ABSTRACT Cucurbits are cultivated worldwide in regions with high concentrations of arsenic (As), a hazardous metalloid, affecting produce quality and increasing the consumer exposure. Cucurbitacins are herbivore‐deterrent secondary metabolites that contribute to the plant defense response. The impact of As exposure on phenotypic and metabolic traits has not been studied in members of the Cucurbitaceae family, such as squash ( Cucurbita pepo L.). To comprehend the effects of As on the root system of C. pepo , we assessed phenotype, cucurbitacin content, and transcriptome under low and high As concentrations. We report that at low dosages, cucurbitacins are decreased, while growth is not significantly affected. Conversely, high dosages impact growth and development altering root phenotype but cucurbitacin content is not significantly different from untreated plants. Furthermore, gene ontology enrichment on results of the RNA‐seq analysis indicate that high dosages of As affect cellular regulatory processes, with genes related to glutathione metabolism being of the most upregulated. Additionally, an in‐depth analysis of orthologs members of the heavy metal–associated (HMA)‐domain superfamily and As‐related transporters suggest a dosage‐dependent participation of key members. WGCNA analysis reveals As‐specific gene co‐expression modules, indicating that low As levels induce adaptive responses in energy and allantoin metabolism, while higher levels trigger intensified oxidative stress responses, including upregulation of MYB transcription factors and heat shock proteins, which may support tolerance to the metalloid. Overall, As influences the root system physiology and metabolism in a concentration‐specific manner, highlighting key defense systems and genes involved in C. pepo response to As exposure. Keywords: arsenic, Cucurbita pepo , Cucurbitacins, plant defense, stress mechanisms 1. Introduction Cucurbitaceae species are major crops. Among them, squash ( Cucurbita pepo L.) was valued at approximately $215 million US dollars in 2023 (USDA, 2023) due to its fruit, flowers, seeds, and the production of its bioactive oil, which is used in traditional medicine systems worldwide (Almeida et al. [36]2023; Sharma et al. [37]2024). In addition, C. pepo is a representative species to study gene duplication events on this family of crops (Montero‐Pau et al. [38]2018). Arsenic (As) is a nonessential hazardous metalloid for life, with toxicity levels determined by its oxidation state (El‐Ghiaty and El‐Kadi [39]2023; Preetha et al. [40]2023). As (V) is the most abundant form (Biswas and Ganesan [41]2024), particularly in aerobic soils where cucurbits are cultivated, followed by As (III), which is 60 times more toxic (Chen et al. [42]2021). In agriculture, the anthropogenic sources of As include pesticides and irrigation water (Mukherjee et al. [43]2024), whereas the geogenic incidence resides in important agricultural places (Zhang et al. [44]2021). For example, Mexico is an important producer of cucurbits, but As concentrations can reach naturally up to 35 times the World Health Organization's (WHO) maximum permissible recommendation of 10 μg L^−1 in water (Mahlknecht et al. [45]2023; Osuna‐Martínez et al. [46]2021). Cucurbitacins are a diverse group of highly oxygenated triterpenoid characteristic of species in the Cucurbitaceae family. Cucurbitacins are categorized into 17 groups (A‐T) where cucurbitacins B, D, E, and I are the most widespread in the Cucurbitaceae family (Almeida et al. [47]2022; Liu et al. [48]2024). Besides their biological function as defense compounds against herbivory and infectious microorganisms (Dong et al. [49]2021), their production is also highly valued for medicinal purposes (Maja et al. [50]2022). Interestingly, abiotic stress, for example, cold and drought, induce cucurbitacin biosynthesis in Lagenaria siceraria , Cucumis sativus, and Luffa acutangula , upregulating genes in the cucurbitacin‐biosynthesis pathway (Liu et al. [51]2019; Mkhize et al. [52]2023; Shang et al. [53]2014). Abiotic stresses including nutrient deficiency and heavy metals have been shown to influence the primary metabolome of cucurbits (Lee et al. [54]2024; Sun et al. [55]2020; Zhao et al. [56]2016). However, the targeted effects of metalloids (e.g., As) on the metabolome, cucurbitacin content, and the cucurbitacin biosynthesis pathway have not been investigated in this plant family (Martínez‐Castillo et al. [57]2022). As exposure in plants results in evident phenotypic effects such as reduced size, decreased biomass, and lower growth ratio (Nabi et al. [58]2021), which are linked to physiological and molecular changes including oxidative stress and protein folding disruption caused by the depletion of antioxidants, for example, glutathione (GSH) and phytochelatins (Rajendran et al. [59]2024). Plants have evolved specialized proteins to regulate the uptake, translocation, and homeostasis of heavy metals (Li et al. [60]2020), and it would follow the same for metalloids as some are essential for plant growth (Kesawat et al. [61]2020). HMA‐domain‐containing proteins share a conserved 30‐amino acid domain containing a crucial CxxC, where two cysteine residues facilitate the binding of essential and nonessential metals such as copper (Cu) and cadmium (Cd) (Takahashi et al. [62]2012; Wong et al. [63]2009). While As has a high affinity for cysteine‐rich peptides (Vergara‐Gerónimo et al. [64]2021), the role of the HMA superfamily in response to As remains poorly understood. In a previous study, we identified that members of the P[1B]‐ATPase clade, a subfamily of HMA‐domain‐containing proteins, differentially expressed when C. pepo was grown in the presence of As (Flores‐Iga et al. [65]2023), but other members of the HMA superfamily, such as heavy metal isoprenylated plant proteins (HIPP), were not studied. Moreover, heavy metalloid uptake, translocation, and extrusion are key mechanism for tolerance that warrant further study in Cucurbits, especially regarding As. For instance, Nodulin 26‐like transporters are recognized to participate in As extrusion and translocation, and ABCC transporters participate in the subcellular sequestration of As in vacuoles (Kamiya et al. [66]2009; Song et al. [67]2010). In this paper, we mechanistically explored the effects of As exposure on the phenotype and cucurbitacin content in the root system of C. pepo , and employed RNA‐seq to gain further insights into the molecular responses of C. pepo roots to As. While at low As concentrations the plant growth is not affected, cucurbitacin concentration decreases, which might negatively impact herbivory defense. Conversely, at high As concentrations, growth is diminished but cucurbitacin production is induced along with activation of oxidative and abiotic‐stress‐coping mechanisms in a concentration‐dependent manner. Additionally, we curated orthologs of the HMA‐domain‐containing superfamily and related As transporters to further explore characterized mechanisms employed to cope with As stress in cucurbits. Our examination of the underlying effects of As in the primary root system of C. pepo suggested preliminary insights that could eventually contribute to the development of healthier Cucurbitaceae crops in As‐contaminated environments. 