545
Views
0
CrossRef citations to date
0
Altmetric
Research Articles

Insights into H2O2-induced signaling in Jurkat cells from analysis of gene expression

ORCID Icon, ORCID Icon, ORCID Icon & ORCID Icon
Pages 666-676 | Received 23 Oct 2022, Accepted 01 Jan 2023, Published online: 11 Jan 2023

Abstract

Hydrogen peroxide (H2O2) is a ubiquitous oxidant produced in a regulated manner by various enzymes in mammalian cells. H2O2 reversibly oxidizes thiol groups of cysteine residues to mediate intracellular signaling. While examples of H2O2-dependent signaling have been reported, the exact molecular mechanism(s) of signaling and the pathways affected are not well understood. Here, the transcriptomic response of Jurkat T cells to H2O2 was investigated to determine global effects on gene expression. With a low H2O2 concentration (10 µM) that did not induce an oxidative stress response or cell death, extensive changes in gene expression occurred after 4 h (6803 differentially expressed genes). Of the genes with a greater then 2-fold change in expression, 85% were upregulated suggesting that in a physiological setting H2O2 predominantly activates gene expression. Pathway analysis identified gene expression signatures associated with FOXO and NTRK signaling. These signatures were associated with an overlapping set of transcriptional regulators. Overall, our results provide a snapshot of gene expression changes in response to H2O2, which, along with further studies, will lead to new insights into the specific pathways that are activated in response to endogenous production of H2O2, and the molecular mechanisms of H2O2 signaling.

Introduction

Hydrogen peroxide (H2O2) can mediate signal transduction by reversible oxidation of redox-sensitive cysteine residues in target proteins and is considered a major reactive oxygen species for redox regulation [Citation1–7]. H2O2 is a suitable second messenger as it is produced in a regulated manner, is small and can diffuse through cell membranes, and is specific in its reactions with thiols and transition metal centers [Citation4,Citation5]. While over 41 enzymes can produce H2O2 or its precursor O2.-, the two main sources of H2O2 are membrane-bound NADPH oxidase (NOX) enzymes and the mitochondrial electron transport chain [Citation6]. NOX enzymes are activated in response to a variety of signals including growth factors and cytokines. They transfer electrons from intracellular NAD(P)H across plasma or endosomal membranes to molecular oxygen-generating superoxide which is then converted to H2O2 [Citation8,Citation9]. The H2O2 enters the cytosol via passive diffusion or facilitated transport by aquaporins to induce a signaling response via direct or indirect protein oxidation. Mitochondrial ROS production is predominantly via reverse electron transport at complex I, and from reaction of oxygen with ubisemiquinone at complex III, generating superoxide (and subsequently H2O2) in either the mitochondrial intermembrane space or the mitochondrial matrix [Citation10].

H2O2 is a pleiotropic molecule that has been reported to affect multiple signaling pathways. For example, it is well documented that in cells stimulated with growth factors, such as epidermal growth factor (EGF), production of H2O2 results in oxidation of a catalytic cysteine residue of protein tyrosine phosphatases (PTPs) leading to their inactivation [Citation11,Citation12]. Inactivation of PTPs is required concurrent with the activation of protein tyrosine kinases (PTKs) for the propagation of an intracellular signal. Often oxidation of specific targets leads to alterations in phosphorylation cascades and subsequent activation of downstream transcription factors leading to changes in gene expression. For example, the MAPKKK ASK1 that activates either p38 or JNK is redox regulated [Citation13–15]. In response to H2O2 treatment, ASK1 forms multimeric disulfide-linked species which are essential for full JNK activation, nuclear translocation, and changes in transcription [Citation14].

There is an abundance of literature implicating H2O2 in cellular signaling, predominantly focusing on investigating specific signaling targets. In addition, microarrays and RNA-Seq have been used to provide a global overview of the transcriptomics response of budding and fission yeasts, and cultured human cells, to H2O2 [Citation16–29]. Collectively these studies support a role for H2O2 in controlling the expression of a range of genes, including p53 target genes, apoptosis genes and signal transduction genes. However, in many of these studies, it is unclear whether changes in gene expression are due to the direct impact of H2O2, because the concentrations used to induce cell death and/or the time point studied is many days after the addition of H2O2 to the cells.

Here, we have used a discovery approach to investigate gene expression changes in response to physiologically relevant concentrations of H2O2. The physiological concentration of H2O2 is estimated to lie between 1–100 nM and to rise transiently to 500–700 nM under oxidative signaling events [Citation30,Citation31]. The majority of studies investigating the transcriptomic response to H2O2 in mammalian cells have focused on conditions that induce a stress response and/or monitor the impact after several days. Here we investigate the early gene expression response to a bolus treatment with H2O2 at a concentration which modeling suggests will result in low nM intracellular H2O2 [Citation31].

We performed RNA-Seq analysis on RNA extracted from human Jurkat T cells four hours after treatment with 10 µM H2O2 and observed widespread changes in gene expression. Using Reactome Pathway Analysis, we identified several clusters of genes involved in intracellular signaling suggesting that H2O2 may play a role in activation of these pathways. Further analysis of these molecular signaling events may provide insight into the role of endogenously produced H2O2 in cell signaling.

Materials and methods

Cell culture and H2O2 treatment

Jurkat T lymphoma cells (TIB-152 clone E6-1 from ATCC) were maintained at a density of less than 1 × 106 cells/mL in RPMI 1640 (Gibco, Invitrogen) with sodium bicarbonate, 10% fetal bovine serum (FBS), and penicillin/streptomycin supplements (Sigma). Cells were maintained at 37 °C in a humidified atmosphere containing 5% CO2. Cell line identity was confirmed by STR analysis and cultures were regularly screened for mycoplasma contamination.

