Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

HIF-1α Activation by Intermittent Hypoxia Requires NADPH Oxidase Stimulation by Xanthine Oxidase

  • Jayasri Nanduri ,

    jnanduri@uchicago.edu

    Affiliation Institute for Integrative Physiology and Center for Systems Biology of O2 Sensing, Biological Sciences Division, University of Chicago, Chicago, Illinois, United States of America

  • Damodara Reddy Vaddi,

    Affiliation Institute for Integrative Physiology and Center for Systems Biology of O2 Sensing, Biological Sciences Division, University of Chicago, Chicago, Illinois, United States of America

  • Shakil A. Khan,

    Affiliation Institute for Integrative Physiology and Center for Systems Biology of O2 Sensing, Biological Sciences Division, University of Chicago, Chicago, Illinois, United States of America

  • Ning Wang,

    Affiliation Institute for Integrative Physiology and Center for Systems Biology of O2 Sensing, Biological Sciences Division, University of Chicago, Chicago, Illinois, United States of America

  • Vladislav Makarenko,

    Affiliation Institute for Integrative Physiology and Center for Systems Biology of O2 Sensing, Biological Sciences Division, University of Chicago, Chicago, Illinois, United States of America

  • Gregg L. Semenza,

    Affiliation Vascular Program, Institute for Cell Engineering; Department of Pediatrics, Medicine, Oncology, Radiation Oncology and Biological Chemistry; and Mckusick-Nathans Institute of Genetic Medicine, The Johns Hopkins University School of Medicine, Baltimore, Maryland, United States of America

  • Nanduri R. Prabhakar

    Affiliation Institute for Integrative Physiology and Center for Systems Biology of O2 Sensing, Biological Sciences Division, University of Chicago, Chicago, Illinois, United States of America

Expression of Concern

Following the publication of this article [1], concerns were raised regarding Figs 4 and 5. Specifically, there appears to be vertical discontinuities in the following locations:

  • Fig 4A, Cytosol p47phox panel, between lanes 1/2 and 2/3.
  • Fig 4A, Membrane Tubulin panel, between lanes 1/2.
  • Fig 5, IP: p47phox panel, between lanes 1/2 and 2/3.

The corresponding author stated that primary data underlying the published results remain available except for the original images underlying Fig 4 and the IP: p47phox panel in Fig 5. In the absence of the original data for these figures, they could not comment on whether the panels listed above were spliced.

The corresponding author provided uncropped images described as replicate data for Fig 4A and individual-level quantitative data underlying the graph in Fig 4B (included with this notice as S1 and S2 Files).

The corresponding author also provided uncropped images described as the original data for Fig 5 IP: p67phox panels (S3 File) and original uncropped images corresponding to the published IP: p47phox, gp91phox, and p67phox panels in Fig 5 (S4 File). For the IP: p67phox panels, it was not possible to be certain that these correspond to the published panel due to poor image quality.

For Figs 4 and 5, the replicate or original data provided lend support for the reported results. However, without the original image data underlying the published figures, the concerns cannot be fully resolved.

The PLOS ONE Editors issue this Expression of Concern to notify readers of the above concerns which may impact the reliability of the conclusions drawn from these figures, and to relay the supporting data provided by the corresponding author.

Supporting information

S2 File. Quantitative data underlying Fig 4B.

https://doi.org/10.1371/journal.pone.0299122.s002

(XLSX)

S3 File. Images described as the original data for Fig 5 IP: p67phox panels.

gp91phox is also known as Nox2.

https://doi.org/10.1371/journal.pone.0299122.s003

(ZIP)

S4 File. Original images underlying Fig 5 IP: p47phox, except the p47phox panel.

gp91phox is also known as Nox2.

https://doi.org/10.1371/journal.pone.0299122.s004

(ZIP)

14 Feb 2024: The PLOS ONE Editors (2024) Expression of Concern: HIF-1α Activation by Intermittent Hypoxia Requires NADPH Oxidase Stimulation by Xanthine Oxidase. PLOS ONE 19(2): e0299122. https://doi.org/10.1371/journal.pone.0299122 View expression of concern

