Abstract
Background
Impaired endothelium-dependent vasodilation has been suggested to be a key component of coronary microvascular dysfunction (CMD). A better understanding of endothelial pathways involved in vasodilation in human arterioles may provide new insight into the mechanisms of CMD.
Objective
To investigate the role of TRPV4, NOX4, and their interaction in human arterioles and examine the underlying mechanisms.
Methods and Results
Arterioles were freshly isolated from adipose and heart tissues obtained from 71 patients without coronary artery disease, and vascular reactivity was studied by videomicroscopy. In human adipose arterioles (HAA), ACh-induced dilation was significantly reduced by TRPV4 inhibitor HC067047 and by NOX 1/4 inhibitor GKT137831, but GKT137831 did not further affect the dilation in the presence of TRPV4 inhibitors. GKT137831 also inhibited TRPV4 agonist GSK1016790A-induced dilation in HAA and human coronary arterioles (HCA). NOX4 transcripts and proteins were detected in endothelial cells of HAA and HCA. Using fura-2 imaging, GKT137831 significantly reduced GSK1016790A-induced Ca2+ influx in the primary culture of endothelial cells and TRPV4-WT-overexpressing human coronary artery endothelial cells (HCAEC). However, GKT137831 did not affect TRPV4-mediated Ca2+ influx in non-phosphorylatable TRPV4-S823A/S824A-overexpressing HCAEC. In addition, treatment of HCAEC with GKT137831 decreased the phosphorylation level of Ser824 in TRPV4. Finally, proximity ligation assay (PLA) revealed co-localization of NOX4 and TRPV4 proteins.
Conclusion
Both TRPV4 and NOX4 contribute to ACh-induced dilation in human arterioles from patients without coronary artery disease. NOX4 increases TRPV4 phosphorylation in endothelial cells, which in turn enhances TRPV4-mediated Ca2+ entry and subsequent endothelium-dependent dilation in human arterioles.
Keywords: NADPH oxidase, transient receptor potential vanilloid, human arterioles, coronary artery disease, vasodilation
Introduction
The coronary microcirculation includes small arteries and arterioles with diameters below 300–400 μm, and plays a critical role in the regulation of vascular resistance and tissue perfusion [8]. Accumulating evidence strongly suggests that coronary microvascular dysfunction (CMD) exists in many patients with coronary heart disease [30, 59], which can account for up to 50–65% of the patients with angina-like symptoms who failed to meet the diagnostic criteria of obstructive coronary artery disease by coronary angiography [42]. Additionally, CMD is associated with an increased risk of major adverse cardiovascular events [51]. Coronary endothelial impairment has been suggested to be a key component of CMD [49, 55]. A recent study has further shown that this impairment is mirrored in the systemic microvasculature, as indicated by reduced relaxation responses to the endothelium-dependent vasodilator acetylcholine (ACh) in adipose arterioles biopsied from patients with microvascular angina [19]. However, the pathogenic mechanisms underlying CMD remain largely unclear. A better understanding of endothelial signaling pathways, especially in human arterioles, may provide important insight into the mechanisms of microvascular dysfunction.
The endothelium regulates vascular tone by releasing various vasoactive factors, including nitric oxide (NO), prostacyclin (PGI2), and endothelium-derived hyperpolarizing (EDH) factors [15] most prominently in resistance arteries and arterioles. A common pathway responsible for the stimulus-induced release of endothelial factors involves an increase in the intracellular Ca2+ concentration ([Ca2+]i) in endothelial cells (EC) [44]. Transient receptor potential (TRP) vanilloid 4 (TRPV4), the fourth member of the TRP vanilloid subfamily, is a non-selective Ca2+-permeable cation channel expressed in vascular ECs of multiple species [18, 70]. It can be activated by diverse chemical and physical stimuli, such as the phorbol derivative 4α-phorbol-12,13-decanoate (4α-PDD) [63], arachidonic acid (AA) and its metabolites, hypotonic cell swelling [35, 57, 65], moderate heat [22], shear stress [20], and the endothelium-dependent vasodilator ACh [71]. Accordingly, TRPV4 has been implicated in a variety of physiological processes and diseases, including endothelium-dependent vasodilation [31, 46]. Previous studies have shown that TRPV4 contributes to ACh-induced dilation in various animal resistance arteries [18, 70], but whether TRPV4 mediates agonist-induced dilation in the human microcirculation remains understudied.
Nicotinamide adenine dinucleotide phosphate oxidase (NADPH oxidases, NOX) is a family of redox enzyme complexes producing the reactive oxygen species (ROS) such as the superoxide anion (O2·-) and hydrogen peroxide (H2O2) in a variety of cell types [13, 58, 75]. It is known that elevated ROS contributes to endothelial dysfunction in various diseases; however, ROS such as H2O2 can also act as physiological signaling molecules in the vascular wall. For example, NOX-derived H2O2 has been shown to regulate ACh-induced dilation in mouse mesenteric arteries [53] and rat intrarenal arteries [41]. Furthermore, Larsen et al. have reported that NOXs are expressed in human coronary endothelial cells and NOX-derived H2O2 is required for bradykinin (BK)-induced dilation in human arterioles [33]. However, the mechanisms by which NOX-derived ROS contribute to agonist-induced dilation are not fully understood. Our recent studies indicate that H2O2 induces robust protein phosphorylation of the TRPV4 channel to regulate its activity in EC [9]. The goal of this study was to determine the role of TRPV4, NOX and their potential interactions in the regulation of endothelium-dependent dilation in human arterioles. We examined the novel hypothesis that NOX-generated H2O2 stimulates phosphorylation and activation of endothelial TRPV4 to evoke Ca2+ entry and subsequent relaxation of the underlying smooth muscle.
Methods
Human tissue acquisition
Fresh human adipose arterioles (HAA) and human coronary arterioles (HCA) with internal diameters of 100–250 μm were isolated from adipose or heart tissues as discarded surgical specimens and unused donor hearts, as we described previously [45]. The study included patients without coronary artery disease (CAD) and with no or ≤ 2 risk factors for CAD, including hypertension, hyperlipidemia, and diabetes. All protocols were approved by the Institutional Review Board of the Medical College of Wisconsin and Froedtert Hospital on the use of human subjects in research. The tissue specimens were placed in ice-cold HEPES buffer (138 mM NaCl, 4 mM KCl, 2 mM MgSO4, 1.6 mM CaCl2, 1.2 mM KH2PO4, 6 mM glucose, and 10 mM HEPES, pH 7.4) and transferred to the lab for vessel isolation. Unless otherwise stated, vasoreactivity and molecular experiments were performed on endothelium-intact arterioles. For endothelium-denuded preparations, the endothelium was removed by slowly injecting 3–5 mL of air through the lumen.
A total of 71 patients who met the inclusion criteria were included in the study. Adipose samples were collected from 49 patients. Heart samples from 22 patients were used for this study. The demographic data of the patients were collected from the Generic Clinical Research Database at the Medical College of Wisconsin and are summarized in Table 1.
Table 1.
Patient demographics
| HAA | HCA | |
|---|---|---|
| Sex | ||
| Male (n) | 16 | 14 |
| Female (n) | 33 | 8 |
| Age (yr) (mean ± SEM) | 48 ± 14 | 58 ± 14 |
| Body Mass Index (mean ± SEM) | 28 ±5 | 26 ± 7 |
| History/Risk Factors | ||
| Hypertension (n) | 8 | 10 |
| Hypolipidemia (n) | 2 | 7 |
| Smoking (n) | 5 | 13 |
| Atrial Fibrillation (n) | 0 | 1 |
| Tumor (n) | 5 | 0 |
Vascular cell dissociation
EC and smooth muscle cells (SMC) were freshly dissociated from arterioles as previously described [38, 69]. In brief, arteriole segments were thoroughly rinsed of luminal blood, cut into small rings, and incubated at room temperature for 10 min in a low-Ca2+ dissociation buffer consisting of (in mM) 145 NaCl, 4.0 KCl, 0.1 CaCl2, 1.0 MgSO4, 10 glucose, and 10 HEPES with 0.1% BSA (pH 7.4). The buffer was carefully removed, followed by sequential incubation with papain (1.0 mg/ml) and DTT (0.5mg/ml) in dissociation buffer for 15 min, and then collagenase (Sigma blend H; 1.0 mg/ml), trypsin inhibitor (1 mg/ml), and elastase (0.5 mg/ml) for 15–30 min at 37 °C. All enzymes and chemicals were purchased from Sigma. Arteriole segments were gently triturated to release EC and SMC. The cells were washed once with the dissociation solution by centrifugation at 250× g for 5 min. EC were separated from the cell suspension by Dynabeads coated with the anti-CD31 antibody (Invitrogen) at 4°C for 1 hour, followed by the beads pull-down by the magnet as previously described [32]. Isolated EC and SMC were stored on ice or at 4 °C and used the same day for the RNA extraction and PCR. For the primary culture of isolated EC, cells were plated on glass-bottom Peri dishes and incubated for 2–3 days in the complete EGM-2MV growth medium (Promo cell) before fura-2 calcium imaging and PLA assay.
Cell culture and transfection
Human coronary artery endothelial cells (HCAEC) purchased from PromoCell (Heidelberg, Germany) were maintained in the complete EGM-2MV growth medium (Promo cell) and incubated at 37°C with 5% CO2. Cells were split at 1:3 to 1:4 ratios when the cells were at 90~95% confluence. HCAEC between passages 6 and 7 were used for the experiments. HEK293 cells between passages 6 to 10, were provided by Dr. David Wilcox (Medical College of Wisconsin) and grown in Dulbecco’s Modified Eagle Medium (DMEM) supplemented with 10% FBS, 1% penicillin G, and streptomycin, and incubated at 37 °C with 5% CO2. HEK293 cells between passages 12 and 15 were used for TRPV4 cell surface and phosphorylation experiments.
The TRPV4-GFP constructs with wildtype and point mutations of TRPV4 protein (S824A, S824E, S823A/S824A) in a pWPTS lentiviral vector were prepared as described previously [9], and the recombinant lentiviruses were produced from HEK293 cells. All constructs were verified by DNA sequencing. For TRPV4 overexpression experiments, HCAEC at passage 6 or 7 were transduced with lentiviruses at an m.o.i (multiplicity of infection) of 5–10. To minimize the potential calcium overload in the cells from the overexpressed TRPV4 channels, the culture medium was replaced with the low calcium culture medium after 16 hours of transduction, with the calcium concentration reduced to ~0.6 mM by the addition of 1.2 mM EDTA. Cells were used 3 to 4 days after transduction for calcium imaging assay, 4–5 days after transduction for immunoblotting experiments.
