Next Article in Journal
Expanding Genotype/Phenotype Correlation in 2p11.2-p12 Microdeletion Syndrome
Previous Article in Journal
Lambda CI Binding to Related Phage Operator Sequences Validates Alignment Algorithm and Highlights the Importance of Overlooked Bonds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Network Analysis of Publicly Available RNA-seq Provides Insights into the Molecular Mechanisms of Plant Defense against Multiple Fungal Pathogens in Arabidopsis thaliana

by
Cynthia Soto-Cardinault
1,†,
Kevin L. Childs
2 and
Elsa Góngora-Castillo
3,*
1
Unidad de Biotecnología, Centro de Investigación Científica de Yucatán, Mérida 97205, Mexico
2
Plant Biology Department, Michigan State University, East Lansing, MI 48824, USA
3
CONAHCYT-Unidad de Biotecnología, Centro de Investigación Científica de Yucatán, Mérida 97205, Mexico
*
Author to whom correspondence should be addressed.
Current address: Liber Institute for Brain Development, Baltimore, MD 21205, USA.
Genes 2023, 14(12), 2223; https://doi.org/10.3390/genes14122223
Submission received: 10 November 2023 / Revised: 6 December 2023 / Accepted: 12 December 2023 / Published: 16 December 2023
(This article belongs to the Section Plant Genetics and Genomics)

Abstract

:
Fungal pathogens can have devastating effects on global crop production, leading to annual economic losses ranging from 10% to 23%. In light of climate change-related challenges, researchers anticipate an increase in fungal infections as a result of shifting environmental conditions. However, plants have developed intricate molecular mechanisms for effective defense against fungal attacks. Understanding these mechanisms is essential to the development of new strategies for protecting crops from multiple fungi threats. Public omics databases provide valuable resources for research on plant–pathogen interactions; however, integrating data from different studies can be challenging due to experimental variation. In this study, we aimed to identify the core genes that defend against the pathogenic fungi Colletotrichum higginsianum and Botrytis cinerea in Arabidopsis thaliana. Using a custom framework to control batch effects and construct Gene Co-expression Networks in publicly available RNA-seq dataset from infected A. thaliana plants, we successfully identified a gene module that was responsive to both pathogens. We also performed gene annotation to reveal the roles of previously unknown protein-coding genes in plant defenses against fungal infections. This research demonstrates the potential of publicly available RNA-seq data for identifying the core genes involved in defending against multiple fungal pathogens.

1. Introduction

Pathogenic fungi are a significant factor in global crop production. Pathogen-related crop damage leads to annual losses ranging from 10% to 23%—equivalent to over USD 200 billion in economic damages [1,2,3]. Climate change has led to unprecedented challenges in food production, including the growing resistance to antifungal agents. In a warmer world, fungal infections are expected to become more widespread among crops, as the life cycle of fungi is directly influenced by temperature and humidity [4,5,6]. According to prediction models, the prevalence of pathogenic fungal infections in commercial crops will rise by 5% to 100% by 2050 as a result of climate change [7,8].
Pathogenic fungi can be classified into three groups: biotrophs, which feed on living tissues; necrotrophs, which kill the host cell and obtain nutrients from dead tissues; and hemibiotrophs, which exhibit an intermediate lifestyle, combining traits of both biotrophs and necrotrophs [9,10,11,12]. One highly important pathogen is Colletotrichum higginsianum, an ascomycete fungus with a hemibiotrophic nature. This fungus causes anthracnose, a disease that affects numerous monocotyledonous and dicotyledonous plants worldwide [9,13,14]. Another important pathogen, Botrytis cinerea, is a necrotrophic fungus with a broad range of hosts that has devastating effects on various plant tissues, including foliage, stems, flowers, and fruits [13,15,16,17].
Plants have developed complex molecular mechanisms to protect themselves from invading pathogens, including the ability to identify pathogens via pathogen-associated molecular patterns (PAMPs) with the help of pattern recognition receptors (PRRs). This leads to a process known as PAMP-triggered immunity (PTI), which limits the progression of disease. This pattern recognition is typically linked with calcium influx, callose deposition, bursts of reactive oxygen species (ROS), activation of miRNA pathways, activation of MAPK cascades, and increased expression of various defense-related genes, such as disease-related proteins (PR) [18,19,20]. In some cases, effectors produced by pathogens are recognized by plant resistance proteins, promoting a response known as effector-triggered immunity (ETI). One of the roles of ETI is to stimulate the hypersensitive response (HR), which results in rapid cell death at the site of pathogen invasion, limiting the spread of the pathogen. In addition, various signal transduction events, such as nitrous oxide, lipids, and various phytohormones that identify pathogen invasions, initiate plant immune defense responses [21,22,23,24]. The PTI and ETI immune responses are interconnected, as the activation of ETI improves the PTI signaling pathway. Both responses share functions related to the recognition of danger signals and the activation of defense mechanisms [18,25,26,27,28].
In the era of next-generation sequencing and machine learning, and considering the need for proactive responses to climate change, it is crucial to implement more comprehensive strategies in modern agriculture. One useful resource is the NCBI SRA, which contains more than 17 petabytes of public sequence data [29]. This underused data repository can be exploited using novel computational methods. A large portion of the sequence data is relevant to the agricultural sector, and opportunities are emerging to implement machine learning strategies that incorporate this immense pool of genomic data.
There is a growing interest in using public RNA-seq datasets to investigate intricate biological processes such as plant defense [30,31,32]. Publicly available RNA-seq expression data are a valuable resource for improving crop quality, particularly when conducting experiments under various environmental stressors or diseases is not feasible. However, the reutilization of public RNA-seq data is limited by the challenge of integrating batch datasets generated from different studies due to experimental variation, which can introduce bias; this must be carefully considered during analysis to ensure accurate and reliable results [30].
Traditionally, transcriptome studies with RNA-seq have been widely used for the identification of specific genes expressed during the fungal infection process, with a focus on specific characteristics of the plant or fungus [9,33,34,35,36,37]. In this study, our objective is to identify the core genes involved in defending against B. cinerea and C. higginsianum pathogens. We define core defense genes as those that confer a broad-spectrum defense mechanism and are not restricted to specific pathogens. We used publicly available RNA-seq expression information from the C. higginsianumA. thaliana and B. cinereaA. thaliana pathosystems. Both of these valuable models have been extensively studied with regard to specific responses during plant–pathogen interaction [14,15,16,17,38,39,40,41]. Additionally, only a limited number of studies have been conducted to understand common defense mechanisms against biotic stress, and our understanding of this topic is still at an early stage [42,43]. To date, the common reactions that these fungal pathogens can elicit in the host remain unclear.
One strategy, Weighted Gene Co-expression Network Analysis (WGCNA) [44], has contributed to crop improvement by revealing complex gene–gene interactions and helping to identify key genes that regulate important traits [45]. For instance, in soybeans, a gene module related to seed development was recognized and was found to include several genes involved in fatty acid biosynthesis [46]. In wheat, a set of genes associated with resistance to Fusarium head blight, a devastating fungal disease, was identified [47]. Furthermore, gene co-expression networks have been constructed using publicly available expression datasets. One of these studies used gene expression data to functionally annotate the rice proteome [48]. Moreover, other studies used gene co-expression networks to identify core genes in A. thaliana that respond to biotic stress using microarray data [42] and those in maize that respond to fungal infection using RNA-seq data [43]. However, neither of these studies considered batch-effect correction.
To identify core genes expressed during C. higginsianum and B. cinerea defense in A. thaliana, publicly available RNA-seq data were preprocessed for batch-effect correction using an in-house method, allowing us to obtain integrated gene expression data that could be used with the WGCNA [44]. Our preprocessing framework includes data integration, normalization and standardization based on variability identification while emphasizing a minimal loss of data, and it is based on statistics such as Gaussian distributions, kernel density estimation (KDE), and quantiles, which assist with the generation of evaluation metrics [49]. A module containing core genes that respond to infection with B. cinerea and C. higginsianum was identified. Using the DAVID agglomeration method [50], 33 genes were annotated by function across six categories, providing new insights into the functions of currently unknown protein-coding genes. Our findings demonstrate that publicly available RNA-seq data are a valuable resource for identifying broad-spectrum genes related to the defense against multiple fungal pathogens.

2. Materials and Methods

2.1. RNA-Seq Data Selection

Twenty-five RNA-seq datasets of fungus-infected leaves of A. thaliana (Col-0) were obtained from the National Centre for Biotechnology Information’s Sequence Read Archive (NCBI SRA) database. The C. higginsianumA. thaliana dataset was composed of 8 samples from BioProject accession number PRJNA148307 [9,51]; the B. cinerea–A. thaliana dataset comprised 6 samples from BioProjects with accession numbers PRJNA315516 and PRJNA593073; the Sclerotinia sclerotiorum–A. thaliana dataset consisted of 3 samples from BioProject accession number PRJNA418121 [52], and this dataset was used as an outlier control. The control dataset comprised 8 samples of healthy plants from BioProjects PRJNA315516 and PRJNA418121. The sample collection was performed from 12 to 30 h for control samples, and from 12 to 40 h for fungus-infected leaf samples (Supplementary Table S1). The transcriptomes were sequenced on Illumina platforms. The number of reads ranged from 10 to 30 million, with read lengths ranging from 93 to 150 bp (Table 1).

2.2. Sequence Quality Filtering and Expression Estimation

Reads from each dataset were processed for quality and filtered using FastQC v0.11.5 [53] and Trimmomatic v.0.38 [54] tools, using a Phred Score ≥20 and a minimum sequence length ≥40 bp. Filtered sequences were aligned to the A. thaliana genome TAIR10 [55] and the latest Araport 11 annotation (GenBank accessions CP002684–CP002688), using STAR alignment tool v2.7.5 [56], with parameters sjdbOverhang = 92 and genomeSAindexNbases = 7 to adjust the tool to the genome size and alignIntronMin = 8 and alignIntronMax = 1999 to adjust the minimum and maximum intron lengths [57]. The alignment quality reported in SAM/BAM files [58] were evaluated with the HTSeq-qa tool v0.11.1 and the expression abundance was estimated using the HTSeq-count tool v0.11.1 [59].