Abstract

Hypoxia-inducible factor 1 (HIF-1) mediates many of the systemic and cellular responses to intermittent hypoxia (IH), which is an experimental model that simulates O2 saturation profiles occurring with recurrent apnea. IH-evoked HIF-1α synthesis and stability are due to increased reactive oxygen species (ROS) generated by NADPH oxidases, especially Nox2. However, the mechanisms by which IH activates Nox2 are not known. We recently reported that IH activates xanthine oxidase (XO) and the resulting increase in ROS elevates intracellular calcium levels. Since Nox2 activation requires increased intracellular calcium levels, we hypothesized XO-mediated calcium signaling contributes to Nox activation by IH. We tested this possibility in rat pheochromocytoma PC12 cells subjected to IH consisting alternating cycles of hypoxia (1.5% O2 for 30 sec) and normoxia (21% O2 for 5 min). Kinetic analysis revealed that IH-induced XO preceded Nox activation. Inhibition of XO activity either by allopurinol or by siRNA prevented IH-induced Nox activation, translocation of the cytosolic subunits p47phox and p67phox to the plasma membrane and their interaction with gp91phox. ROS generated by XO also contribute to IH-evoked Nox activation via calcium-dependent protein kinase C stimulation. More importantly, silencing XO blocked IH-induced upregulation of HIF-1α demonstrating that HIF-1α activation by IH requires Nox2 activation by XO.

Introduction

Sleep disordered breathing with recurrent apnea is characterized by transient (~10–15 sec in adults) repetitive cessations of breathing, resulting in periodic decreases in arterial blood O2 or intermittent hypoxia (IH). Patients with recurrent apnea develop several co-morbidities including hypertension and breathing abnormalities [1, 2]. Hypoxia-inducible factor-1 (HIF-1) is the master transcriptional activator that regulates gene expression during hypoxia [3]. IH activates HIF-1 mediated transcription in cell cultures and rodents [46]. Mice with heterozygous deficiency of HIF-1α exhibit remarkable absence of IH-induced hypertension, and breathing abnormalities [5, 6], suggesting that activation of HIF-1 contributes to the cardio-respiratory abnormalities caused by IH. Analysis of the mechanisms underlying HIF-1 activation showed that IH activates NADPH oxidase (Nox), especially Nox2 [7]. Disruption of Nox2 function, prevents HIF-1 activation by IH in cell cultures and in mice [4, 8], suggesting that Nox2 is critical for HIF-1 activation by IH. However, the mechanism by which IH activates Nox2 has not been examined.

Nox2 is a multi-enzyme protein complex consisting of a membrane-bound gp91phox and cytosolic p67phox and p47phox subunits [911]. Nox2 activation requires phosphorylation of cytosolic subunits by Ca2+-activated protein kinase C (PKC) and translocation to the membrane to form a complex with gp91phox [12]. We recently reported that IH activates xanthine oxidase (XO), which is a major cellular source of reactive oxygen species (ROS) [13]. IH-induced XO activation leads to ROS-dependent elevation of intracellular calcium ([Ca2+]i). Given that Nox2 activation critically depends on Ca2+-activated PKC [14], we hypothesized that [Ca2+]i elevation by XO mediates IH-induced Nox2 activation and subsequent stimulation of HIF-1. We examined this possibility in rat pheochromocytoma 12 (PC12) cell cultures exposed to IH.

Materials and Methods

Exposure of cell cultures to IH

PC12 cells (original clone from Dr. Lloyd Greene, Columbia University Medical Center, New York) [15] were cultured in Dulbecco’s modified Eagle’s medium (DMEM) supplemented with 10% horse serum, 5% fetal bovine serum (FBS), penicillin (100 U/ml), and streptomycin (100 μg/ml) under 10% CO2 and 90% air (20% O2) at 37°C [16]. Experiments were performed on cells serum starved for 16 h in antibiotic free medium. In the experiments involving treatment with drugs, cells were pre-incubated for 30 min with either drug or vehicle. Cell cultures were exposed to IH (1.5% O2 for 30 sec followed by 20% O2 for 5 min at 37°C) as described [16]. Ambient O2 levels in the IH chamber were monitored by an O2 analyzer (Alpha Omega Instruments).

Studies with short interfering RNA (siRNA)

PC12 cells (5x105) were plated on culture dishes coated with type IV collagen (BD Biosciences, Bedford, MA) and transfected with siRNA (Santa Cruz) targeted to XDH, gp91phox (Nox2), or a scrambled sequence at a concentration of 100 pmol/mL using DharmaFECT 2 (Dharmacon Research). Transfected cells were cultured in complete medium for 48 h before being exposed to IH.

Measurement of XO activity

Amplex Red Xanthine/Xanthine oxidase assay kit (Molecular Probes) was used to monitor to XO activity as described [13]. Cell lysates were incubated with a reaction mixture containing hypoxanthine, horseradish peroxidase (HRP) and Amplex Red. H2O2 reacts with Amplex Red in the presence of HRP to generate the red fluorescent oxidation product resorufin. Fluorescence was measured by excitation at 530 nm and emission at 590 nm. The concentration of XO in samples was determined from a standard curve and expressed as XO mU/mg protein.