2. Methods 2.1. C. pepo Growth and Root Phenotype Cucurbita pepo var. golden glory seeds were used, and four subsequent steps were employed for seed sterilization. First was a 20‐min wash in 1% Mild Cream Soap (ABENA, Denmark), then a 1‐min rinse with 70% ethanol with a further 15‐min soak in 1.5% NaClO. Finally, the seeds were rinsed with sterile water and placed on a plate for surface drying. Because As(V) is the most prevalent form of As present in aerobic soils (Panda et al. [68]2010; Khalid et al. [69]2017) in which C. pepo would be cultivated, Murashige and Skoog media, with 0, 50, 100, and 200 μM of As(V) ion from Na[2]HAsO[4] (Sigma‐Aldrich, St Louis, MO, USA), was used for C. pepo growth in square culture vessels. Growth conditions were light at 79 μE, 25°C, 80% humidity, and 16‐h photoperiod in Fitotron type SGC120‐H (Weiss Technik UK Ltd., Epinal Way, Loughborough, UK). Root phenotypes including length, area, diameter, volume and number of tips, forks, and crossing were analyzed in WinRhizo ([70]http://regentinstruments.com). 2.2. Metabolic Profile Analysis by LC–MS Frozen powdered materials (100 mg) of 2‐week‐old C. pepo var. golden glory roots grown under ½ MS agar supplemented with 0, 50, 100, and 200 μM As were extracted (in triplicates) with 600 μL of methanol containing 50 μM of protopanaxatriol (Extrasynthese, Lyon, France) as internal standard. Samples were ultrasonicated for 20 min at room temperature, and further centrifuged for 10 min at 13,000 rpm. The supernatant was filtered employing 0.22 μm centrifugal filters (UFC30GV, Merck Millipore). Cucurbitacin quantification was performed as previously reported (Almeida et al. [71]2022). We focused exclusively on roots because they are the site of cucurbitacin production at the seedling stage, and it is reported that cucurbitacin biosynthetic genes are not expressed in cotyledons (Almeida et al. [72]2022; Brzozowski et al. [73]2020). 2.3. RNA‐Sequencing and Analysis Total RNA from 12 samples: three biological replicates of 2‐week‐old C. pepo var. golden glory roots grown under ½ MS agar supplemented with 0, 50, 100, and 200 μM As, were isolated with the Spectrum Plant Total RNA Kit (Sigma‐Aldrich, St Louis, MO, USA) and treated with RQ1 RNase‐free DNase I treatment (Promega, Madison, WI, USA). RNA‐seq library preparation of RNA samples involved NEBNext Ultra II RNA Library Prep Kit for Illumina with NEBNext Poly (A) mRNA Magnetic Isolation Module (NEB, Ipswich, MA, USA). Quality and quantity of total RNA samples and their generated libraries were subjected to Bioanalyzer 2100 (Agilent, Santa Clara, CA, USA) and Qubit 3.0 fluorimeter (Invitrogen, Waltham, MA, USA). Paired‐end sequencing (2 × 75 bp) of libraries were carried out with the Illumina NextSeq 500 platform (Illumina, San Diego, CA, USA). The bcl2fastq tool (Illumina, San Diego, CA, USA) was used to convert the BCL files generated from NextSeq 500 to FASTQ files with 2 × 75 bp reads. Quality of reads was assessed using the fastqc tool ([74]https://www.bioinformatics.babraham.ac.uk/projects/fastqc). Adapters and low‐quality reads (Phred score QV < 30) were removed with Trimmomatic (Bolger et al. [75]2014). The quality‐filtered reads were mapped to the C. pepo genome ([76]http://cucurbitgenomics.org/v2/ftp/genome/Cucurbita_pepo/), excluding chromosome 0, with STAR RNA‐Seq aligner ([77]https://github.com/alexdobin/STAR). BAM and GFF were used to generate the count table with HTSeq ([78]https://github.com/htseq/htseq). Fragments per kilobase per million mapped reads (FPKMs) were calculated with the count table and length of each gene (Corchete et al. [79]2020). To identify differentially expressed genes (DEGs) between control and 50, 100, and 200 μM As treatments, DESeq2 was employed (Love et al. [80]2014). DEGs with a minimum log2fold change ≥ 1 and an adjusted p value ≤ 0.05 were further considered. PyWGCNA (Rezaie et al. [81]2023) was used to identify co‐expressed gene clusters through weighted gene co‐expression network analysis. Gene annotation, classification, gene ontology (GO), and pathway enrichment were determined using the CuGenDB tools ([82]http://cucurbitgenomics.org/). 2.4. Identification, Phylogenetics, and Structural Analysis of HMA Superfamily Members and As Transporters in C. pepo To identify HMA‐domain‐containing superfamily, aquaporin and ABCC transporters in C. pepo , characterized Arabidopsis protein sequences (Kamiya et al. [83]2009; Li et al. [84]2020; Song et al. [85]2010), were retrieved from TAIR ([86]https://www.arabidopsis.org/), and subjected to BLAST in the Cucurbit Genomics Database v2 (CuGenDBv2; [87]http://cucurbitgenomics.org/v2) to collect C. pepo protein sequences as well as Cucumis melo , Citrullus lanatus , and Cucurbita maxima when mentioned. To ensure the presence of conserved domains (HMA, ATPase, Hydrolase, MIP, and ABC), the conserved domain database (CDD) tool from NCBI was used ([88]https://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi) with the Pfam database as reference. In the case of HMA‐domain‐containing proteins, we named identified sequences according to chromosomal number. For identified Aquaporins and ABCC transporters, we used the orthologous name. MEGA11 (Tamura et al. [89]2021) was used for phylogenetic analysis. Multiple sequence alignments were performed using the Muscle (Edgar [90]2004) algorithm. Phylogenetic tree was built with the maximum likelihood method with the Jones–Taylor–Thronton (JTT) matrix–based model, using partial deletions and gamma distributed rates, further tested with bootstrap analysis (1000 replicates). Phylogenetic trees were exported as newick format for display in iTOL version 6.9 ([91]https://itol.embl.de/). TBtools (Chen et al. [92]2023) was used to display gene and protein structure as well as to perform Ka/Ks analysis. PlantCARE database ([93]https://bioinformatics.psb.ugent.be/webtools/plantcare/html/) was used to survey cis‐acting elements in 2000 bp upstream gene promoter sequences. 