2.3. Data Preprocessing and Batch-Effect Correction

Central tendency and statistical dispersion measures were calculated, including the mean (µ), standard deviation (σ), and range (R), and histograms with KDE of distributions and percentiles were created using Seaborn [49] (Figure 1).

2.3.1. Integration of Raw Counts and Data Normalization (Figure 1a,b)

A single expression matrix was generated using the raw count files for each dataset. Protein-coding genes (henceforth referred to as genes) with expression values equal to zero were eliminated. Transcripts per million (TPM) [60] were used to perform data normalization with the TPM normalization tool v0.9.1 using Gene_length_extraction_from_GTF.ipynb [61]. For TPM normalization, the gene length was obtained from the A. thaliana genome GTF/GFF file [55]. The script 1_Step1_integrating_raw_counts.ipynb and script 2_Step2_TPM_normalization.ipynb were used to organize the gene expression data and perform data normalization, respectively (Table S2).

2.3.2. Data Standardization and Outlier Identification (Figure 1c,d)

The gene expression values were divided into percentiles, and top and bottom extreme tails were calculated. Genes with TPM values between the 1st and 99th percentile were retained, while genes beyond this range were discarded. After filtering, TPM values were transformed to log2 (TPM + 1). To evaluate the efficiency of our methods in detecting samples with atypical distributions, we included 3 samples with known atypical distribution as negative controls. The 3_Step3_TPM_standardization.ipynb and 4_Step4_Log2_scale.ipynb scripts were used (Table S2).

2.4. Weighted Gene Co-Expression Networks and Identification of Core Genes

Weighted gene co-expression network analysis (WGCNA v1.69-81) was performed using R package [44]. The pickSoftThreshold function according to the Scale Free Topology model fit was used to define the appropriate soft-thresholding power. The adjacency matrix was generated using a signed topological overlap matrix (TOM), along with dissimilarity values (1-TOM). Gene clusters were detected using a dynamic cutting algorithm with a minimum module size of 20 and a default cutoff height (0.99). Merged networks were derived at 0.1 distances. Pearson’s correlation analysis (r2 > 0.75) was used to identify modules linked to phenotype. The moduleEigenes, cor, corPvalueStudent functions were used to define module membership (MM) and gene significance (GS) with a cutoff of r2 > 0.75, r2 < 0.75 and p-value < 0.05.
To identify the core genes, genetic modules from healthy-control and infected-plant gene networks were compared. Modules in the infected-plant gene network with a difference >75% in genes (unique genes) were selected as exclusive gene modules for the infected-plant gene co-expression network. The correlation cutoff was set at R2 > 0.50. Gene modules that fit these criteria were selected as core genes.

2.5. Functional Annotation of the Gene Co-Expression Networks and Core Genes

A Gene Ontology (GO) overrepresentation test was performed using PANTHER v17.0 [62,63]. Gene modules with coefficients of R2 > 0.75 were selected, and a binomial test with Bonferroni correction was used to identify GO-Slim terms for biological process and molecular function GO modules [64,65] (http://geneontology.org/, accessed on 11 July 2023).
The core genes were analyzed using The Database for Annotation, Visualization and Integrated Discovery (DAVID, https://david.ncifcrf.gov/home.jsp, accessed on 15 July 2023). Nine of the fourteen DAVID annotation categories were used in this study. To identify closely related genes within the module, the DAVID Gene Functional Annotation tool applies a kappa score distribution and an agglomeration method based on heuristic fuzzy multiple-linkage partitioning. These methods are implemented through the tool’s web-based interface, and allow the user to select different levels of stridency, ranging from low to very high [66] (in this study, medium and high levels were selected). The selected gene groups were those repeated across the selection criteria. To identify the functional gene groups most relevant to the study (e.g., those of plants infected with pathogenic fungi), the DAVID tool uses the EASE score, which is a modified Fisher’s exact test, to assign an enrichment score to each annotated gene group. The cutoff was >0.80 [50,66].
The Arabidopsis Information Portal (Araport) through ThaleMine webtool (https://bar.utoronto.ca/thalemine/, accessed on 4 September 2023) was used to identify each A. thaliana gene [67,68].

2.6. Implementation

Data processing was implemented using Jupyter Notebooks [69] for Python v3.8.10 and WGCNA v1.69-81 [44] for R v4.1.0. The R scripts and code are freely available via GitHub at https://github.com/cyntsc/RNA-Seq-raw-integration (accessed on 28 September 2023) (DOI 10.5281/zenodo.7076416) (Table S2).

3. Results

To improve our understanding of the core genes involved in defending against multiple fungal infections, we conducted a comparative study using publicly accessible transcriptome data from C. higginsianum– and B. cinereaA. thaliana pathosystems (Table 1 and Table S1). First, we controlled the batch effects of public data to reduce data variability using an in-house framework (Figure 1). Next, we performed a weighted gene co-expression network analysis (WGCNA) [44] of the preprocessed data to identify the core genes. These core genes were then subjected to further agglomerative analysis using DAVID [50] to identify functions in closely related gene groups.

3.1. Controlling the Batch Effect in Public RNA-seq Data to Construct Gene Co-Expression Networks (GCN)

RNA-seq sequences were filtered for quality, recovering more than 92% of the reads. These were aligned to the A. thaliana TAIR10 genome [55], achieving an alignment coverage rate of 97.5% to 98.3% for the healthy samples and 77.4% to 98.2% for the infected samples. S. sclerotiorum samples (Ss30, Ss30.1 and Ss30.2), with alignment coverages ranging from 29.5% to 36.5%, were included as an outlier control to track data changes (Table 1, Figure 2).
After read mapping, two gene expression matrices were obtained, one containing more than 221 thousand expression values (27,655 genes × 8 samples) derived from healthy-control samples and another containing more than 470 thousand expression values (27,655 genes × 17 samples) from infected-plant samples. Genes with an expression value equal to zero were eliminated (Figure S1), resulting in a total of 22,426 and 24,239 genes in the healthy-control and infected expression matrices, respectively.
The expression values were normalized to TPM to set the means of the distributions to the same point, and the evaluation metrics showed a µ = 44.6, σ = 333 ± 6 and R = [0 to 36,543] for the healthy-control gene expression matrix and a µ = 41.2, σ = 284.5 ± 84.5 and R = [0 to 30,000] for the infected-plant gene expression matrix (Figure 3a). Extreme values in both expression matrices were identified and removed to reduce data variability. In the healthy-control expression matrix, 2262 genes were below the 1st percentile, in which TPM values were smaller than 0.1, and 372 genes were above the 99th percentile, with TPM values larger than 840. A total of 2634 genes were removed. In the infected-plant expression matrix, a total of 3965 genes were filtered. Of these, 3495 genes were below the 1st percentile and 470 genes were above the 99th percentile, where the TPM values were smaller than 0.1 and larger than 845, respectively. In total, 19,792 and 20,274 were obtained for the healthy-control and infected-plant expression matrices, respectively (Figure 3b). Data were transformed to log2(TPM + 1) to reduce the scale, and the metrics for the healthy-control samples were σ = 2.1 ± 0.1, µ = 3.4 ± 0.2 and R = [0 to 9.28] while for the infected-plant samples, the metrics were σ = 2.1 ± 0.1, µ = 2.9 ± 0.8 and R = [0 to 9.72] (Figure 3c).
To understand and demonstrate the effect of highly variable gene expression, Ss30, Ss30.1 and Ss30.2 samples were added to introduce variability in the data collection, with the aim of monitoring the efficacy of the workflow corrections. The histograms with KDE reveal a clear contrast between the preponderant and the outlier distributions of the Ss30, Ss30.1 and Ss30.2 samples (Figure 3c). The violin plots with KDE show individual distributions and reveal a pronounced difference in the distributions of the Ss30, Ss30.1 and Ss30.2 samples compared to other samples (Figure S2). Once the Ss30, Ss30.1 and Ss30.2 samples were eliminated, a more homogeneous dataset was obtained for the co-expression analysis (Figure 3d and Figure S2).

3.2. Gene Co-Expression Network Analyses and Gene Ontology Overrepresentation Test for Healthy and Infected Plants with Fungal Pathogens

Two gene co-expression networks were built. In the healthy-control gene network, a coefficient R2 = 0.80 was reached, and 237 clusters were identified with a Node-Connectivity Mean (NCM) of 374 genes per cluster (β = 28; R2 = 0.78). The merged network provided 23 clusters (Figure 4a, Table 2). In the infected-plant gene network, a coefficient R2 = 0.83 was reached, and 100 clusters with a NCM of 270 genes per cluster were identified (β = 27; R2 = 0.79). The merged network contained 36 clusters (Figure 4b, Table 2). A correlation of module eigengenes to disease trait data was performed, with a significant cutoff value of |GS| > 0.75 and p-value <0.05 (Datas S1 and S2).
To confirm the biological relevance of the identified networks, biological and molecular functions for each gene network were identified via a Gene Ontology (GO) overrepresentation gene test on the genetic modules with coefficients R2 > 0.75. Three and four genetic modules were analyzed to identify molecular function and biological processes from the healthy-control and infected-plant gene networks, respectively (Table S3). The number of genes in each module of the healthy-control sample gene network ranged from 79 to 2024, and 34–96% of these were assigned to ontology classes. In contrast, the infected-plant gene network modules contained 72 to 818 genes, 25–30% of which were classified into ontology classes (Table S4).
The results for the healthy-control gene network showed overrepresentation of 16 molecular functions mainly related to binding (GO:0005488) and translation (GO:0045182) in the “Coral3” module, which contains 1991 genes. The “Navajowhite3” module has 489 genes, and two biological processes related to vesicle-mediated transport (GO:0016192) and Golgi vesicle transport (GO:0048193) were overrepresented. The “Blue3” module contains 79 genes, and response to light (GO:0009416) and radiation (GO:0009314) are the main functions associated with this module (Figure 5a,b). In contrast the infected-plant gene network modules showed overrepresentation of 19 molecular functions in the 805-gene “Chocolate” module, in which protein binding (GO: 0005515) and kinase activity (GO: 0004672) are the main molecular functions. The “Chocolate2” and “Green3” modules each had only one molecular function class. The molecular function for the “Chocolate2” module is related to RNA binding (GO:0003723), while the “Green3” module is associated with unfolded protein binding (GO:0051082) (Figure 5c, Data S3).