Vascular reactivity
Freshly isolated arterioles (≈100–200 μm) were cannulated and pressurized under 60 mm Hg in Krebs – physiological salt solution (PSS) that contains (in mM): 123 NaCl, 4.7 KCl, 1.2 MgSO4, 2.5 CaCl2, 19 NaHCO3, 1.2 KH2PO4, 0.026 EDTA, and 11 glucose, gassed with 21% O2 – 5% CO2 at 37 °C. Changes in internal diameter were measured by a videomicroscopy system as we described previously [45]. Vessels were preconstricted with endothelin-1 (0.1–0.3 nM) by ~30% of the baseline internal diameter. Two dose-response curves were performed on each arteriole. For endothelium-dependent vasodilation, dose-dependent relaxation responses to the endothelium-dependent vasodilator acetylcholine (ACh, log 10−9 – 10−5 M), or TRPV4 agonist GSK1016790A (1–100 nM) were determined in the presence or absence of the NOX1/4 inhibitor GKT137831 (1 μM), or TRPV4 inhibitor HC067046 (2 μM) in the vessel bath. To further identify the endothelial-dependent role of NOX4, responses to sodium nitroprusside (log 10−10- 10−4 M), an endothelium-independent vasodilator, were performed with or without GKT137831 (1 μM) in the bath. At the end of each experiment, papaverine (100 μM), an endothelium-independent vasodilator, was added to the bath for maximal vasodilation. The vessel dilation response was calculated using the equations: percent maximal dilation = [((diameter with test compound) – (Basal diameter)) / ((Maximum diameter) – (Basal diameter))] ×100. The percentage of maximal dilation was averaged and plotted from at least 6 independent experiments. The efficacy and specificity of the above pharmacological agonists and inhibitors have been reported in previous studies [4, 16, 26, 61].
RNA extraction and reverse transcription-polymerase chain reaction
Freshly dissected human adipose arterioles (200–400 μm) were snap-frozen in liquid nitrogen and stored at −80 °C until use. Enzymatically dissociated EC and SMC were used the same day. Total RNA was extracted with the RNeasy Plus Universal Mini Kit (Qiagen), and the cDNA was synthesized with SuperScript III reverse transcriptase kit (Invitrogen), according to the manufacturer’s instructions. For regular PCR, 2 ng of cDNA was amplified using Platinum PCR Supermix (Invitrogen) in the thermal cycler with a 40-cycle touch-down protocol. The negative control group was amplified with primers without reverse transcription (RT- group), or without template (H2O group). For quantitative PCR (qPCR), 20 μL reaction volume containing 2–5 ng of cDNA,10 μL SYBR Green qPCR Mix (Qiagen), 1 μL of reverse and forward primer each and nuclease-free water was performed using CFX96 C1000 Thermal Cycler. Cycle threshold (Ct) values were determined by CFX Manager 3.1 software. H2O was used instead of the template for negative control. The thermal profile employed 1 cycle of initial denaturation at 95°C for 5 min, 40 cycles of denaturation at 95°C for 10 sec, annealing at 60°C for 30 sec, and extension at 72°C for 15 sec. Melting curve analysis was performed immediately after the 40th cycle at 72 to 95°C with 0.5 °C increments for 5 sec. NOX1 to 5 primers were synthesized by Integrated DNA Technologies Inc. and sequence information is provided in Table 2. Since all NOX primers are designed to span an exon-intron boundary, this avoids the amplification of the possible small amount of contaminating genomic DNA. 18S rRNA primers as an internal control were obtained from Qiagen (Cat No. QT00199367), and the primer sequences are not publicly available.
Table 2.
Primer sequences
| Target gene | Orientation | Size (bp) | Primer Sequence |
|---|---|---|---|
| NOX1_HUMAN | Forward (22) | 102 | 5’-GGT TTT ACC GCT CCC AGC AGA A-3’ |
| Reverse (22) | 5’-CTT CCA TGC TGA AGC CAC GCT T-3’ | ||
| NOX2_HUMAN | Forward (23) | 132 | 5’-CTC TGA ACT TGG AGA CAG GCA AA-3’ |
| Reverse (22) | 5’-CAC AGC GTG ATG ACA ACT CCA G-3’ | ||
| NOX3_HUMAN | Forward (22) | 160 | 5’-CCT GGA AAC ACG GAT GAG TGA G-3’ |
| Reverse (22) | 5’-CCT CCC ATA GAA GGT CTT CTG C-3’ | ||
| NOX4_HUMAN | Forward (23) | 124 | 5’-CCA AGC AGG AGA ACC AGG AGA TT-3’ |
| Reverse (24) | 5’-AGA AGT TGA GGG CAT TCA CCA GAT-3’ | ||
| NOX5_HUMAN | Forward (22) | 127 | 5’-CCA CCA TTG CTC GCT ATG AGT G-3’ |
| Reverse (22) | 5’-GCC TTG AAG GAC TCA TAC AGC C-3’ |
RNA sequencing
Total RNA was extracted from arterioles with the RNeasy Plus Universal Mini Kit (Qiagen) and libraries prepared at the Genomic Science and Precision Medical Center (GSPMC at the Medical College of Wisconsin) according to the SMART-Seq Stranded Kit (Takara) utilizing 10ng of input. Briefly, RNA was reverse transcribed into cDNA using random priming and Illumina adapters and barcodes added by PCR. Degradation of ribosomal cDNA was completed with scZapR and scR-Probes as part of the kit protocol and final libraries amplified for sequencing. Final assessment, quantification, and pooling of the RNAseq libraries were completed with qPCR (Kapa Library Quantification Kit, Kapa Biosystems) and DNA high sensitivity fragment analysis (Agilent). The GSPMC completed next-generation sequencing on the Illumina NovaSeq600 with paired-end 100 base pair reads generated at >50 million reads per sample. Sequencing reads were aligned to the human Gencode v32 (based on Ensembl v98) and processed through the MAPR-Seq Workflow [29] with differential expression analysis completed using Bioconductor, edgeR v 3.8.6 software. Genes with a false discovery rate (FDR) less than 5% and an absolute fold change ≥ 4 will initially be filtered and considered significantly differentially expressed.
Immunoblotting
Fresh human arterioles were homogenized in ice-cold lysis buffer containing 25 mM Tris-HCL (pH 7.4), 150 mM NaCl, 0.1% SDS, and 1% NP-40, supplemented with protease and phosphatase inhibitors, and centrifuged at 12,000 g for 10 min at 4°C. Protein concentrations of the supernatants were measured by a BCA kit according to the manufacturer’s instructions. Samples were aliquoted and frozen in liquid N2, stored at −80°C before use. Protein samples (20 μg) were separated by SDS-PAGE on 10% Bio-Rad TGX mini gels and transferred to polyvinylidene difluoride membranes. Membranes were blocked by 5% non-fat dry milk (NFDM) or bovine serum albumin (BSA) for 1 hour at room temperature and incubated overnight at 4 °C with a primary monoclonal antibody against NADPH oxidase 4 (kindly given by Dr. Doroshow at the NIH/NCI) (1:500 dilution in TBST) in TBST with 5% NFDM, or a monoclonal antibody against β-actin (1:40,000 dilution in TBST) in TBST with 2% NFDM. Blots were washed with TBST and then incubated with HRP-conjugated secondary antibody (1:20,000 dilution in TBST with 2% NFDM) for 1 hour at room temperature. Membranes were developed using ECL prime reagent and KONICA developer. The films were scanned using an EPSON scanner.
Biotinylation of cell surface proteins
Plasma membrane proteins from HCAEC were isolated by using a cell surface protein biotinylation kit (Pierce Co.). Cells were incubated with PBS containing 0.5 mg/ml sulfosuccinimidyl 2-(biotiamido) ethyl-1.3-dithiopropionate (Sulfo-NHS-SS-Biotin) for 30 min at 4 °C to label the surface protein. The biotinylation reaction was stopped by a quenching solution. After washing the cells with ice-cold Tris-buffered saline (TBS), a protein lysis buffer described above was added to extract total proteins. To isolate biotin-labeled surface proteins, total protein samples (150–200 μg) were gently mixed with 50 μl of NeutrAvidin agarose beads (supplied as 50% slurry) and rotated for 3 hours at 4 °C. The beads were collected after washing and centrifugation. Protein was then eluted from the beads use 100 μl of 1× Laemmli sample buffer with 50 mM DTT by heating the samples for 5 min at 95 °C. Western blot was followed to analyze the eluted protein.
Proximity ligation assay
The protein-protein interactions between TRPV4 and NOX4 in situ (distances < 40 nm) were detected by a proximity ligation assay (PLA) kit (Sigma-Aldrich #DUO92101). Briefly, primary EC isolated from HAA and HCAEC transduced with TRPV4-GFP were plated in 35-mm glass-bottom Petri dishes and grown to 60 – 70% confluence. Cells were fixed with 4% PFA in PBS for 20 min at room temperature (RT), followed by 0.1% Triton-X/0.3M glycine in PBS for 30 min at RT. Cells were blocked with block solution for 1 hour at RT, and two specific primary antibodies from different species that recognize and bind to TRPV4 (or TRPV4-GFP), and NOX4 proteins, including mouse anti-TRPV4 (Sigma MABS466) or mouse anti-GFP, and rabbit anti-NOX4 (47–6), were added and incubated overnight at 4°C. After 3 rinses in PBS, cells were incubated with anti-rabbit PLUS and anti-mouse MINUS PLA probes, which match the host of the primary antibodies, for 20 min at RT, followed by PCR amplification of fluorescent probes, according to the manufacturer’s instructions. Cells were then stained with DAPI and fluorescent images were captured using a fluorescence microscope (model TE200, Nikon) with a 20× objective. The cytoplasm size was estimated according to cellular fluorescence and DAPI-stained nuclei. PLA signals (red puncta) in the cell were counted and analyzed by Image J software.
Fura-2 calcium imaging
HECEC were plated and transduced with TRPV4-GFP wildtype or mutant in 35-mm glass-bottom Petri dishes and grown to 60 – 70% confluence. Cells were incubated with fura-2 AM (5 μM) (Molecular Probes) and 0.02% Pluronic F-127 at 37 °C for 30 min in the modified Hank’s balanced salt solution (HBSS) that contained (in mM): 123 NaCl, 5.4 KCl, 1.6 CaCl2, 0.5 MgCl2, 0.4 MgSO4, 4.2 NaHCO3, 0.3 Na2HPO4, 0.4 KH2PO4, 5.5 glucose, and 20 HEPES (pH 7.4 with NaOH). Fura-2 assay was used to monitor cytosolic Ca2+ signals as described previously [12, 36]. The emitted fura-2 fluorescence at 510 nm in cells that are alternately exposed to 340 and 380 nm excitation wavelength was recorded and analyzed by the MetaFluor software (Molecular Devices). Background fluorescence was subtracted before the study. The images of cell fluorescence were acquired every 3 seconds during the experiment. Changes in the intracellular Ca2+ concentration were presented as the ratio of the fluorescence intensity at 340 nm versus 380 nm excitation (F340/F380). The F340/380 ratio was averaged and plotted from at least 6 independent experiments. For the single experiment, 20 −40 cells were selected with the MetaFluor software based on the basal [Ca2+]i. The cells with high basal [Ca2+]i (F340/F380 ratio > 2.0) were excluded to minimize potential variations in response to TRPV4 agonists.