2.5. RT‐qPCR Equivalent concentrations of total RNA used for sequencing were subjected to first‐strand cDNA synthesis with SuperScript™ IV First‐Strand Synthesis System (Invitrogen, Waltham, MA, USA). Primers from selected genes (Table [94]S1) were designed with the CLC Genomics workbench software (QIAGEN, Aarhus, Denmark). The actin gene was used as endogenous control to normalize the relative gene expression. The reaction mixture contained 1 μL of cDNA template, 2 μL of forward and reverse primer at 10 μM, 10 μL of PowerTrack SYBR Green Master Mix (Applied Biosystems, Waltham, MA, USA), and 7 μL of nuclease‐free water for a final volume of 20 μL. Three technical replicates were assessed per each biological replicate. The qPCR was run in a QuantStudio 6 Pro (Applied Biosystems, Waltham, MA, USA). Melting curves were assessed to confirm the presence of a single amplicon and the absence of primer dimers. Ct values were further analyzed with the 2−ΔΔCt method (Livak and Schmittgen [95]2001). 2.6. Statistical Analysis Statistical analysis for phenotyping, cucurbitacins content and RT‐qPCR was carried out using R version 4.2.0 ([96]http://www.r‐project.org/). Normality was assessed using the Shapiro–Wilk test, and means were compared using a one‐way ANOVA with Tukey post hoc test for group comparison. The principal component analyses (PCA) were carried out with metabolomics profile (Table [97]S2), and RNA‐seq count table (Table [98]S3), to build a DESeq object in the DESeq package ([99]https://bioconductor.org/packages/DESeq2/). 3. Results 3.1. As Stress Deteriorates Root Growth and Modulates Cucurbitacin Production in C. pepo In order to understand the detrimental effects of As on C. pepo , we analyzed the root phenotype and cucurbitacin production in the roots of seedlings. High As concentrations primarily affect root development (Figure [100]1). The root length, projected and surface area, root volume, tips, and forks were significantly decreased with increasing As concentrations, especially in the 100‐ and 200‐μM treatments (p value < 0.001) (Table [101]S4). Whereas in plants grown at 50 μM, the overall root phenotype did not differ significantly from those of control plants, but there was an increase in the number of crossings and forks 103% (p value < 0.05). Finally, the average root diameter increased by 17% in the 200 μM As treatment (p value < 0.001); this was the only root phenotype showing a significant increase due to high levels of As. FIGURE 1. FIGURE 1 [102]Open in a new tab High arsenic concentrations are detrimental for Cucurbita pepo root growth. (a–d) Representative pictures of Cucurbita pepo roots grown under different Arsenic concentrations for 2 weeks. Differences between control and As treatments on (e) primary root length, (f) root projected area, (g) root surface area, (h) root diameter, (i) root volume, (j) number of root tips, (k) number of root forks, and (l) number of root crossings. Data are means of the biological replicates with SEM represented as error bars. “ns,” “*,” “**,” and “***” indicate statistical significance between control and treatments by one‐way ANOVA with post hoc Tukey HSD at the levels of p values ≥ 0.05, ≤ 0.05, ≤ 0.01, and ≤ 0.001, respectively. In terms of C. pepo root metabolism, it can be observed that their metabolic profile is affected by varying As concentrations, as summarized by the principal component analysis (PCA) generated from the LC–MS analysis (Figure [103]2a). The 200‐μM treatment deviates most from the control, primarily along the first principal component (PC1), which accounts for 64.8% of the variance. In the particular case of cucurbitacins, there is a sharp decrease of cucurbitacin B (p value < 0.001), E (p value < 0.05), and I (p value < 0.01) content in the 50‐μM treatment (Figure [104]2b). In the case of cucurbitacin B, the 100‐μM As treatment also showed lower content than controls (p value < 0.01), although the decrease was less pronounced than at the 50‐μM treatment. Conversely, in the 200‐μM As treatment, the levels of all three evaluated cucurbitacins were not significantly different when compared with those in controls. These observations suggest As exposure influences C. pepo root metabolism, showing the greatest shift at 200 μM, while cucurbitacin content is affected only at low As levels. FIGURE 2. FIGURE 2 [105]Open in a new tab Arsenic affects the metabolic profile of Cucurbita pepo roots modulating their cucurbitacin content. (a) Distribution of individual roots exposed to different concentrations of Arsenic in the space of the first two principal components of the principal component analysis on LC–MS analysis. (b) Cucurbitacin B, E, and I concentrations in 50, 100, and 200 μM As treatments. Data are means of the biological replicates with their SEM represented as error bars. “ns,” “*,”“**,” and “***” indicate statistical significance between control and treatments by one‐way ANOVA with post hoc Tukey HSD at the levels of p values ≥ 0.05, ≤ 0.05, ≤ 0.01, and ≤ 0.001, respectively. 3.2. Expression Levels of Cucurbitacin Pathway Genes Do Not Explain the Observed Cucurbitacin Content Regulation in As Treatments The observed differences in the C. pepo root metabolism prompted us to also investigate the transcriptomic profile, the molecular mechanisms underlying the coping responses to As at different concentrations, as well as the associated phenotypic changes and cucurbitacin modulation observed in C. pepo under As stress, by conducting RNA‐sequencing and confirming the expression of specific genes using RT‐qPCR. Similar to the metabolic profile, a PCA reveals As influences the C. pepo transcriptome in a concentration‐dependent manner (Figure [106]3a). The transcriptomic profile shows distinct responses depending on the concentration, with the 200‐μM As treatment diverging the most from the control, primarily along PC1 (77% variance). Additionally, the proximity of the 50 and 100‐μM treatment clusters suggests similar transcriptomic profiles at these concentrations. Moreover, we identified common top upregulated and downregulated genes across As treatments, along with a clear clustering pattern of highly variable genes based on the presence or absence of As (Figure [107]S1). Notably, upregulated genes primarily encode stress‐related proteins, such as members of the heat‐shock family and MYB transcription factors, while downregulated genes are associated with growth and pathogenesis‐related responses. FIGURE 3. FIGURE 3 [108]Open in a new tab Arsenic impact on the transcriptomic responses of Cucurbita pepo . (a) Distribution of individual roots exposed to different concentrations of arsenic in