3.3. Identification and Functional Annotation of Core Defense Genes in A. thaliana

To identify defense core genes related to multi-fungal infections, healthy-control and infected-plant networks modules were compared, and modules in the infected-plant gene network with at least 75% unique or different genes were identified (Figure S3). Finally, modules that showed a positive correlation (R2 > 0.50) were selected. It was determined that the “Darkmagenta” module fit our criteria, with a correlation coefficient of R2 = 0.75 for B. cinerea at 24 hpi and a R2 = 0.59 for C. higginsianum at 22 hpi (Figure 6). It contains 113 genes, and of these, 103 (77.4%) were exclusively in this module (Figure S3). The functional annotation of this module was performed using the Database for Annotation, Visualization and Integrated Discovery (DAVID, https://david.ncifcrf.gov/home.jsp, accessed on 28 September 2023), which helps to identify closely related gene groups [50,66].
The 113 genes in the “Darkmagenta” module were mapped to nine DAVID functional categories. To categorize highly related genes into functional groups relevant to the pathogenic fungal infection, the DAVID gene functional annotation tool [66] employs a kappa distribution score, an agglomeration method and an EASE enrichment score. The web tool implementation enables users to select different stringency levels. We identified gene groups that remained consistent at medium and high stringency. Thus, 14 enriched functional groups were obtained and then filtered using an EASE score >0.80 and p-value < 0.05, resulting in six enriched functional groups (Table 3 and Table S5). Of these six groups, seven genes were classified in the “WD40/YVTN repeat-like-containing” group, three genes in the “sterol metabolism” group, ten genes in the “glycosyltransferase” group, seven genes in the “intracellular protein transport” group, four genes in “Zinc finger, FYVE/PHD-type”, and six genes in “Methyltransferase” (Table 3). Four of these thirty-seven genes have multiple functions (AT5G06050, AT5G13960, AT5G40870 and AAT5G45660) and are associated with at least two functional groups (Table 4). Finally, 33 defense-related genes and their roles in B. cinerea and C. higginsianum infection in A. thaliana at 22 and 24 hpi were identified through DAVID functional annotation (Table 4).

4. Discussion

Pathogenic fungi-induced diseases are among the main causes of losses in commercial crops. In light of major challenges such as climate change and fungicide resistance, it is imperative to develop global strategies to tackle these losses. Conducting experiments under a wide range of environmental stresses can be both challenging and time consuming. However, the advent of NGS technologies has allowed the collection of large amounts of data that could help to address the challenges associated with modern agriculture. The utilization of publicly available RNA-seq expression data holds great potential for further study of complex plant–pathogen interactions, leading to the identification of candidate genes and molecular markers and optimization of breeding strategies by targeting specific genes. However, utilizing publicly available RNA-seq data remains challenging. To tackle this, we based our study on computational approaches using public data from RNA-seq transcriptome studies of A. thalian-B. cinerea and A. thaliana-C. higginsianum pathosystems to identify core genes that respond to multiple types of fungal infection using GCN. Reanalysis of public RNA-seq expression data requires the minimization of technical biases, known as the batch effect, to obtain high-quality co-expression networks [30]. Although several tools may be used for batch-effect correction (such as ComBat-Seq [70], ComBat [71], svaseq [72]), further efforts may be required, particularly when sample sizes are limited (<30) [30]. Therefore, we developed a straightforward workflow to preprocess data in order to correct the batch effect that allowed us to recover over 80% of genes for WGCNA (Figure 1).
Previous studies revealed that gene modules generated by WGCNA from condition-dependent gene expression experiments are more informative than gene modules identified by combining the entire dataset regardless of condition [48]. Thus, two gene co-expression networks were constructed using filtered data from healthy (control) and infected A. thaliana plants with B. cinerea and C. higginsianum. The network for healthy-control plants included 23 gene modules with an NCM of 374, while for the network for infected plants, 36 gene modules with an NCM of 270 were obtained (Figure 4, Table 2). Additionally, the quality of each network was confirmed through GO overrepresentation analysis, resulting in biological processes and functions that are overrepresented in each network and linked to each phenotype. Modules with an R2 values greater than 0.75 were selected for both networks, and it was determined that within the healthy-control plant network, genetic modules coral3, blue3 and navajowhite3, modules with positive correlation (R2 of 0.98, 0.96, and 0.79, respectively) were related to plant development and growth, exhibiting functions such as translation and protein binding (GO:0005488), vesicle-mediated protein transport (GO:0016192), as well as response to light and radioactivity (GO:0016192) (Figure 5, Table S2). The gene modules from the plant-infected network, chocolate, dodgerblue1, green and chocolate2 with positive correlations (R2 of 0.98, 0.96, 0.95, 0.91, respectively), displayed overrepresented molecular functions of genes associated with regulating gene expression and signal transduction, which are the primary stress response mechanisms. These functions included protein binding (GO:0005515), kinase activity (GO:0004672), RNA binding (GO:000373) and unfolded protein binding (GO:0051082) (Figure 5, Table S2) [73,74].
To identify core genes that respond to B. cinerea and C. higginsianum infection in A. thaliana, the genetic modules in the network of infected plants should differ by at least 75% compared to the network of healthy-control plants with a positive correlation (R2 > 0.5) for the infection conditions. Therefore, we selected the “Darkmagenta” module due to its differentiation in 77.4% of genes compared to the control. This module displayed an R2 = 0.57 with the B. cinerea treatment at 24 hpi and an R2 = 0.59 with the C. higginsianum treatment at 22 hpi, as shown in the heatmap (Figure 6). This module contained 113 genes that were functionally annotated using the DAVID tool [66], revealing 33 genes distributed across six functional categories (Table 3, Table 4 and Table S5).
The first group of genes in the “Darkmagenta” module consisted of seven genes (AT5G23430, AT3G13340, AT3G18060, AT5G51980, AT1G79990, VPS and VSC) that contained the WD40/YVTN repeat-like domain (IPR015943). In A. thaliana, 237 proteins have been reported to contain four or more copies of the WD40 domain and participate in diverse biological processes, including defense against pathogens [75,76]. The most extensively studied of the WDR proteins are the Gβ proteins, which are associated with type-1 membrane receptors in the plant innate immunity signaling pathway. Interactions between Gβ and type-1 membrane receptors convert extracellular signals into intracellular chemical defense responses, including reactive oxygen species (ROS) production, callose deposition, MAPK activation, defense gene activation and programmed cell death [77]. In this group, we identified at least 3 Gβ-type genes (AT5G23430, AT5G51980, AT1G79990) (Table 4).
In the second group, we identified three CPG for key enzymes in sterol biosynthesis, SMT1, SMO1-1, and 3BETAHSD/D1 (Table 4). Sterols are a complex mixture of organic compounds that are synthesized in plants and function as structural components of cell membranes. They play a vital role in processes such as channel regulation, protein trafficking, signal transduction and plant–pathogen interactions [78]. The SMT1, 4-methyl sterol oxidase (SMO) and 3-BETAHSD enzymes participate in the biosynthesis of c24-alkyl sterol and cholesterol in plants [79]. The role of sterols in plant defense has been studied more comprehensively in relation to abiotic stressors, such as drought, salinity, and cold, than against biotic stress. However, certain sterols such as cholesterol activate biotic stress response genes [78], while pathogenic bacteria and ROS stimulate biosynthesis of stigmasterol. Moreover, PATHOGENESIS-RELATED PROTEIN 1 (PR-1) can bind various sterols, including stigmasterol, to inhibit pathogen growth by sequestering sterols from pathogens [80]. Although there is evidence suggesting that sterols might function as a crucial element in plant defense, further studies are needed to demonstrate their efficacy.
The glycosyltransferase genes in group three consist of 10 CPG (ALG3, FUT13, CALS1, UK/UPRT1, AT4G38040, GUT2, PARVUS, SETH2, AT5G45660, AT1G34270). Glycosylation is one of the major posttranslational modifications in plants, playing an important role in defense against pathogens by inactivating toxic microbial compounds and strengthening the cell wall. Glycosylation also regulates secondary metabolite levels by increasing their water solubility to facilitate metabolite transport throughout the plant, increasing the metabolites’ stability to avoid degradation and even altering their biological activity to increase toxicity to pathogens or decrease bioactivity for plants. An example of this is the glycosylation of salicylic acid (SA), a phytohormone that functions as a signal molecule during plant defense against pathogens, activating systemic acquired resistance (SAR). This glycosylation is believed to facilitate detoxification of SA, allowing it to be safely stored and moved within the plant, which helps to avoid excessive activation of defense responses that could harm the plant’s growth and development [81,82,83]. Interestingly, within this group of genes, we have identified CALS1 (AT1G05570), which responds to SA (Table 4).
Four genes were identified as Zinc FYVE/PHD group (EMB1135, ATX2, AT5G12350, and AT1G5062). The zinc finger protein (ZFP) transcription factors (TFs) constitute a vast family in the plant kingdom and play a crucial role in stress tolerance. The ZFP TFs mainly regulate genes involved in antioxidation activity, which reduce ROS accumulation. Controlling the ROS level is critical, as it helps to avoid damage in molecules, including DNA, proteins and lipids, and to enhance stress tolerance. AT1G50620 and EMB1135 (alternatively known as the FGT 1 gene, AT1G79350), containing the RING/FYVE/PHD-type (RFP) domain (IPR011011) [84,85], have a zinc ion binding function, and recent studies indicate that this function may help with ROS scavenging [85,86]. Additionally, rapid accumulation of ROS during the development of fungal appressoria has been observed as a defense mechanism to prevent penetration of the pathogenic fungi [87,88]. This is consistent with the infection period of C. higginsianum, since at 22 h post inoculation, the fungus is in the prepenetration stage and is engaged in forming the appressorium [9]. Additionally, FGT 1 (EMB1135) encodes the FORGETTER 1 protein that binds directly to a specific class of heat-inducible genes through nucleosome remodeling [89]. These types of ZFP TFs regulate cellular responsiveness by binding to promoter regions of actively expressed genes in a heat-dependent fashion, such as heat shock proteins (HSP). The HSPs are induced by a variety of stresses, including oxidative stress; however, the crosstalk between abiotic and biotic stress responses in the ROS network remains poorly understood [90].
A group of six methyltransferases were identified (AT3G15530, SUVH4, AT2G34300, SMT1, ATX2, AT5G06050); all of these genes are involved in methylation. However, to date, there have been few studies on the role of methylation in plant–pathogen interactions, particularly with regard to infections caused by pathogenic fungi. The available evidence suggests that epigenetic control of gene expression contributes to the fine-tuning of plant defenses in response to pathogens [91,92]. Crespo-Salvador et al. (2018) demonstrated that differentially expressed genes responding to B. cinerea infection were induced 24 h after infection and displayed similar histone modification patterns, suggesting that epigenetics marks may impact the transcriptional regulation in this particular pathosystem [93]. A. thaliana hypomethylated mutants showed resistance against the biotrophic pathogen Hyaloperonospora arabidopsidis, whereas hypermethylated mutants were more susceptible to this pathogen [92,94]. Additionally, research on geminivirus infections has shown that the SUVH4 protein is a virus-silencing target that aims to evade host defenses [95]. Interestingly, our results highlighted the presence of SUVH4 (AT5G13960) and ATX2 (AT1G05830), which are involved in the maintenance of epigenetic control [96].