Data and statistical analysis
All data are presented as means ± SEM. Molecular experiments were repeated at least three times unless otherwise stated. For vascular reactivity, data from the independent experiments were averaged, and significant differences in vasodilatory responses were evaluated by two-way (factor) repeated measures ANOVA (factor 1, doses; factor 2, different treatments such as a vehicle or an inhibitor), followed by the Holm-Sidak test for pairwise comparisons or comparisons versus control. For other studies, statistical comparisons were made by Student’s t-test or one-way ANOVA. All statistical analysis was performed using SigmaPlot (version 12.0). Significance in figures was depicted as *P < 0.05; ** P < 0.01.
Reagents
GSK1016790A, GSK2795039, SNP, and Endothelin-1 (ET-1) were obtained from Sigma-Aldrich. GKT1372831 was obtained from MCE MedChemExpress. HC067047 was obtained from Tocris Bioscience. All other chemicals were purchased from Sigma. Stock solutions were prepared in DMSO (1000× or higher). ET-1 stock was prepared in saline solution supplemented with 1% BSA. Working stocks were freshly prepared in distilled water.
Results
NOX4 mRNA and protein expression in human adipose and coronary arterioles
To characterize the NOX isoforms expressed in human arterioles and isolated vascular cells, we first examined their mRNA expression. RT-PCR analysis revealed that mRNA transcripts of both NOX2 (amplicon size: 132 bp) and NOX4 (amplicon size: 124 bp) are most abundantly expressed in HAA and HCA, as well as in the freshly isolated endothelial cells (EC) and smooth muscle cells (SMC; Fig. 1a). The purity of isolated cells has been confirmed with the endothelial cell marker, platelet endothelial cell adhesion molecule-1 (PECAM-1), and smooth muscle cell marker, transgelin (SM22), as we reported recently [32]. Using quantitative PCR (qPCR), NOX4 mRNA expression tends to be lower in endothelium-denuded HAA compared with endothelium-intact HAA (Fig. 1b). Although both NOX2 and NOX4 have been detected in human arterioles by RT-PCR, NOX4 is thought to be constitutively active and produce H2O2 rather than O2·- [17, 58]; thus, we focused on the function of NOX4 in the subsequent studies. Consistent with RT-PCR results, NOX4 protein (~55 kD) was detected in both HAA (Fig. 1c) and HCA (Fig 1d). The NOX4 protein band is slightly smaller than the expected size of 67 kD for the canonical isoform 1 of NOX4 (Uniprot ID Q9NPH5–1), which is not further explored in the present study but may represent some splicing variants of NOX4, such as isoform 2 (58 kD; Uniprot ID Q9NPH5–2). As controls, NOX4 was detected in NOX4 overexpressing but not in non-transfected HEK293 cells, confirming the specificity of the anti-NOX4 antibody [37] (Fig. 1c). These results demonstrate that NOX4 is expressed at both mRNA and protein levels in the endothelial cells of human arterioles.
Fig 1. NOX4 mRNA and protein expression in human adipose and coronary arterioles and isolated endothelial and smooth muscle cells.

a) RT-PCR analysis shows that NOX4, in addition to NOX2, is expressed in HAA and HCA (n=3 patients for each), as well as in endothelial cells (EC) and smooth muscle cells (SMC) freshly isolated from HAA (n=3 patients). b) NOX4 mRNA was further analyzed with qPCR in both endothelium-intact and denuded HAA (n=3), 18s was used as housekeeping gene (n=3). c) Representative gel images of NOX4 protein expression in HAA from 2 patients, HEK293 non-transfected cells (293 NT) as a negative control, and NOX4-overexpressed HEK293 cells (293 NOX4 O/E) as a positive control. β-actin was included as an internal reference and loading control. d) Representative gel images of NOX4 protein expression in endothelium-intact and -denuded HCA from 2 patients. Compared with endothelium-intact HCA, NOX4 protein expression is reduced in endothelium-denuded HAA; β-actin was included as an internal reference and loading control.
TRPV4 and NOX4 in acetylcholine-induced dilation in human adipose arterioles
To understand the vasomotor function of TRPV4 and NOX4 in the human microvasculature, we examined their roles in acetylcholine (ACh)-induced vasodilation in freshly isolated human adipose arterioles (HAA). In endothelin-1-preconstricted adipose arterioles, acetylcholine (log 10−9 to 10−5 M), an endothelium-dependent vasodilator, induced potent dilation in a concentration-dependent manner (Fig. 2). This dilation was significantly attenuated by preincubation of arterioles with the TRPV4-selective inhibitor HC067047 (2 μM; Fig. 2a). ACh-induced dilation was similarly inhibited by the NOX1/4 inhibitor GKT137831 (1 μM; Fig. 2b). In the presence of HC067047, however, GKT137831 did not significantly reduce ACh-induced vasodilation (Fig. 2c). Together, these results indicate that both TRPV4 and NOX4 contribute to ACh-induced dilation in human arterioles, likely acting in series through a similar signaling pathway. To further confirm the role of endogenous H2O2 in ACh-induced dilation, adipose arterioles were pretreated with peg-catalase (500 U/ml), a cell-permeable H2O2 scavenger. Similar to NOX1/4 inhibitor GKT137831, peg-catalase markedly inhibited ACh-induced dilation at log −7 M of ACh (Fig S1a). In the presence of peg-catalase, NOX1/4 inhibitor GKT137831, as well as TRPV4 inhibitor HC067047, did not significantly further reduce the dilation (Fig S1b–c) as compared to peg-catalase pretreated arterioles. These results provide additional support that NOX4, H2O2, and TRPV4 likely act through a common signaling pathway in the regulation of human arteriolar relaxation.
Fig 2. Role of TRPV4 and NOX4 in acetylcholine (ACh)-induced dilation in human adipose arterioles (HAA).

In arterioles preconstricted with endothelin-1 (ET-1), the TRPV4-selective blocker HC067047 (2 μM) significantly reduced ACh-induced vasodilation at the concentrations of log 10−7 – 10−5 M. b) The NOX1/4 inhibitor GKT137831 (1 μM) reduced ACh-induced dilation at the concentrations of log 10−8 – 10−7 M. c) In the presence of HC067047, GKT137831 did not further reduce ACh -induced dilation. Data a to c, * P < 0.05 vs control or HC067047 (n=5 vessels for each group).
Consistent with previous studies [6], ACh-induced dilation in adipose arterioles was reduced but not abolished by NG-nitro-L-arginine methyl ester (L-NAME, 100 μM), an inhibitor of NO synthase, combined with indomethacin (Indo, 10 μM), an inhibitor of cyclooxygenase, confirming that NO and prostacyclin (PGI2) partially mediate ACh-induced dilation (Fig. S1d). Interestingly, in the presence of L-NAME and indomethacin, GKT137831 further inhibited ACh-induced dilation (Fig. S1e). These results suggest that NOX4 is mainly involved in non-NO and non-PGI2-mediated dilation.
NOX4 in TRPV4 agonist-induced dilation in human adipose and coronary arterioles
NOX4 has been recently shown to contribute to endothelial-dependent dilation in intrarenal arterioles [41]. To further investigate whether NOX4 participates in TRPV4-mediated vasodilation in human adipose and coronary arterioles, TRPV4 agonist GSK1016790A (GSK) was used to dilate the arterioles. As a synthetic agonist of TRPV4, GSK activates TRPV4 by directly binding to its transmembrane domains in a ligand-like manner [62], which in turn increases the intracellular Ca2+ concentration of endothelial cells, leading to endothelial-dependent vasodilation. As shown in Fig 3a–b, GSK (1 – 100 nM) dilated the arterioles in a dose-dependent manner. In HAA, after preincubation with NOX1/4 inhibitor GKT137831 (1 μM), the percentage of maximum dilation of GSK was significantly reduced (Fig 3a). In human coronary arterioles (HCA), GSK-induced vasodilation was also reduced after preincubation with GKT137831 (Fig. 3b). GKT137831 did not affect the response to the direct smooth muscle vasorelaxing agent, sodium nitroprusside (SNP), indicating an endothelium-dependent effect of GKT137831 (Fig 3c). In contrast to ACh, preincubation with L-NAME (100 μM) and indomethacin (10 μM) eliminated GSK-induced dilation, indicating that NO and PGI2 are the main mediators for GSK-induced vasodilation (Fig S2a). The specificity of GSK1016790A was confirmed by TRPV4 selective antagonist HC0670147 (2 μM; Fig S2b). Together, these results indicate that NOX4 is involved in TRPV4-mediated dilation in human adipose and coronary arterioles.
Fig 3. Role of NOX4 in the TRPV4 agonist GSK1016790A-induced dilation in human adipose and coronary arterioles.

a) In HAA, preincubation with GKT137831 (1 μM) significantly reduced GSK1016790A-induced dilation at 10 to 100 nM. b) In human coronary arterioles (HCA), preincubation with GKT137831 also reduced GSK1016790A-induced dilation at 10 to 100 nM. c) For the endothelium-independent vasodilator sodium nitroprusside (SNP)-induced dilation, there is no significant difference between control and GKT137831-treated vessels. * P < 0.05 vs control (Data in a and b, n=6 vessels for each group; data in c, n=4 vessels).
NOX4 regulates GSK-induced Ca2+ influx in primary endothelial cells of human adipose arterioles
With the finding that GKT137831 largely inhibited TRPV4 agonist-induced dilation, it is reasonable to consider that NOX4 may act downstream of TRPV4 activation; however, NOX4 regulation of TRPV4 activation cannot be excluded and may also play an essential role. Since the vasomotor function of TRPV4 is mainly due to Ca2+ influx through the channel, we then examined whether NOX4 regulates TRPV4-mediated vasodilation by affecting TRPV4-mediated Ca2+ influx. Endothelial cells were freshly isolated from HAA and briefly cultured for 2–3 days before Ca2+ assay. The intracellular Ca2+ concentrations were monitored by the ratiometric calcium indicator fura-2 and the results are presented as F340/F380 ratios, with the corresponding fluorescence pseudo-colored from blue to red on a scale of 0.1 to 1. As shown in Fig 4a, the red fluorescence increased after additions of TRPV4 agonist 3 to 10 nM GSK, and with the addition of TRPV4-selective antagonist HC067047 (2 μM), the red fluorescence decreased rapidly. Representative traces in Fig 4b show that GSK induced Ca2+ increase in a concentration-dependent manner and this increase was largely reversed by HC067047. After preincubation with NOX1/4 inhibitor GKT137831 (1 μM), GSK-induced Ca2+ influx was reduced, and the remaining response was further inhibited by HC067047 (Fig 4c). In summarized data, Ca2+ influx induced by 3 and 10 nM GSK was significantly reduced by GKT137831 (Fig 4d), indicating that NOX4 regulates Ca2+ influx through the TRPV4 channel.
