ, Boyd AE, Parada CA, Chen X, Reichling DB, Levine JD. Hypotonicity induces TRPV4-mediated nociception in rat. Neuron. 2003; 39: 497eC511.

    Vriens J, Watanabe H, Janssens A, Droogmans G, Voets T, Nilius B. Cell swelling, heat, and chemical agonists use distinct pathways for the activation of the cation channel TRPV4. Proc Natl Acad Sci U S A. 2004; 101: 396eC401.

    Gao X, Wu L, O’Neil RG. Temperature-modulated diversity of TRPV4 channel gating: Activation by physical stresses and phorbol ester derivatives through protein kinase C-dependent and -independent pathways. J Biol Chem. 2003; 278: 27129eC22737.

    Liedtke W, Friedman JM. Abnormal osmotic regulation in trpv4eC/eC mice. Proc Natl Acad Sci U S A. 2003; 100: 13698eC13703.

    Suzuki M, Mizuno A, Kodaira K, Imai M. Impaired pressure sensation in mice lacking TRPV4. J Biol Chem. 2003; 278: 22664eC22668.

    Alessandri-Haber N, Dina OA, Yeh JJ, Parada CA, Reichling DB, Levine JD. Transient receptor potential vanilloid 4 is essential in chemotherapy-induced neuropathic pain in the rat. J Neurosci. 2004; 24: 4444eC4452.

    Guler AD, Lee H, Iida T, Shimizu I, Tominaga M, Caterina M. Heat-evoked activation of the ion channel, TRPV4. J Neurosci. 2002; 22: 6408eC6414.

    Watanabe H, Vriens J, Suh SH, Benham CD, Droogmans G, Nilius B. Heat-evoked activation of TRPV4 channels in a HEK293 cell expression system and in native mouse aorta endothelial cells. J Biol Chem. 2002; 277: 47044eC47051.

    Lee H, Iida T, Mizuno A, Suzuki M, Caterina MJ. Altered thermal selection behavior in mice lacking transient receptor potential vanilloid 4. J Neurosci. 2005; 25: 1304eC1310.

    Watanabe H, Davis JB, Smart D, Jerman JC, Smith GD, Hayes P, Vriens J, Cairns W, Wissenbach U, Prenen J, Flockerzi V, Droogmans G, Benham CD, Nilius B. Activation of TRPV4 channels (hVRL-2/mTRP12) by phorbol derivatives. J Biol Chem. 2002; 277: 13569eC13577.

    Watanabe H, Vriens J, Prenen J, Droogmans G, Voets T, Nilius B. Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channels. Nature. 2003; 424: 434eC438.

    Zhao X, Yamamoto T, Newman JW, Kim IH, Watanabe T, Hammock BD, Stewart J, Pollock JS, Pollock DM, Imig JD. Soluble epoxide hydrolase inhibition protects the kidney from hypertension-induced damage. J Am Soc Nephrol. 2004; 15: 1244eC1253.

    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: 493eC497.

    Campbell WB, Gebremedhin D, Pratt PF, Harder DR. Identification of epoxyeicosatrienoic acids as endothelium-derived hyperpolarizing factors. Circ Res. 1996; 78: 415eC423.

    Lu T, Katakam PV, VanRollins M, Weintraub NL, Spector AA, Lee HC. Dihydroxyeicosatrienoic acids are potent activators of Ca(2+)-activated K(+) channels in isolated rat coronary arterial myocytes. J Physiol. 2001; 534: 651eC667.

    Campbell WB, Deeter C, Gauthier KM, Ingraham RH, Falck JR, Li PL. 14,15-Dihydroxyeicosatrienoic acid relaxes bovine coronary arteries by activation of K(Ca) channels. Am J Physiol Heart Circ Physiol. 2002; 282: H1656eCH1664.

    Suh SH, Vennekens R, Manolopoulos VG, Freichel M, Schweig U, Prenen J, Flockerzi V, Droogmans G, Nilius B. Characterisation of explanted endothelial cells from mouse aorta: electrophysiology and Ca2+ signaling. Pflugers Arch. 1999; 438: 612eC620.

    Edgell CJ, McDonald CC, Graham JB. Permanent cell line expressing human factor VIII-related antigen established by hybridization. Proc Natl Acad Sci U S A. 1983; 80: 3734eC3737.

    Vriens J, Janssens A, Prenen J, Nilius B, Wondergem R. TRPV channels and modulation by hepatocyte growth factor/scatter factor in human hepatoblastoma (HepG2) cells. Cell Calcium. 2004; 36: 19eC28.