5. Conclusions

Core gene clusters that respond to multiple fungal infections were identified. However, further studies are needed to elucidate the role of several important genes in plant immune responses. Identifying coexpressed gene associations may provide insights into the functions of currently unknown coding genes. In the context of climate change, genomic data repositories are important resources for developing new tools for crop improvement and disease prevention, positively influencing productivity, sustainability and food security.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/genes14122223/s1: Figure S1: Venn diagrams for nonexpressed genes; Figure S2: Distribution of atypical samples in infected-plant gene networks; Figure S3: Unique genes identified in infected-plant gene modules; Table S1: Sample information obtained from NCBI SRA; Table S2: Databases, Bioinformatic tools and source code; Table S3: Significant gene modules identified in healthy-control and fungal-infected samples; Table S4: Gene Ontology overrepresentation test summary; Table S5: Genes functional groups related to multiple fungal infections identified by DAVID; Data S1: Normalized expression data for control samples; Data S2: Normalized expression for infected-plant samples; and Data S3: Gene Ontology overrepresentation analysis results.

Author Contributions

Conceptualization, E.G.-C.; methodology, software, validation, data curation, and visualization, C.S.-C.; formal analysis, C.S.-C. and E.G.-C.; investigation, C.S.-C. and E.G.-C.; resources, E.G.-C. and K.L.C.; writing—original draft preparation, E.G.-C. and C.S.-C.; writing—review and editing, E.G.-C. and K.L.C.; supervision, E.G.-C.; project administration, E.G.-C.; funding acquisition, E.G.-C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Consejo Nacional de Humanidades, Ciencia y Tecnología (CONAHCYT), Mexico. Grant number INFR201601-269833, and scholarship 552999.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The code developed for this study is freely available via GitHub at https://github.com/cyntsc/RNA-Seq-raw-integration (accessed on 28 September 2023) (DOI 10.5281/zenodo.7076416).