Fig 4. GSK1016790A-induced Ca2+ responses in the primary culture of endothelial cells isolated from human adipose arterioles.

a) Representative fura-2 fluorescence images in control cells (baseline), and after sequential additions of TRPV4 agonist GSK1016790A (GSK) 1 nM, GSK 3 nM, GSK 10 nM, and HC067047 (HC) 2 μM, respectively. The F340/F380 ratio is set on a scale of 0.1 to 1, with the corresponding color changes from blue to red. b-c) Representative traces. GSK1016790A induced Ca2+ increase in a concentration-dependent manner, and this increase was largely reversed by TRPV4-selective antagonist HC067047 (2 μM). After preincubation with NOX1/4 inhibitor GKT137831 (GKT137, 1 μM), GSK-induced Ca2+ response was reduced at GSK 3 and 10 nM, and the remaining response was further inhibited by HC067047. d) Summary data of concentration-dependent Ca2+ response induced by GSK1016790A in the absence and presence of GKT137831. GSK-induced response is reversible by HC under both conditions. * P < 0.05 vs control (n=6 independent experiments in each group).
NOX4 regulates GSK-induced Ca2+ influx and Ser824 phosphorylation in TRPV4-overexpressing human coronary artery endothelial cells
To further investigate the role and underlying mechanisms of NOX4 regulation, wild-type (WT) TRPV4- overexpressing human coronary artery endothelial cell (HCAEC) was used as in our previous studies [27, 73]. In control TRPV4-WT overexpressing HCAEC, GSK- induced Ca2+ influx in a concentration-dependent manner (Fig 5a). Consistent with the results obtained from the primary EC (Fig 4), preincubation with NOX1/4 inhibitor GKT137831 (1 μM) significantly inhibited GSK-induced Ca2+ response (Fig 5b–c). Given that TRPV4 as a channel protein requires expression on the plasma membrane, we confirmed the cell surface expression of transfected WT and the mutants by biotinylation cell surface assay (Supplemental Fig S3).
Fig 5. Effect of NOX1/4 inhibitor GKT137831 on GSK1016790A-induced Ca2+ influx in TRPV4 wild-type (WT)-overexpressing human coronary artery endothelial cells (HCAEC).

a) In control HCAEC overexpressed with TRPV4-WT, GSK1016790A induced Ca2+ influx in a concentration-dependent manner, the arrows indicate the administration of GSK1016790A (GSK). The intracellular Ca2+ concentrations were monitored by the ratiometric calcium indicator fura-2 and the results are presented as F340/F380 ratios. b) In WT TRPV4-overexpressing HCAEC, GSK1016790A-induced Ca2+ response at 1 nM and 3 nM was inhibited by the preincubation with GKT137831 (1 μM). c) Summary data of concentration-dependent Ca2+ responses induced by GSK1016790A in the absence and presence of GKT137831. * P < 0.05 vs control (n=6 independent experiments in each group).
Our previous work had shown that TRPV4 is regulated by Ser824 phosphorylation through the H2O2-PKA pathway [9]. NOX4 as an endogenous source of H2O2 may regulate the phosphorylation of the TRPV4 channel at Ser824. Therefore, the non-phosphorylatable TRPV4 mutant (S823A/S824A) was used to further investigate whether Ser824 phosphorylation is involved. As shown in Fig 6, GSK activated TRPV4 S823A/S824A mutant in control, but this activation was no longer affected by preincubation with NOX1/4 inhibitor GKT137831 (1 μM). To directly examine S824 phosphorylation, we performed immunoblotting using a phosphoserine motif antibody against the motif RXRXXS*/T* as previously reported [9]. As shown in Figs. 6d–e, GKT137831 significantly reduced TRPV4 phosphorylation at Ser824. The specific detection of Ser824 phosphorylation was verified by the non-phosphorylatable mutant-expressing cells in the absence and presence of PKC activator PMA (1 μM) stimulation (Supplemental Fig S4). The above results suggest that S824 phosphorylation is required for TRPV4 regulation by NOX4.
Fig 6. Role of Ser824 phosphorylation in the regulation of TRPV4-mediated Ca2+ influx by NOX4.

a) In control HCAEC overexpressing TRPV4-S823A/824A (3A/4A) mutant, GSK10167890A induced Ca2+ influx in a dose-dependent manner. b) After the preincubation with GKT137831, a similar Ca2+ response to GSK1016790A was observed in TRPV4-S823A/824A mutant-overexpressing HCAEC. c) Summary data of concentration-dependent Ca2+ responses induced by GSK1016790A. There is no significant difference in GSK1016790A responses between control and GKT137831-treated cells. d) Representative immunoblots of phospho-Ser824 (p-S824) of TRPV4 in the vehicle, NOX2 inhibitor GSK2795039 (GSK279, 1 μM)- and NOX1/4 inhibitor GKT137831 (GKT137, 1 μM)-pretreated TRPV4-WT overexpressing HCAEC. TRPV4 Ser824 phosphorylation was analyzed with a phosphoserine motif antibody against the motif RXRXXS*/T* (p-S824 antibodies), and the same blot was re-probed with GFP antibodies to detect total cellular TRPV4 proteins. E) Quantitative analysis of immunoblots. * P < 0.05 vs control (n=3 independent experiments in each group).
The protein-protein interactions of TRPV4 and NOX4
Based on the above findings that NOX4 and TRPV4 contribute to a similar pathway of the endothelial agonist ACh-induced dilation and NOX4 regulates TRPV4 channel phosphorylation and Ca2+ influx, we next examined whether these two proteins are colocalized in endothelial cells. Duolink proximity ligation assay (PLA assay) was performed using both primary EC isolated from HAA and TRPV4-GFP-overexpressing HCAEC.
Mouse- anti-TRPV4 (or mouse-anti-GFP) and rabbit-anti-NOX4 antibodies were used as primary antibodies to recognize and bind TRPV4 (or TRPV4-GFP) and NOX4 protein, respectively. After ligation and amplification, the PLA signal was detected in primary cultured HAA-EC (Fig 7a) and TRPV4-GFP- overexpressing HCAEC (Fig 7b) as red puncta using fluorescence microscopy, with the nuclei signal of DAPI shown in blue. No PLA signal was detected in control cells treated with no primary, but only secondary antibodies. The mean red puncta per cell were counted and analyzed. Compared with control, the PLA signal was significantly higher in cells incubated with TRPV4 and NOX4 antibodies, for both primary HAA-EC and TRPV4-GFP-overexpressing HCAEC, indicating that NOX4 and TRPV4 are closely localized in endothelial cells.
Fig 7. Detection of protein-protein interactions between TRPV4 and NOX4 in primary cultured EC isolated from HAA and TRPV4-GFP overexpressed HCAEC by proximity ligation assay (PLA).

PLA puncta are present in both a) primary cultured EC isolated from HAA and b) TRPV4-GFP overexpressed HCAEC incubated with anti-TRPV4 and anti-NOX4 antibodies, and with anti-GFP and anti-NOX4 antibodies, respectively. No PLA signal was detected in the control group without the addition of two primary antibodies. Anti-CD31 beads used to isolate primary HAA-EC show low-level red fluorescence. In panels a-b, images from left to right indicate PLA probe fluorescence in cells (red), cell nuclei stained by DAPI (blue), and overlays of two images. The mean PLA signal per cell was counted and quantified by Image J software, and summarized as bar graphs on the right of each panel. * P < 0.05 vs no antibody control (Neg Ctrl) (n=3 independent experiments in each group)
Discussion
To our knowledge, this represents the first study to characterize the role of TRPV4 and its interaction with NOX4 in endothelium-dependent vasodilation in human arterioles. The major findings of this study are as follows: first, both TRPV4 and NOX4 contribute to ACh-induced dilation in adipose arterioles from subjects without CAD or significant risk factors, and both act through a similar vasodilation pathway. Second, as one of the major NOX isoforms expressed in arteriolar EC, NOX4 colocalizes with TRPV4 in arteriolar EC and regulates TRPV4 agonist-induced dilation in adipose and coronary arterioles. Third, NOX4 increases Ser824 phosphorylation of TRPV4, which in turn facilitates Ca2+ influx through TRPV4 to regulate TRPV4-mediated vasodilation. Together, these results demonstrate a novel interaction between TRPV4 and NOX4 in regulating endothelium-dependent vasodilation in human arterioles. A schematic diagram of the proposed mechanisms is shown in Figure 8.
Fig 8. Schematic diagram illustrating the proposed mechanism by which NOX4 regulates TRPV4-mediated vasodilation in human arterioles.

In endothelial cells of human arterioles, NOX4 and TRPV4 proteins are functionally expressed and closely localized on or near the cell membrane. Through this close localization, NOX4 enhances TRPV4 agonist GSK1016790A (GSK)-induced channel activation by increasing protein phosphorylation of TRPV4 at Ser824, which in turn facilitates Ca2+ influx through TRPV4 channels to induce NO-mediated vasodilation. The protein-protein interaction of NOX and TRPV4 also contributes to the physiological agonist acetylcholine (ACh)-induced vasodilation in human arterioles, which involves NO and prostaglandin I2 (PGI2), as well as a non-NO, non-prostanoid component of dilation.
Role of TRPV4 and NOX4 in human arteriolar vasodilation
As one of the most extensively studied TRP channels, TRPV4 has been implicated in receptor agonist- and flow-induced endothelium-dependent vasodilation in various animal vascular beds [2, 31, 38, 46, 48, 52, 71]. Our previous studies further indicate that TRPV4 mediates flow-induced dilation in coronary arterioles from subjects with CAD [5]. However, TRPV4 does not seem to contribute to flow-mediated dilation in subjects without CAD (unpublished observations), raising an intriguing question about the role of TRPV4 under normal conditions. In this study, we used adipose arterioles freshly isolated from non-CAD subjects with no or ≤2 risk factors, and vasodilator responses to receptor agonist ACh were examined in cannulated preparations to mimic in vivo physiological conditions. After treatment with TRPV4-selective inhibitor HC067047, the maximum dilation to ACh was reduced, indicating that TRPV4 contributes to ACh-induced dilation in non-CAD arterioles. This proposed role is further supported by the findings that the direct TRPV4 agonist GSK1016790A induced significant Ca2+ entry in arteriolar EC and concentration-dependent dilation in adipose and coronary arterioles. These results are in line with those of earlier studies from our and other groups that TRPV4 contributes to ACh-induced dilation in normal mouse arteries [48, 71]. In contrast, Heathcote et al. have recently reported that, although TRPV4 channels modulate vascular tone by increasing endothelial Ca2+, ACh-induced vasodilation in rat mesenteric arteries was preserved after TRPV4 inhibition by HC067047 [23]. The reasons for this discrepancy remain unclear but underscore potential differential regulation of TRPV4 channels or compensation for loss of TRPV4 channels in different species and vascular beds. Similarly, whether these mechanisms contribute to the contrasting roles of TRPV4 in ACh- vs. flow-induced vasodilation of non-CAD arterioles remains to be determined.