Measurement of NADPH oxidase activity

NADPH oxidase activity in the membrane-enriched protein fractions was measured by superoxide dismutase-inhibitable rate of cytochrome c reduction as described [17]. Briefly, the assay medium contained 100 μg membrane protein, 150 μm cytochrome c and 100 μm NADPH in 25 mM HEPES buffer (pH 7.0). The assay was performed in the presence and absence of superoxide dismutase (200 units/ml) at 37°C for 30 min. Cytochrome c reduction was measured by reading the absorbance at 550 nm. NADPH oxidase activity was calculated based on extinction coefficient [21mmol/(L)] per cm and expressed as nmol/min/mg protein.

Immunoprecipitation and Immunoblot assays

Membrane enriched fractions from PC12 cells were isolated using a plasma membrane protein extraction kit (Abcam). Briefly, cells were homogenized in the buffer mix provided and centrifuged at 700 g for 10 min at 4°C. The supernatant was further centrifuged at 10,000 g for 30 minutes at 4°C. The pellet contained the plasma membrane and cellular organelle membrane proteins. The purified membrane enriched proteins were resuspended in lysis buffer and fractionated by polyacrylamide-SDS gel electrophoresis and immunoblotted with gp91phox (sc-7663), p67phox (sc-7660) or p47phox (sc-3678) antibodies (Santa Cruz Biotechnology Inc, Dallas, USA). Antibody binding was detected using HRP-conjugated secondary antibodies followed by enhanced chemiluminescence detection system (Bio-Rad). Immunoprecipitation experiments were performed using protein A/G magnetic beads (Millipore). To detect phosphorylated p47phox and p67phox, total cell lysates were incubated with p47phox and p67phox antibodies and immune complexes were isolated using protein A/G magnetic beads. Phosphorylation of p47phox and p67phox was analyzed by immunoblots with an anti-phosphoserine antibody at a 1:500 dilution.

Measurement of PKC activity

The PKC activity was determined using an ELISA kit (Enzo Life sciences, Farmingdale, NY, USA), according to the manufacturer's instructions. Briefly, lysates from control or IH exposed cells were added to a microplate that was precoated with a PKC substrate, followed by the addition of ATP to initiate the reaction. PKC activity was determined by adding a phospho-specific substrate antibody conjugated to peroxidase and tetramethylbenzidine as a substrate. Color development, which is proportional to PKC activity was measured at 450-nm absorbance.

Measurement of [Ca2+]i

Intracellular calcium levels were monitored in PC12 cells using Fura-2-AM as described previously [18]. Background fluorescence was subtracted from signals. Image intensity at 340 nm was divided by 380-nm image intensity to obtain the ratiometric image. Ratios were converted to free [Ca2+]i using calibration curves constructed in vitro by adding Fura-2 (50 μM, free acid) to solutions containing known concentrations of Ca2+ (0–2000 nM).

Statistical analysis

Data were expressed as mean ± S.E.M from 3–5 independent experiments with each performed in triplicate. Statistical analysis was performed by analysis of variance (ANOVA) and p values < 0.05 were considered significant.

Results

XO mediates IH-induced Nox activation

PC12 cells were exposed to increasing cycles of IH with each cycle consisting of 30 sec of hypoxia followed by 5 min of normoxia. Analysis of the time course revealed that XO activity increased progressively with increasing number of IH cycles with a significant activation after 5 cycles, whereas Nox activation required a minimum of 10 cycles of IH (Fig. 1A). Because XO activation by IH precedes that of Nox, we hypothesized that it triggers Nox activation. To test this possibility, PC12 cells were treated with 10 μM allopurinol (ALLO), an inhibitor of XO and then were exposed to IH. ALLO blocked Nox activation in cells exposed to either IH10 or IH60 (Fig. 1B). Conversely, treating control cells with Xanthine/XO, (250 μM/0.01 U/mL), stimulated Nox activity and ALLO blocked this effect (Fig. 1C).

thumbnail
Fig 1. XO activation is required for increased Nox activity in response to IH.