the space of the first two principal components of the principal component analysis on gene expression profiles derived from RNA‐seq analysis; (b) expression of cucurbitacin pathway‐related genes; (c) gene ontology (GO) enrichment analysis of biological process (BP), cellular component (CC), and molecular function (MF); and (d) pathway enrichment analysis of the common 592 DEGs (cut‐off p value ≤ 0.05). Data are means of the biological replicates with their SEM represented as error bars. “ns,” “*,”“**,”and “***” indicate statistical significance between control and treatments by one‐way ANOVA with post hoc Tukey HSD at the levels of p values ≥ 0.05, ≤ 0.05, ≤ 0.01, and ≤ 0.001, respectively. Remarkably, we observed an upregulation in the expression of cucurbitacin‐biosynthetic pathway genes at 200 μM As, including cucurbitadienol synthase, CpCPQ (Cp4.1LG17g04420), characterized cytochromes P450, CpCYP81BF1 (Cp4.1LG12g10100), and CpCYP81BB1 (Cp4.1LG17g09460), but not for the acyl transferase CpACT1 (Cp4.1LG17g09410) (Figure [109]3b). Nevertheless, expression of these genes was not significantly different at the 50‐ and 100‐μM As treatments when compared with controls, explaining why these genes did not pass the thresholds of the downstream RNA‐seq analysis to be classified as differentially expressed genes (DEGs) and therefore are not listed in the DEGs (Table [110]S5–[111]S9). Although the expression profiles of cucurbitacin pathway genes do not fully explain the changes in cucurbitacin content observed across the treatments, the metabolic profiles suggest that As exposure modulates the root metabolism, which could indirectly cause the changes in cucurbitacin content. Furthermore, gene ontology (Figure [112]3c; Table [113]S10) and pathway (Figure [114]3d; Table [115]S11) analyses were conducted to gain insights into the mechanisms affected by As. Responses to chemicals, toxic substances, and other stresses were the most enriched biological processes (BP). Accordingly, the glutathione‐mediated detoxification pathway in response to oxidative stress was commonly enriched among the upregulated genes in the treatments with As exposure. The extracellular region was the only cellular component (CC) enriched, while for molecular function (MF), DNA‐binding transcription factor activity and glutathione transferase activity were the most enriched. On the other hand, we found downregulated common peroxidase genes involved in lignan biosynthesis such as justidicin and matairesinol. Additionally, the correlation between gene expression and As concentration was assessed using WGCNA (Figure [116]S2). After assigning genes into modules, we found that fundamental biosynthesis and gene regulation processes, as well as the glutathione‐mediated detoxification II pathway, were coregulated with the 100 and 200‐μM As treatments, whereas signaling and energy process and the urate degradation to allantoin pathway were coregulated with the 50‐μM As treatment (Table [117]S12). These results suggest that exposure to different concentrations of As leads to the upregulation of detoxification and stress tolerance mechanisms, while also affecting fundamental processes in a concentration‐specific pattern, which correlates with its phenotypic traits. 3.3. Genes Encoding HMA‐Domain‐Containing Proteins and As‐Transporters Are Differentially Expressed Upon As Stress We studied in more detail genes encoding members in the HMA‐domain superfamily, aquaporin and ABCC transporter families due to their association to metalloid tolerance, As transport, and their GO terms matching with those enriched in this study (response to oxidative stress, inorganic substance, and detoxification). Moreover, we aimed to curate the orthologs of these known metal (loid)‐tolerance related genes in this non‐model crop to enable further in‐depth study within the Cucurbitaceae family. Importantly, we did not observe the modulation of As(V) transporters. For instance, PHT1:1 (AT5G43350) is known to participate in As(V) entry in Arabidopsis thaliana (Catarecha et al. [118]2007). In our dataset, the putative homologs Cp4.1LG03g10070, Cp4.1LG12g01500, and Cp4.1LG15g00730 were not found to be differentially regulated across conditions (Table [119]S5–[120]S9). Members of the HMA‐domain superfamily fall into different subfamilies, including HIPP (Heavy metal‐associated Isoprenylated Plant Protein), HPP (Heavy metal‐associated Plant Protein), P[1B]‐ATPases, and ATX1 (antioxidant protein 1) differing in protein structure and functionality (De Abreu‐Neto et al. [121]2013; Li et al. [122]2020; Ye et al. [123]2022). Using a comparative genomics approach, we identified 77 genes encoding HMA‐domain‐containing proteins, 53 HIPP, 14 P[1B]‐type ATPases, nine HPP, and one ATX1 across the genome of C. pepo (Table [124]S13; Figure [125]S3). We further studied the relationships of C. pepo orthologs to characterized HMA‐domain‐containing proteins in A. thaliana through phylogenetic analysis of amino acid sequences. The HIPP family in A. thaliana is known to be involved in the coping mechanisms against different metals and metalloids (Gao et al. [126]2009; Tehseen et al. [127]2010; Wei et al. [128]2023). After curation, the resulting phylogenetic tree of the HIPP protein family revealed CpHIPP42 (Cp4.1LG15g05950) and CpHIPP17 (Cp4.1LG04g04920) as the C. pepo orthologs of AtHIPP26 and CpHIPP36 (Cp4.1LG13g03670) as an ortholog of AtHIPP22 (Figure [129]4a). Construction of a phylogenetic tree including amino acid sequences of HIPP homologs from other Cucurbitaceae species (Figure [130]S4) suggests CpHIPP17 is an ortholog that arose from a duplication event unique to the Cucurbiteae tribe (squash, pumpkin) as no sequences from members of the Benincaseae tribe (cucumber, melon and watermelon) or A. thaliana group with it. Although being a segmental duplication from Chromosome 4 to 15, the Ka/Ks ratio of this event is less than 1 (0.048), which implies purifying selection, also shown in the protein structure. Nevertheless, the promoter of both genes differ, as a survey of their cis‐acting elements revealed that salicylic acid responsiveness and specific MYB binding sites were found only for CpHIPP42 (Table [131]S14). Accordingly, gene expression of these HIPP paralogs is different in response to As. Whereas the gene expression of CpHIPP17 significantly decreased in the 100 μM As treatment compared with controls, for CpHIPP42, the expression increased six‐fold in the 200‐μM As treatment (Figure [132]4a). As for CpHIPP36, its expression was upregulated as As concentrations increased, reaching up to a 40‐fold increase in the 200‐μM As treatment. Despite these differences in regulation, the structure of these three HIPP proteins share similar features (Figure [133]4a). FIGURE 4. FIGURE 4 [134]Open in a new tab Expression of HMA‐containing‐protein genes is upregulated under As stress in Cucurbita pepo . (a) (top) Maximum‐likelihood tree constructed on deduced amino acid sequences of 13 plant HIPPs aligned by ClustalW spanning 157 positions using the MEGAX program. The statistical significance of each node was tested by the bootstrap method using 1000 iterations. Accession numbers and representative names are given in the figure. (bottom‐left) Relative expression profiles of C. pepo HIPP‐protein genes under different arsenic concentrations. (bottom‐right) Protein and gene structure of C. pepo HIPP proteins. (b) (top) Maximum‐likelihood tree constructed on deduced amino acid sequences of 13 plant HMA‐containing proteins aligned by ClustalW spanning 717 positions using the MEGAX program. The statistical significance of each node was tested by the bootstrap method using 1000 iterations. Accession numbers and representative names are given in the figure. (bottom‐left) Relative expression profiles of C. pepo HMA‐containing‐protein genes under different arsenic concentrations. (bottom‐right) Protein and gene structure of C. pepo HMA‐containing proteins. Gene and protein structural features are shown in the right side of the expression bar plot for each gene. The different colored rectangles represent CDS, UTRs, and structural domains mentioned in the color legend. The below scale indicates the length in kb and aa, accordingly. Data are means of the biological replicates with their SEM represented as error bars. “ns,” “*,” “**,” and “***” indicate statistical significance between control and treatments by one‐way ANOVA with post hoc Tukey HSD at the levels of p values ≥ 0.05, ≤ 0.05, ≤ 0.01, and ≤ 0.001, respectively. Next, we studied members of the P[1B]‐ATPase family as they also contain HMA domain(s) in their protein structure (Axelsen and Palmgren [135]2001). Phylogenetic analysis of a curated set of amino acid sequences including functionally characterized P[1B]‐ATPases of A. thaliana shows that CpHMA2 (Cp4.1LG05g03530) and CpHMA8 (Cp4.1LG10g10980) group in the clade of AtHMA5, while CpHMA11 groups with that of AtHMA6. Whereas CpHMA8 appears to be the true ortholog of AtHMA5, CpHMA2 seems to have arisen from a duplication event in the common ancestor of the Benincaseae and Cucurbiteae tribes (Figure [136]4b), which was consistent with our previous study (Flores‐Iga et al. [137]2023). AtHMA5, AtHMA7, CpHMA8, and CpHMA2 all have a similar structure containing three HMA domains in the first 300 aa of the N‐terminal, while AtHMA6 and CpHMA11 contain a single HMA domain in their N‐terminal. In regard to gene expression, in general, the three C. pepo P[1B]‐ATPases are upregulated in the presence of As, but their profiles vary significantly (Figure [138]4b). Specifically, CpHMA8 increases more than five‐fold in all As treatments but being most highly expressed at the 50‐μM As treatment. Interestingly, despite belonging to different phylogenetic clades, CpHMA2 and CpHMA11 follow a similar trend of increased gene expression with ascending As concentration in the 50‐ and 100‐μM treatments but having a decrease at 200 μM As. Furthermore, we surveyed aquaporins in the NIP group, which are key As transporters involved in its extrusion and translocation (Kamiya et al. [139]2009). There are two proteins in C. pepo , named here as CpNIP1:1 (Cp4.1LG19g06570) and CpNIP1:2 (Cp4.1LG10g05550), that group in a clade with AtNIP1:1 and AtNIP1:2 (Figure [140]5a). CpNIP1:1 appears to be the result of a duplication event exclusive to the Cucurbiteae tribe, whereas CpNIP1:2 groups with other aquaporins of the Benincaseae tribe (Figure [141]S5). Interestingly, expression of CpNIP1:1 remains unaffected at low As concentrations but increases 10‐fold in the 200 μM As treatment, while expression of CpNIP1:2 decreases nearly half under 50 and 100 μM of As (Figure [142]5a). FIGURE 5. FIGURE 5 [143]Open in a new tab Orthologous transporters are regulated on Cucurbita pepo in high As conditions. (a) (left) Maximum‐likelihood tree constructed on deduced amino acid sequences of eight NIPs aligned by ClustalW spanning 218 positions using the MEGAX program. The statistical significance of each node was tested by the bootstrap method using 1000 iterations. Accession numbers and representative names are given in the figure. (middle) Relative expression profiles of C. pepo NIP‐protein‐encoding genes under different arsenic concentrations. (right) Protein and gene structure of C. pepo NIP proteins. (b) (left) Maximum‐likelihood tree constructed on deduced amino acid sequences of 10 ABCC transporters aligned by ClustalW spanning 936 positions using the MEGAX program. The statistical significance of each node was tested by the bootstrap method using 1000 iterations. Accession numbers and representative names are given in the figure. (middle) Relative expression profiles of C. pepo ABCC transporter genes under different Arsenic concentrations. (right) Protein and gene structure of C. pepo ABCC transporters. The different colored rectangles represent CDS, UTRs, and structural domains according to the color legend. The below scale indicates the length in kb and aa, accordingly. Data are means of the biological replicates with their SEM represented as error bars. “ns,” “*,” “**,” and “***” indicate statistical significance between control and treatments by one‐way ANOVA with post hoc Tukey HSD at the levels of p values ≥ 0.05, ≤ 0.05, ≤ 0.01, and ≤ 0.001, respectively. Finally, we studied ABCC transporters because they have a functionally characterized role in transporting As‐complexes in A. thaliana (Song et al. [144]2010) and are recognized to participate in the subcellular sequestration of As (Zhang et al. [145]2021). Our focus was on two identified C. pepo ABCC transporters, namely, CpABCC1 (Cp4.1LG15g02900) and CpABCC2 (Cp4.1LG19g05820). These C. pepo ABCC transporters exhibited distinct expression patterns across As treatments (Figure [146]5b), with a significant up‐regulation at 200 μM: with CpABCC1 increased 12‐fold, whereas CpABCC2 showed a three‐fold (Figure [147]5b). In summary, the phylogenetic analysis showed that evolution of the HIPP, P1[B]‐ATPase, and aquaporin families in the Cucurbiteae tribe, specifically C. pepo , appears to have diverged from that of A. thaliana and members of the Benincaseae tribe, possibly as a result of duplication events, while the ABCC transporter family remains relatively conserved. 4. Discussion 4.1. As Exposure Is Detrimental to Plant Growth and Plant Defense of C. pepo Previous studies on model and non‐model plants have established plant responses to As stress, a hazardous nonessential metalloid (Gracia‐Rodriguez et al. [148]2024). Before this study, little was known about the molecular responses of cucurbits to As and how exposure to this metalloid influenced C. pepo metabolism. In the Cucurbitaceae family, exposure to As significantly diminishes root and shoot length in Cucurbita moschata and C. sativus (Hong et al. [149]2011; Mushtaq et al. [150]2020). In our study, we noticed As exposure negatively impacts the growth of the root system in C. pepo , particularly at 100 and 200 μM, indicating its vulnerability at high As levels that have been previously studied in A. thaliana (Navarro et al. [151]2024; Peña‐Garcia et al. [152]2021; Zou et al. [153]2024). Damage to the root system likely triggers a cascade of negative effects on plant health. For instance, reduced root function hinders water and nutrient uptake, ultimately impacting plant metabolism (Kosakivska et al. [154]2021). Although members of the Benincaseae tribe, C. melo and C. lanatus exhibited a species‐specific pattern of As bioaccumulation in response to As stress (Pineda‐Chacón and Alarcón‐Herrera [155]2016), both species presented significant biomass decrease and plant weakening. As the root diameter is the only trait that increased in size in response to As exposure, it warrants further investigations. Cucurbitacins are bitter secondary metabolites recognized for medicinal purposes (Abdelmonsef et al. [156]2024); within the plant system, they serve as a defense mechanism against biotic stress such as herbivores and microbial pathogens (Brzozowski et al. [157]2020). Our LC–MS analysis indicates that As exposure modulates the metabolome of C. pepo , particularly the cucurbitacin profile, dramatically decreasing the concentrations of cucurbitacin B, E, and I in the 50 μM As treatment. Interestingly, the expression profile of the cucurbitacin biosynthetic pathway genes was only modulated in the 200‐μM As, with the exception of CpACT1, which catalyzes the final step of cucurbitacin E biosynthesis (Ma et al. [158]2022). Our observations suggest that, under low As concentrations, cucurbitacin levels are indirectly influenced by changes in metabolic pathways or regulatory networks rather than by direct transcriptional regulation. Although no significant changes in cucurbitacin content were detected at 200 μM, the increased gene expression at this concentration suggests a metabolomic shift that redirects isoprenoid precursors and/or intermediates towards cucurbitacin biosynthesis. According to our PCA, this shift was most pronounced at this high As concentration, warranting further investigation. A comprehensive untargeted metabolomic analysis using MS/MS data would be crucial to gain deeper insights into the metabolic changes affecting cucurbitacin accumulation. To our knowledge, this is the first report linking the levels of defense compounds, such as cucurbitacins, in Cucurbitacea crops to As exposure. Analysis of our RNA‐seq data showed downregulation of genes associated with immune and stress signaling pathway genes, providing further evidence that metalloid exposure, specifically As, not only affects oxidative stress pathways but also disrupts C. pepo biotic defense mechanisms, weakening the plant system (Labidi et al. [159]2023; Labidi et al. [160]2021). Interestingly, we found degradation of Allantoin correlated with the 50‐μM As treatment in our WGCNA. Allantoine is a nitrogen‐rich compound that participates in the resistance mechanism of C. melo to powdery mildew, a major fungal disease (Dun et al. [161]2024). These results provide insights into how As stress influences plant growth and plant defense, potentially contributing to an increased susceptibility of crops to infections. In addition, upregulated genes showed GO enrichment of toxic substance and oxidative stress responses in the biological process category across As treatments. Notably, genes related to the gluthathione‐mediated detoxification II pathway were enriched, which is well‐known for its role in metal and metalloid stress (Kumar and Trivedi [162]2018). When we assessed WGCNA to understand the co‐expression networking, we found regulatory processes governing transcription and biosynthetic pathways being disturbed, particularly in the 100‐ and 200‐μM treatments, which concur with previous results in A. thaliana (Abercrombie et al. [163]2008; Peña‐Garcia et al. [164]2021). Other genes with expression patterns worth mentioning are the top up‐regulated DEGS, specifically, three heat shock‐like proteins (Cp4.1LG08g07390, Cp4.1LG03g14960, Cp4.1LG01g11230) and a MYB transcription factor (Cp4.1LG02g16350). In Arabidopsis, genes in the MYB family have been linked to metalloid tolerance, regulating key players such as members of the HMA family (Cao et al. [165]2024; Chen et al. [166]2021), and we found MYB regulatory sites in the promoter of CpHIPP42, which was upregulated as concentrations increased; further studies could reveal a link between this MYB factor and HMA regulation or other factors in the As response. On the other hand, biotic stress signaling pathway genes such as jasmonic acid (Cp4.1LG18g02900, Cp4.1LG03g14930, Cp4.1LG11g01710) and pathogenesis related signaling (Cp4.1LG15g04120) were found as the most down‐regulated genes. Consistently, jasmonic acid biosynthesis genes have been found to be down‐regulated when C. pepo is exposed to trace heavy metals (Labidi et al. [167]2023). 4.2. HMA‐Domain‐Containing Proteins and As‐Related Genes Undergo Different Genetic Expressions Under As Stress Our study applies comparative genomics to explore metalloid tolerance in C. pepo by examining mechanisms from A. thaliana and Oryza sativa , specifically focusing on the HMA‐domain superfamily, whose members play key roles in the transport, binding, homeostasis, and detoxification of metals such as Cu, Ag, Pb, and Cd (Jamla et al. [168]2021; Li et al. [169]2020). In contrast to 55 and 46 members in A. thaliana and O. sativa , respectively (De Abreu‐Neto et al. [170]2013; Li et al. [171]2020), we identified 77 HMA‐domain‐containing superfamily members across the C. pepo genome. The known high affinity to thiol groups (Sun et al. [172]2023) made us suspect that As could interact with cysteine residues in the HMA‐domain of members in this superfamily and regulation of these genes could be influenced. Our prior analysis of the P[1B]‐ATPase family in Cucurbitaceae pointed this was a hypothesis worthwhile testing as we had observed gene expression differences in C. pepo under As stress, particularly in the Cu/Ag clade (Flores‐Iga et al. [173]2023). Indeed, initial observations on the gene expression of members of this superfamily in the RNA‐seq analysis suggested their regulation was different in response to As (Figure [174]S6), which was confirmed with RT‐qPCR (Figures [175]4, [176]5). In the HIPP family, we found that CpHIPP17 and CpHIPP42 are orthologs of AtHIPP26, and CpHIPP36 of AtHIPP22. Both AtHIPP26 and AtHIPP22 are known and required for heavy metal tolerance such as Cd (Manara et al. [177]2020; Tehseen et al. [178]2010). In this study, we observed a substantial expression of CpHIPP36 only, especially in the highest As concentration. Moreover, intron length and regulation of CpHIPP17 and CpHIPP42 is clearly different but the protein structure is not, such differences in transcription activity of duplicated genes are recognized (Panchy et al. [179]2016). In a survey of their cis‐acting elements, salicylic acid responsiveness elements were unique for the promoter region of CpHIPP42. Salicilyc acid is a ROS‐stimulated plant hormone that has been reported to activate antioxidant defense systems (Emamverdian et al. [180]2020). In the P1B‐ATPase family, CpHMA8 and CpHMA2 were found to be orthologs of AtHMA5 and AtHMA7, respectively. AtHMA5 participates in Cu detoxification in roots, while AtHMA7 is essential for Cu transport to ethylene receptor ETR1 (Andrés‐Colás et al. [181]2006; Puig et al. [182]2007; Migocka [183]2015). CpHMA11 is orthologue of AtHMA6, which contains a lone HMA domain and is involved in delivery of Cu to the chloroplastic compartment (Mayerhofer et al. [184]2016). Elevated transcription of CpHMA8 and CpHMA11 across As treatments suggests that gene structure, specifically HMA domain quantity, does not determine transcriptional response, but instead transcription may be influenced by functional factors (Abercrombie et al. [185]2008), which is also observable in different tissues at later developmental stages of C. pepo (Flores‐Iga et al. [186]2023). AtATX1 (AtCuSOD) is a Cu chaperone for superoxide dismutase (SOD) with a role of homeostasis under heavy metal stress, for example, Cu (Shin et al. [187]2012). Besides delivering Cu to Cu/Zn SOD, it participates in Cu trafficking for homeostasis networking with Cu/Ag members of the P[1B]‐ATPase subfamily (Zhang et al. [188]2018). Interestingly, expression levels of CpCuSOD were higher in all As treatments. Whether the higher expression of CpCuSOD is caused directly by As, P[1B]‐ATPase induction or indirectly due to ROS production by As remains to be determined. The genome‐wide identification of HMA members and gene expression analysis in this study narrow down the members that are likely activated in the primary stage due to As exposure. The HMA investigation needs further functional characterization to understand what are the triggering requirements of different HMA‐domain‐containing proteins to act synergistically towards heavy metal and metalloid tolerance and its related oxidative stress, which is proposed as an auxiliary mechanism for plants to cope with a wide range of metals and metalloid stressors (Rono et al. [189]2022). A common mechanism of As tolerance is the reduction of As (V) into a higher affinity and likely extrusive form, As (III) (Chao et al. [190]2014; Wang et al. [191]2018). Non‐enzymatic and enzymatic antioxidants are responsible for this mechanism, which is observed in our RNA‐seq analysis with the upregulation of glutathione detoxification pathway and glutathione metabolic processes. In this sense, aquaporins are recognized to participate in As exclusion and translocation, particularly of the As (III) form (Sun et al. [192]2024). Here, high expression levels of CpNIP1:2 were observed in the 200 As μM treatment, but not in CpNIP1:1. AtNIP1:1 and AtNIP1:2 are determinants of As transport and permeability across root cells, and higher expression of their encoding genes has been reported in earlier reports (Ji et al. [193]2017; Kamiya et al. [194]2009). Similarly, OsNIP1:1 participates in root‐shoot transport, and accumulation (Sun et al. [195]2018). Therefore, the studied aquaporins may play a supportive role in the plant's response to high cellular As (III) toxicity after the first As(V) to As (III) reduction step (Dhankher et al. [196]2006; Zou et al. [197]2024), potentially facilitating its release from the root system after prolonged intracellular accumulation. We ultimately investigated C. pepo orthologues of AtABCC1 and AtABCC2, which are well‐characterized vacuolar transporters that facilitate the cellular sequestration of As by transporting As (III)‐phytochelatin and As (III)‐GS complexes (Song et al. [198]2010). We observed that CpABCC1 and CpABCC2 have sequence divergence in the leading ABC membrane superfamily domain, and their expression levels differ. Notably, only CpABCC2 exhibits high transcription rates, especially at 200 μM (Figure [199]5b), consistent with observations that the expression of these transporters is solely induced when chelate action and vacuolar As transport are coordinated (Briat [200]2010). Divergence in the function of CpABCC1 reflected by difference in its structure could be a factor contributing to the lack of transcription induction for its gene under As stress. To study the effect on gene activation, further investigation is needed to explore the regulatory mechanisms behind the transcriptional control of CpABCC2, such as upstream signaling pathways and cis‐binding regulators providing insights into how this plant orchestrates its stress response at the molecular level. 5. Conclusion Cucurbita pepo plants exposed to As stress exhibit root damage and altered cucurbitacin levels, possibly weakening the plant to biotic stress. Specifically, our GO enrichment analysis of DEGs suggests that to cope with As toxicity, the plant activates oxidative stress responses with regulatory and metabolomic disruption at differing As levels. This could include a possible increase of glutathione production and other thiol‐rich peptides. Furthermore, WGCNA analysis of As‐concentration‐specific gene co‐expression modules indicate that low As levels stimulate energy and signaling pathways, along with allantoin metabolism, while higher As concentrations intensify oxidative stress responses. This includes upregulation of MYB transcription factors and heat shock proteins, which may play a role in regulating As tolerance. In addition, by curating C. pepo orthologs of A. thaliana , we were able to gain insights into possible downstream mechanisms of As transport, extrusion, and sequestration, highlighting potential pathways for As detoxification. Collectively, these preliminary findings suggest a shift from pathogen defense towards metalloid stress management in As ‐polluted environments, providing a foundation for future research aimed at understanding the underlying mechanisms of the responses observed in in this study and developing As‐tolerant cucurbits in the future. Author Contributions G.F.‐I.: writing – original draft preparation, writing – review and editing, visualization, software, investigation, formal analysis, data curation, conceptualization. C.L.‐O.: writing – review and editing, validation, formal Analysis, conceptualization. P.N.: software. P.N.: project administration, funding acquisition. U.K.R.: writing – review and editing, supervision, project administration, funding acquisition. N.B.: writing – review and editing, supervision, project administration, funding acquisition. A.A.: writing – original draft preparation, writing – review and editing, investigation, validation, supervision, software, project administration, methodology, funding acquisition, formal analysis, conceptualization. Conflicts of Interest The authors declare no conflicts of interest. Peer Review The peer review history for this article is available in the [201]Supporting Information for this article. Supporting information Data S1. Peer review. [202]PLD3-9-e70074-s009.pdf^ (108.7KB, pdf) Table S1. Primer sequences of selected Cucurbita pepo genes evaluated by RT‐qPCR under arsenic stress. [203]PLD3-9-e70074-s010.xlsx^ (10.6KB, xlsx) Table S2. Metabolomic profile of Cucurbita pepo roots under different As treatments (raw data). [204]PLD3-9-e70074-s005.xlsx^ (32.1KB, xlsx) Table S3. RNA‐seq count table of Cucurbita pepo roots under different As treatments. [205]PLD3-9-e70074-s014.xlsx^ (1.9MB, xlsx) Table S4. Data of Cucurbita pepo phenotyping under different As concentrations. [206]PLD3-9-e70074-s015.xlsx^ (11.3KB, xlsx) Table S5. FPKM values of control and As treatment RNA‐seq samples from Cucurbita pepo . [207]PLD3-9-e70074-s003.xlsx^ (4.4MB, xlsx) Table S6. DEGs between control and 50 μM As treatment in Cucurbita pepo . [208]PLD3-9-e70074-s008.xlsx^ (125.6KB, xlsx) Table S7. DEGs between control and 100 μM As treatment in Cucurbita pepo . [209]PLD3-9-e70074-s012.xlsx^ (117.4KB, xlsx) Table S8. DEGs between control and 200 μM As treatment in Cucurbita pepo . [210]PLD3-9-e70074-s011.xlsx^ (130.7KB, xlsx) Table S9. Common DEGs between control and As treatments in Cucurbita pepo . [211]PLD3-9-e70074-s007.xlsx^ (48.4KB, xlsx) Table S10. enriched GO terms of common and specific DEGs in different As treatments. [212]PLD3-9-e70074-s001.xlsx^ (15.8KB, xlsx) Table S11. Enriched pathways of common and specific DEGs in different As treatments. [213]PLD3-9-e70074-s002.xlsx^ (10.7KB, xlsx) Table S12. Enrichment of genes found in modules from WGCNA in Cucurbita pepo to As stress. [214]PLD3-9-e70074-s004.xlsx^ (68.1KB, xlsx) Table S13. Identified HMA domain containing proteins in Cucurbita pepo to As stress and their log2(FPKM) expression. [215]PLD3-9-e70074-s006.xlsx^ (15.9KB, xlsx) Table S14. Cis‐acting elements present in CpHIPP17 and CpHIPP42. [216]PLD3-9-e70074-s016.xlsx^ (30.1KB, xlsx) Figure S1 Overview of RNA‐seq analysis in Cucurbita pepo under different arsenic treatments. Volcano plots highlighting top differentially expressed genes from 50 μM As (a), 100 μM As (b), and 200 μM As (c) with thresholds: minimum log2 fold change = 1, p value ≤ 0.05 and lcshrink() function for low counts. Heatmap of the 25 common top and bottom genes with the highest expression variation (d). Dendograms explain the clustering pattern of samples and gene expression values. (e) Venn diagram showing the number of DEGs shared and unique for each of the three As treatments. Figure S2. WGCNA analysis of Cucurbita pepo exposed to As. WGCNA preprocessing results (a). In the left panel, scale‐free fit index. In the right panel, mean connectivity. The panels display the relationship and function with respect to the soft‐thresholding power (X‐axis), respectively. Clustering dendrogram of genes (b), with dissimilarity based on topological overlap, together with assigned merged module colors and the original module colors. Module–trait relationship heatmap generated from WGCNA for C. pepo ‐As concentration (c). Rows represent different As concentrations, and columns represent modules. Values in the squares indicate correlation module‐treatment according to Pearson correlations, p values are in parenthesis (null hypothesis; probability to find the current results if the correlation coefficients were zero). Color fade indicate positive (red) and negative (blue) correlation. Figure S3. Phylogenetic analysis and gene expression heatmap of the HMA‐containing proteins superfamily in Cucurbita pepo . HMA‐containing proteins were clustered with a Maximum‐likelihood dendogram. Heatmap was generated with normalized values of expression (Table S14). From inside to outside, expression of HMA‐containing proteins in 50, 100, and 200 μM As treatments are displayed. Figure S4. Maximum likelihood phylogenetic tree of orthologous HIPP members with CuSOD (ATX1) sequences used as outgroup. Arabidopsis thaliana, Cucurbita pepo, Cucurbita maxima, Cucumis sativus, Citrullus lanatus, and Cucumis melo sequences were used to understand the orthologs and duplication events between the Cucurbitae and Benincaseae tribes. Figure S5. Maximum likelihood phylogenetic tree of orthologous Aquaporin members with AtNIP5:1 and AtNIP6:1 used as outgroup. Arabidopsis thaliana , Solanum lycopersicum , Cucurbita pepo , Cucurbita maxima , Citrullus lanatus , and Cucumis melo sequences were used to understand the orthologs, conservation and duplication events between the Cucurbitae and Benincaseae tribes. Figure S6. Gene expression RNA‐seq values for genes of interest. P1B‐ATPase (a), HIPP (b), aquaporin (c), and ABCC (d) families. Data are means of the biological replicates with their SEM represented as error bars. “ns,” “*,”“**,”and “***” indicate statistical significance between control and treatments by one‐way ANOVA with post hoc Tukey HSD at the level of p values ≥ 0.05, ≤ 0.05, ≤ 0.01, and ≤ 0.001, respectively. [217]PLD3-9-e70074-s013.pdf^ (2MB, pdf) Funding: The authors received no specific funding for this work. Gerardo Flores‐Iga and Aldo Almeida contributed equally to this work. Data Availability Statement The RNA sequencing project is available in the NCBI database under the BioProject accession number PRJNA1177526. Scripts to analyze and display the data can be found at [218]https://github.com/gerapskx/Cucurbita‐pepo_Arsenic. References