The NOX family is an endogenous source of ROS such as H2O2 in the vasculature [34]. NOXs contain 7 isoforms, including NOX1–5, DUOX1, and DUOX2. Munoz et al. reported that NOX2 and NOX4, as endothelial sources of H2O2, contribute to endothelium-dependent vasodilation in intrarenal arteries [40, 41]. The studies from Zinkevich et al. and Larsen et al. have also shown that inhibiting NOX2 reduces flow-induced and bradykinin-induced vasodilation in HCA from subjects with CAD [33, 74]. In this study, we examined the novel role of NOX in non-CAD human arterioles. By characterizing the mRNA and protein expression of NOX isoforms in non-CAD arterioles and freshly isolated vascular cells, we found that NOX4, which constitutively produces H2O2 [47, 58], is abundantly expressed in adipose and coronary arterioles and arteriolar EC. Furthermore, inhibition of NOX4 reduced ACh-induced but not SNP-induced vasodilation in non-CAD arterioles, providing the first evidence that NOX4, like TRPV4, contributes to endothelium-dependent vasodilation under normal conditions. For ACh-induced dilation, while inhibition of TRPV4 and NOX4 each reduced the dilation, inhibition of both TRPV4 and NOX4 did not further reduce the response, indicating TRPV4 and NOX4 act through a similar vasodilatory pathway. Consistent with our previous results that H2O2 regulates TRPV4 function in cultured EC [9], we also found that NOX4 inhibition largely abolished TRPV4 agonist-induced dilation in HCA and HAA. Together, these results demonstrate a novel role of NOX4 and its interaction with TRPV4 in the regulation of endothelium-mediated dilation in non-CAD arterioles. Given the findings that NOX isoforms have similar expression in HAA and HCA, and NOX4 regulates TRPV4-mediated dilation in both vascular beds, HAA may serve as a surrogate of coronary function. This further supports the concept that the regulation mechanism in human arterioles is often conserved across the arteriolar beds [3].
Since NOX4 mainly produces H2O2, we hypothesize that the regulation of ACh- and TRPV4 agonist-induced dilation by NOX4 is at least partially mediated through NOX4-derived H2O2. Such NOX4-H2O2 pathway has been reported by others in that H2O2 production in NOX4 siRNA-treated cells was significantly reduced as compared with the control group [66, 67]. Besides, Munoz et al. had demonstrated that NOX4 is involved in endothelium-dependent relaxations and ROS production in renal arterioles [41]. In this study, similar to NOX4 inhibition, H2O2 scavenger peg-catalase inhibited the ACh-induced dilation in HAA, and in the presence of peg-catalase, ACh-induced dilation was not further reduced by NOX4 inhibitor GKT137831, as well as by TRPV4 inhibitor HC067047, as compared with peg-catalase alone (Supplemental Fig S1a–c), which further supported that a shared signaling pathway involved in the NOX4, H2O2, and TRPV4 in regulation of human arteriolar relaxation.
NOX1 has also been reported as an important source of O2·- in endothelial and smooth muscle cells of some vascular beds[48, 60]. Recent studies have shown that NOX1 deletion reduces the oxidant stress and restores microvascular health in a mouse model of diet-induced obesity [60]. In the present study, NOX1 mRNA is below the detection limit in non-CAD human arterioles, which is consistent with RNA sequencing results of the NOX family in human coronary arteries according to the GTEx database, as well as with our pilot sequencing in human adipose arterioles (Fig. S5). Thus, NOX1 may not play a significant role in the observed effects of NOX1/4 inhibitor GKT137831 on ACh- and TRPV4 agonist-induced dilation in human arterioles. However, it will be interesting to determine whether the expression of NOX1 is upregulated in human arterioles under pathological conditions such as CAD.
Since TRPV4 is a Ca2+-permeable cation channel, the mechanism of TRPV4 activation by receptor agonists such as ACh likely involves receptor-operated Ca2+ (ROC) entry following the binding of the agonist to the membrane receptor in EC [21, 28, 43, 44]. To investigate the endothelial factors released in responses to ACh and TRPV4 agonist GSK1016790A, arterioles were pretreated with NO synthase and cyclooxygenase inhibitors, L-NAME, and indomethacin, respectively. ACh-induced dilation was partially inhibited by L-NAME and indomethacin and the remaining dilation could be further inhibited by the NOX4 inhibitor GKT137831, indicating that NOX4 may contribute to vasodilation independent of NO and PGI2. In contrast to ACh, TRPV4 agonist-induced dilation was largely abolished by L-NAME and indomethacin, as well as by GKT137831. This indicates that NOX4 can also contribute to NO- and PGI2-mediated dilation depending on the endothelial stimuli employed. How this differential coupling to vasodilator factors occurs remains to be explored in future studies.
NOX4 regulation of TRPV4 function in EC
There is abundant evidence that NOX-derived H2O2 acts as a diffusible EDH factor to relax smooth muscle cells in human arterioles, especially in disease conditions such as CAD [33, 74]. However, in some vascular beds under normal conditions, exogenous application of H2O2 can also enhance EDH-type relaxation by elevating Ca2+ release from intracellular stores in EC [11, 14]. Given the finding that NOX4 inhibitors markedly reduced the TRPV4-mediated vasodilation in non-CAD arterioles, NOX4 may contribute to vasodilation by modulating TRPV4-mediated Ca2+ entry. To directly test this hypothesis, we first examined TRPV4-mediated Ca2+ responses in the primary culture of endothelial cells freshly isolated from human adipose arterioles, and then further verified the results in TRPV4-overexpressing HCAEC. In both endothelial systems, inhibition of NOX4 significantly reduced GSK1016790A (3 nM)-induced Ca2+ entry. Together with the findings of vasodilation studies, these results suggest that NOX4 can regulate endothelium-mediated dilation by enhancing TRPV4-mediated Ca2+ signaling.
Protein phosphorylation is one of the endogenous mechanisms that regulate TRPV4 channels. Fan et al. have reported that TRPV4 channel activation is enhanced by phosphorylation via different serine/threonine protein kinases [17]. Located at the C-terminal domain of TRPV4 channels, Ser824 residue is an important protein kinase A (PKA)-dependent phosphorylation site responsible for TRPV4 sensitization [17]. Our previous studies indicate that H2O2 at pathophysiological concentrations induces robust TRPV4 phosphorylation at Ser824 and enhances TRPV4-mediated Ca2+ entry, and both effects can be inhibited by PKA inhibitors. In addition, physiological lipid mediator arachidonic acid (AA)-induced TRPV4 activation is strongly regulated by Ser824 phosphorylation through the H2O2-PKA signaling pathway [9]. In this study, we examined Ca2+ response to the increasing dose of GSK1016790A in Ser823A/Ser824A double-mutant, non-phosphorylatable TRPV4-expressing HCAEC. Unlike the results seen in WT-TRPV4-expressing HCAEC discussed above, NOX4 inhibitors did not significantly affect Ca2+ responses to GSK1016790A in Ser823A/Ser824A-TRPV4-expressing HCAEC. Furthermore, in WT-TRPV4-overexpressing HCAEC, PKI, an inhibitor of PKA, shifted the Ca2+ dose-response curve induced by GSK1019790A (Supplemental Fig S6) to the right. Thus, PKA-mediated protein phosphorylation of TRPV4 at Ser823/Ser824 is likely required for the regulation of TRPV4-mediated Ca2+ entry by NOX4.
It is surprising that the TRPV4 channel response to synthetic agonists such as GSK101 was affected by protein phosphorylation. However, Cenac et al. have reported that PKC activators increased the phosphorylation of TRPV4 and its responsiveness to TRPV4 agonists 4αPDD [10]. Mohapatra et al. have also demonstrated that TRPV1 phosphorylation by PKA regulates the channel sensitivity, and consistent with our GSK101 data, PKA-mediated protein phosphorylation increased the sensitivity of TRPV1 channels to capsaicin but did not affect the maximal response to this agonist [39]. It is of note that TRPV4-mutant (S823A/S824A)-expressing HCAEC showed more Ca2+ influx than that in TRPV4-WT-overexpressing HCAECs (Fig 6), which may be due to mutation-related differences in the phosphorylation of other sites that regulate TRPV4 channel response to the agonists [17].
To further corroborate the results of Ca2+ experiments, Ser824 phosphorylation was directly measured by Western blotting with p-Ser824 antibodies [9]. Consistent with previous studies by Fan et al. [17], we found that TRPV4 channels exhibit basal S824 phosphorylation. Interestingly, as compared to the no-treatment control, pretreatment of NOX4 inhibitor GKT137831 significantly reduced the basal phosphorylation of TRPV4 at Ser824 in WT-TRPV4 expressing HCAEC. Since EC were not stimulated by PKA or PKC activators, these results indicate that TRPV4 channels exhibit a basal level of phosphorylation, and this basal phosphorylation is likely maintained by NOX4-dependent processes so that NOX4 inhibitors can down-regulate this basal S824 phosphorylation to regulate TRPV4 activity. This notion is supported by the results of the PLA assay indicating that TRPV4 and NOX4 may colocalize with the opportunity for functional interactions on or near the plasma membrane of EC. In the present study, a siRNA approach was also used to understand the role of NOX4 in S824 phosphorylation (Supplemental Fig S7). Overall, results obtained with NOX siRNA further support the role of NOX4 (and possibly NOX2) in the regulation of TRPV4 function, although some discrepancies in the relative effectiveness of NOX4/2 siRNA and inhibitors were also noted. The underlying causes of this discrepancy remain to be determined but may be related to an upregulation of other NOX or ROS enzymes by NOX4/2 siRNA [50] or a disruption of protein-protein interactions by siRNA [64]. For example, given that TRPV4 and NOX4 proteins are functionally closely localized, using NOX4 siRNA to knock down NOX4 protein expression may disrupt the proposed interaction between NOX4 and TRPV4, which in turn may reduce the inhibitory effects of NOX4 siRNA on S824 phosphorylation. In contrast, GKT137831 as a small-molecule inhibitor blocks the enzyme activity of NOX4, while the scaffold for TRPV4 and NOX4 interaction is likely preserved.