Acknowledgments

The authors thank Adriana Quiroz-Moreno for providing technical advice and facilities throughout the project.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Avery, S.V.; Singleton, I.; Magan, N.; Goldman, G.H. The Fungal Threat to Global Food Security. Fungal Biol. 2019, 123, 555–557. [Google Scholar] [CrossRef] [PubMed]
  2. Kettles, G.J.; Luna, E. Food Security in 2044: How Do We Control the Fungal Threat? Fungal Biol. 2019, 123, 558–564. [Google Scholar] [CrossRef] [PubMed]
  3. Steinberg, G.; Gurr, S.J. Fungi, Fungicide Discovery and Global Food Security. Fungal Genet. Biol. 2020, 144, 103476. [Google Scholar] [CrossRef] [PubMed]
  4. Miedaner, T.; Juroszek, P. Climate Change Will Influence Disease Resistance Breeding in Wheat in Northwestern Europe. Theor. Appl. Genet. 2021, 134, 1771–1785. [Google Scholar] [CrossRef] [PubMed]
  5. Fisher, M.C.; Hawkins, N.J.; Sanglard, D.; Gurr, S.J. Worldwide Emergence of Resistance to Antifungal Drugs Challenges Human Health and Food Security. Science 2018, 360, 739–742. [Google Scholar] [CrossRef] [PubMed]
  6. Almeida, F.; Rodrigues, M.L.; Coelho, C. The Still Underestimated Problem of Fungal Diseases Worldwide. Front. Microbiol. 2019, 10, 214. [Google Scholar] [CrossRef] [PubMed]
  7. Chaloner, T.M.; Gurr, S.J.; Bebber, D.P. Plant Pathogen Infection Risk Tracks Global Crop Yields under Climate Change. Nat. Clim. Change 2021, 11, 710–715. [Google Scholar] [CrossRef]
  8. Bregaglio, S.; Donatelli, M.; Confalonieri, R. Fungal Infections of Rice, Wheat, and Grape in Europe in 2030–2050. Agron. Sustain. Dev. 2013, 33, 767–776. [Google Scholar] [CrossRef]
  9. O’Connell, R.J.; Thon, M.R.; Hacquard, S.; Amyotte, S.G.; Kleemann, J.; Torres, M.F.; Damm, U.; Buiate, E.A.; Epstein, L.; Alkan, N.; et al. Lifestyle Transitions in Plant Pathogenic Colletotrichum Fungi Deciphered by Genome and Transcriptome Analyses. Nat. Genet. 2012, 44, 1060–1065. [Google Scholar] [CrossRef]
  10. Mendgen, K.; Hahn, M. Plant Infection and the Establishment of Fungal Biotrophy. Trends Plant Sci. 2002, 7, 352–356. [Google Scholar] [CrossRef]
  11. Laluk, K.; Mengiste, T. Necrotroph Attacks on Plants: Wanton Destruction or Covert Extortion? Arab. Book 2010, 8, e0136. [Google Scholar] [CrossRef] [PubMed]
  12. Fatima, U.; Senthil-Kumar, M. Plant and Pathogen Nutrient Acquisition Strategies. Front. Plant Sci. 2015, 6, 750. [Google Scholar] [CrossRef] [PubMed]
  13. Dean, R.; Van Kan, J.A.L.; Pretorius, Z.A.; Hammond-Kosack, K.E.; Di Pietro, A.; Spanu, P.D.; Rudd, J.J.; Dickman, M.; Kahmann, R.; Ellis, J.; et al. The Top 10 Fungal Pathogens in Molecular Plant Pathology. Mol. Plant Pathol. 2012, 13, 414–430. [Google Scholar] [CrossRef] [PubMed]
  14. Yan, Y.; Yuan, Q.; Tang, J.; Huang, J.; Hsiang, T.; Wei, Y.; Zheng, L. Colletotrichum higginsianum as a Model for Understanding Host–Pathogen Interactions: A Review. Int. J. Mol. Sci. 2018, 19, 2142. [Google Scholar] [CrossRef] [PubMed]
  15. Muckenschnabel, I.; Goodman, B.A.; Williamson, B.; Lyon, G.D.; Deighton, N. Infection of Leaves of Arabidopsis thaliana by Botrytis cinerea: Changes in Ascorbic Acid, Free Radicals and Lipid Peroxidation Products. J. Exp. Bot. 2002, 53, 207–214. [Google Scholar] [CrossRef] [PubMed]
  16. Windram, O.; Madhou, P.; McHattie, S.; Hill, C.; Hickman, R.; Cooke, E.; Jenkins, D.J.; Penfold, C.A.; Baxter, L.; Breeze, E.; et al. Arabidopsis Defense against Botrytis cinerea: Chronology and Regulation Deciphered by High-Resolution Temporal Transcriptomic Analysis. Plant Cell 2012, 24, 3530–3557. [Google Scholar] [CrossRef]
  17. Veillet, F.; Gaillard, C.; Lemonnier, P.; Coutos-Thévenot, P.; La Camera, S. The Molecular Dialogue between Arabidopsis thaliana and the Necrotrophic Fungus Botrytis cinerea Leads to Major Changes in Host Carbon Metabolism. Sci. Rep. 2017, 7, 17121. [Google Scholar] [CrossRef]
  18. Jones, J.D.G.; Dangl, J.L. The Plant Immune System. Nature 2006, 444, 323–329. [Google Scholar] [CrossRef]
  19. Zipfel, C.; Robatzek, S.; Navarro, L.; Oakeley, E.J.; Jones, J.D.G.; Felix, G.; Boller, T. Bacterial Disease Resistance in Arabidopsis through Flagellin Perception. Nature 2004, 428, 764–767. [Google Scholar] [CrossRef]
  20. Navarro, L.; Jay, F.; Nomura, K.; He, S.Y.; Voinnet, O. Suppression of the MicroRNA Pathway by Bacterial Effector Proteins. Science 2008, 321, 964–967. [Google Scholar] [CrossRef]
  21. Vleeshouwers, V.G.A.A.; Oliver, R.P. Effectors as Tools in Disease Resistance Breeding Against Biotrophic, Hemibiotrophic, and Necrotrophic Plant Pathogens. MPMI 2014, 27, 196–206. [Google Scholar] [CrossRef] [PubMed]
  22. Landete, J.M. Effector Molecules and Regulatory Proteins: Applications. Trends Biotechnol. 2016, 34, 777–780. [Google Scholar] [CrossRef] [PubMed]
  23. Khan, M.; Seto, D.; Subramaniam, R.; Desveaux, D. Oh, the Places They’ll Go! A Survey of Phytopathogen Effectors and Their Host Targets. Plant J. 2018, 93, 651–663. [Google Scholar] [CrossRef] [PubMed]
  24. Koeck, M.; Hardham, A.R.; Dodds, P.N. The Role of Effectors of Biotrophic and Hemibiotrophic Fungi in Infection: Effectors of Biotrophic Fungi. Cell. Microbiol. 2011, 13, 1849–1857. [Google Scholar] [CrossRef] [PubMed]
  25. Bigeard, J.; Colcombet, J.; Hirt, H. Signaling Mechanisms in Pattern-Triggered Immunity (PTI). Mol. Plant 2015, 8, 521–539. [Google Scholar] [CrossRef]
  26. Thomma, B.P.H.J.; Nürnberger, T.; Joosten, M.H.A.J. Of PAMPs and Effectors: The Blurred PTI-ETI Dichotomy. Plant Cell 2011, 23, 4–15. [Google Scholar] [CrossRef]
  27. Tsuda, K.; Katagiri, F. Comparing Signaling Mechanisms Engaged in Pattern-Triggered and Effector-Triggered Immunity. Curr. Opin. Plant Biol. 2010, 13, 459–465. [Google Scholar] [CrossRef]
  28. Dodds, P.N.; Rathjen, J.P. Plant Immunity: Towards an Integrated View of Plant–Pathogen Interactions. Nat. Rev. Genet. 2010, 11, 539–548. [Google Scholar] [CrossRef]
  29. GenBank and WGS Statistics. Available online: https://www.ncbi.nlm.nih.gov/genbank/statistics/ (accessed on 30 August 2022).
  30. Vandenbon, A. Evaluation of Critical Data Processing Steps for Reliable Prediction of Gene Co-Expression from Large Collections of RNA-Seq Data. PLoS ONE 2022, 17, e0263344. [Google Scholar] [CrossRef]
  31. Crandall, S.G.; Gold, K.M.; del Mar Jiménez-Gasco, M.; Filgueiras, C.C.; Willett, D.S. A Multi-Omics Approach to Solving Problems in Plant Disease Ecology. PLoS ONE 2020, 15, e0237975. [Google Scholar] [CrossRef]
  32. Subramanian, I.; Verma, S.; Kumar, S.; Jere, A.; Anamika, K. Multi-Omics Data Integration, Interpretation, and Its Application. Bioinform. Biol. Insights 2020, 14, 1177932219899051. [Google Scholar] [CrossRef] [PubMed]
  33. Bhadauria, V.; Vijayan, P.; Wei, Y.; Banniza, S. Transcriptome Analysis Reveals a Complex Interplay between Resistance and Effector Genes during the Compatible Lentil-Colletotrichum Lentis Interaction. Sci. Rep. 2017, 7, 42338. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, L.; Huang, X.; He, C.; Zhang, Q.-Y.; Zou, X.; Duan, K.; Gao, Q. Novel Fungal Pathogenicity and Leaf Defense Strategies Are Revealed by Simultaneous Transcriptome Analysis of Colletotrichum Fructicola and Strawberry Infected by This Fungus. Front. Plant Sci. 2018, 9, 434. [Google Scholar] [CrossRef] [PubMed]
  35. Pérez-Torres, C.-A.; Ibarra-Laclette, E.; Hernández-Domínguez, E.-E.; Rodríguez-Haas, B.; Pérez-Lira, A.-J.; Villafán, E.; Alonso-Sánchez, A.; de Jesus García-Ávila, C.; Ramírez-Pool, J.-A.; Sánchez-Rangel, D. Molecular Evidence of the Avocado Defense Response to Fusarium Kuroshium Infection: A Deep Transcriptome Analysis Using RNA-Seq. PeerJ 2021, 9, e11215. [Google Scholar] [CrossRef] [PubMed]
  36. Xu, X.-H.; Wang, C.; Li, S.-X.; Su, Z.-Z.; Zhou, H.-N.; Mao, L.-J.; Feng, X.-X.; Liu, P.-P.; Chen, X.; Hugh Snyder, J.; et al. Friend or Foe: Differential Responses of Rice to Invasion by Mutualistic or Pathogenic Fungi Revealed by RNAseq and Metabolite Profiling. Sci. Rep. 2015, 5, 13624. [Google Scholar] [CrossRef] [PubMed]
  37. Takahara, H.; Dolf, A.; Endl, E.; O’Connell, R. Flow Cytometric Purification of Colletotrichum higginsianum Biotrophic Hyphae from Arabidopsis Leaves for Stage-specific Transcriptome Analysis. Plant J. 2009, 59, 672–683. [Google Scholar] [CrossRef]
  38. Mulema, J.M.K.; Denby, K.J. Spatial and Temporal Transcriptomic Analysis of the Arabidopsis thalianaBotrytis cinerea Interaction. Mol. Biol. Rep. 2012, 39, 4039–4049. [Google Scholar] [CrossRef]
  39. La Camera, S.; L’Haridon, F.; Astier, J.; Zander, M.; Abou-Mansour, E.; Page, G.; Thurow, C.; Wendehenne, D.; Gatz, C.; Métraux, J.-P.; et al. The Glutaredoxin ATGRXS13 Is Required to Facilitate Botrytis cinerea Infection of Arabidopsis thaliana Plants: Role of ATGRXS13 during B. cinerea Infection. Plant J. 2011, 68, 507–519. [Google Scholar] [CrossRef]