    Van Baelen K, Vanoevelen J, Callewaert G, Parys JB, De Smedt H, Raeymaekers L, Rizzuto R, Missiaen L, Wuytack F. The contribution of the SPCA1 Ca2+ pump to the Ca2+ accumulation in the Golgi apparatus of HeLa cells assessed via RNA-mediated interference. Biochem Biophys Res Commun. 2003; 306: 430eC436.

    Morisseau C, Goodrow MH, Newman JW, Wheelock CE, Dowdy DL, Hammock BD. Structural refinement of inhibitors of urea-based soluble epoxide hydrolases. Biochem Pharmacol. 2002; 63: 1599eC1608.

    Fisslthaler B, Hinsch N, Chataigneau T, Popp R, Kiss L, Busse R, Fleming I. Nifedipine increases cytochrome P4502C expression and endothelium-derived hyperpolarizing factor-mediated responses in coronary arteries. Hypertension. 2000; 36: 270eC275.

    Brian WR, Srivastava PK, Umbenhauer DR, Lloyd RS, Guengerich FP. Expression of a human liver cytochrome P-450 protein with tolbutamide hydroxylase activity in Saccharomyces cerevisiae. Biochemistry. 1989; 28: 4993eC4999.

    Sai Y, Dai R, Yang TJ, Krausz KW, Gonzalez FJ, Gelboin HV, Shou M. Assessment of specificity of eight chemical inhibitors using cDNA-expressed cytochromes P450. Xenobiotica. 2000; 30: 327eC343.

    Mancy A, Dijols S, Poli S, Guengerich P, Mansuy D. Interaction of sulfaphenazole derivatives with human liver cytochromes P450 2C: molecular origin of the specific inhibitory effects of sulfaphenazole on CYP 2C9 and consequences for the substrate binding site topology of CYP 2C9. Biochemistry. 1996; 35: 16205eC16212.

    Yu Z, Xu F, Huse LM, Morisseau C, Draper AJ, Newman JW, Parker C, Graham L, Engler MM, Hammock BD, Zeldin DC, Kroetz DL. Soluble epoxide hydrolase regulates hydrolysis of vasoactive epoxyeicosatrienoic acids. Circ Res. 2000; 87: 992eC998.

    Morisseau C, Goodrow MH, Dowdy D, Zheng J, Greene JF, Sanborn JR, Hammock BD. Potent urea and carbamate inhibitors of soluble epoxide hydrolases. Proc Natl Acad Sci U S A. 1999; 96: 8849eC8854.

    Carroll MA, Garcia MP, Falck JR, McGiff JC. Cyclooxygenase dependency of the renovascular actions of cytochrome P450-derived arachidonate metabolites. J Pharmacol Exp Ther. 1992; 260: 104eC109.

    Smith GD, Gunthorpe MJ, Kelsell RE, Hayes PD, Reilly P, Facer P, Wright JE, Jerman JC, Walhin JP, Ooi L, Egerton J, Charles KJ, Smart D, Randall AD, Anand P, Davis JB. TRPV3 is a temperature-sensitive vanilloid receptor-like protein. Nature. 2002; 418: 186eC190.

    Muraki K, Iwata Y, Katanosaka Y, Ito T, Ohya S, Shigekawa M, Imaizumi Y. TRPV2 is a component of osmotically sensitive cation channels in murine aortic myocytes. Circ Res. 2003; 93: 829eC838.

    Grimm C, Kraft R, Sauerbruch S, Schultz G, Harteneck C. Molecular and functional characterization of the melastatin-related cation channel TRPM3. J Biol Chem. 2003; 278: 21493eC21501.

    Fleming I. Cytochrome p450 and vascular homeostasis. Circ Res. 2001; 89: 753eC762.

    Popp R, Brandes RP, Ott G, Busse R, Fleming I. Dynamic modulation of interendothelial gap junctional communication by 11,12-epoxyeicosatrienoic acid. Circ Res. 2002; 90: 800eC806.

    Michaelis UR, Falck JR, Schmidt R, Busse R, Fleming I. Cytochrome P4502C9-derived epoxyeicosatrienoic acids induce the expression of cyclooxygenase-2 in endothelial cells. Arterioscler Thromb Vasc Biol. 2005; 25: 321eC326.

    Jung O, Brandes RP, Kim IH, Schweda F, Schmidt R, Hammock BD, Busse R, Fleming I. Soluble epoxide hydrolase is a main effector of angiotensin II-induced hypertension. Hypertension. 2005; 45: 759eC765.

    Fang X, Weintraub NL, McCaw RB, Hu S, Harmon SD, Rice JB, Hammock BD, Spector AA. Effect of soluble epoxide hydrolase inhibition on epoxyeicosatrienoic acid metabolism in human blood vessels. Am J Physiol Heart Circ Physiol. 2004; 287: H2412eCH2420.(J. Vriens, G. Owsianik, B)