Study limitations
We could not control all patient characteristics since tissues used in the study were obtained as surgical discards. The demographic information, like medication history, presence of non-cardiovascular diseases, results of clinical tests of the patients, or other components of the medical history is limited. Our non-CAD patient cohort includes subjects without clinically diagnosed CAD and has no or ≤2 risk factors for CAD, and thus not all may be considered as true healthy subjects. For fura-2 Ca2+ assay, we used the primary culture of EC isolated from HAA as a surrogate for native EC. We have tried to measure TRPV4-mediated Ca2+ responses in EC in situ of freshly isolated human arterioles, but unlike mouse vessels that we have reported previously [71], responses of EC in situ of HCA and HAA were small and it is thus difficult to compare control vs treatments. Further study is needed to answer whether this difference in TRPV4-mediated Ca2+ responses is intrinsic to human arterioles or due to other technical variables.
In the present study, NOX1/4 inhibitor GKT137831 partially inhibited TRPV4-mediated Ca2+ response but largely blocked vasodilation, indicating that other underlying mechanisms could be involved. For example, besides TRPV4 phosphorylation examined in this study, there is evidence that during angiogenesis, NOX4 activates eNOS by eNOS phosphorylation [12]. In addition, NOX-derived H2O2 may act as a diffusible endothelial factor to relax smooth muscle, especially during CAD or other disease conditions [7, 33]. Since the same concentration of GKT137 (1 μM) was used in both the relaxation (HAA) and Ca2+ imaging (HAA-EC, HCAEC) studies, we expect that GKT137831 would be as effective, if not more effective, in a simpler EC system as compared to intact HAA. Although the NOX1/4 inhibitor GKT137831 (Setanaxib) has been tested extensively in the literature and has also entered into clinical trials [54], current vasorelaxation studies obtained with this inhibitor could be strengthened by a complementary siRNA approach.
Using PLA assay on both primary cultured EC and cells overexpressing TRPV4-GFP, our results indicate that TRPV4 is localized in close proximity to NOX4 in endothelial cells. However, further studies are required to determine how TRPV4 and NOX4 (or NOX2) proteins functionally interact with each other in endothelial cells. We have employed a co-immunoprecipitation technique to further test the protein-protein interaction between TRPV4 and NOX4 in HCAEC, HAA, and HCA, but the results were inconclusive (data not shown). It is possible that the expression level of TRPV4 or NOX4 proteins in the arterioles is relatively low and the interaction of these proteins may be transient or disrupted during sample preparation. Alternatively, NOX4 could be adjacent to but not necessarily physically interact with TRPV4 to regulate TRPV4 channel function in EC.
Perspective on clinical applications
There is increasing consensus that CMD plays a pivotal role in myocardial ischemia and predicts outcomes in patients with coronary heart disease, even in the absence of obstructive CAD. Coronary microvascular impairment as a role of adjunct cardioprotection target has been gradually explored in recent years [24, 72], however, due to the limited knowledge of the underlying mechanisms of microvascular dysfunction, the clinical efficacy of such therapy remains suboptimal. Although animal models of CMD recapitulate certain characteristics of microvascular dysfunction seen in patients, a perfect translation of experimental findings obtained from animal models to the clinical setting remains an important challenge in the field [56]. Therefore, a better understanding of the mechanisms of microvascular regulation in human arterioles may provide needed insight into the pathogenic processes in patients with CMD.
Currently, both TRPV4 and NOXs have been implicated in many physiological and pathological processes in cardiovascular and other systems. For example, studies have shown that TRPV4 channels play an important role in a broad range of diseases such as cardiac remodeling [1], systemic hypertension [48], and pulmonary hypertension [36, 68]. It has been recently proposed that increased vascular expression of NOXs may be involved in heart failure with preserved ejection fraction (HFpEF) [25]. Our previous studies have shown that TRPV4 channels contribute to flow-mediated dilation in CAD arterioles [73]. The present data revealed a novel role for TRPV4 channels and its interplay with NOX4 in the regulation of the receptor agonist ACh-induced dilation in non-CAD arterioles. Given that ACh-induced dilation in adipose arterioles may serve as a more clinically accessible surrogate for coronary endothelial function in patients with CMD [19], our new findings will guide future research to elucidate the precise role and mechanism of TRPV4 and NOXs, as well as their therapeutic potential, in CMD and other cardiovascular diseases.
Supplementary Material
Acknowledgment
We thank Elmbrook Memorial Hospital and Froedtert Hospital for providing human tissues for the study. We also thank Dr. James H. Doroshow (NCI, NIH)) for kindly providing NOX4 antibody and NOX4 overexpressed HEK293 cell lysate and Dr. Charles K. Thodeti (NEOMED, OH) for providing PCR primer sequence information of NOX isoforms. This work was supported by the National Heart, Lung and Blood Institute Grant RO1-HL 096647 (to D.X.Z.), a generous gift from John B. and Judith A. Gardetto to the Children’s Research Foundation (to D.A.W.), and the funding for a joint Ph.D. program from the China Scholarship Council (to Y.X, contract 201708340067)
Funding
National Heart, Lung and Blood Institute Grant RO1-HL 096647 (to D.X.Z.), a generous gift from John B. and Judith A. Gardetto to the Children’s Research Foundation (to D.A.W.), and the funding for a joint Ph.D. program from the China Scholarship Council (to Y.X, contract 201708340067).
Footnotes
Conflicts of interest/Competing interests
The authors declare that they have no competing interests
Ethics approval
All protocols were approved by the Institutional Review Board of the Medical College of Wisconsin and Froedtert Hospital on the use of human subjects in research.
Availability of data and material
All data generated or analyzed during this study are included in this published article.
References
- 1.Adapala RK, Kanugula AK, Paruchuri S, Chilian WM, Thodeti CK (2020) TRPV4 deletion protects heart from myocardial infarction-induced adverse remodeling via modulation of cardiac fibroblast differentiation. Basic Res Cardiol 115:14 doi: 10.1007/s00395-020-0775-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Adapala RK, Talasila PK, Bratz IN, Zhang DX, Suzuki M, Meszaros JG, Thodeti CK (2011) PKCalpha mediates acetylcholine-induced activation of TRPV4-dependent calcium influx in endothelial cells. Am J Physiol-Heart Circul Physiol 301:H757–765 doi: 10.1152/ajpheart.00142.2011 [DOI] [Google Scholar]
- 3.Allaqaband H, Gutterman DD, Kadlec AO (2018) Physiological Consequences of Coronary Arteriolar Dysfunction and Its Influence on Cardiovascular Disease. Physiology 33:338–347 doi: 10.1152/physiol.00019.2018 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Aoyama T, Paik YH, Watanabe S, Laleu B, Gaggini F, Fioraso-Cartier L, Molango S, Heitz F, Merlot C, Szyndralewiez C, Page P, Brenner DA (2012) Nicotinamide adenine dinucleotide phosphate oxidase in experimental liver fibrosis: GKT137831 as a novel potential therapeutic agent. Hepatology 56:2316–2327 doi: 10.1002/hep.25938 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Bubolz AH, Mendoza SA, Zheng X, Zinkevich NS, Li R, Gutterman DD, Zhang DX (2012) Activation of endothelial TRPV4 channels mediates flow-induced dilation in human coronary arterioles: role of Ca2+ entry and mitochondrial ROS signaling. Am J Physiol-Heart Circul Physiol 302:H634–642 doi: 10.1152/ajpheart.00717.2011 [DOI] [Google Scholar]
- 6.Buus NH, Simonsen U, Pilegaard HK, Mulvany MJ (2000) Nitric oxide, prostanoid and non-NO, non-prostanoid involvement in acetylcholine relaxation of isolated human small arteries. Br J Pharmacol 129:184–192 doi: 10.1038/sj.bjp.0703041 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Cai H, Griendling KK, Harrison DG (2003) The vascular NAD(P)H oxidases as therapeutic targets in cardiovascular diseases. Trends Pharmacol Sci 24:471–478 doi: 10.1016/S0165-6147(03)00233-5 [DOI] [PubMed] [Google Scholar]
- 8.Camici PG, Crea F (2007) Coronary microvascular dysfunction. N Engl J Med 356:830–840 doi: 10.1056/NEJMra061889 [DOI] [PubMed] [Google Scholar]
- 9.Cao S, Anishkin A, Zinkevich NS, Nishijima Y, Korishettar A, Wang Z, Fang J, Wilcox DA, Zhang DX (2018) Transient receptor potential vanilloid 4 (TRPV4) activation by arachidonic acid requires protein kinase A-mediated phosphorylation. J Biol Chem 293:5307–5322 doi: 10.1074/jbc.M117.811075 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Cenac N, Altier C, Motta JP, d’Aldebert E, Galeano S, Zamponi GW, Vergnolle N (2010) Potentiation of TRPV4 signalling by histamine and serotonin: an important mechanism for visceral hypersensitivity. Gut 59:481–488 doi: 10.1136/gut.2009.192567 [DOI] [PubMed] [Google Scholar]
- 11.Chidgey J, Fraser PA, Aaronson PI (2016) Reactive oxygen species facilitate the EDH response in arterioles by potentiating intracellular endothelial Ca(2+) release. Free Radic Biol Med 97:274–284 doi: 10.1016/j.freeradbiomed.2016.06.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Craige SM, Chen K, Pei Y, Li C, Huang X, Chen C, Shibata R, Sato K, Walsh K, Keaney JF Jr. (2011) NADPH oxidase 4 promotes endothelial angiogenesis through endothelial nitric oxide synthase activation. Circulation 124:731–740 doi: 10.1161/CIRCULATIONAHA.111.030775 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Djordjevic T, Pogrebniak A, BelAiba RS, Bonello S, Wotzlaw C, Acker H, Hess J, Gorlach A (2005) The expression of the NADPH oxidase subunit p22phox is regulated by a redox-sensitive pathway in endothelial cells. Free Radic Biol Med 38:616–630 doi: 10.1016/j.freeradbiomed.2004.09.036 [DOI] [PubMed] [Google Scholar]