H2O2 concentration was determined spectrophotometrically (ϵ240 = 43.6 M−1 cm −1). For treatment with H2O2, cells were harvested by centrifugation and resuspended in PBS at a density of 1 × 106 cells/mL to prevent H2O2 interacting with components in RPMI media and to slow the reduction of oxidized proteins [Citation32]. Following treatment, cells were pelleted, washed in 1 mL PBS, resuspended in 1 mL TRIzol (Life Technologies), and stored at −80 °C.

Analysis of cell death

Jurkat cells were resuspended at a density of 1 × 106 cells/mL in 2 mL PBS and equilibrated to 37 °C/5% CO2 before treatment with 0, 10, 20 or 500 µM H2O2 per 106 cells. After incubation at 37 °C for 4 h, the cells were pelleted and resuspended in 2 mL RPMI containing 10% FBS. After 24 h at 37 °C/5% CO2, the cells were pelleted and then washed twice in 1 mL PBS. Subsequently, the cells were resuspended in 100 µL PBS with 4 µL of propidium iodide (PI) stain (50 µg/mL). The cells were incubated in the dark for 10 min before analysis by flow cytometry using the Guava® easyCyte™ 5HT HPL Benchtop flow cytometer (Merck). Data were analyzed using FlowJo™ v10.2 Software (BD Life Sciences).

RNA sequencing

RNA extractions were carried out using TRIzol according to the manufacturer’s instructions. RNA quality was assessed by running an RNA 6000 Nano chip on a 2100 Bioanalyzer (Agilent). TruSeq (Illumina) libraries for polyA + RNA-Seq were prepared from 250 ng RNA per sample using the TruSeq RNA Library Preparation Kit v2. The libraries were multiplexed (24 samples per pool) and sequenced to obtain 50 bp single-end reads on the Illumina HiSeq 3000 platform (UCLA Technology Center for Genomics & Bioinformatics). Library preparation and sequencing was performed in two batches, each with four biological replicates of control and treated cells.

Read mapping and differential expression analysis

The quality of the raw data was evaluated using FastQC [Citation33] and reads were mapped to human genome version GRCh38 using STAR v2.7.1a [Citation34], resulting in an overall mapping efficiency of >80%. The GENCODE hv35 was used as the annotation file and the reads per gene were counted using the HTseq package v0.6.1 [Citation35], with parameters -m union (recommended) -s no. Differential gene expression was performed with limma [Citation36,Citation37]. Genes with low expression in the samples (<160 reads across 16 samples) were excluded prior to differential expression analysis. Principal Component Analysis (PCA) and the heatmap.2 functions in the gplots package [Citation38] for R (v4.0.2) were used to visualize similarities and differences among the control and treated samples [Citation39]. The PCA plot of the samples showed that approximately 70% of the variation in the data was explained by H2O2 treatment, however, the remaining variation was a batch effect associated with the two separate library preparations and sequencing runs (Figure S1, Supplementary Material). For this reason, batch correction was completed using limma. Adjusted p-values were calculated using the Benjamini–Hochberg method [Citation40] and a value of 0.05 was set as threshold for statistical significance of differential gene expression. A volcano plot to visualize differential gene expression was generated using the R package EnrichedVolcano [Citation41].

Gene ontology and pathway analysis

Significantly differentially expressed gene ensemble IDs were converted to entrez IDs using the AnnotationDb package [Citation42] in R. Gene set analysis was performed using GOseq [Citation43] with Reactome pathways [Citation44] to correct for gene length, as well as via a non-length corrected hypergeometric model. The overlap between the output from GOSeq and the hypergeometric model suggested there was little or no length bias in the data. Therefore, gene set enrichment analysis (GSEA) [Citation45] was performed for Reactome pathways [Citation44] using the ReactomePA (v1.9.4) [Citation46] package in R. Of the 6803 DEGs, 3356 were suitable for pathway analysis (2780 were not annotated with a Reactome pathway ID and the remainder didn’t have a corresponding Entrez ID). The heatmap.2 function was used to plot the proportion of overlapping gene content between pathways that were found to be significantly enriched in the data.

RT-qPCR

Selected differentially expressed genes were analyzed by RT-qPCR. RNA (1 μg) was reverse transcribed to cDNA using SuperScript IV VILO Master Mix with ezDNAse enzyme (ThermoFisher) according to the manufacturer’s instructions. One-fifth of the volume of cDNA reaction mixture was used in the qPCR reaction containing 1 × SYBR Green Master Mix (ThermoFisher) in a 384-well-plate. Cycling conditions were as follows: 2 min at 50 °C and 2 min at 95 °C then 40 cycles of amplification at 95 °C for 15 s and 60 °C for 1 min. Primer specificity was initially tested using agarose gel electrophoresis and was subsequently monitored using melting curve analysis. RPL27 (the most stable of five candidate reference genes analyzed using Normfinder [Citation47]) was used as the reference gene and gene expression relative to RPL27 was calculated as 2–ΔCt [Citation48]. Primer information is provided in Table S1 (Supplementary Material).