A) Time course of IH on XO and Nox activity. Squares represent significant increase in XO and Nox activity at 5 and 10 cycles of IH respectively. B) Effect of XO inhibitor, Allopurinol (ALLO; 10 μM) on Nox activity in PC12 cells exposed to 10 cycles (IH10) and 60 cycles (IH60) of IH or normoxia (N). C) Effect of Xanthine/XO (250 μM/0.01 U/mL) treatment on Nox activity under normoxic conditions with and without ALLO. Data is presented as mean ± S.E.M from four independent experiments. ** p<0.01; * p< 0.05; ns: not significant.

https://doi.org/10.1371/journal.pone.0119762.g001

To further establish the role of XO, PC12 cells were transfected with small interfering RNA (siRNA) targeted to xanthine dehydrogenase (XDH), which is the precursor of XO. IH-induced XO and Nox activation were absent in cells transfected with XDH siRNA as compared with cells treated with scrambled siRNA (Fig. 2A-B).

thumbnail
Fig 2. Effect of silencing XO on Nox activation.

A-B) XO and Nox activity in PC12 cells transfected with siRNA targeted against XDH, the precursor of XO or scrambled (scr) siRNA and exposed to IH10, IH60 or normoxia (N). Data is presented as mean ± S.E.M from four independent experiments. **p<0.01; ns: not significant.

https://doi.org/10.1371/journal.pone.0119762.g002

Ca2+ dependent PKC activation by IH requires XO

The mechanism(s) underlying XO-dependent Nox2 activation by IH were determined. Nox2 activation requires Ca2+-dependent PKC activation [14, 19]. Cells exposed to IH10 exhibited two-fold elevation in [Ca2+]i levels and this effect was blocked by ALLO (Fig. 3A). We next determined whether IH10 elevated calcium levels activate PKC. In response to IH10, PKC activity increased 2.5 fold compared to control cells (P<0.01), and this effect was prevented by either ALLO, an inhibitor of XO; BAPTA-AM (10μM), a calcium chelator; or bisindolymaleimide (Bis-1; 10 μM), a pan PKC isoform inhibitor (Fig. 3B).

thumbnail
Fig 3. XO augments Nox stimulation via Ca2+ dependent PKC activation.

A) [Ca2+]i levels were monitored in PC12 cells exposed to normoxia (N) or IH10 with or without ALLO treatment. B) Measurement of PKC activity in cells exposed to normoxia (N) or IH10 in presence of ALLO, bisindolymaleimide (Bis-1; inhibitor of multiple PKC isoforms, 10 μM) or BAPTA (intracellular calcium chelator, 10 μM). C) Representative immunoblot showing phosphorylation of p67phox and p47phox in presence of Bis-1 and BAPTA. D) Nox activity in cells exposed to normoxia (N) or IH10 in presence of BAPTA and Bis-1. Data are presented as mean ± S.E.M. from four experiments. ** p< 0.01; * p< 0.05; ns: not significant.

https://doi.org/10.1371/journal.pone.0119762.g003

XO mediates IH-evoked phosphorylation and translocation of p47phox and p67phox

Nox2 activation requires PKC-dependent phosphorylation and subsequent translocation of the cytosolic p47phox and p67phox subunits [9] to the membrane. As shown in Fig. 3C, IH-exposed cells exhibited increased levels of phosphorylated p47phox and p67phox levels. Treating cells with either BAPTA-AM or Bis-1 prevented IH-induced phosphorylation of cytosolic subunits (Fig. 3C) as well as Nox activation (Fig. 3D).

To demonstrate the translocation of the cytosolic subunits to the plasma membrane, the distribution of p67phox and p47phox proteins was analyzed in cytosolic and plasma membrane fractions of PC12 cells exposed to normoxia or IH along with gp91phox. The cytosol and membrane fractions were confirmed with the relative abundance of α-tubulin and pan-cadherin proteins, respectively (Fig. 4A). In normoxic cells, gp91phox was localized predominantly in the membrane fraction, whereas p67phox and p47phox were abundant in the cytosolic fraction (Fig. 4A). While IH had no effect on gp91phox, it increased p67phox and p47phox levels in the membrane fraction with concomitant reduction in the cytosolic fraction (Fig. 4A-B). Pre-treatment with ALLO prevented the effects of IH on the distribution of p67phox and p47phox proteins to the membrane (Fig. 4A-B).

thumbnail
Fig 4. XO mediates IH-evoked translocation of p47phox and p67phox.