  40. Narusaka, Y.; Narusaka, M.; Park, P.; Kubo, Y.; Hirayama, T.; Seki, M.; Shiraishi, T.; Ishida, J.; Nakashima, M.; Enju, A.; et al. RCH1, a Locus in Arabidopsis That Confers Resistance to the Hemibiotrophic Fungal Pathogen Colletotrichum higginsianum. MPMI 2004, 17, 749–762. [Google Scholar] [CrossRef]
  41. Takahara, H.; Hacquard, S.; Kombrink, A.; Hughes, H.B.; Halder, V.; Robin, G.P.; Hiruma, K.; Neumann, U.; Shinya, T.; Kombrink, E.; et al. Colletotrichum higginsianum Extracellular LysM Proteins Play Dual Roles in Appressorial Function and Suppression of Chitin-triggered Plant Immunity. New Phytol. 2016, 211, 1323–1337. [Google Scholar] [CrossRef]
  42. Amrine, K.C.H.; Blanco-Ulate, B.; Cantu, D. Discovery of Core Biotic Stress Responsive Genes in Arabidopsis by Weighted Gene Co-Expression Network Analysis. PLoS ONE 2015, 10, e0118731. [Google Scholar] [CrossRef] [PubMed]
  43. Kumar, A.; Kanak, K.R.; Arunachalam, A.; Dass, R.S.; Lakshmi, P.T.V. Comparative Transcriptome Profiling and Weighted Gene Co-Expression Network Analysis to Identify Core Genes in Maize (Zea mays L.) Silks Infected by Multiple Fungi. Front. Plant Sci. 2022, 13, 985396. [Google Scholar] [CrossRef] [PubMed]
  44. Langfelder, P.; Horvath, S. WGCNA: An R Package for Weighted Correlation Network Analysis. BMC Bioinform. 2008, 9, 559. [Google Scholar] [CrossRef] [PubMed]
  45. Serin, E.A.R.; Nijveen, H.; Hilhorst, H.W.M.; Ligterink, W. Learning from Co-Expression Networks: Possibilities and Challenges. Front. Plant Sci. 2016, 7, 444. [Google Scholar] [CrossRef] [PubMed]
  46. Han, X.; Zhang, Y.-W.; Liu, J.-Y.; Zuo, J.-F.; Zhang, Z.-C.; Guo, L.; Zhang, Y.-M. 4D Genetic Networks Reveal the Genetic Basis of Metabolites and Seed Oil-Related Traits in 398 Soybean RILs. Biotechnol. Biofuels 2022, 15, 92. [Google Scholar] [CrossRef] [PubMed]
  47. Sari, E.; Cabral, A.L.; Polley, B.; Tan, Y.; Hsueh, E.; Konkin, D.J.; Knox, R.E.; Ruan, Y.; Fobert, P.R. Weighted Gene Co-Expression Network Analysis Unveils Gene Networks Associated with the Fusarium Head Blight Resistance in Tetraploid Wheat. BMC Genom. 2019, 20, 925. [Google Scholar] [CrossRef] [PubMed]
  48. Childs, K.L.; Davidson, R.M.; Buell, C.R. Gene Coexpression Network Analysis as a Source of Functional Annotation for Rice Genes. PLoS ONE 2011, 6, e22196. [Google Scholar] [CrossRef]
  49. Waskom, M. Seaborn: Statistical Data Visualization. JOSS 2021, 6, 3021. [Google Scholar] [CrossRef]
  50. Sherman, B.T.; Hao, M.; Qiu, J.; Jiao, X.; Baseler, M.W.; Lane, H.C.; Imamichi, T.; Chang, W. DAVID: A Web Server for Functional Enrichment Analysis and Functional Annotation of Gene Lists (2021 Update). Nucleic Acids Res. 2022, 50, gkac194. [Google Scholar] [CrossRef]
  51. Robin, G.P.; Kleemann, J.; Neumann, U.; Cabre, L.; Dallery, J.-F.; Lapalu, N.; O’Connell, R.J. Subcellular Localization Screening of Colletotrichum higginsianum Effector Candidates Identifies Fungal Proteins Targeted to Plant Peroxisomes, Golgi Bodies, and Microtubules. Front. Plant Sci. 2018, 9, 562. [Google Scholar] [CrossRef]
  52. Badet, T.; Voisin, D.; Mbengue, M.; Barascud, M.; Sucher, J.; Sadon, P.; Balagué, C.; Roby, D.; Raffaele, S. Parallel Evolution of the POQR Prolyl Oligo Peptidase Gene Conferring Plant Quantitative Disease Resistance. PLOS Genet. 2017, 13, e1007143. [Google Scholar] [CrossRef] [PubMed]
  53. Andrews, S. FastQC. A Quality Control Tool for High Throughput Sequence Data. 2010. Available online: https://www.bioinformatics.babraham.ac.uk/projects/fastqc/ (accessed on 8 February 2021).
  54. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A Flexible Trimmer for Illumina Sequence Data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [PubMed]
  55. Lamesch, P.; Berardini, T.Z.; Li, D.; Swarbreck, D.; Wilks, C.; Sasidharan, R.; Muller, R.; Dreher, K.; Alexander, D.L.; Garcia-Hernandez, M.; et al. The Arabidopsis Information Resource (TAIR): Improved Gene Annotation and New Tools. Nucleic Acids Res. 2012, 40, D1202–D1210. [Google Scholar] [CrossRef] [PubMed]
  56. Dobin, A.; Davis, C.A.; Schlesinger, F.; Drenkow, J.; Zaleski, C.; Jha, S.; Batut, P.; Chaisson, M.; Gingeras, T.R. STAR: Ultrafast Universal RNA-Seq Aligner. Bioinformatics 2013, 29, 15–21. [Google Scholar] [CrossRef] [PubMed]
  57. Chang, N.; Sun, Q.; Hu, J.; An, C.; Gao, A.H. Large Introns of 5 to 10 Kilo Base Pairs Can Be Spliced out in Arabidopsis. Genes 2017, 8, 200. [Google Scholar] [CrossRef]
  58. Li, H.; Handsaker, B.; Wysoker, A.; Fennell, T.; Ruan, J.; Homer, N.; Marth, G.; Abecasis, G.; Durbin, R. 1000 Genome Project Data Processing Subgroup The Sequence Alignment/Map Format and SAMtools. Bioinformatics 2009, 25, 2078–2079. [Google Scholar] [CrossRef] [PubMed]
  59. Anders, S.; Pyl, P.T.; Huber, W. HTSeq—A Python Framework to Work with High-Throughput Sequencing Data. Bioinformatics 2015, 31, 166–169. [Google Scholar] [CrossRef]
  60. Zhao, Y.; Li, M.-C.; Konaté, M.M.; Chen, L.; Das, B.; Karlovich, C.; Williams, P.M.; Evrard, Y.A.; Doroshow, J.H.; McShane, L.M. TPM, FPKM, or Normalized Counts? A Comparative Study of Quantification Measures for the Analysis of RNA-Seq Data from the NCI Patient-Derived Models Repository. J. Transl. Med. 2021, 19, 269. [Google Scholar] [CrossRef]
  61. Renesh, B. Reneshbedre/Bioinfokit: Bioinformatics Data Analysis and Visualization Toolkit|Zenodo. Available online: https://zenodo.org/record/3965241 (accessed on 21 January 2022).
  62. Mi, H.; Thomas, P. PANTHER Pathway: An Ontology-Based Pathway Database Coupled with Data Analysis Tools. Methods Mol. Biol. 2009, 563, 123–140. [Google Scholar] [CrossRef]
  63. Mi, H.; Muruganujan, A.; Ebert, D.; Huang, X.; Thomas, P.D. PANTHER Version 14: More Genomes, a New PANTHER GO-Slim and Improvements in Enrichment Analysis Tools. Nucleic Acids Res. 2019, 47, D419–D426. [Google Scholar] [CrossRef]
  64. The Gene Ontology Consortium. Expansion of the Gene Ontology Knowledgebase and Resources. Nucleic Acids Res. 2017, 45, D331–D338. [Google Scholar] [CrossRef] [PubMed]
  65. The Gene Ontology Consortium. The Gene Ontology Resource: Enriching a GOld Mine. Nucleic Acids Res. 2021, 49, D325–D334. [Google Scholar] [CrossRef] [PubMed]
  66. Huang, D.W.; Sherman, B.T.; Tan, Q.; Collins, J.R.; Alvord, W.G.; Roayaei, J.; Stephens, R.; Baseler, M.W.; Lane, H.C.; Lempicki, R.A. The DAVID Gene Functional Classification Tool: A Novel Biological Module-Centric Algorithm to Functionally Analyze Large Gene Lists. Genome Biol. 2007, 8, R183. [Google Scholar] [CrossRef] [PubMed]
  67. Pasha, A.; Subramaniam, S.; Cleary, A.; Chen, X.; Berardini, T.; Farmer, A.; Town, C.; Provart, N. Araport Lives: An Updated Framework for Arabidopsis Bioinformatics. Plant Cell 2020, 32, 2683–2686. [Google Scholar] [CrossRef] [PubMed]
  68. Krishnakumar, V.; Hanlon, M.R.; Contrino, S.; Ferlanti, E.S.; Karamycheva, S.; Kim, M.; Rosen, B.D.; Cheng, C.-Y.; Moreira, W.; Mock, S.A.; et al. Araport: The Arabidopsis Information Portal. Nucleic Acids Res. 2015, 43, D1003–D1009. [Google Scholar] [CrossRef] [PubMed]
  69. Perkel, J.M. Why Jupyter Is Data Scientists’ Computational Notebook of Choice. Nature 2018, 563, 145–146. [Google Scholar] [CrossRef] [PubMed]
  70. Zhang, Y.; Parmigiani, G.; Johnson, W.E. ComBat-Seq: Batch Effect Adjustment for RNA-Seq Count Data. NAR Genom. Bioinform. 2020, 2, lqaa078. [Google Scholar] [CrossRef]
  71. Leek, J.T.; Johnson, W.E.; Parker, H.S.; Jaffe, A.E.; Storey, J.D. The Sva Package for Removing Batch Effects and Other Unwanted Variation in High-Throughput Experiments. Bioinformatics 2012, 28, 882–883. [Google Scholar] [CrossRef]
  72. Leek, J.T. Svaseq: Removing Batch Effects and Other Unwanted Noise from Sequencing Data. Nucleic Acids Res. 2014, 42, e161. [Google Scholar] [CrossRef]
  73. Backlund, M.; Stein, F.; Rettel, M.; Schwarzl, T.; Perez-Perri, J.I.; Brosig, A.; Zhou, Y.; Neu-Yilik, G.; Hentze, M.W.; Kulozik, A.E. Plasticity of Nuclear and Cytoplasmic Stress Responses of RNA-Binding Proteins. Nucleic Acids Res. 2020, 48, 4725–4740. [Google Scholar] [CrossRef]
  74. Liang, X.; Zhang, J. Regulation of Plant Responses to Biotic and Abiotic Stress by Receptor-like Cytoplasmic Kinases. Stress Biol. 2022, 2, 25. [Google Scholar] [CrossRef] [PubMed]
  75. van Nocker, S.; Ludwig, P. The WD-Repeat Protein Superfamily in Arabidopsis: Conservation and Divergence in Structure and Function. BMC Genom. 2003, 4, 50. [Google Scholar] [CrossRef] [PubMed]
  76. Li, Q.; Zhao, P.; Li, J.; Zhang, C.; Wang, L.; Ren, Z. Genome-Wide Analysis of the WD-Repeat Protein Family in Cucumber and Arabidopsis. Mol. Genet. Genom. 2014, 289, 103–124. [Google Scholar] [CrossRef] [PubMed]
  77. Miller, J.C.; Chezem, W.R.; Clay, N.K. Ternary WD40 Repeat-Containing Protein Complexes: Evolution, Composition and Roles in Plant Immunity. Front. Plant Sci. 2016, 6, 1108. [Google Scholar] [CrossRef] [PubMed]