Motif analysis using ChEA3 and TRAP

Two algorithms, TRAP (Transcription Factor Affinity Prediction) and ChEA3 (ChIP-X Enrichment Analysis 3) were used to identify transcription factors which may be responsible for regulating gene sets identified by pathway analysis. TRAP uses a biophysical model to predict which transcription factors have the highest affinity for a given set of DNA sequences [Citation49,Citation50]. TRAP (multiple sequences) at http://trap.molgen.mpg.de/cgi-bin/home.cgi was used. The 1000 bp upstream flanking region of each significantly differentially expressed gene in each gene set underlying identified Reactome Pathways was generated using BioMart and saved as a .fasta file. This file was submitted to TRAP and enriched binding motifs were predicted and compared with random promoters using the settings “jaspar_vertebrates” and “human_promoters.” Motifs for significant hits were obtained from JASPAR CORE Vertebrates clustering (https://jaspar.genereg.net/matrix-clusters/vertebrates/?detail=true). ChEA3 ranks TFs based on the overlap between gene input and annotated sets of TF targets in background databases (available at https://maayanlab.cloud/chea3/) [Citation51]. Gene symbols for each gene set to be analyzed were submitted. As a control, the sample function in R was used to generate a random set of 1050 human genes from the genes expressed in Jurkat cells. Transcriptional regulators identified that had low expression in the Jurkat cells (<160 reads across 16 samples) were removed from the results.

Statistical analysis

For analysis of cell death, changes in variables were analyzed by one-way ANOVA with Dunnett’s multiple comparison test assuming single pooled variance (GraphPad Prism v9.3.1). For analysis of qPCR data, changes in variables were analyzed by two-way ANOVA with Tukey’s posthoc test for multiple comparisons (GraphPad Prism v9.3.1). In both cases, differences were considered statistically significant at p < 0.05.

Results

Treatment with 10 µM H2O2 does not induce cell death

In Jurkat cells, H2O2 is reported to induce apoptotic or necroptotic cell death at concentrations of 25–200 µM and greater than 200 µM respectively [Citation52]. At 10 µM H2O2/106 cells/mL there was no significant cell death at 24 h, and only a small increase with 20 µM H2O2/106 cells/mL (). As expected, treatment with 500 µM H2O2/106 cells/mL killed almost all the cells. Therefore, treatment with 10 µM H2O2/106 cells/mL and harvesting after 4 h was used for the analysis of gene expression (referred to as 10 μM hereafter).

Figure 1. Cell death of Jurkat T cells after treatment with varying concentrations of H2O2. Cells were treated with H2O2 in PBS and after four hours returned to RPMI/FCS for a further 24 h before staining with PI and analysis by flow cytometry. U = cells that remained in RPMI/FCS for the full 28 h. ap < 0.05, bp < 0.0001 compared to U (n = 3 ± SD).

Figure 1. Cell death of Jurkat T cells after treatment with varying concentrations of H2O2. Cells were treated with H2O2 in PBS and after four hours returned to RPMI/FCS for a further 24 h before staining with PI and analysis by flow cytometry. U = cells that remained in RPMI/FCS for the full 28 h. ap < 0.05, bp < 0.0001 compared to U (n = 3 ± SD).

H2O2 treatment of Jurkat cells leads to changes in gene expression

RNA-Seq was performed on RNA isolated from control cells, and cells harvested 4 h after treatment with 10 μM H2O2, with 8 biological replicates for each. Of the ∼15,000 genes identified by RNA Seq, 6,803 were significantly differentially expressed genes (DEG, defined by padj < 0.05), the majority with low fold-change (). When considering the 1,050 DEG with greater than 2-fold change in expression (|log2 fold-change| (|log2FC|) >1), 86% had increased expression. There was limited induction of oxidative stress response genes (as defined by Desaint et al. [Citation20]) in our dataset (all |log2FC| < 0.3). Overall, this indicates that at a low concentration H2O2 predominantly induces gene expression.

Figure 2. Volcano plot of differentially expressed genes following H2O2 treatment. The black dots are DEG with |log2FC| >1 and the grey dots are DEG with –1 ≤ log2FC ≤ 1. The list of DEG is in Table S2, Supplementary Material.

Figure 2. Volcano plot of differentially expressed genes following H2O2 treatment. The black dots are DEG with |log2FC| >1 and the grey dots are DEG with –1 ≤ log2FC ≤ 1. The list of DEG is in Table S2, Supplementary Material.

Reactome pathway analysis identified gene sets which respond to H2O2

To identify pathways enriched in the list of DEG, and thereby provide insight into the biological impact of H2O2-mediated signaling, we used the Homo sapiens-focused Reactome Knowledgebase [Citation44]. Focusing on the 1050 genes with >2-fold change in expression, 391 were unique and able to be mapped in Reactome (Table S2, Supplementary Material). This analysis identified 21 Reactome pathways with a range of significance (Padj from 8.3 × 10−6 to 0.04) (). As expected from the analysis of cell death and expression of oxidative stress genes, none of the identified pathways were within the superpathways “Programmed Cell Death” or “DNA Repair,” confirming that 10 μM H2O2 induces a signaling response independent of a classic oxidative stress response. An additional 330 Reactome pathways were identified using the mappable genes from the full set of DEG (Table S4, Supplementary Material).

Figure 3. Enriched Reactome pathways in Jurkat T cells treated with H2O2. Pathway analysis was completed using genes with Padj < 0.05 and >2-fold change in expression. (A) Bar graph is ordered based on Padj values (highest to lowest), and sizes of bars indicate the number of DEGs within each pathway. Table S3 (Supplementary Material) contains more detailed information for each pathway. (B) Plot showing the overlap in gene content between each Reactome pathway regulated by H2O2. The five clusters are annotated as NTRK, UPR, HSP, FOXO, and COMP. The values in the plot are computed by the number of genes appearing in both pathways (A and B) as a proportion of the total size of either pathway A (upper diagonal) or pathway B (lower diagonal). The scale is from 1 (100% gene overlap between Reactome pathways) to 0 (0% gene overlap between Reactome pathways).

Figure 3. Enriched Reactome pathways in Jurkat T cells treated with H2O2. Pathway analysis was completed using genes with Padj < 0.05 and >2-fold change in expression. (A) Bar graph is ordered based on Padj values (highest to lowest), and sizes of bars indicate the number of DEGs within each pathway. Table S3 (Supplementary Material) contains more detailed information for each pathway. (B) Plot showing the overlap in gene content between each Reactome pathway regulated by H2O2. The five clusters are annotated as NTRK, UPR, HSP, FOXO, and COMP. The values in the plot are computed by the number of genes appearing in both pathways (A and B) as a proportion of the total size of either pathway A (upper diagonal) or pathway B (lower diagonal). The scale is from 1 (100% gene overlap between Reactome pathways) to 0 (0% gene overlap between Reactome pathways).

There is frequently gene membership overlap between Reactome pathways [Citation53]. Therefore, a plot showing the overlap in gene content between pathways was used to gain a more precise understanding of biological consequences of H2O2 signaling (). Of the 21 significant Reactome pathways, 17 were grouped into five clusters: NTRK signaling, FOXO signaling, unfolded protein response (UPR), heat shock response, and complement (Comp). Two Reactome pathways, “metallothioneins bind metals” and “IL4 & IL14 signaling,” exhibited only minimal gene-sharing with other enriched pathways.

FOxO signaling and NTRK signaling are H2O2-dependent

Reactome pathways that respond to H2O2 were identified, however, there was a caveat to this result. The H2O2-treated RNA samples were harvested from Jurkat cells incubated in PBS for four hours. Therefore, the pathways identified may be H2O2-dependent and/or PBS-dependent. To determine which responses were H2O2 dependent, and to gain insights into the pattern of activation, RT-qPCR of two genes from each of the NTRK, FOXO and UPR clusters was carried out over a time-course at 0, 5, 10, and 20 μM H2O2.

Expression of the NTRK cluster genes FOS and FOSB was H2O2 dependent, with expression increasing over time in response to 10 and 20 μM H2O2 (). The FOXO cluster genes BTG1 and CITED2 showed similar H2O2-dependent increases over time at 10 and 20 μM H2O2 (). There was no change in the expression of these four genes in the absence of H2O2. In contrast, the expression of the UPR cluster genes BIP and CHOP was not dependent on H2O2, confirming that the UPR is activated due to culture in PBS (). Interestingly, 20 μM H2O2 appeared to protect the cells from PBS-induced UPR activation. Overall, these results confirm that H2O2 induces dose- and time-dependent changes in gene expression consistent with the activation of FOXO and NTRK signaling.

Figure 4. RT-qPCR showing changes in expression of selected genes from Reactome Pathways. (A) NTRK cluster genes FOS and FOSB. (B) FOXO cluster genes BTG1 and CITED2. (C) UPR cluster genes BIP and CHOP. For each concentration of H2O2 a two-way ANOVA with multiple comparisons testing was used to determine statistical significance. Note that 0 h for each gene is the same data. ap < 0.05, bp < 0.01 and cp < 0.0001 (n = 3–12 ± SD). The results at 4 h with 10 μM H2O2 were consistent with the RNA-Seq data (Table S5, Supplementary Material).

Figure 4. RT-qPCR showing changes in expression of selected genes from Reactome Pathways. (A) NTRK cluster genes FOS and FOSB. (B) FOXO cluster genes BTG1 and CITED2. (C) UPR cluster genes BIP and CHOP. For each concentration of H2O2 a two-way ANOVA with multiple comparisons testing was used to determine statistical significance. Note that 0 h for each gene is the same data. ap < 0.05, bp < 0.01 and cp < 0.0001 (n = 3–12 ± SD). The results at 4 h with 10 μM H2O2 were consistent with the RNA-Seq data (Table S5, Supplementary Material).

Transcriptional regulator enrichment analysis

Having confirmed that the expression of genes in the NTRK and FOXO clusters was H2O2-dependent, we next hypothesized that the DEG within each cluster would share common upstream regulatory elements. Identifying these elements could provide new insights into the mechanism(s) by which H2O2 regulates gene expression. The DEG within the Reactome pathways contributing to the NTRK and FOXO clusters (Table S6, Supplementary Material) were first analyzed with TRAP. TRAP uses the JASPAR database to predict binding motifs based on the 1000 bp upstream flanking region of submitted genes. Analysis with TRAP identified 11 and 15 transcriptional regulators for the FOXO and NTRK clusters respectively ( and ). We next obtained the motifs for these significant hits. Five JASPAR motifs were identified for the FOXO DEG and eight for the NTRK DEG ( and ). JASPAR motifs 21 and 27 (Figure S2, Supplementary Material) accounted for the majority of identified transcription regulators in both gene sets. We then compared these results with those obtained from ChEA3, which ranks transcriptional regulators associated with specific gene sets based on ChIPSeq and RNA-Seq data. The six databases within ChEA3 vary in the number of transcriptional regulators represented (118–1628, compared to 841 in JASPAR). Therefore, for each transcriptional regulator identified with TRAP, we extracted the Padj values across the ChEA3 databases ( and ). This allowed us to determine whether putative H2O2 responsive transcriptional regulators identified by TRAP could be confirmed with a different analysis method. Reassuringly all significant hits from TRAP were also significant in at least 1, and up to 6, ChEA3 databases. This suggests the identified DNA binding motifs and associated transcriptional regulators are bone-fide targets of H2O2.

Table 1. Potential transcriptional regulators of genes in the FOXO cluster.

Table 2. Potential transcriptional regulators of genes in the NTRK cluster.

A small number of transcriptional regulators identified in the ChEA3 analysis of the NTRK and FOXO gene sets were not present in JASPAR (Tables S9 and 10, Supplementary Material). In particular, HBP1 and ZBTB10 were significantly enriched from the FOXO gene set, with CSRNP1 and GTF2I significantly enriched from the NTRK gene set, suggesting that these transcriptional regulators may also be responsive to H2O2.

We also performed TRAP and ChEA3 with the 1050 DEG with fold-change >2, and with a control set of 1050 randomly selected genes. Of the 1050 genes with fold-change >2 the 1000 bp upstream sequence could be mapped for 904 genes for use in TRAP, and 791 had a gene symbol for use in ChEA3. For the 1050 control genes, the 1000 bp upstream sequence could be mapped for all. TRAP analysis of DEG with fold-change >2 identified 32 transcriptional regulators that were also significant in one or more of the ChEA3 databases (Table S11, Supplementary Material). This included 14 that were identified in the analysis of the gene sets associated with the FOXO and/or NTRK Reactome pathway clusters. The 32 transcriptional regulators were represented by 23 JAPSAR clusters, including clusters 21 (16%) and 27 (9%), consistent with the FOXO and NTRK data. Using the control gene set, 38 transcriptional regulators had a significant combined P value in the TRAP analysis of the control set of genes. However, none were significant in any of the ChEA3 databases (Table S12, Supplementary Material). The combined use of the TRAP and ChEA3 data is therefore a strong approach for identifying transcriptional regulators sensitive to H2O2 or other stimuli.

Discussion

H2O2 is produced in a regulated manner in response to a wide range of stimuli. Here we mimicked stimulus-induced H2O2 production by a plasma membrane NADPH oxidase. By carrying out RNA-Seq followed by pathway and transcription factor motif analyses we have characterized the effect of a physiologically relevant dose of H2O2 on gene expression, and gained insights into the pathways and transcription factors mediating gene expression changes.

Only a small number of studies have investigated the short-term gene expression response to sub-lethal H2O2 concentrations. An early microarray study identified 828-1097 DEG 5 h after treatment of MCF-7a, MCF-7b or MRC-9 cells with 100 μM H2O2, with many of these genes related to genotoxic stress [Citation20]. Treatment of PIG1 cells with 100 μM H2O2 resulted in only 661 DEG after 6 h [Citation26], but no analysis was presented on the functional classes of these genes. In HUVEC treated with 10 μM H2O2, no DEG were identified up to 900 min post-treatment, although 300 μM H2O2 resulted in maximum DEG (3540) at 4.5 h [Citation29]. In contrast, we identified 6803 DEG in Jurkat cells treated with a 10 µM bolus dose of H2O2. Our greater sensitivity is most likely due to the strong statistical power from analyzing eight biological replicates compared to three or less in other studies [Citation54]. The majority of DEG (85%) have less than 2-fold change in expression and within this set, there were similar numbers of up and down-regulated genes. For the DEG with a robust >2-fold change in expression, the majority were upregulated. This result suggests that a physiologically relevant concentration of H2O2 predominantly activates gene expression. There are at least two possible mechanisms by which H2O2 alters gene expression, and both are likely responsible for our observations. Firstly, H2O2 is a well-described activator of signaling cascades leading to transcription factor activation [Citation2,Citation6,Citation55]. Second, a low dose of H2O2 decreases DNA methylation 4 h after treatment [Citation56], possibly due to inhibition of DNA methyltransferases during DNA replication [Citation57].

To identify signaling cascades that may be activated by H2O2 we performed pathway analysis focusing on genes with >2-fold change in expression. Two pathways were convincingly identified, FOXO signaling and NTRK signaling. There are four mammalian FOXO transcription factors, FOXO1, FOXO3, FOXO4, and FOXO6, which are characterized by the presence of a “forkhead” DNA binding domain [Citation58]. In response to various stimuli, these transcription factors induce the expression of genes involved in a variety of cellular processes (apoptosis, DNA damage repair, cell cycle) [Citation58,Citation59]. There is a myriad of evidence that FOXO transcription factors are regulated indirectly by ROS, by modulation of post-translational modifications such as phosphorylation and acetylation [Citation60]. For example, Essers et al. showed that treatment of A14 cells with 20–200 μM H2O2 led to JNK-dependent phosphorylation of FOXO4, nuclear translocation, and transcriptional activation [Citation61]. In the present study, the FOXO transcriptional response was induced using a lower dose of H2O2. There were two interesting observations: FOXO3 and FOXO4 were upregulated in response to H2O2 treatment, and TRAP and ChEA3 predicted that FOXO3 was a transcriptional regulator controlling the expression of genes from the FOXO cluster. Previously, it has been shown that FOXO3 is able to upregulate FOXO1 and FOXO4 expression in a positive feedback loop [Citation62]. It is possible that FOXO3 is the key transcription factor involved in activating the FOXO transcriptional response to a physiologically relevant dose of H2O2.

H2O2 also specifically activated a gene expression signature consistent with NTRK pathway activation. The NTRK genes encode transmembrane tyrosine kinase receptors, TRKA, TRKB, and TRKC [Citation63]. Upon binding of their ligands (NGF, BDNF, and NT-3) the TRKA, TRKB, and TRKC receptors dimerize and initiate phosphorylation cascades via PI3K, Ras, and PLCγ [Citation64–66]. Our data suggest that one or more of these branches of the NTRK pathway are activated by a low, physiologically relevant dose of H2O2. Many of the upregulated genes within the NTRK signaling pathway identified in this analysis were MAP kinases, and the transcription factors FOS, FOSB, JUND, and JUNB were also upregulated. It is well-established that heterodimeric AP-1 complexes, made up of FOS and JUN proteins, are both directly and indirectly regulated by concentrations of H2O2 that would induce oxidative stress [Citation67–69]. Both TRAP and ChEA3 predicted that the genes in the NTRK cluster may be activated by RELA/REL, subunits of the NF-κB transcription factor [Citation70]. NF-κB was first found to be regulated by H2O2 in 1991 [Citation71], however, the exact molecular mechanisms are still unclear. A recent study highlighted a link between NTRK signaling and NF-κB activation. Treatment of the KM12SM cell line with NTRK inhibitors, LOXO-101, entrectinib, and dovitinib, led to reduced protein expression of phosphorylated NF-κB [Citation72]. Therefore, it is possible that the activation of NF-κB could be leading to regulation of the NTRK signaling pathway.

Taken together, these results suggest that at physiologically relevant levels, H2O2 acts as a bona fide signaling molecule to stimulate widespread changes in gene expression, with the FOXO and NTRK signaling pathways identified as specific targets. It was perhaps surprising that additional pathways were not identified. However, pathway identification is limited to the extent to which the DEG can be mapped. Indeed only 50% of all DEG, and 37% of DEG with fold-change greater than 2, were available for analysis by Reactome Pathways. Thus, a significant amount of information is lost, and there will be a bias toward well-studied pathways. It is highly likely that H2O2 activates additional pathways, and as the number of mappable genes increases, reanalysis of the DEG identified in this study will be informative.

Supplemental material

Supplemental Material

Download Zip (819.4 KB)

Disclosure statement

No potential conflict of interest was reported by the author(s).

Data availability statement

The RNA-Seq data have been deposited in NCBI’s Gene Expression Omnibus [1] and are accessible through GEO Series accession number GSE206509 (https://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE206509).

Additional information

Funding

This work was supported by an University of Otago Doctoral scholarship and a Freemasons NZ postgraduate scholarship to MFT, and a Royal Society of New Zealand Marsden Fund Grant [17-UOO-086] to MBH.

References

  • Barrett T, Wilhite SE, Ledoux P, et al. NCBI GEO: archive for functional genomics data sets – update. Nucleic Acids Res. 2013;41(Database issue):D991–995.
  • Marinho HS, Real C, Cyrne L, et al. Hydrogen peroxide sensing, signaling and regulation of transcription factors. Redox Biol. 2014;2:535–562.
  • García-Santamarina S, Boronat S, Hidalgo E. Reversible cysteine oxidation in hydrogen peroxide sensing and signal transduction. Biochemistry. 2014;53(16):2560–2580.
  • Sies H. Role of metabolic H2O2 generation: redox signaling and oxidative stress. J Biol Chem. 2014;289(13):8735–8741.
  • Winterbourn CC. Biological production, detection, and fate of hydrogen peroxide. Antioxid Redox Signal. 2018;29(6):541–551.
  • Sies H, Jones DP. Reactive oxygen species (ROS) as pleiotropic physiological signaling agents. Nat Rev Mol Cell Biol. 2020;21(7):363–383.
  • Sies H, Belousov VV, Chandel NS, et al. Defining roles of specific reactive oxygen species (ROS) in cell biology and physiology. Nat Rev Mol Cell Biol. 2022;23(7):499–515.
  • Oakley FD, Abbott D, Li Q, et al. Signaling components of redox active endosomes: the redoxosomes. Antioxid Redox Signal. 2009;11(6):1313–1333.
  • Begum R, Thota S, Abdulkadir A, et al. NADPH oxidase family proteins: signaling dynamics to disease management. Cell Mol Immunol. 2022;19(6):660–686.
  • Murphy MP. How mitochondria produce reactive oxygen species. Biochem J. 2009;417(1):1–13.
  • Sundaresan M, Yu ZX, Ferrans VJ, et al. Requirement for generation of H2O2 for platelet-derived growth factor signal transduction. Science. 1995;270(5234):296–299.
  • Bae YS, Kang SW, Seo MS, et al. Epidermal growth factor (EGF)-induced generation of hydrogen peroxide. Role in EGF receptor-mediated tyrosine phosphorylation. J Biol Chem. 1997;272(1):217–221.
  • Ichijo H, Nishida E, Irie K, et al. Induction of apoptosis by ASK1, a mammalian MAPKKK that activates SAPK/JNK and p38 signaling pathways. Science. 1997;275(5296):90–94.
  • Nadeau PJ, Charette SJ, Toledano MB, et al. Disulfide bond-mediated multimerization of Ask1 and its reduction by thioredoxin-1 regulate H2O2-induced c-Jun NH2-terminal kinase activation and apoptosis. Mol Biol Cell. 2007;18(10):3903–3913.
  • Jarvis RM, Hughes SM, Ledgerwood EC. Peroxiredoxin 1 functions as a signal peroxidase to receive, transduce, and transmit peroxide signals in mammalian cells. Free Radic Biol Med. 2012;53(7):1522–1530.
  • Gasch AP, Spellman PT, Kao CM, et al. Genomic expression programs in the response of yeast cells to environmental changes. Mol Biol Cell. 2000;11(12):4241–4257.
  • Chuang YY, Chen Y, Gadisetti Chandramouli VR, et al. Gene expression after treatment with hydrogen peroxide, menadione, or t-butyl hydroperoxide in breast cancer cells. Cancer Res. 2002;62(21):6246–6254.
  • Chen D, Toone WM, Mata J, et al. Global transcriptional responses of fission yeast to environmental stress. Mol Biol Cell. 2003;14(1):214–229.
  • Murray JI, Whitfield ML, Trinklein ND, et al. Diverse and specific gene expression responses to stresses in cultured human cells. Mol Biol Cell. 2004;15(5):2361–2374.
  • Desaint S, Luriau S, Aude JC, et al. Mammalian antioxidant defenses are not inducible by H2O2. J Biol Chem. 2004;279(30):31157–31163.
  • Vandenbroucke K, Robbens S, Vandepoele K, et al. Hydrogen peroxide-induced gene expression across kingdoms: a comparative analysis. Mol Biol Evol. 2008;25(3):507–516.
  • Fomenko DE, Koc A, Agisheva N, et al. Thiol peroxidases mediate specific genome-wide regulation of gene expression in response to hydrogen peroxide. Proc Natl Acad Sci USA. 2011;108(7):2729–2734.
  • Purcell M, Kruger A, Tainsky MA. Gene expression profiling of replicative and induced senescence. Cell Cycle. 2014;13(24):3927–3937.
  • Wang R, Wei B, Wei J, et al. Caspase-related apoptosis genes in gliomas by RNA-seq and bioinformatics analysis. J Clin Neurosci. 2016;33:259–263.
  • Barandalla M, Shi H, Xiao H, et al. Global gene expression profiling and senescence biomarker analysis of hESC exposed to H2O2 induced non-cytotoxic oxidative stress. Stem Cell Res Ther. 2017;8(1):160.
  • Sastry KS, Naeem H, Mokrab Y, et al. RNA-seq reveals dysregulation of Novel melanocyte genes upon oxidative stress: implications in vitiligo pathogenesis. Oxid Med Cell Longev. 2019;2019:2841814.
  • Huang X, He C, Hua X, et al. Oxidative stress induces monocyte-to-myofibroblast transdifferentiation through p38 in pancreatic ductal adenocarcinoma. Clin Transl Med. 2020;10(2):e41.
  • Chen X, Wang X, Shen M, et al. Combined RNA-seq and molecular biology technology revealed the protective effect of Cyclocarya paliurus polysaccharide on H2O2-induced oxidative damage in L02 cells thought regulating mitochondrial function, oxidative stress and PI3K/Akt and MAPK signaling pathways. Food Res Int. 2022;155:111080.
  • Müller N, Warwick T, Noack K, et al. Reactive oxygen species differentially modulate the metabolic and transcriptomic response of endothelial cells. Antioxidants. 2022;11(2):434.
  • Stone JR, Yang S. Hydrogen peroxide: a signaling messenger. Antioxid Redox Signal. 2006;8(3–4):243–270.
  • Antunes F, Brito PM. Quantitative biology of hydrogen peroxide signaling. Redox Biol. 2017;13:1–7.
  • Pace PE, Peskin AV, Konigstorfer A, et al. Peroxiredoxin interaction with the cytoskeletal-regulatory protein CRMP2: investigation of a putative redox relay. Free Radic Biol Med. 2018;129:383–393.
  • Andrews S. 2010. FastQC: a quality control tool for high throughput sequence data. http://www.bioinformatics.babraham.ac.uk/projects/fastqc/.
  • Dobin A, Davis CA, Schlesinger F, et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics. 2013;29(1):15–21.
  • Anders S, Pyl PT, Huber W. HTSeq–a Python framework to work with high-throughput sequencing data. Bioinformatics. 2015;31(2):166–169.
  • Law CW, Chen Y, Shi W, et al. Voom: precision weights unlock linear model analysis tools for RNA-seq read counts. Genome Biol. 2014;15(2):R29.
  • Ritchie ME, Phipson B, Wu D, et al. Limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 2015;43(7):e47.
  • Warnes GR, Bolker B, Bonebakker L, et al. 2009. Various R programming tools for plotting data. https://CRAN.R-project.org/package=gplots.
  • R Core Team. 2017. A language and environment for statistical computing. In R Foundation for Statistical Computing, Vienna, Austria. https://www.R-project.org/
  • Benjamini Y, Hochberg Y. Controlling the false discovery rate – a practical and powerful approach to multiple testing. J. R. Stat. Soc. B. 1995;57(1):289–300.
  • Blighe K, Rana S, Lewis M. 2021. EnhancedVolcano: publication-ready volcano plots with enhanced colouring and labeling. https://github.com/kevinblighe/EnhancedVolcano.
  • Pagès H, Carlson M, Falcon S, et al. 2021. AnnotationDbi: Manipulation of SQLite-based annotations in Bioconductor., R package version 1.54.1 Ed., https://bioconductor.org/packages/AnnotationDbi.
  • Young MD, Wakefield MJ, Smyth GK, et al. Gene ontology analysis for RNA-seq: accounting for selection bias. Genome Biol. 2010;11(2):R14.
  • Gillespie M, Jassal B, Stephan R, et al. The reactome pathway knowledgebase 2022. Nucleic Acids Res. 2022;50(D1):D687–D692.
  • Subramanian A, Tamayo P, Mootha VK, et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc Natl Acad Sci USA. 2005;102(43):15545–15550.
  • Yu G, He QY. ReactomePA: an R/bioconductor package for reactome pathway analysis and visualization. Mol Biosyst. 2016;12(2):477–479.
  • Andersen CL, Jensen JL, Orntoft TF. Normalization of real-time quantitative reverse transcription-PCR data: a model-based variance estimation approach to identify genes suited for normalization, applied to bladder and Colon cancer data sets. Cancer Res. 2004;64(15):5245–5250.
  • Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2–ΔΔCT method. Methods. 2001;25(4):402–408.
  • Roider HG, Kanhere A, Manke T, et al. Predicting transcription factor affinities to DNA from a biophysical model. Bioinformatics. 2007;23(2):134–141.
  • Thomas-Chollier M, Hufton A, Heinig M, et al. Transcription factor binding predictions using TRAP for the analysis of ChIP-seq data and regulatory SNPs. Nat Protoc. 2011;6(12):1860–1869.
  • Keenan AB, Torre D, Lachmann A, et al. ChEA3: transcription factor enrichment analysis by orthogonal omics integration. Nucleic Acids Res. 2019;47(W1):W212–W224.
  • Hampton MB, Orrenius S. Dual regulation of caspase activity by hydrogen peroxide: implications for apoptosis. FEBS Lett. 1997;414(3):552–556.
  • Pita-Juárez Y, Altschuler G, Kariotis S, et al. The pathway coexpression network: revealing pathway relationships. PLoS Comput Biol. 2018;14(3):e1006042.
  • Trostle AJ, Wang J, Li L, et al. Most high throughput expression data sets are underpowered. BioRxiv. 2022. DOI: 10.1101/2022.08.03.502688
  • Lennicke C, Cocheme HM. Redox metabolism: ROS as specific molecular regulators of cell signaling and function. Mol Cell. 2021;81(18):3691–3707.
  • Seddon AR, Liau Y, Pace PE, et al. Genome-wide impact of hydrogen peroxide on maintenance DNA methylation in replicating cells. Epigenetics Chromatin. 2021;14:17.
  • O’Connor KM, Das AB, Winterbourn CC, et al. Inhibition of DNA methylation in proliferating human lymphoma cells by immune cell oxidants. J Biol Chem. 2020;295(23):7839–7848.
  • Brown AK, Webb AE. Regulation of FOXO factors in mammalian cells. Curr Top Dev Biol. 2018;127:165–192.
  • Huang H, Tindall DJ. Dynamic FoxO transcription factors. J Cell Sci. 2007;120(Pt 15):2479–2487.
  • Klotz LO, Sanchez-Ramos C, Prieto-Arroyo I, et al. Redox regulation of FoxO transcription factors. Redox Biol. 2015;6:51–72.
  • Essers MAG, Weijzen S, de Vries-Smits AMM, et al. FOXO transcription factor activation by oxidative stress mediated by the small GTPase Ral and JNK. EMBO J. 2004;23(24):4802–4812.
  • Essaghir A, Dif N, Marbehant CY, et al. The transcription of FOXO genes is stimulated by FOXO3 and repressed by growth factors. J Biol Chem. 2009;284(16):10334–10342.
  • Hechtman JF. NTRK insights: best practices for pathologists. Mod Pathol. 2022;35(3):298–305.
  • Wang H, Wang R, Thrimawithana T, et al. The nerve growth factor signaling and its potential as therapeutic target for glaucoma. Biomed Res Int. 2014;2014:759473.
  • Vaishnavi A, Le AT, Doebele RC. TRKing down an old oncogene in a new era of targeted therapy. Cancer Discov. 2015;5(1):25–34.
  • Lange AM, Lo HW. Inhibiting TRK proteins in clinical cancer therapy. Cancers. 2018;10(4):105.
  • Xanthoudakis S, Miao G, Wang F, et al. Redox activation of Fos-Jun DNA binding activity is mediated by a DNA repair enzyme. EMBO J. 1992;11(9):3323–3335.
  • Beiqing L, Chen M, Whisler RL. Sublethal levels of oxidative stress stimulate transcriptional activation of c-jun and suppress IL-2 promoter activation in Jurkat T cells. J. Immunol. 1996;157(1):160–169.
  • Klatt P, Molina EP, De Lacoba MG, et al. Redox regulation of c-Jun DNA binding by reversible S-glutathiolation. FASEB J. 1999;13(12):1481–1490.
  • Ghosh S, May MJ, Kopp EB. NF-κB and Rel proteins: evolutionarily conserved mediators of immune responses. Annu Rev Immunol. 1998;16:225–260.
  • Schreck R, Rieber P, Baeuerle PA. Reactive oxygen intermediates as apparently widely used messengers in the activation of the NF-κB transcription factor and HIV-1. EMBO J. 1991;10(8):2247–2258.
  • Sohn SH, Sul HJ, Kim B, et al. TRK inhibitors block NFKB and induce NRF2 in TRK fusion-positive colon cancer. J Cancer. 2021;12(21):6356–6362.