- 14.Edwards DH, Li Y, Griffith TM (2008) Hydrogen peroxide potentiates the EDHF phenomenon by promoting endothelial Ca2+ mobilization. Arterioscler Thromb Vasc Biol 28:1774–1781 doi: 10.1161/ATVBAHA.108.172692 [DOI] [PubMed] [Google Scholar]
- 15.Edwards G, Feletou M, Weston AH (2010) Endothelium-derived hyperpolarising factors and associated pathways: a synopsis. Pflugers Arch 459:863–879 doi: 10.1007/s00424-010-0817-1 [DOI] [PubMed] [Google Scholar]
- 16.Everaerts W, Zhen X, Ghosh D, Vriens J, Gevaert T, Gilbert JP, Hayward NJ, McNamara CR, Xue F, Moran MM, Strassmaier T, Uykal E, Owsianik G, Vennekens R, De Ridder D, Nilius B, Fanger CM, Voets T (2010) Inhibition of the cation channel TRPV4 improves bladder function in mice and rats with cyclophosphamide-induced cystitis. Proc Natl Acad Sci U S A 107:19084–19089 doi: 10.1073/pnas.1005333107 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Fan HC, Zhang X, McNaughton PA (2009) Activation of the TRPV4 ion channel is enhanced by phosphorylation. J Biol Chem 284:27884–27891 doi: 10.1074/jbc.M109.028803 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Filosa JA, Yao X, Rath G (2013) TRPV4 and the regulation of vascular tone. J Cardiovasc Pharmacol 61:113–119 doi: 10.1097/FJC.0b013e318279ba42 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Ford TJ, Rocchiccioli P, Good R, McEntegart M, Eteiba H, Watkins S, Shaukat A, Lindsay M, Robertson K, Hood S, Yii E, Sidik N, Harvey A, Montezano AC, Beattie E, Haddow L, Oldroyd KG, Touyz RM, Berry C (2018) Systemic microvascular dysfunction in microvascular and vasospastic angina. Eur Heart J 39:4086–4097 doi: 10.1093/eurheartj/ehy529 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Gao X, Wu L, O’Neil RG (2003) 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 278:27129–27137 doi: 10.1074/jbc.M302517200 [DOI] [PubMed] [Google Scholar]
- 21.Groschner K, Graier WF, Kukovetz WR (1994) Histamine induces K+, Ca2+, and Cl- currents in human vascular endothelial cells. Role of ionic currents in stimulation of nitric oxide biosynthesis. Circ Res 75:304–314 doi: 10.1161/01.res.75.2.304 [DOI] [PubMed] [Google Scholar]
- 22.Guler AD, Lee H, Iida T, Shimizu I, Tominaga M, Caterina M (2002) Heat-evoked activation of the ion channel, TRPV4. J Neurosci 22:6408–6414 doi: 10.1523/JNEUROSCI.22-15-06408.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Heathcote HR, Lee MD, Zhang X, Saunter CD, Wilson C, McCarron JG (2019) Endothelial TRPV4 channels modulate vascular tone by Ca(2+) -induced Ca(2+) release at inositol 1,4,5-trisphosphate receptors. Br J Pharmacol 176:3297–3317 doi: 10.1111/bph.14762 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Heusch G (2019) Coronary microvascular obstruction: the new frontier in cardioprotection. Basic Res Cardiol 114:45 doi: 10.1007/s00395-019-0756-8 [DOI] [PubMed] [Google Scholar]
- 25.Heusch G (2022) Coronary blood flow in heart failure: cause, consequence and bystander. Basic Res Cardiol 117:1 doi: 10.1007/s00395-022-00909-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Hirano K, Chen WS, Chueng AL, Dunne AA, Seredenina T, Filippova A, Ramachandran S, Bridges A, Chaudry L, Pettman G, Allan C, Duncan S, Lee KC, Lim J, Ma MT, Ong AB, Ye NY, Nasir S, Mulyanidewi S, Aw CC, Oon PP, Liao S, Li D, Johns DG, Miller ND, Davies CH, Browne ER, Matsuoka Y, Chen DW, Jaquet V, Rutter AR (2015) Discovery of GSK2795039, a Novel Small Molecule NADPH Oxidase 2 Inhibitor. Antioxid Redox Signal 23:358–374 doi: 10.1089/ars.2014.6202 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Ho WS, Zheng X, Zhang DX (2015) Role of endothelial TRPV4 channels in vascular actions of the endocannabinoid, 2-arachidonoylglycerol. Br J Pharmacol 172:5251–5264 doi: 10.1111/bph.13312 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Jow F, Numann R (2000) Histamine increases [Ca(2+)](in) and activates Ca-K and nonselective cation currents in cultured human capillary endothelial cells. J Membr Biol 173:107–116 doi: 10.1007/s002320001012 [DOI] [PubMed] [Google Scholar]
- 29.Kalari KR, Nair AA, Bhavsar JD, O’Brien DR, Davila JI, Bockol MA, Nie J, Tang X, Baheti S, Doughty JB, Middha S, Sicotte H, Thompson AE, Asmann YW, Kocher JP (2014) MAP-RSeq: Mayo Analysis Pipeline for RNA sequencing. BMC Bioinformatics 15:224 doi: 10.1186/1471-2105-15-224 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Khuddus MA, Pepine CJ, Handberg EM, Bairey Merz CN, Sopko G, Bavry AA, Denardo SJ, McGorray SP, Smith KM, Sharaf BL, Nicholls SJ, Nissen SE, Anderson RD (2010) An intravascular ultrasound analysis in women experiencing chest pain in the absence of obstructive coronary artery disease: a substudy from the National Heart, Lung and Blood Institute-Sponsored Women’s Ischemia Syndrome Evaluation (WISE). J Interv Cardiol 23:511–519 doi: 10.1111/j.1540-8183.2010.00598.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Kohler R, Heyken WT, Heinau P, Schubert R, Si H, Kacik M, Busch C, Grgic I, Maier T, Hoyer J (2006) Evidence for a functional role of endothelial transient receptor potential V4 in shear stress-induced vasodilatation. Arterioscler Thromb Vasc Biol 26:1495–1502 doi: 10.1161/01.ATV.0000225698.36212.6a [DOI] [PubMed] [Google Scholar]
- 32.Korishettar AM, Nishijima Y, Wang Z, Xie Y, Fang J, Wilcox DA, Zhang DX (2021) Endothelin-1 potentiates TRPV1-mediated vasoconstriction of human adipose arterioles in a protein kinase C-dependent manner. Br J Pharmacol 178:709–725 doi: 10.1111/bph.15324 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Larsen BT, Bubolz AH, Mendoza SA, Pritchard KA Jr., Gutterman DD (2009) Bradykinin-induced dilation of human coronary arterioles requires NADPH oxidase-derived reactive oxygen species. Arterioscler Thromb Vasc Biol 29:739–745 doi: 10.1161/ATVBAHA.108.169367 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Lassegue B, Clempus RE (2003) Vascular NAD(P)H oxidases: specific features, expression, and regulation. Am J Physiol Regul Integr Comp Physiol 285:R277–297 doi: 10.1152/ajpregu.00758.2002 [DOI] [PubMed] [Google Scholar]
- 35.Liedtke W, Choe Y, Marti-Renom MA, Bell AM, Denis CS, Sali A, Hudspeth AJ, Friedman JM, Heller S (2000) Vanilloid receptor-related osmotically activated channel (VR-OAC), a candidate vertebrate osmoreceptor. Cell 103:525–535 doi: 10.1016/s0092-8674(00)00143-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Martin E, Dahan D, Cardouat G, Gillibert-Duplantier J, Marthan R, Savineau JP, Ducret T (2012) Involvement of TRPV1 and TRPV4 channels in migration of rat pulmonary arterial smooth muscle cells. Pflugers Arch 464:261–272 doi: 10.1007/s00424-012-1136-5 [DOI] [PubMed] [Google Scholar]
- 37.Meitzler JL, Makhlouf HR, Antony S, Wu Y, Butcher D, Jiang G, Juhasz A, Lu J, Dahan I, Jansen-Durr P, Pircher H, Shah AM, Roy K, Doroshow JH (2017) Decoding NADPH oxidase 4 expression in human tumors. Redox Biol 13:182–195 doi: 10.1016/j.redox.2017.05.016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Mendoza SA, Fang J, Gutterman DD, Wilcox DA, Bubolz AH, Li R, Suzuki M, Zhang DX (2010) TRPV4-mediated endothelial Ca2+ influx and vasodilation in response to shear stress. Am J Physiol-Heart Circul Physiol 298:H466–476 doi: 10.1152/ajpheart.00854.2009 [DOI] [Google Scholar]
- 39.Mohapatra DP, Nau C (2005) Regulation of Ca2+-dependent desensitization in the vanilloid receptor TRPV1 by calcineurin and cAMP-dependent protein kinase. J Biol Chem 280:13424–13432 doi: 10.1074/jbc.M410917200 [DOI] [PubMed] [Google Scholar]
- 40.Munoz M, Lopez-Oliva ME, Rodriguez C, Martinez MP, Saenz-Medina J, Sanchez A, Climent B, Benedito S, Garcia-Sacristan A, Rivera L, Hernandez M, Prieto D (2020) Differential contribution of Nox1, Nox2 and Nox4 to kidney vascular oxidative stress and endothelial dysfunction in obesity. Redox Biol 28:101330 doi: 10.1016/j.redox.2019.101330 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Munoz M, Martinez MP, Lopez-Oliva ME, Rodriguez C, Corbacho C, Carballido J, Garcia-Sacristan A, Hernandez M, Rivera L, Saenz-Medina J, Prieto D (2018) Hydrogen peroxide derived from NADPH oxidase 4- and 2 contributes to the endothelium-dependent vasodilatation of intrarenal arteries. Redox Biol 19:92–104 doi: 10.1016/j.redox.2018.08.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Murthy VL, Naya M, Foster CR, Hainer J, Gaber M, Di Carli G, Blankstein R, Dorbala S, Sitek A, Pencina MJ, Di Carli MF (2011) Improved cardiac risk assessment with noninvasive measures of coronary flow reserve. Circulation 124:2215–2224 doi: 10.1161/CIRCULATIONAHA.111.050427 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Nilius B, Schwartz G, Oike M, Droogmans G (1993) Histamine-activated, non-selective cation currents and Ca2+ transients in endothelial cells from human umbilical vein. Pflugers Arch 424:285–293 doi: 10.1007/BF00384354 [DOI] [PubMed] [Google Scholar]
- 44.Nilius B, Viana F, Droogmans G (1997) Ion channels in vascular endothelium. Annu Rev Physiol 59:145–170 doi: 10.1146/annurev.physiol.59.1.145 [DOI] [PubMed] [Google Scholar]
- 45.Nishijima Y, Cao S, Chabowski DS, Korishettar A, Ge A, Zheng X, Sparapani R, Gutterman DD, Zhang DX (2017) Contribution of KV1.5 Channel to Hydrogen Peroxide-Induced Human Arteriolar Dilation and Its Modulation by Coronary Artery Disease. Circ Res 120:658–669 doi: 10.1161/CIRCRESAHA.116.309491 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Nishijima Y, Zheng X, Lund H, Suzuki M, Mattson DL, Zhang DX (2014) Characterization of blood pressure and endothelial function in TRPV4-deficient mice with l-NAME- and angiotensin II-induced hypertension. Physiol Rep 2:e00199 doi: 10.1002/phy2.199 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Nisimoto Y, Diebold BA, Cosentino-Gomes D, Lambeth JD (2014) Nox4: a hydrogen peroxide-generating oxygen sensor. Biochemistry 53:5111–5120 doi: 10.1021/bi500331y [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Ottolini M, Hong K, Cope EL, Daneva Z, DeLalio LJ, Sokolowski JD, Marziano C, Nguyen NY, Altschmied J, Haendeler J, Johnstone SR, Kalani MY, Park MS, Patel RP, Liedtke W, Isakson BE, Sonkusare SK (2020) Local Peroxynitrite Impairs Endothelial Transient Receptor Potential Vanilloid 4 Channels and Elevates Blood Pressure in Obesity. Circulation 141:1318–1333 doi: 10.1161/CIRCULATIONAHA.119.043385 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Parrinello R, Sestito A, Di Franco A, Russo G, Villano A, Figliozzi S, Nerla R, Tarzia P, Stazi A, Lanza GA, Crea F (2014) Peripheral arterial function and coronary microvascular function in patients with variant angina. Cardiology 129:20–24 doi: 10.1159/000362380 [DOI] [PubMed] [Google Scholar]
- 50.Pendyala S, Gorshkova IA, Usatyuk PV, He D, Pennathur A, Lambeth JD, Thannickal VJ, Natarajan V (2009) Role of Nox4 and Nox2 in hyperoxia-induced reactive oxygen species generation and migration of human lung endothelial cells. Antioxid Redox Signal 11:747–764 doi: 10.1089/ARS.2008.2203 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Pepine CJ, Anderson RD, Sharaf BL, Reis SE, Smith KM, Handberg EM, Johnson BD, Sopko G, Bairey Merz CN (2010) Coronary microvascular reactivity to adenosine predicts adverse outcome in women evaluated for suspected ischemia results from the National Heart, Lung and Blood Institute WISE (Women’s Ischemia Syndrome Evaluation) study. J Am Coll Cardiol 55:2825–2832 doi: 10.1016/j.jacc.2010.01.054 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Rath G, Saliez J, Behets G, Romero-Perez M, Leon-Gomez E, Bouzin C, Vriens J, Nilius B, Feron O, Dessy C (2012) Vascular hypoxic preconditioning relies on TRPV4-dependent calcium influx and proper intercellular gap junctions communication. Arterioscler Thromb Vasc Biol 32:2241–2249 doi: 10.1161/ATVBAHA.112.252783 [DOI] [PubMed] [Google Scholar]
- 53.Ray R, Murdoch CE, Wang M, Santos CX, Zhang M, Alom-Ruiz S, Anilkumar N, Ouattara A, Cave AC, Walker SJ, Grieve DJ, Charles RL, Eaton P, Brewer AC, Shah AM (2011) Endothelial Nox4 NADPH oxidase enhances vasodilatation and reduces blood pressure in vivo. Arterioscler Thromb Vasc Biol 31:1368–1376 doi: 10.1161/ATVBAHA.110.219238 [DOI] [PubMed] [Google Scholar]
- 54.Reutens AT, Jandeleit-Dahm K, Thomas M, Salim A, De Livera AM, Bach LA, Colman PG, Davis TME, Ekinci EI, Fulcher G, Hamblin PS, Kotowicz MA, MacIsaac RJ, Morbey C, Simmons D, Soldatos G, Wittert G, Wu T, Cooper ME, Shaw JE (2020) A physician-initiated double-blind, randomised, placebo-controlled, phase 2 study evaluating the efficacy and safety of inhibition of NADPH oxidase with the first-in-class Nox-1/4 inhibitor, GKT137831, in adults with type 1 diabetes and persistently elevated urinary albumin excretion: Protocol and statistical considerations. Contemp Clin Trials 90:105892 doi: 10.1016/j.cct.2019.105892 [DOI] [PubMed] [Google Scholar]
- 55.Rigo F, Pratali L, Palinkas A, Picano E, Cutaia V, Venneri L, Raviele A (2002) Coronary flow reserve and brachial artery reactivity in patients with chest pain and “false positive” exercise-induced ST-segment depression. Am J Cardiol 89:1141–1144 doi: 10.1016/s0002-9149(02)02292-0 [DOI] [PubMed] [Google Scholar]
- 56.Sorop O, van de Wouw J, Chandler S, Ohanyan V, Tune JD, Chilian WM, Merkus D, Bender SB, Duncker DJ (2020) Experimental animal models of coronary microvascular dysfunction. Cardiovasc Res 116:756–770 doi: 10.1093/cvr/cvaa002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Strotmann R, Harteneck C, Nunnenmacher K, Schultz G, Plant TD (2000) OTRPC4, a nonselective cation channel that confers sensitivity to extracellular osmolarity. Nat Cell Biol 2:695–702 doi: 10.1038/35036318 [DOI] [PubMed] [Google Scholar]
- 58.Takac I, Schroder K, Zhang L, Lardy B, Anilkumar N, Lambeth JD, Shah AM, Morel F, Brandes RP (2011) The E-loop is involved in hydrogen peroxide formation by the NADPH oxidase Nox4. J Biol Chem 286:13304–13313 doi: 10.1074/jbc.M110.192138 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Taqueti VR, Hachamovitch R, Murthy VL, Naya M, Foster CR, Hainer J, Dorbala S, Blankstein R, Di Carli MF (2015) Global coronary flow reserve is associated with adverse cardiovascular events independently of luminal angiographic severity and modifies the effect of early revascularization. Circulation 131:19–27 doi: 10.1161/CIRCULATIONAHA.114.011939 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Thompson JA, Larion S, Mintz JD, Belin de Chantemele EJ, Fulton DJ, Stepp DW (2017) Genetic Deletion of NADPH Oxidase 1 Rescues Microvascular Function in Mice With Metabolic Disease. Circ Res 121:502–511 doi: 10.1161/CIRCRESAHA.116.309965 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Thorneloe KS, Sulpizio AC, Lin Z, Figueroa DJ, Clouse AK, McCafferty GP, Chendrimada TP, Lashinger ES, Gordon E, Evans L, Misajet BA, Demarini DJ, Nation JH, Casillas LN, Marquis RW, Votta BJ, Sheardown SA, Xu X, Brooks DP, Laping NJ, Westfall TD (2008) N-((1S)-1-{[4-((2S)-2-{[(2,4-dichlorophenyl)sulfonyl]amino}−3-hydroxypropanoyl)-1 -piperazinyl]carbonyl}−3-methylbutyl)-1-benzothiophene-2-carboxamide (GSK1016790A), a novel and potent transient receptor potential vanilloid 4 channel agonist induces urinary bladder contraction and hyperactivity: Part I. J Pharmacol Exp Ther 326:432–442 doi: 10.1124/jpet.108.139295 [DOI] [PubMed] [Google Scholar]
- 62.Vriens J, Owsianik G, Janssens A, Voets T, Nilius B (2007) Determinants of 4 alpha-phorbol sensitivity in transmembrane domains 3 and 4 of the cation channel TRPV4. J Biol Chem 282:12796–12803 doi: 10.1074/jbc.M610485200 [DOI] [PubMed] [Google Scholar]
- 63.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 (2002) Activation of TRPV4 channels (hVRL-2/mTRP12) by phorbol derivatives. J Biol Chem 277:13569–13577 doi: 10.1074/jbc.M200062200 [DOI] [PubMed] [Google Scholar]
- 64.Weiss WA, Taylor SS, Shokat KM (2007) Recognizing and exploiting differences between RNAi and small-molecule inhibitors. Nat Chem Biol 3:739–744 doi: 10.1038/nchembio1207-739 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Wissenbach U, Bodding M, Freichel M, Flockerzi V (2000) Trp12, a novel Trp related protein from kidney. FEBS Lett 485:127–134 doi: 10.1016/s0014-5793(00)02212-2 [DOI] [PubMed] [Google Scholar]
- 66.Xie J, Hong E, Ding B, Jiang W, Zheng S, Xie Z, Tian D, Chen Y (2020) Inhibition of NOX4/ROS Suppresses Neuronal and Blood-Brain Barrier Injury by Attenuating Oxidative Stress After Intracerebral Hemorrhage. Front Cell Neurosci 14:578060 doi: 10.3389/fncel.2020.578060 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Xue K, Wang Y, Wang Y, Fang H (2021) Advanced Oxidation Protein Product Promotes Oxidative Accentuation in Renal Epithelial Cells via the Soluble (Pro)renin Receptor-Mediated Intrarenal Renin-Angiotensin System and Nox4-H2O2 Signaling. Oxid Med Cell Longev 2021:5710440 doi: 10.1155/2021/5710440 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Yang XR, Lin AH, Hughes JM, Flavahan NA, Cao YN, Liedtke W, Sham JS (2012) Upregulation of osmo-mechanosensitive TRPV4 channel facilitates chronic hypoxia-induced myogenic tone and pulmonary hypertension. Am J Physiol Lung Cell Mol Physiol 302:L555–568 doi: 10.1152/ajplung.00005.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Zhang DX, Borbouse L, Gebremedhin D, Mendoza SA, Zinkevich NS, Li R, Gutterman DD (2012) H2O2-induced dilation in human coronary arterioles: role of protein kinase G dimerization and large-conductance Ca2+-activated K+ channel activation. Circ Res 110:471–480 doi: 10.1161/CIRCRESAHA.111.258871 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Zhang DX, Gutterman DD (2011) Transient receptor potential channel activation and endothelium-dependent dilation in the systemic circulation. J Cardiovasc Pharmacol 57:133–139 doi: 10.1097/FJC.0b013e3181fd35d1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Zhang DX, Mendoza SA, Bubolz AH, Mizuno A, Ge ZD, Li R, Warltier DC, Suzuki M, Gutterman DD (2009) Transient receptor potential vanilloid type 4-deficient mice exhibit impaired endothelium-dependent relaxation induced by acetylcholine in vitro and in vivo. Hypertension 53:532–538 doi: 10.1161/hypertensionaha.108.127100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Zhang Y, Wernly B, Cao X, Mustafa SJ, Tang Y, Zhou Z (2021) Adenosine and adenosine receptor-mediated action in coronary microcirculation. Basic Res Cardiol 116:22 doi: 10.1007/s00395-021-00859-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Zheng X, Zinkevich NS, Gebremedhin D, Gauthier KM, Nishijima Y, Fang J, Wilcox DA, Campbell WB, Gutterman DD, Zhang DX (2013) Arachidonic acid-induced dilation in human coronary arterioles: convergence of signaling mechanisms on endothelial TRPV4-mediated Ca2+ entry. J Am Heart Assoc 2:e000080 doi: 10.1161/JAHA.113.000080 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Zinkevich NS, Fancher IS, Gutterman DD, Phillips SA (2017) Roles of NADPH oxidase and mitochondria in flow-induced vasodilation of human adipose arterioles: ROS-induced ROS release in coronary artery disease. Microcirculation 24 doi: 10.1111/micc.12380 [DOI] [Google Scholar]
- 75.Zinkevich NS, Gutterman DD (2011) ROS-induced ROS release in vascular biology: redox-redox signaling. Am J Physiol Heart Circ Physiol 301:H647–653 doi: 10.1152/ajpheart.01271.2010 [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All data generated or analyzed during this study are included in this published article.