  78. Rogowska, A.; Szakiel, A. The Role of Sterols in Plant Response to Abiotic Stress. Phytochem. Rev. 2020, 19, 1525–1538. [Google Scholar] [CrossRef]
  79. Du, Y.; Fu, X.; Chu, Y.; Wu, P.; Liu, Y.; Ma, L.; Tian, H.; Zhu, B. Biosynthesis and the Roles of Plant Sterols in Development and Stress Responses. Int. J. Mol. Sci. 2022, 23, 2332. [Google Scholar] [CrossRef] [PubMed]
  80. Aboobucker, S.I.; Suza, W.P. Why Do Plants Convert Sitosterol to Stigmasterol? Front. Plant Sci. 2019, 10, 354. [Google Scholar] [CrossRef]
  81. Zhang, W.; Wang, S.; Yang, J.; Kang, C.; Huang, L.; Guo, L. Glycosylation of Plant Secondary Metabolites: Regulating from Chaos to Harmony. Environ. Exp. Bot. 2022, 194, 104703. [Google Scholar] [CrossRef]
  82. Lim, E.-K.; Doucet, C.J.; Li, Y.; Elias, L.; Worrall, D.; Spencer, S.P.; Ross, J.; Bowles, D.J. The Activity of ArabidopsisGlycosyltransferases toward Salicylic Acid, 4-Hydroxybenzoic Acid, and Other Benzoates. J. Biol. Chem. 2002, 277, 586–592. [Google Scholar] [CrossRef]
  83. Huang, X.; Zhu, G.; Liu, Q.; Chen, L.; Li, Y.; Hou, B. Modulation of Plant Salicylic Acid-Associated Immune Responses via Glycosylation of Dihydroxybenzoic Acids. Plant Physiol. 2018, 176, 3103–3119. [Google Scholar] [CrossRef]
  84. Noman, A.; Aqeel, M.; Khalid, N.; Islam, W.; Sanaullah, T.; Anwar, M.; Khan, S.; Ye, W.; Lou, Y. Zinc Finger Protein Transcription Factors: Integrated Line of Action for Plant Antimicrobial Activity. Microb. Pathog. 2019, 132, 141–149. [Google Scholar] [CrossRef] [PubMed]
  85. Zang, D.; Li, H.; Xu, H.; Zhang, W.; Zhang, Y.; Shi, X.; Wang, Y. An Arabidopsis Zinc Finger Protein Increases Abiotic Stress Tolerance by Regulating Sodium and Potassium Homeostasis, Reactive Oxygen Species Scavenging and Osmotic Potential. Front. Plant Sci. 2016, 7, 1272. [Google Scholar] [CrossRef] [PubMed]
  86. Davletova, S.; Schlauch, K.; Coutu, J.; Mittler, R. The Zinc-Finger Protein Zat12 Plays a Central Role in Reactive Oxygen and Abiotic Stress Signaling in Arabidopsis. Plant Physiol. 2005, 139, 847–856. [Google Scholar] [CrossRef] [PubMed]
  87. Brown, S.H.; Yarden, O.; Gollop, N.; Chen, S.; Zveibil, A.; Belausov, E.; Freeman, S. Differential Protein Expression in Colletotrichum Acutatum: Changes Associated with Reactive Oxygen Species and Nitrogen Starvation Implicated in Pathogenicity on Strawberry. Mol. Plant Pathol. 2008, 9, 171–190. [Google Scholar] [CrossRef] [PubMed]
  88. Foley, R.C.; Kidd, B.N.; Hane, J.K.; Anderson, J.P.; Singh, K.B. Reactive Oxygen Species Play a Role in the Infection of the Necrotrophic Fungi, Rhizoctonia Solani in Wheat. PLoS ONE 2016, 11, e0152548. [Google Scholar] [CrossRef]
  89. Brzezinka, K.; Altmann, S.; Czesnick, H.; Nicolas, P.; Gorka, M.; Benke, E.; Kabelitz, T.; Jähne, F.; Graf, A.; Kappel, C.; et al. Arabidopsis FORGETTER1 Mediates Stress-Induced Chromatin Memory through Nucleosome Remodeling. eLife 2016, 5, e17061. [Google Scholar] [CrossRef]
  90. Driedonks, N.; Xu, J.; Peters, J.L.; Park, S.; Rieu, I. Multi-Level Interactions Between Heat Shock Factors, Heat Shock Proteins, and the Redox System Regulate Acclimation to Heat. Front. Plant Sci. 2015, 6, 999. [Google Scholar] [CrossRef]
  91. Ding, B.; Wang, G.-L. Chromatin versus Pathogens: The Function of Epigenetics in Plant Immunity. Front. Plant Sci. 2015, 6, 675. [Google Scholar] [CrossRef]
  92. Alonso, C.; Ramos-Cruz, D.; Becker, C. The Role of Plant Epigenetics in Biotic Interactions. New Phytol. 2019, 221, 731–737. [Google Scholar] [CrossRef]
  93. Crespo-Salvador, Ó.; Escamilla-Aguilar, M.; López-Cruz, J.; López-Rodas, G.; González-Bosch, C. Determination of Histone Epigenetic Marks in Arabidopsis and Tomato Genes in the Early Response to Botrytis cinerea. Plant Cell Rep. 2018, 37, 153–166. [Google Scholar] [CrossRef]
  94. López Sánchez, A.; Stassen, J.H.M.; Furci, L.; Smith, L.M.; Ton, J. The Role of DNA (de)Methylation in Immune Responsiveness of Arabidopsis. Plant J. 2016, 88, 361–374. [Google Scholar] [CrossRef] [PubMed]
  95. Castillo-González, C.; Liu, X.; Huang, C.; Zhao, C.; Ma, Z.; Hu, T.; Sun, F.; Zhou, Y.; Zhou, X.; Wang, X.-J.; et al. Geminivirus-Encoded TrAP Suppressor Inhibits the Histone Methyltransferase SUVH4/KYP to Counter Host Defense. eLife 2015, 4, e06671. [Google Scholar] [CrossRef] [PubMed]
  96. Malagnac, F. An Arabidopsis SET Domain Protein Required for Maintenance but Not Establishment of DNA Methylation. EMBO J. 2002, 21, 6842–6852. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Framework for integrating publicly available RNA-seq data. (a) Step 1. Raw counts are integrated in a single expression matrix and evaluation metrics (standard deviation (σ), mean (µ) and range (R)) are calculated. Genes with zero expression values are removed. (b) Step 2. Data arrays are normalized by TPM and evaluation metrics are recalculated. (c) Step 3. Percentiles and distribution are calculated to identify tail values within the distribution. TPM values below the 1st and above the 99th percentile are removed, and evaluation metrics are recalculated. (d) Step 4. Normalized and filtered TPM values are transformed to log2(TMP + 1) to identify atypical distributions, and evaluation metrics are recalculated.
Figure 1. Framework for integrating publicly available RNA-seq data. (a) Step 1. Raw counts are integrated in a single expression matrix and evaluation metrics (standard deviation (σ), mean (µ) and range (R)) are calculated. Genes with zero expression values are removed. (b) Step 2. Data arrays are normalized by TPM and evaluation metrics are recalculated. (c) Step 3. Percentiles and distribution are calculated to identify tail values within the distribution. TPM values below the 1st and above the 99th percentile are removed, and evaluation metrics are recalculated. (d) Step 4. Normalized and filtered TPM values are transformed to log2(TMP + 1) to identify atypical distributions, and evaluation metrics are recalculated.
Genes 14 02223 g001
Figure 2. Rates of read alignments against the Arabidopsis thaliana genome. Percentage of unique reads mapped to the A. thaliana genome (TAIR11) from healthy-control and infected-plant samples.
Figure 2. Rates of read alignments against the Arabidopsis thaliana genome. Percentage of unique reads mapped to the A. thaliana genome (TAIR11) from healthy-control and infected-plant samples.
Genes 14 02223 g002
Figure 3. Integration of RNA-seq datasets. (a) The expression matrix was normalized to TPM and evaluation metrics were calculated, µ = 41.2, σ = 284.5 ± 84.5, and R = [0 to 30,000]. (b) Percentiles and distribution were calculated to identify tail values within the distribution. TPM values below the 1st percentile (TPM < 0.1) and above the 99th percentile (TPM > 845) were removed. Evaluation metrics were calculated (µ = 27.21 ± 6.25, σ = 62.5 ± 4.5 and R = [0 to 845]). (c) TPM values were transformed to log2(TMP + 1) to reduce the scale and to identify atypical distributions. Evaluation metrics were calculated (σ = 2.93 ± 0.78, µ = 2.13 ± 0.13, and R = [0 to 9.72]). (d) Fungal-infected samples with atypical distributions were removed (Ss30, Ss30.1 and Ss30.2) and evaluation metrics were calculated (σ = 2.075 ± 0.075, µ = 3.55 ± 0.16, and R = [0 to 9.72]). The colored lines represent the RNA-seq data used in this study; colored bars represent percentiles.
Figure 3. Integration of RNA-seq datasets. (a) The expression matrix was normalized to TPM and evaluation metrics were calculated, µ = 41.2, σ = 284.5 ± 84.5, and R = [0 to 30,000]. (b) Percentiles and distribution were calculated to identify tail values within the distribution. TPM values below the 1st percentile (TPM < 0.1) and above the 99th percentile (TPM > 845) were removed. Evaluation metrics were calculated (µ = 27.21 ± 6.25, σ = 62.5 ± 4.5 and R = [0 to 845]). (c) TPM values were transformed to log2(TMP + 1) to reduce the scale and to identify atypical distributions. Evaluation metrics were calculated (σ = 2.93 ± 0.78, µ = 2.13 ± 0.13, and R = [0 to 9.72]). (d) Fungal-infected samples with atypical distributions were removed (Ss30, Ss30.1 and Ss30.2) and evaluation metrics were calculated (σ = 2.075 ± 0.075, µ = 3.55 ± 0.16, and R = [0 to 9.72]). The colored lines represent the RNA-seq data used in this study; colored bars represent percentiles.
Genes 14 02223 g003
Figure 4. Co-expression network and clustering. The Scale-Free-Topology model was used to assess healthy-control samples and infected-plant gene networks. A signed network was built using Pearson’s method. (a) Control -sample gene network reached a maximum of R2 = 0.80. The soft threshold cutoff was set at β = 28 for network discovery, which resulted in R2 = 0.78 and a Node Mean Connectivity (NMC) of 374 genes. The merged network comprised 23 gene clusters. (b) Infected-plant gene network reached a maximum of R2 = 0.83. The soft threshold cutoff was β = 27, which resulted in a model fit of R2 = 0.79 with an NCM of 270 genes. The merged network contained 36 gene clusters.