A) Representative immunoblot showing the distribution of gp91phox, p67phox and p47phox proteins in cytosolic and membrane fractions from cells exposed to normoxia (N) or IH10 with or without ALLO. B) Densitometric analysis of three immunoblots presented as mean ± S.E.M. Membrane to cytosol ratio was calculated for each protein after normalizing to tubulin and pan-cadherin respectively. Data is presented as mean ± S.E.M from three independent experiments. ** p<0.01; ns: not significant.

https://doi.org/10.1371/journal.pone.0119762.g004

Nox2 complex formation by IH requires XO

The effect of IH on Nox2 complex formation was determined. To this end, p67phox and p47phox were immunoprecipitated and immunoblot assays were performed with anti-gp91phox, p67phox and p47phox antibodies. As shown in Fig. 5, amount of gp91phox co-immunoprecipitated with p67phox and p47phox increased in IH-exposed cells as compared to cells exposed to normoxia, indicating formation of the active Nox2 complex. Likewise, p67phox was detected in p47phox immunoprecipitates and vice versa demonstrating the association of all three subunits. Treatment with ALLO completely abolished Nox2 complex formation in cells subjected to IH (Fig. 5).

thumbnail
Fig 5. Nox2 complex formation by IH requires XO.

Nox2 complex immunoprecipitated from membrane fraction of cells exposed to normoxia (N) or IH10 with p67phox or p47phox antibody and analyzed for gp91phox, p67phox and p47phox proteins by immunoblot.

https://doi.org/10.1371/journal.pone.0119762.g005

XO contributes to HIF-1α upregulation by IH

Our previous studies showed that Nox2 is required for HIF-1α activation by IH [4]. The results described above demonstrate that XO mediates IH-evoked Nox activation. Therefore, we tested the effects of silencing XO activation on HIF-1α activation by IH. Since significant HIF-1α activation was seen with IH60 but not IH10, HIF-1α protein expression was analyzed in IH60 exposed cells transfected with siRNAs targeted to either XDH or gp91phox. IH-evoked HIF-1α activation was completely prevented in cells treated with XDH or gp91phox siRNA (Fig. 6A-B).

thumbnail
Fig 6. XO contributes to HIF-1α upregulation by IH.

A) Top panel: Representative immunoblot showing the effects of XO silencing on HIF-1α protein expression in cells exposed to IH60. Bottom panel: densitometric analysis of three immunoblots presented as mean ± S.E.M. B) Representative immunoblot showing HIF-1α and Nox2 expression in cells transfected with scrambled (scr) siRNA or Nox2 siRNA and exposed to normoxia (N) or IH60 with or without ALLO treatment. Data are presented as mean ± S.E.M. from four experiments. ** p < 0.01; ns: not significant.

https://doi.org/10.1371/journal.pone.0119762.g006

Discussion

In the present study, we delineated the signaling mechanisms underlying Nox2 activation by IH. The following findings demonstrate that XO activation is the initial trigger for Nox2 activation by IH: first, IH-induced XO activation precedes Nox2; second, pharmacological blockade or genetic silencing of XO abolished Nox2 activation by IH; and third, XO activation increased Nox activity under normoxia, mimicking the effects of IH.

Our results provide insight into how XO stimulates Nox activity. In both vascular endothelial and smooth muscle cells, Nox2 activation requires translocation of the cytosolic subunits to the membrane via [Ca2+]i-activated PKC [19, 20]. Consistent with this possibility, we found that IH increases Ca2+-dependent PKC activation in PC12 cells and this effect was prevented by blockade of XO activity. More importantly, our results further demonstrated robust PKC–dependent phosphorylation of p47phox and p67phox, and translocation to the plasma membrane in IH-exposed cells. Co-immunoprecipitation experiments provided direct evidence for Nox2 complex formation during IH, which was dependent on XO activation. PKC α, β, δ, and ζ are all known to phosphorylate p47phox (14), whereas only PKC ζ is known to target p67phox [21]. Further studies are necessary to delineate the relative contribution of PKC isoforms to the phosphorylation of Nox2 cytosolic subunits during IH. Interestingly, IH10 not only increased phosphorylation of p67phox and p47phox but also increased their relative expression. One possible explanation could be that complexing with gp91phox at the membrane or changing the phosphorylation states of the individual components may increase their stability as previously proposed [20]. Pharmacological treatments that inhibited the subunits from translocating to the membrane also prevented their increased expression, supporting the proposed hypothesis. Further investigation is required to confirm this model. An important finding of this study is that IH-induced XO activity also contributes to increased levels of HIF-1α. We previously reported that IH-evoked HIF-2α degradation is mediated by Ca2+-dependent calpain activation via XO-generated ROS [13]. Current results demonstrate that the same Ca2+ signaling triggered by XO also mediates IH-evoked HIF-1α activation by stimulating Nox2. XO activation by IH is mediated by proteolytic conversion of XDH to XO which is blocked by inhibiting proteolysis with a trypsin inhibitor [13]. These findings are reminiscent of an early study showing that XDH can be converted to XO within minutes in response to ischemia [22] and suggests that similar rapid proteolytic conversion of XDH to XO during the initial cycles of the IH paradigm may play a regulatory role in triggering Nox2 activation. Previous studies with ischemia re-perfusion models showed that XO gets activated during the re-oxygenation phase rather than during ischemia. Similar to ischemia-reperfusion, IH is associated with alternating cycles of hypoxia and re-oxygenation. Since such re-oxygenation periods are absent during sustained hypoxia, we believe XO-NOX-HIF-α pathway is selective to IH.

It was proposed that increased generation of ROS mediates the pathological effects of IH [23]. Indeed, ROS levels were elevated in the carotid body [24], adrenal medulla [5, 25, 26] and central nervous system [27, 28] of rats and mice exposed to chronic IH. Remarkably, the effects of IH on catecholamine secretion, blood pressure elevation, long-term facilitation of respiratory motor activity, and sensory long-term facilitation in the carotid body can all be blocked by treating rodents with a ROS scavenger [2933]. Obstructive sleep apnea patients exhibit elevated ROS levels [3436] with impaired vasodilation, and antioxidant treatment restores the vascular responses [37]. These studies suggest that ROS play a critical role in mediating the cardiovascular pathology associated with IH. We previously reported that IH increases HIF-1α and decreases HIF-2α protein levels [25]. HIF-1 mediates transcriptional activation of Nox2, a pro-oxidant, whereas HIF-2 is a potent activator of genes encoding anti-oxidants including superoxide dismutase-2 [38]. Taken together, results from the current and previous studies, it is likely that XO-dependent ROS mediates imbalance between HIF-1 and HIF-2 via Nox2 and calpain activation, which leads to further long-lasting oxidative stress during IH that results in autonomic morbidities as summarized in Fig. 7.

thumbnail
Fig 7. Schematic representation of feed-forward ROS-induced ROS mechanism by XO during IH.

Exposing PC12 cells to IH10 activates XO, which triggers a series of events that lead to more persistent ROS production by increasing NOX and calpain activity and causes an imbalance in the HIFs that favors transactivation of pro-oxidant enzymes over anti-oxidant ones. Increased ROS levels lead to breathing abnormalities and hypertension.

https://doi.org/10.1371/journal.pone.0119762.g007

Author Contributions

Conceived and designed the experiments: JN NRP GLS. Performed the experiments: DRV SAK NW VM. Analyzed the data: JN DRV SAK VM. Wrote the paper: JN NRP GLS.

References

  1. 1. Nieto FJ, Young TB, Lind BK, Shahar E, Samet JM, Redline S, et al. Association of sleep-disordered breathing, sleep apnea, and hypertension in a large community-based study. Sleep Heart Health Study. JAMA. 2000;283(14):1829–1836. pmid:10770144
  2. 2. Young T, Finn L, Peppard PE, Szklo-Coxe M, Austin D, Nieto FJ, et al. Sleep disordered breathing and mortality: eighteen-year follow-up of the Wisconsin sleep cohort. Sleep. 2008;31(8):1071–1078. pmid:18714778
  3. 3. Prabhakar NR, Semenza GL. Adaptive and maladaptive cardiorespiratory responses to continuous and intermittent hypoxia mediated by hypoxia-inducible factors 1 and 2. Physiol Rev. 2012;92(3):967–1003. pmid:22811423
  4. 4. Yuan G, Nanduri J, Khan S, Semenza GL, Prabhakar NR. Induction of HIF-1alpha expression by intermittent hypoxia: involvement of NADPH oxidase, Ca2+ signaling, prolyl hydroxylases, and mTOR. J Cell Physiol. 2008;217(3):674–685. pmid:18651560
  5. 5. Peng YJ, Yuan G, Ramakrishnan D, Sharma SD, Bosch-Marce M, Kumar GK, et al. Heterozygous HIF-1alpha deficiency impairs carotid body-mediated systemic responses and reactive oxygen species generation in mice exposed to intermittent hypoxia. J Physiol. 2006;577(Pt 2):705–716. pmid:16973705
  6. 6. Prabhakar NR, Kumar GK, Nanduri J. Intermittent hypoxia augments acute hypoxic sensing via HIF-mediated ROS. Respir Physiol Neurobiol. 2010;174(3):230–234. pmid:20804864
  7. 7. Yuan G, Khan SA, Luo W, Nanduri J, Semenza GL, Prabhakar NR. Hypoxia-inducible factor 1 mediates increased expression of NADPH oxidase-2 in response to intermittent hypoxia. J Cell Physiol. 2011;226(11):2925–233. pmid:21302291
  8. 8. Peng YJ, Nanduri J, Yuan G, Wang N, Deneris E, Pendyala S, et al. NADPH oxidase is required for the sensory plasticity of the carotid body by chronic intermittent hypoxia. J Neurosci. 2009;29(15):4903–4910. pmid:19369559
  9. 9. Block K, Gorin Y. Aiding and abetting roles of NOX oxidases in cellular transformation. Nat Rev Cancer. 2012;12(9):627–637. pmid:22918415
  10. 10. Leto TL, Morand S, Hurt D, Ueyama T. Targeting and regulation of reactive oxygen species generation by Nox family NADPH oxidases. Antioxid Redox Signal. 2009;11(10):2607–2619. pmid:19438290
  11. 11. Bedard K, Krause KH. The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev. 2007;87(1):245–313. pmid:17237347
  12. 12. Rotrosen D, Leto TL. Phosphorylation of neutrophil 47-kDa cytosolic oxidase factor. Translocation to membrane is associated with distinct phosphorylation events. J Biol Chem. 1990;265(32):19910–19915. pmid:2246268
  13. 13. Nanduri J, Vaddi DR, Khan SA, Wang N, Makerenko V, Prabhakar NR. Xanthine oxidase mediates hypoxia-inducible factor-2α degradation by intermittent hypoxia. PLoS One. 2013;8(10):e75838. pmid:24124516
  14. 14. Fontayne A, Dang PM, Gougerot-Pocidalo MA, El-Benna J. Phosphorylation of p47phox sites by PKC alpha, beta II, delta, and zeta: effect on binding to p22phox and on NADPH oxidase activation. Biochemistry. 2002;41(24):7743–7750. pmid:12056906
  15. 15. Yuan G, Adhikary G, McCormick AA, Holcroft JJ, Kumar GK, Prabhakar NR. Role of oxidative stress in intermittent hypoxia-induced immediate early gene activation in rat PC12 cells. J Physiol. 2004;557(Pt 3):773–783. pmid:15107478
  16. 16. Yuan G, Nanduri J, Bhasker CR, Semenza GL, Prabhakar NR. Ca2+/calmodulin kinase-dependent activation of hypoxia inducible factor 1 transcriptional activity in cells subjected to intermittent hypoxia. J Biol Chem. 2005;280(6):4321–4328. pmid:15569687
  17. 17. Khan SA, Nanduri J, Yuan G, Kinsman B, Kumar GK, Joseph J, et al. NADPH oxidase 2 mediates intermittent hypoxia-induced mitochondrial complex I inhibition: relevance to blood pressure changes in rats. Antioxid Redox Signal. 2011;14(4):533–542. pmid:20618070
  18. 18. Makarenko VV, Nanduri J, Raghuraman G, Fox AP, Gadalla MM, Kumar GK, et al. Endogenous H2S is required for Hypoxic Sensing by Carotid Body Glomus Cells. Am J Physiol Cell Physiol. 2012;301(9):C916–923. pmid:22744006
  19. 19. Cosentino-Gomes D, Rocco-Machado N, Meyer-Fernandes JR. Cell Signaling through Protein Kinase C Oxidation and Activation. Int J Mol Sci. 2012;13(9):10697–10721. pmid:23109817
  20. 20. Li JM, Shah AM. Intracellular localization and preassembly of the NADPH oxidase complex in cultured endothelial cells. J Biol Chem. 2002;277(22):19952–19960. pmid:11893732
  21. 21. Zhao X, Xu B, Bhattacharjee A, Oldfield CM, Wientjes FB, Feldman GM, et al. Protein kinase Cdelta regulates p67phox phosphorylation in human monocytes. J Leukoc Biol. 2005;77(3):414–420. pmid:15591124
  22. 22. Parks DA, Granger DN, Bulkley GB, Shah AK. Soybean trypsin inhibitor attenuates ischemic injury to the feline small intestine. Gastroenterology. 1985;89(1):6–12. pmid:4007413
  23. 23. Prabhakar NR, Kumar GK, Nanduri J, Semenza GL. ROS signaling in systemic and cellular responses to chronic intermittent hypoxia. Antioxid Redox Signal. 2007;9(9):1397–1403. pmid:17627465
  24. 24. Prabhakar NR, Peng YJ, Kumar GK, Pawar A. Altered carotid body function by intermittent hypoxia in neonates and adults: relevance to recurrent apneas. Respir Physiol Neurobiol. 2007;157(1):148–153. pmid:17317339
  25. 25. Nanduri J, Wang N, Yuan G, Khan SA, Souvannakitti D, Peng YJ, et al. Intermittent hypoxia degrades HIF-2alpha via calpains resulting in oxidative stress: implications for recurrent apnea-induced morbidities. Proc Natl Acad Sci U S A. 2009;106(4):1199–1204. pmid:19147445
  26. 26. Kumar GK, Rai V, Sharma SD, Ramakrishnan DP, Peng YJ, Souvannakitti D, et al. Chronic intermittent hypoxia induces hypoxia-evoked catecholamine efflux in adult rat adrenal medulla via oxidative stress. J Physiol. 2006;575(Pt 1):229–239. pmid:16777938
  27. 27. Raghuraman G, Rai V, Peng YJ, Prabhakar NR, Kumar GK. Pattern-specific sustained activation of tyrosine hydroxylase by intermittent hypoxia: role of reactive oxygen species-dependent downregulation of protein phosphatase 2A and upregulation of protein kinases. Antioxid Redox Signal. 2009;11(8):1777–1789. pmid:19335094
  28. 28. Sharma SD, Raghuraman G, Lee MS, Prabhakar NR, Kumar GK. Intermittent hypoxia activates peptidylglycine alpha-amidating monooxygenase in rat brain stem via reactive oxygen species-mediated proteolytic processing. J Appl Physiol. 2009;106(1):12–9. pmid:18818385
  29. 29. Troncoso Brindeiro CM, da Silva AQ, Allahdadi KJ, Youngblood V, Kanagy NL. Reactive oxygen species contribute to sleep apnea-induced hypertension in rats. Am J Physiol Heart Circ Physiol. 2007;293(5):H2971–2976. pmid:17766485
  30. 30. Peng YJ, Prabhakar NR. Reactive oxygen species in the plasticity of respiratory behavior elicited by chronic intermittent hypoxia. J Appl Physiol. 2003;94(6):2342–2349. pmid:12533494
  31. 31. Prabhakar NR, Peng YJ, Yuan G, Kumar GK. Reactive oxygen species facilitate oxygen sensing. Novartis Found Symp. 2006;272:95–9; discussion 100–5, 31–40. pmid:16686431
  32. 32. Kuri BA, Khan SA, Chan SA, Prabhakar NR, Smith CB. Increased secretory capacity of mouse adrenal chromaffin cells by chronic intermittent hypoxia: involvement of protein kinase C. J Physiol. 2007;584(Pt 1):313–319.
  33. 33. Souvannakitti D, Kumar GK, Fox A, Prabhakar NR. Neonatal intermittent hypoxia leads to long-lasting facilitation of acute hypoxia-evoked catecholamine secretion from rat chromaffin cells. J Neurophysiol. 2009;101(6):2837–2846. pmid:19339466
  34. 34. Dyugovskaya L, Lavie P, Lavie L. Increased adhesion molecules expression and production of reactive oxygen species in leukocytes of sleep apnea patients. Am J Respir Crit Care Med. 2002;165(7):934–939. pmid:11934717
  35. 35. Christou K, Markoulis N, Moulas AN, Pastaka C, Gourgoulianis KI. Reactive oxygen metabolites (ROMs) as an index of oxidative stress in obstructive sleep apnea patients. Sleep Breath. 2003;7(3):105–110. pmid:14569521
  36. 36. Suzuki YJ, Jain V, Park AM, Day RM. Oxidative stress and oxidant signaling in obstructive sleep apnea and associated cardiovascular diseases. Free Radic Biol Med. 2006;40(10):1683–1692. pmid:16678006
  37. 37. Grebe M, Eisele HJ, Weissmann N, Schaefer C, Tillmanns H, Seeger W, et al. Antioxidant vitamin C improves endothelial function in obstructive sleep apnea. Am J Respir Crit Care Med. 2006;173(8):897–901. pmid:16439717
  38. 38. Yuan G, Peng YJ, Reddy VD, Makarenko VV, Nanduri J, Khan SA, et al. Mutual antagonism between hypoxia-inducible factors 1α and 2α regulates oxygen sensing and cardio-respiratory homeostasis. Proc Natl Acad Sci U S A. 2013;110(19):E1788–1796. pmid:23610397