Figure 4. Co-expression network and clustering. The Scale-Free-Topology model was used to assess healthy-control samples and infected-plant gene networks. A signed network was built using Pearson’s method. (a) Control -sample gene network reached a maximum of R2 = 0.80. The soft threshold cutoff was set at β = 28 for network discovery, which resulted in R2 = 0.78 and a Node Mean Connectivity (NMC) of 374 genes. The merged network comprised 23 gene clusters. (b) Infected-plant gene network reached a maximum of R2 = 0.83. The soft threshold cutoff was β = 27, which resulted in a model fit of R2 = 0.79 with an NCM of 270 genes. The merged network contained 36 gene clusters.
Genes 14 02223 g004
Figure 5. Gene Ontology (GO) overrepresentation test. (a) Bar plots of overrepresented molecular function GO-slim classes in Coral3 gene module and overrepresented biological process GO-slim classes in Navajowhite3 gene module obtained for control sample gene network. (b) Bar plot of overrepresented biological process GO-slim classes in Blue3 gene module obtained for control sample gene network. (c) Bar plots of overrepresented molecular function GO-slim classes in Chocolate, Chocolate2 and Green3 gene modules obtained for infected-plant gene expression network.
Figure 5. Gene Ontology (GO) overrepresentation test. (a) Bar plots of overrepresented molecular function GO-slim classes in Coral3 gene module and overrepresented biological process GO-slim classes in Navajowhite3 gene module obtained for control sample gene network. (b) Bar plot of overrepresented biological process GO-slim classes in Blue3 gene module obtained for control sample gene network. (c) Bar plots of overrepresented molecular function GO-slim classes in Chocolate, Chocolate2 and Green3 gene modules obtained for infected-plant gene expression network.
Genes 14 02223 g005
Figure 6. Core gene module. (a) The heatmap shows the different treatments analyzed in the study. The rectangle highlights the treatments with significant correlation. B_24 hpi represents the sample with B. cinerea at 24 h pos infection. Ch_22 hpi corresponds to the sample with C. higginsianum at 22 h postinfection. (b) Selected modules are displayed. The ellipses represent the consensus module “Darkmagenta”, with a R2 > 0.59 for C. higginsianum and a R2 > 0.57 for B. cinerea.
Figure 6. Core gene module. (a) The heatmap shows the different treatments analyzed in the study. The rectangle highlights the treatments with significant correlation. B_24 hpi represents the sample with B. cinerea at 24 h pos infection. Ch_22 hpi corresponds to the sample with C. higginsianum at 22 h postinfection. (b) Selected modules are displayed. The ellipses represent the consensus module “Darkmagenta”, with a R2 > 0.59 for C. higginsianum and a R2 > 0.57 for B. cinerea.
Genes 14 02223 g006
Table 1. Description of public RNA-Seq datasets downloaded from the SRA.
Table 1. Description of public RNA-Seq datasets downloaded from the SRA.
BioProjectSample IDHPI 1RunLayoutReads (M 2)% Clean Reads% Aligned Reads
PRJNA148307Ch2222SRR364389SE12.692.1287.17
Ch22.1SRR364390SE12.492.0377.45
Ch22.2SRR364391SE12.492.1877.4
Ch22.3SRR364392SE12.292.0887.29
Ch4040SRR364400SE11.991.4082.08
Ch40.1SRR364401SE11.991.4082.25
Ch40.2SRR364398SE13.292.5580.25
Ch40.3SRR364399SE13.292.7380.25
PRJNA315516
PRJNA593073
Bc1212SRR3383696SE12.1100.0098.2
Bc12.1SRR3383697SE15100.0098.14
Bc1818SRR3383779SE10.397.3793.55
Bc18.1SRR3383780SE13.697.2695.2
Bc2424SRR10586397PE22.295.5389.07
Bc24.1SRR10586399PE2295.7088.69
PRJNA418121Ss3030SRR6283146SE20.895.0136.48
Ss30.1SRR6283147SE20.996.0932.35
Ss30.2SRR6283148SE21.292.9029.5
PRJNA315516
PRJNA418121
healthy1212SRR3383640SE10.997.1298.31
healthy12.1SRR3383641SE22.697.5098.34
healthy1818SRR3383782SE29.897.6298.13
healthy18.1SRR3383783SE14.297.6998.34
healthy2424SRR3383821SE1597.1898.09
healthy24.1SRR3383822SE10.295.9897.87
healthy3030SRR6283144SE22.195.3197.51
healthy30.1SRR6283145SE19.795.7297.99
1 Hours Post Inoculation; 2 Million reads.
Table 2. TOM network results and the Node-Connectivity Mean (NCM).
Table 2. TOM network results and the Node-Connectivity Mean (NCM).
ConditionR2# Clusters of
Standard Network
NCM# Clusters of
Merged Network
Healthy-control samplesβ = 28; 0.7823737423
Infected-plant samples β = 27; 0.7910027036
Table 3. Enriched functional groups from “Darkmagenta” gene module.
Table 3. Enriched functional groups from “Darkmagenta” gene module.
CategoryTerm (Group #)p ValueP Benjamin#GenesGeneE. Score
IPR015943WD40/YVTN repeat-like-containing
(group 1)
0.00290.53077AT5G23430, AT3G13340, AT3G18060, AT5G51980, VPS11, AT1G79990, VCS1.5527
KW-1207
KW-0752
GO:0016126
Sterol metabolism
(group 2)
0.0125
0.0275
0.0277
0.4613
0.7459
0.9995
3SMT1, SMO1-1, 3BETAHSD/D11.3511
GO:0016757
KW-0328
Glycosyltransferase
(group 3)
0.0113
0.0346
0.8358
0.7467
10ALG3, FUT13, CALS1, UK/UPRT1, AT4G38040, GUT2, PARVUS, SETH2, AT5G45660, AT1G342701.3489
KW-0968
GO:0006886
Intracellular protein transport
(group 4)
0.0066
0.0428
0.1592
0.9999
7PLA2-α, PAT2, AT1G60070, AT4G13730, AT1G14910, VPS11, AT1G799901.1112
IPR011011
IPR019787
Zinc finger, FYVE/PHD-type
(group 5)
0.0190
0.0477
0.9928
0.9999
4EMB1135, ATX2, AT5G12350, AT1G506201.0173
KW-0489Methyltransferase
(group 6)
0.04030.79916AT3G15530, SUVH4, AT2G34300, SMT1, ATX2, AT5G060500.8795
Table 4. Description of core genes involved in multiple fungal infections.
Table 4. Description of core genes involved in multiple fungal infections.
GroupTAIR IDGeneChrDescription
1AT5G23430AT5G23430Chr5Transducin/WD40 repeat-like superfamily protein.
1AT3G13340AT3G13340Chr3Transducin/WD40 repeat-like superfamily protein
1AT3G18060AT3G18060Chr3Transducin/WD-40 repeat family protein
1AT5G51980AT5G51980Chr5Transducin/WD40 repeat-like superfamily protein
1–4AT1G79990AT1G79990Chr1Coatomer subunit β-2. WD repeat COPB2 family
1–4AT2G05170VPS11Chr2Vacuolar protein sorting 11
1AT3G13300VCSChr3VARICOSE. Transducin/WD40 repeat-like superfamily protein
2–6AT5G13710SMT1Chr5Sterol methyltransferase 1
2AT4G12110SMO1-1Chr4Sterol-4alpha-methyl oxidase 1-1
2AT1G472903BETAHSD/D1Chr13beta-hydroxysteroid-dehydrogenase/decarboxylase isoform 1
3AT2G47760ALG3Chr2Asparagine-linked glycosylation 3
3AT1G71990FUT13Chr1Fucosyltransferase 13
3AT1G05570CALS1Chr1Callose synthase 1
3AT5G40870UK/UPRT1Chr5Uridine kinase/uracil phosphoribosyltransferase 1
3AT1G27440GUT2Chr1Exostosin family protein
3AT5G12350AT5G12350Chr5Exostosin family protein(AT4G38040)
3AT1G05570CALS1Chr1Uridine kinase/uracil phosphoribosyltransferase 1(UK/UPRT1)
3AT1G19300PARVUSChr1Nucleotide-diphospho-sugar transferases superfamily
3AT3G45100SETH2Chr3UDP-Glycosyltransferase superfamily protein
3AT4G38040AT4G38040Chr4Exotosin family protein. Glycosyltransferase 47 family
3AT5G45660AT5G45660Chr5Adenine phophoribosyltransferase
3AT1G34270AT1G34270Chr1Exotosin family protein
4AT2G06925PLA2-αChr2Phospholipase A2 family protein
4AT3G55480PAT2Chr3Protein affected trafficking 2
4AT1G60070AT1G60070Chr1Adaptor protein complex AP-1
4AT4G13730AT4G13730Chr4Ypt/Rab-GAP domain of gyp1p superfamily protein
4AT1G14910AT1G14910Chr1ENTH/ANTH/VHS superfamily protein
5AT1G79350EMB1135Chr1RING/FYVE/PHD zinc finger superfamily protein
5–6AT1G05830ATX2Chr1Trithorax-like protein 2
5AT5G12350AT5G12350Chr5RCC1 family with FYVE zinc finger domain-containing protein
5AT1G50620AT1G50620Chr1RING/FYVE/PHD zinc finger superfamily protein
6AT5G13960SUVH4Chr5Histone-lysine N-methyltransferase, H3 lysine-9
6AT3G15530AT3G15530Chr3S-adenosyl-L-methionine-dependent methyltransferases superfamily
6AT2G34300AT2G34300Chr2S-adenosyl-L-methionine-dependent methyltransferases superfamily
6AT5G06050AT5G06050Chr5Putative methyltransferase family protein
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Soto-Cardinault, C.; Childs, K.L.; Góngora-Castillo, E. Network Analysis of Publicly Available RNA-seq Provides Insights into the Molecular Mechanisms of Plant Defense against Multiple Fungal Pathogens in Arabidopsis thaliana. Genes 2023, 14, 2223. https://doi.org/10.3390/genes14122223

AMA Style

Soto-Cardinault C, Childs KL, Góngora-Castillo E. Network Analysis of Publicly Available RNA-seq Provides Insights into the Molecular Mechanisms of Plant Defense against Multiple Fungal Pathogens in Arabidopsis thaliana. Genes. 2023; 14(12):2223. https://doi.org/10.3390/genes14122223

Chicago/Turabian Style

Soto-Cardinault, Cynthia, Kevin L. Childs, and Elsa Góngora-Castillo. 2023. "Network Analysis of Publicly Available RNA-seq Provides Insights into the Molecular Mechanisms of Plant Defense against Multiple Fungal Pathogens in Arabidopsis thaliana" Genes 14, no. 12: 2223. https://doi.org/10.3390/genes14122223

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop