Abstract Background Alfalfa (Medicago sativa L.) is a vital forage crop with substantial economic and ecological significance in agriculture and animal husbandry. However, atrazine, a widely used herbicide, negatively impacts the growth and yield of alfalfa due to its residual presence in the environment. Transcriptomic analysis was performed to investigate the differences in tolerance and uncover the potential molecular regulatory mechanisms between the tolerant variety JN5010 and the sensitive variety WL363 when subjected to atrazine stress, using RNA-seq on pooled samples. Results Based on the analysis of gene expression profiles, significant differences were observed between the tolerant variety JN5010 and the sensitive variety WL363 under atrazine stress: 2,297 upregulated and 3,167 downregulated in the shoot parts, and 3,232 upregulated and 4,907 downregulated in the roots of JN5010. In WL363, 2,937 genes were upregulated and 4,237 genes were downregulated in the shoot parts, while 5,316 genes were upregulated and 7,977 genes were downregulated in the roots. The DEGs in the shoot parts were mainly involved in biological regulation, metabolic processes, and cellular processes, including proline metabolic processes and S-adenosylmethionine cycle. The DEGs in the roots were predominantly associated with nitric oxide synthesis and metabolism, as well as processes related to cell wall biosynthesis and degradation. In the shoot parts of JN5010, six DEGs were mapped onto the proline metabolic pathway, including four upregulated genes involved in proline biosynthesis and two downregulated genes involved in proline catabolism. In the roots of WL363, eleven DEGs were mapped onto the phenylpropanoid biosynthesis pathway, including seven upregulated genes involved in flavonoid biosynthesis and four downregulated genes associated with lignin biosynthesis. These findings highlight the distinct genetic responses of the two alfalfa varieties to atrazine stress, with JN5010 exhibiting more consistent gene expression patterns compared to the sensitive variety WL363. Conclusions The tolerant variety JN5010 shows improved tolerance to atrazine stress by maintaining stable gene expression and precise regulation in various pathways, such as antioxidant processes, signaling, photosynthesis, and toxin removal. This differential gene expression helps JN5010 maintain stability in its functions under stress, demonstrating better adaptability. These findings enhance our understanding of how alfalfa tolerates atrazine stress and provide important insights for developing atrazine-tolerant varieties. Supplementary Information The online version contains supplementary material available at 10.1186/s12870-025-07129-x. Keywords: Alfalfa, Atrazine, Abiotic stress, Oxidative damage, Comparative transcriptomics Introduction Alfalfa (Medicago sativa L.), often referred to as the “king of forages,” holds immense economic and ecological value in agriculture and animal husbandry. Renowned for its ability to provide high-quality feed for livestock, it also enhances soil fertility through nitrogen fixation [[44]1]. However, research has demonstrated that alfalfa monoculture systems cultivated for over six years display a notable decline in yield, accompanied by a degradation of soil microbial community structure and an increase in pathogens [[45]2]. Furthermore, alfalfa releases allelopathic compounds to inhibit subsequent alfalfa growth, resulting in continuous cropping challenges [[46]3, [47]4]. Consequently, crop rotation with corn (Zea mays L.) has been widely adopted. Studies revealed that alfalfa-corn rotation improves soil pH, reduces salinity, and enhances organic matter content, thereby boosting soil fertility [[48]5]. The rotation system also significantly increased soil nitrogen utilization through heightening soil nitrogen content, enhancing phosphorus solubility, and stimulating of microbial activity. These benefits not only reduce the reliance on chemical fertilizers but also improve the yield and quality of corn. The effectiveness of this rotational system becomes more pronounced with the increasing duration of alfalfa cultivation, markedly enhancing land use efficiency and economic returns [[49]6]. Nonetheless, the widespread use of atrazine, a primary herbicide for controlling corn weeds, poses a challenge. Atrazine residues can hinder the growth and development of subsequent crops, ultimately leading to substantial yield losses. Atrazine is a commonly utilized triazine herbicide, primarily applied for weed control in crops such as corn and sugarcane. Despite its notable effectiveness in agricultural production, the environmental and health risks it poses are significant and warrant attention. Research has shown that atrazine accumulation in drinking water may pose reproductive and developmental risks to children [[50]7]. Moreover, with a half-life of up to 57 weeks in soil, atrazine leaves persistent residues that can harm subsequent crops and contaminate groundwater and surface water, endangering aquatic ecosystems [[51]8]. While the European Union banned atrazine in 2003, it continues to be widely used in agricultural powerhouse nations such as China and Brazil due to its high efficacy and versatile application methods [[52]9]. Studies have demonstrated that atrazine residues exhibit toxic effects on various crops, including cucumber, soybean, and peanut [[53]10]. Atrazine disrupts photosystem II (PS II), inhibits the electron transport chain, and reduces photosynthetic efficiency, which leads to the accumulation of reactive oxygen species (ROS) [[54]11]. This accumulation of ROS can cause oxidative stress and cellular damage. In recent years, advancements in transcriptomics and metabolomics have provided novel insights into the toxic mechanisms of atrazine on plants. Research has revealed that, under atrazine stress, genes involved in antioxidant defense (such as peroxidase (POD) and superoxide dismutase (SOD))are significantly upregulated in rice as a response to ROS accumulation [[55]12]. Metabolomic analyses have identified substantial alterations in amino acid metabolism and secondary metabolites, such as flavonoids and phenolic compounds, which appear to contribute to the plant’s detoxification and adaptive mechanisms [[56]13]. Further studies indicate that under atrazine stress, genes encoding cysteine synthase and spermidine synthase in alfalfa seedlings were significantly upregulated. This upregulation likely enhances the antioxidant capacity of alfalfa, aiding in its response to atrazine-induced oxidative stress [[57]14]. However, there are considerable variances in the tolerance of different alfalfa varieties to atrazine, and the specific molecular mechanisms underlying these differences remain poorly understood. Thus, elucidating the molecular response mechanisms in various alfalfa varieties under atrazine stress is crucial for a deeper understanding of atrazine detoxification and degradation strategies. In this study, we selected two alfalfa varieties with contrasting tolerance characteristics: JN5010, a tolerant variety, and WL363, a sensitive variety, both identified through previous field trials. Leveraging high-throughput RNA-Seq technology, we conducted a systematic comparison and analysis of their growth parameters and transcriptomic responses under atrazine stress. The objective is to integrate physiological indicators (including SOD, MDA, chlorophyll content, and soluble sugar levels) with transcriptomic sequencing and qRT-PCR validation to identify differentially expressed genes (DEGs) between the tolerant and sensitive varieties exposed to atrazine. The molecular response mechanisms of alfalfa tolerance to atrazine were elucidated through the integrated analysis of transcriptomic and phenotypic data. Furthermore, we sought to elucidate dynamic changes in key metabolic and signal transduction pathways, as well as to clarify the molecular response mechanisms distinguishing these tolerant varieties. This research provides valuable theoretical insights into the molecular mechanisms underlying alfalfa tolerance to atrazine, establishing a crucial foundation for molecular breeding practices aimed at developing tolerant varieties. Materials and methods Experimental materials and pretreatment Atrazine (suspension concentrate with a purity of 38%) was obtained from Qiaochang Modern Agriculture Company. The atrazine-tolerant alfalfa variety JN5010 was purchased from Zhengzhou Huafeng Grass Industry Technology Co., Ltd., and the sensitive variety WL363 was purchased from Beijing Zhengdao Seed Industry Co., Ltd. Seeds of both varieties were disinfected with 5% sodium hypochlorite and thoroughly rinsed with distilled water after 10 min of treatment. Petri dishes were lined with 10 cm × 10 cm filter papers, to which an adequate amount of sterile water was added to facilitate seed germination. The dishes were then placed in a light-incubated chamber. After five days, seedlings exhibiting uniform growth were selected, with six pots allocated per variety and 20 seedlings planted in each pot. The atrazine culture system used a modified Hoagland nutrient solution at half strength, which was replaced every three days. After three weeks, three pots from each of the six pots per variety were subjected to stress treatment with 2.0 mg/L atrazine (38%) for six days, during which the nutrient solution was not replaced. From each pot, 10 uniformly growing alfalfa plants were carefully removed from the atrazine solution, quickly rinsed with sterilized DEPC water, and blotted dry with filter paper to remove excess moisture. Measurements were taken for plant height and the natural length of the primary roots. Subsequently, the shoot and root biomass of each plant were recorded. The stems and roots of the plants were then cut into segments roughly 1 cm in length, rapidly frozen using liquid nitrogen, and stored at -80 °C for subsequent total RNA extraction. Measurement of physiological indicators Chlorophyll content in leaves and soluble sugar levels of two alfalfa varieties were determined using Solarbio kits (BC0995 and BC0035; Solarbio, Beijing, China). Superoxide dismutase (SOD) activity and malondialdehyde (MDA) content were measured with Abbkine kits (KTB1030 and KTB1050; Abbkine, Wuhan, China). Chlorophyll was quantified only in the shoot parts (stems and leaves), while all other parameters were assayed separately in both roots and aerial tissues (stems and leaves). To ensure accuracy and reliability, each physiological parameter was assessed in triplicate. RNA extraction and quality control Total RNA was separately extracted from the shoot parts (stems and leaves) and roots of 20 alfalfa seedlings using TRIzol^® Reagent (Invitrogen, California, USA) following the manufacturer’s instructions (Pooled tissues), to reduce the impact of individual variation. Genomic DNA contamination was eliminated through treatment with DNase I (TaKaRa). The quality and concentration of the extracted RNA were evaluated using a 2100 Bioanalyzer (Agilent Technologies, California, USA) and a NanoDrop ND-2000 spectrophotometer (NanoDrop Technologies, Delaware, USA). Only RNA samples meeting stringent quality criteria (OD260/280 = 1.8–2.2, OD260/230 ≥ 2.0, RIN ≥ 6.5, 28 S:18 S ≥ 1.0, and > 1 µg yield) were selected for library construction and sequencing. Library Preparation and illumina sequencing RNA-seq libraries were prepared using the TruSeq™ RNA Sample Preparation Kit (Illumina, San Diego, CA) with 1 µg of total RNA. Messenger RNA was isolated via the poly(A) selection method with oligo(dT) beads and subsequently fragmented. Double-stranded cDNA was synthesized using the SuperScript Double-Stranded cDNA Synthesis Kit (Invitrogen, California, USA) with random hexamer primers (Illumina). The resulting cDNA underwent end-repair, phosphorylation, and ‘A’ base addition, followed by size selection for 300 bp fragments using a 2% Low Range Ultra Agarose gel. The libraries were then amplified through 15 PCR cycles using Phusion DNA polymerase (NEB). Once quantified, the libraries were sequenced on the Illumina HiSeq X Ten/NovaSeq 6000 platform, producing paired-end reads with a length of 150 bp. See Fig. [58]S1 for the sequencing data statistics table. The raw sequencing data were uploaded to the CNCB SRA database (Shoot: CRA024145; Root: CRA024142). Read mapping and transcript assembly The raw paired-end reads were trimmed and quality-controlled using SeqPrep ([59]https://github.com/jstjohn/SeqPrep) and Sickle ([60]https://github.com/najoshi/sickle) with default parameters. The clean reads were then aligned to the reference genome using the HISAT2 software ([61]http://ccb.jhu.edu/software/hisat2/index.shtml) in an orientation-based mode. The mapped reads of each sample were assembled using StringTie ([62]https://ccb.jhu.edu/software/stringtie/index.shtml?t=example) in a reference-based approach. Differential expression analysis and functional enrichment The expression levels of individual transcripts were quantified using the fragments per kilobase of transcript per million mapped reads (FPKM), calculated via RSEM ([63]http://deweylab.biostat.wisc.edu/rsem/). Differential expression analysis was conducted using DESeq2, DEGseq, or EdgeR, with a q-value ≤ 0.05 and|log[2]FC| >1 considered thresholds for identifying significantly DEGs. Functional enrichment analyses, including Gene Ontology (GO) classification and Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway evaluation, were performed using Goatools ([64]https://github.com/tanghaibao/Goatools) and KOBAS ([65]http://kobas.cbi.pku.edu.cn/home.do), respectively, applying a Bonferroni-corrected p-value threshold of ≤ 0.05. Weighted gene Co-expression network analysis and Protein-Protein interaction network analysis Weighted gene co-expression network analysis (WGCNA) was applied to build networks of co-expressed genes and to delineate modules whose members share highly similar expression profiles. Modules exhibiting significant correlations with phenotypic traits were then selected for in-depth investigation of how core module genes relate to the observed phenotypes. For full implementation details see the [WGCNA package] ([66]https://cran.r-project.org/web/packages/WGCNA/index.html). Separately, proteins encoded by the set of DEGs were queried in the STRING database (v11.5; [67]https://string-db.org) to assemble a protein–protein interaction map, thereby uncovering putative functional links among these genes. qRT-PCR analysis To validate the accuracy of our RNA-seq data, eight DEGs were selected for Real-Time Quantitative PCR (qRT-PCR) analysis. These genes were chosen from our dataset and have also been previously reported to be involved in stress responses. For the analysis, total RNA was first treated with gDNA Eraser to remove genomic DNA contamination and subsequently reverse-transcribed into cDNA using the PrimeScript™ RT Reagent Kit (TaKaRa, Cat# RR047A, Tokyo, Japan). The qRT-PCR was then performed using the resulting cDNA as a template on a CFX96 Real-Time PCR Detection System (Bio-Rad, USA). Each gene was assayed in three independent biological replicates, with MsACTIN2 used as the endogenous control for normalization. Gene-specific primers (Table. S2) were designed using Primer-BLAST (NCBI) and Oligo 7 software, with the following parameters: 18–24 nt length, 40–60% GC content, 58–62 °C melting temperature (Tm), and producing an 80–200 bp amplicon. The specificity of each primer pair was confirmed by melting curve analysis and subsequent Sanger sequencing of the PCR products. Data analysis This study employed a triplicate experimental design, with each reported result representing the average of three replicate treatments, each consisting of 20 seedlings. Statistical analyses were conducted to identify DEGs across the libraries. Gene expression levels, or transcript abundance, were normalized to fragments per kilobase of transcript per million mapped reads (FPKM). DEGs between the libraries were identified using a Student’s t-test, with p-values adjusted for multiple testing to control the false discovery rate (FDR). The correlation among detected counts in parallel libraries was evaluated using the Pearson correlation coefficient. Biochemical and molecular response assessments involved determining significant differences between treatments through standard deviation and analysis of variance (ANOVA). ANOVA was also used for statistical comparisons across treatment groups, and when results showed significance at P < 0.05, a least significant difference (LSD) test was performed. All statistical analyses were carried out using SPSS version 27.0. Result Effects of atrazine stress on alfalfa growth and physiology In our previous field experiments, we identified the atrazine-tolerant variety ‘JN5010’ and atrazine-sensitive variety ‘WL363’ from 20 alfalfa varieties (Fig. [68]S1). In this study, the atrazine-tolerant variety ‘JN5010’ and atrazine-sensitive variety ‘WL363’ were exposed to atrazine stress, with subsequent quantification of critical growth parameters. The experimental results demonstrate that atrazine stress significantly impacts the growth and development of alfalfa seedlings (Fig. [69]1A). The atrazine-tolerant variety JN5010 and the sensitive variety WL363 exhibited markedly different responses to atrazine exposure. No significant change in plant height was observed in either variety (Fig. [70]1B, P > 0.05). However, root length was significantly inhibited in both varieties under atrazine stress (Fig. [71]1C). The reduction was more pronounced in WL363 (19.5%) compared to JN5010 (12.1%) (P < 0.05). A clear distinction in biomass accumulation was observed between the two varieties (Fig. [72]1D, E). For the tolerant variety JN5010, atrazine exposure led to a significant reduction in biomass, with a 42.5% decrease in shoot weight and a 50.0% decrease in root weight (P < 0.05). In contrast, the biomass of the sensitive variety WL363 remained statistically unchanged, showing no significant alterations in either shoot or root weight (P > 0.05). Fig. 1. [73]Fig. 1 [74]Open in a new tab Effects of atrazine on the growth and physiological characteristics of two alfalfa varieties. Shown are the overall phenotype (A), plant height (B), root length (C), shoot weight (D), root weight (E), malondialdehyde (MDA) content (F), superoxide dismutase (SOD) activity (G), soluble sugar content (H), and chlorophyll content (I) of JN5010 (tolerant) and WL363 (sensitive) seedlings. Data are presented as the mean ± standard deviation from three biological replicates. Different letters denote significant differences at P < 0.05 Regarding antioxidant defense and oxidative stress, the two varieties showed different response patterns (Fig. [75]1F, G). Under atrazine stress, malondialdehyde (MDA) content increased significantly (P < 0.05) in both the shoots and roots of both varieties. However, the sensitive variety WL363 suffered more severe oxidative damage. After stress treatment, its shoot MDA content reached 84.76 nmol/g, a level significantly higher (P < 0.05) than the 57.79 nmol/g observed in the tolerant variety JN5010. Relative to their respective control levels, the MDA content increased by 20.0% in WL363 and 24.8% in JN5010. For superoxide dismutase (SOD) activity, a key distinction was observed in the roots: JN5010 root SOD activity was significantly induced by 40.5%, from 142.85 to 200.72 U/g (P < 0.05), whereas the root SOD activity in WL363 showed no significant change (P > 0.05). No significant alterations in shoot SOD activity were observed for either variety (P > 0.05). In terms of soluble sugar content (Fig. [76]1H), both varieties exhibited a slight but significant decrease in their shoots (P < 0.05). In the roots, a significant increase of 47.4% was recorded for JN5010 (P < 0.05), while no significant change was observed for WL363 (P > 0.05). Photosynthetic pigment content also differed substantially (Fig. [77]1I). In the tolerant variety JN5010, chlorophyll a content remained stable (P > 0.05), while chlorophyll b and total chlorophyll content significantly decreased by 12.1% and 12.6%, respectively (P < 0.05). In contrast, the photosynthetic system of the sensitive variety WL363 was severely compromised. Its chlorophyll a content decreased sharply by 36.4% (from 0.77 to 0.49 mg/g), and its chlorophyll b content decreased by 45.5% (from 0.55 to 0.30 mg/g), leading to a 37.6% reduction in total chlorophyll content (P < 0.05). Gene segregation in tolerant alfalfa under atrazine stress A principal component analysis (PCA) was conducted on the four gene libraries derived from the shoot tissues of two varieties, JN5010 and WL363, under both atrazine-treated and control conditions. The analysis revealed a distinct separation between the atrazine treatment group and the control group (Fig. [78]2A). Notably, in the PCA, the gene libraries subjected to atrazine stress-JN5010_ATZ and WL363_ATZ-clustered more closely together compared to the control groups, JN5010_CK and WL363_CK. This pattern suggests that under atrazine stress, a significant number of shared genes in the shoot tissues of both varieties exhibited differential expression, resulting in greater similarity in gene expression profiles under stress conditions. Venn diagram analysis further revealed a total of 71,338 genes identified across the four gene libraries (Fig. [79]2C). Among these, 16,430 genes were commonly expressed in the treatment and control groups of JN5010, while16,759 genes were uniquely expressed in both treatment and control groups of WL363. In contrast, 13,607 uniquely expressed genes were identified in JN5010_CK and WL363_CK, compared to 11,688 uniquely expressed genes in JN5010_ATZ and WL363_ATZ. These findings indicate that the number of specifically expressed genes between the two treatment groups exceeds the number of uniquely expressed genes between the two varieties. Additionally, under atrazine treatment, the number of uniquely expressed genes between the two varieties was lower than that observed under control conditions. Fig. 2. [80]Fig. 2 [81]Open in a new tab Comprehensive analysis of gene expression patterns in alfalfa under treatment conditions. (A) Principal component analysis (PCA) of shoot and root transcriptomes showing distinct clustering by treatment. (B) Root-specific PCA demonstrating treatment group separation. (C) Venn diagram of shoot genes with shared and unique responses. (D) Venn diagram of root genes highlighting treatment-specific signatures Principal component analysis revealed a clear separation between the atrazine treatment group and the control group in the root samples (Fig. [82]2B). Unlike the shoot tissues, the separation between the treatment groups (JN5010_ATZ and WL363_ATZ) on the principal coordinate plot was more pronounced than that of their respective control groups (JN5010_CK and WL363_CK). Additionally, the distance between the treatment and control groups of the WL363 variety was greater than that observed in the JN5010 variety. The Venn diagram analysis identified a total of 70,989 genes commonly detected across the four root transcriptome libraries (Fig. [83]2D). Among these, 15,034 genes were specifically expressed in the treatment and control groups of the JN5010 variety, while 18,163 specific genes were detected in WL363. Under control conditions, 11,657 genes were specifically expressed between the two varieties, whereas this number increased to 12,322 under atrazine treatment. These findings demonstrate that the number of specifically expressed genes between the treatment groups of the two varieties exceeded that observed between their respective control groups. Moreover, under atrazine treatment, the number of specific genes expressed between the two varieties was higher than under control conditions. This evidence further suggests that under atrazine-induced stress, a significant number of genes in the roots of both alfalfa varieties exhibited specific expression, with WL363 exhibiting a stronger induction of gene expression compared to JN5010. Under atrazine stress, notable shifts in gene expression were observed in the shoot and root tissues of two alfalfa varieties, JN5010 and WL363. In the shoot tissues of JN5010, 2,297 genes were upregulated, while 3,167 genes were downregulated. Meanwhile, in WL363, 2,937 genes were upregulated, with 4,237 genes downregulated (Fig. [84]3A, B). In the root tissues, JN5010 showcased 3,232 upregulated genes and 4,907 downregulated genes, whereas WL363 exhibited 5,316 upregulated genes alongside 7,977 downregulated genes (Fig. [85]3C, D). These findings suggest that gene expression in both shoot and root tissues of the alfalfa varieties underwent significant alterations under atrazine stress. Plants respond to abiotic stress by modulating gene expression to mitigate adverse conditions. PCA and Venn diagram analyses revealed that atrazine stress influenced the transcriptomic profiles of both alfalfa varieties in their roots and shoot tissues. Intriguingly, gene expression in the shoot parts became more uniform under atrazine stress, while the root tissues exhibited a noticeable degree of differentiation. This indicates that, despite changes in gene expression patterns in both shoot and root tissues, the mechanisms of response varied between these regions. Furthermore, a substantial number of common genes were affected in the shoot tissues, in contrast to the roots, which demonstrated distinct gene expression changes between JN5010 and WL363 under atrazine stress. Fig. 3. [86]Fig. 3 [87]Open in a new tab Heatmaps and bar charts of DEGs in shoot and root tissues between two alfalfa varieties under treatment and control conditions. (A) Heatmap of shoot tissues; (B) Bar chart of shoot tissues showing the number of up- and down-regulated DEGs; (C) Heatmap of root tissues; (D) Bar chart of root tissues showing the number of up- and down-regulated DEGs GO annotation and enrichment analysis of DEGs under atrazine stress GO annotation analysis was conducted on the DEGs identified from the comparisons between shoot and root libraries, with the results presented in Fig. [88]4A. Under atrazine stress, the DEGs in the shoot and root tissues of alfalfa were predominantly categorized into three main groups: Biological Process (BP), Cellular Component (CC), Molecular Function (MF). Significant biological processes included biological regulation, metabolic processes, and cellular processes. The affected cellular components encompassed protein-containing complexes, organelle parts, organelles, membrane parts, and cell parts. Key molecular functions primarily involved catalytic activity and binding functions. Fig. 4. [89]Fig. 4 [90]Open in a new tab GO annotation and functional enrichment of DEGs between two alfalfa cultivars (JN5010 and WL363). (A) GO classification of DEGs into Biological Process, Cellular Component, and Molecular Function categories, showing results for root (R) and shoot (S) tissues. (B) Functional enrichment analysis of DEGs in shoot tissues. (C) Functional enrichment analysis of DEGs in root tissues A functional enrichment analysis was conducted on the DEGs in the shoot parts and roots of two alfalfa varieties under atrazine stress. The top 10 enriched functions were compared, as illustrated in Fig. [91]4B, C. In the shoot parts, the proline metabolic process was significantly enriched in both JN5010 and WL363. Notably, JN5010 exhibited unique enrichments, including the S-adenosylmethionine cycle, meristem development, establishment of plastid localization, chloroplast localization, chloroplast migration, glycine biosynthesis, plastid localization, glycine biosynthesis via the glyoxylic acid transaminase pathway, and iron ion transport. Conversely, WL363 demonstrated unique enrichments among its top 10 significantly enriched functions, such as pre-replicative complex assembly involved in nuclear DNA replication, pre-replicative complex assembly involved in cell cycle DNA replication, pre-replicative complex components, initiation of DNA replication during mitosis, branched-chain amino acid catabolism, initiation of nuclear DNA replication, initiation of cell cycle DNA replication, and the proline catabolic process converting proline to glutamate. In the roots, the synthesis and metabolism of nitric oxide were significantly enriched in both JN5010 and WL363 within the DEGs. Among the top 10 enriched DEGs, JN5010 exhibited root-specific functions such as cysteine biosynthesis from cystathionine, the conversion of proline to glutamate, proline catabolism, seed development, β-glucan metabolic processes in the cell wall, and the biosynthesis and metabolism of cellulose in the plant cell wall. In contrast, WL363 displayed distinct functions among the top 10 enriched DEGs, including meiosis II, meiotic cell cycle processes in male meiosis II, saponin biosynthesis, glucoside biosynthesis, as well as pre-replicative complex formation and assembly essential for cell cycle DNA replication. KEGG pathway enrichment analysis of DEGs under atrazine stress Employing KEGG database for functional annotation analysis facilitates the classification of DEGs based on their roles in metabolic pathways and biological functions. As shown in Fig. [92]S2, the DEGs from the shoot parts and roots of two alfalfa varieties subjected to atrazine stress are primarily grouped into the following categories: Organismal Systems, Genetic Information Processing, Cellular Processes, Environmental Information Processing, and Metabolism. Within the Genetic Information Processing category, the DEGs are involved in pathways linked to Environmental Adaptation, Transcription, Replication and Repair, Folding, Sorting and Degradation, and Translation. In the Cellular Processes category, the DEGs are associated with Transport and Catabolism. For the Environmental Information Processing category, they are connected to pathways such as Membrane Transport and Signal Transduction. The Metabolism category spans several pathways, including Carbohydrate Metabolism, Amino Acid Metabolism, Biosynthesis of Secondary Metabolites, Energy Metabolism, Lipid Metabolism, Terpenoid Metabolism, Polyketide Metabolism, Other Amino Acid Metabolism, Cofactor and Vitamin Metabolism, Glycine Biosynthesis and Metabolism, as well as Nucleotide Metabolism. The functional enrichment analysis of DEGs in the shoot and root parts of two alfalfa varieties revealed both shared features and distinct differences among the top 20 significantly enriched pathways (Fig. [93]5). Pathways significantly enriched in both varieties include β-alanine metabolism, amino sugar and nucleotide sugar metabolism, cysteine and methionine metabolism, galactose metabolism, alanine-aspartate-glutamate metabolism, propionate metabolism, starch and sucrose metabolism, glycine-serine-threonine metabolism, photosynthesis-antenna proteins, glycolysis/gluconeogenesis, selenocompound metabolism, valine-leucine-isoleucine degradation, and glyoxylic acid and dicarboxylic acid metabolism. Distinctly, the JN5010 variety demonstrates unique enrichment in pathways such as pyruvate metabolism, carbon fixation in photosynthetic organisms, peroxisome metabolism, tyrosine metabolism, phenylpropanoid biosynthesis, fructose and mannose metabolism, and fatty acid degradation. Conversely, the WL363 variety shows unique pathway enrichment related to ribosome function, ascorbate and aldarate metabolism, phenylpropanoid biosynthesis, pyrimidine metabolism, cyanogenic amino acid metabolism, phagosome formation, and histidine metabolism. Fig. 5. [94]Fig. 5 [95]Open in a new tab Presents the KEGG enrichment analysis of DEGs in alfalfa. (A) shows the distribution of enriched metabolic pathways in shoot tissues, combining data from two alfalfa varieties. (B) shows the distribution of enriched metabolic pathways in root tissues, combining data from two alfalfa varieties. In the enrichment plots, dot size represents the number of genes associated with each pathway, and dot color indicates the level of significance (P-value) Gene co-expression network Under atrazine stress, alfalfa demonstrates a multifaceted genetic response. Transcriptome analysis of various atrazine-tolerant alfalfa varieties identified over 60,000 DEGs. By employing Weighted Gene Co-expression Network Analysis (WGCNA), correlations between DEGs and atrazine stress-related physiological traits were examined (Fig. [96]6A). Twelve distinct gene co-expression modules were identified, and module-trait association analyses revealed significant correlations between gene expression levels in specific modules and particular traits (Fig. [97]6B). For instance, the brown module exhibited a strong positive correlation with plant height, with correlation coefficients ranging from 0.825. Additionally, it showed significant negative correlations with root length and root weight, with coefficients ranging from − 0.822 to -0.910. In contrast, the turquoise module demonstrated significant negative correlations with plant height, shoot biomass, and chlorophyll B content, with correlation coefficients ranging from − 0.808 to -0.825. Conversely, it exhibited a significant positive correlation with superoxide dismutase (SOD), with a correlation coefficient of 0.839 (Fig. [98]6B). These findings suggest that genes within the brown module play a crucial role in promoting plant height while suppressing root length and root weight under atrazine stress. Similarly, in various atrazine-tolerant alfalfa varieties, plant height, shoot biomass, and chlorophyll B content were strongly negatively correlated with gene expression in the turquoise module, while a significant positive correlation with SOD was observed. The brown module comprises 12,771 standard genes, whereas the turquoise module contains 20,419 standard genes. Among these, the top 30 hub genes, identified by connectivity, were recognized as critical genes representing the primary functions of these modules (Fig. [99]6C, D). Visualization of these hub genes using Cytoscape 3.9.1 ([100]https://cytoscape.org/) [[101]15] revealed their involvement in numerous physiological processes, including photosynthesis, respiration, carbon metabolism, cytokinin signaling, stress responses, and root and embryo development (Fig. [102]6C, D). These findings underscore the vital roles played by these genes under atrazine stress conditions. Fig. 6. [103]Fig. 6 [104]Open in a new tab WGCNA reveals gene-physiology correlations in alfalfa under atrazine stress. (A) A module classification tree uses colors to represent different modules, with grey indicating genes that have not been assigned to a specific module. (B) Module-phenotype correlations heatmap depicting relationships between modules and physiological parameters. (C) Gene Co-Expression Network of DEGs in MEbrown. (D) Gene Co-Expression Network of DEGs in MEturquoise Validation of RNA-seq identified DEGs via qRT-PCR in transcriptional profiling To further validate the expression trends of DEGs, we selected eight genes for qRT-PCR experiments. These genes primarily include those involved in ethylene synthesis (MS.gene045407, Fig. [105]7A) and abscisic acid (ABA) synthesis (MS.gene39848, Fig. [106]7B), both of which play crucial roles in regulating plant stress responses. Additionally, the ASMT gene (MS.gene77494, Fig. [107]7C) mediates the synthesis of melatonin, which scavenges reactive oxygen species (ROS), while the APX6 gene (MS.gene67066, Fig. [108]7D) enhances oxidative defense through its antioxidant activity. The CHS gene (MS.gene02598, Fig. [109]7E) and F3H gene (MS.gene04220, Fig. [110]7F) are responsible for synthesizing flavonoids and anthocyanins through phenylpropanoid metabolism. The ABCB1 gene (MS.gene005430, Fig. [111]7G) utilizes the ATP-binding cassette (ABC) transport system to expel toxins or regulate hormone distribution, while the GA2ox gene (MS.gene70788, Fig. [112]7H) regulates the degradation of gibberellins. The relative expression levels indicate that the qRT-PCR results are consistent with the expression trends obtained from RNA sequencing, demonstrating the accuracy of our sequencing results and their applicability for subsequent transcriptomic analyses. To ensure the reliability of the transcriptome data, we validated it by calculating the linear regression coefficient between the RNA-seq data (log2(FC)) and the qRT-PCR results (ΔΔCt method). The results showed a strong positive correlation (R² >0.6) between the two methods, indicating the reliability of our transcriptome results. The correlation is illustrated in Fig. [113]S3. Fig. 7. [114]Fig. 7 [115]Open in a new tab The results of the qRT-PCR analysis involved eight key genes: (A) ACCO, (B) AAO3, (C) ASMT, (D) APX6, (E) CHS, (F) F3H, (G) ABCB1, and (H) GA2ox. To ensure reproducibility and reliability, qRT-PCR analyses were conducted on three independent biological replicates. The relative expression levels were calculated using the 2^−∆∆Ct method, and the values presented represent the means of the three biological replicates. Duncan’s test was employed to assess statistical significance, with different lowercase letters indicating significant differences between treatments (P < 0.05) Discussion Atrazine has a soil half-life of up to 57 weeks, and its residues can affect alfalfa-maize rotation systems for as long as two years, resulting in yield losses that cannot be underestimated [[116]16, [117]17]. In our field trials, the root biomass of the sensitive variety WL363 was completely suppressed, whereas JN5010 retained 98% of its root biomass (Fig. [118]1E). Elucidating the molecular mechanisms of atrazine tolerance not only sheds light on the fundamental nature of plant herbicide tolerance but also offers practical guidance for improving productivity in rotational cropping systems. Exogenous stress-induced differential gene expression in plants highlights their role in stress response mechanisms. A scatter plot of DEGs demonstrates a substantial presence of DEGs in the shoots and roots of both plant varieties under atrazine stress. To validate the RNA-seq results and identify key pathways involved in atrazine tolerance, we performed GO enrichment analysis on the DEGs, validated by qRT-PCR. The root DEGs of JN5010 are primarily associated with cell wall functionality and the metabolism of proline and cysteine. The cell wall, as the first line of defense against external stressors, undergoes functional changes that play a critical role in plant adaptation to stress. Proline functions as an osmo-protectant, contributing to growth and development while scavenging reactive oxygen species; its levels fluctuate under various stress conditions [[119]18–[120]24]. Interestingly, the ALDH7A1 gene, involved in proline metabolism, showed significant downregulation in the roots of JN5010 under atrazine stress. This suggests a potential shift in proline metabolism within the roots of the tolerant variety under stress. Cysteine supports stress response mechanisms by catalyzing H[2]S production, which, as a gaseous signaling molecule, mitigates oxidative damage and enhances stress tolerance through synergistic effects [[121]25–[122]27]. In contrast, the root DEGs of WL363 are primarily enriched in cell cycle-related functions. Research indicates that heavy metal stress can trigger oxidative damage, leading to cell cycle arrest and subsequent inhibition of plant growth [[123]28–[124]30]. Notably, nitric oxide (NO), an essential signaling molecule, plays a pivotal role in seed germination, root development, and stress responses [[125]31, [126]32]. The biosynthesis and metabolic processes of NO are enriched in the root DEGs of both plant varieties. Functional analysis of the shoot DEGs in WL363 similarly highlights pathways related to the cell cycle and proline metabolism. However, the shoot DEGs of JN5010 are enriched in chloroplast- and plastid-related functions. Furthermore, distinct, significant metabolic pathways were identified in both varieties under atrazine stress. In JN5010, DEGs are predominantly enriched in pathways linked to pyruvate metabolism, photosynthetic carbon fixation, and peroxisome functions. In WL363, shoot DEGs are significantly enriched in pathways associated with ascorbate and aldehyde acid metabolism, as well as phenylpropanoid biosynthesis, while its root DEGs are enriched in the citric acid cycle (TCA cycle) and phosphoinositide metabolism pathways. For example, the ABCB1 gene, which encodes an ABC transporter, has been shown in other species to confer herbicide tolerance by pumping toxins out of the cell. These specific pathways, including ascorbate synthesis, plant hormone signal transduction, the MAPK signaling pathway, photosynthetic carbon fixation, the Calvin cycle, glycolysis, and ABC transporters, offer valuable insights into the variations in tolerance between the two varieties. Examining the response patterns of key genes within these pathways to atrazine stress offers valuable insights into the tolerance differences among various alfalfa varieties. Ascorbate synthesis pathway Ascorbate (AsA) is a vital water-soluble antioxidant in plants, playing a central role in various physiological processes, such as photosynthesis, respiration, cell division, growth regulation, hormone signal integration, and senescence [[127]33–[128]35]. Its homeostasis is maintained through mechanisms including synthesis, recycling, degradation, and intracellular and intercellular transport. Among these, the galactose metabolic pathway constitutes a major route for AsA production [[129]36–[130]38]. In this biosynthetic pathway, GDP-D-mannose-3’,5’-epimerase (GME) acts as a key mediator in the interconversion between GDP-D-mannose, GDP-L-galactose, and GDP-L-glucose, while also participating in the L-galactose and L-gulose pathways [[131]39, [132]40]. Research shows that GME expression levels typically decrease under exogenous stress, and GME inactivation can result in reduced cell division and impaired development [[133]41]. In this experiment, the expression levels of three genes encoding GME in both the shoot parts and roots of WL363 were downregulated under atrazine stress (Fig. [134]8). In the roots of JN5010, a similar downward trend was observed, but the fold change in downregulation was less pronounced than in WL363. Meanwhile, in the shoot parts of JN5010, two of the three GME-encoding genes were differentially upregulated, with MS.gene44007 showing no significant change (Table. S3). The enzyme L-galactose-1-phosphate phosphatase (GPP), encoded by VTC4, dephosphorylates L-galactose-1-phosphate to L-galactose, making it an essential component of galactose metabolism for AsA synthesis [[135]42]. Studies have revealed that mutations in this gene may lead to a 50% reduction in ascorbate content [[136]43]. In this experiment, the expression of three VTC4-encoding genes in the shoot parts and roots of WL363 was downregulated under atrazine stress. In comparison, only one gene in the shoot parts and roots of JN5010 exhibited downregulation, with this decrease being less pronounced than in WL363. Similar expression patterns were observed in genes encoding L-galactose dehydrogenase (GalDH), gulonolactone oxidase (Gulo), and dehydroascorbate reductase (DHAR) in both varieties (Fig. [137]8). L-galactose dehydrogenase catalyzes L-galactonate-1,4-lactone, representing the final step in AsA synthesis through the galactose pathway. The transcription levels and enzyme activity of L-galactose dehydrogenase are positively correlated with L-ascorbate content [[138]44–[139]46]. Gulo catalyzes L-gulonolactone-1,4-lactone, a critical step in AsA synthesis through the gulose pathway; plants genetically transformed to express this enzyme exhibit enhanced tolerance to abiotic stress [[140]47]. Dehydroascorbate reductase, which is an essential antioxidant enzyme, plays a pivotal role in the ascorbate-glutathione (AsA-GSH) cycle, ensuring normal levels of AsA in plants. Extensive research indicates that dehydroascorbate reductase regulates reactive oxygen species levels, improves leaf photosynthesis, and boosts stress tolerance, thereby mitigating plant damage [[141]48–[142]51]. Plants’ synthesis and metabolic capacity of AsA significantly influence their antioxidant capacity and ability to withstand stress [[143]52–[144]54]. For example, research by Mirza Hasanuzzaman et al. highlighted that nitrogen supplementation in wheat notably increased the activity of AsA and related enzymes, such as SOD, CAT, GR, and DHAR, under salt stress, thereby bolstering the antioxidant defense system [[145]52]. Moreover, atrazine stress significantly induced root SOD activity in JN5010 (142.85 to 200.72 U/g) but not in WL363 (Fig. [146]1G). While both varieties showed increased shoot MDA content under atrazine stress, the increase was less pronounced in JN5010 (46.31 to 57.79 nmol/g) compared to WL363 (70.61 to 84.76 nmol/g) (Fig. [147]1F), which indicated that both varieties activate stress response pathways. However, the tolerant variety JN5010 exhibited greater root SOD induction and a smaller increase in shoot MDA, potentially contributing to its atrazine tolerance (Fig. [148]1F, G). In summary, atrazine stress significantly suppresses the expression of genes involved in the AsA synthesis pathways in the shoot parts and roots of WL363. By contrast, the highly tolerant variety JN5010 experiences less inhibition, with certain genes even exhibiting upregulated expression in response to atrazine stress. This more stable AsA synthesis in JN5010 under atrazine stress may serve as a key factor contributing to its robust atrazine tolerance (Fig. [149]8). Fig. 8. [150]Fig. 8 [151]Open in a new tab Molecular regulatory network for shoot and root in response to Atrazine stress in atrazine-tolerant and sensitive alfalfa varieties Plant hormone signal transduction and the MAPK signaling pathway Ethylene serves as a pivotal regulatory factor in the processes of plant growth and development. It plays a crucial role in enhancing plant tolerance to various abiotic stresses by synthesizing secondary metabolites, interacting with other hormones, and regulating metabolic functions. Research has demonstrated that suppressing ethylene production and its signal transduction can adversely affect tolerance to abiotic stress [[152]55–[153]59]. The ethylene response factor (ERF) family, positioned downstream in the ethylene signaling pathway, contributes to abiotic stress responses by modulating the expression of various functional genes. Extensive studies have revealed that ERF genes are influenced by external stressors, with their expression closely linked to plant tolerance to conditions such as salinity, drought, and heat [[154]60–[155]62]. For instance, research by Zhang et al. [[156]63]. found that overexpression of the GmEFR3 gene in soybeans increased levels of free proline and soluble carbohydrates, thereby improving drought tolerance. Similarly, Wu et al. [[157]64]. showed that overexpression of ERF genes in tomatoes enhanced tolerance to salt, drought, and cold stresses, highlighting the pivotal role of ERF gene expression in stress resilience. In this experiment, two genes encoding ERF1 exhibited downregulated expression under atrazine stress, with the downregulation in WL363 being more pronounced than in JN5010 (Fig. [158]8). EIN3, a key component of the ethylene signaling pathway, functions downstream of EIN2 and MAPK3/6 and is negatively regulated by the ubiquitination/proteasome pathway controlled by EBF1/2. This pathway transmits regulatory signals to ethylene response factors (ERFs) [[159]65, [160]66]. In the current experiment, EIN3 was upregulated in both varieties of alfalfa under atrazine stress; however, expression differences in the regulators MPK3 and EBF1/2 exhibited varietal specificity under atrazine stress (Table. S3). Thus, variations in expression levels of the key ERF genes within the ethylene signaling pathway under atrazine stress may constitute one of the factors influencing atrazine tolerance in alfalfa. H₂O₂ is a reactive oxygen species with significant roles in diverse biological systems [[161]67]. It not only participates in essential developmental processes such as photosynthesis, stomatal movement, and leaf senescence but also functions as a pivotal signaling molecule in plant growth, development, and stress responses [[162]68–[163]71]. Studies suggest that increasing endogenous H₂O₂ levels or modulating the expression of genes associated with H₂O₂ production can enable plants to better adapt to abiotic and biotic stresses [[164]72]. The activation of MPK4 by H₂O₂ requires MEKK1, highlighting its central role in maintaining H₂O₂ mediated reactive oxygen balance through interaction with its downstream target, MPK4. Research has revealed a correlation between reactive oxygen species and salicylic acid, with mekk1 and mpk4 mutants exhibiting elevated salicylic acid levels [[165]73, [166]74]. Moreover, MEKK1 is capable of activating two highly homologous MAPKs, MPK4 and MPK6, facilitating signal transduction related to reactive oxygen, as well as biotic and abiotic stresses via downstream MPK4 and MPK6. The gene expression profiles of MEKK1 and MPK4 show considerable overlap in common stress-response genes [[167]73, [168]75, [169]76]. Consistent with previous findings, the gene expression patterns of MEKK1 and MPK4 in the shoot parts and roots of both varieties under atrazine stress were generally similar in this experiment. Significant upregulation was observed in WL363, while in JN5010, only the expression levels of MPK4 in the roots were upregulated, albeit to a lesser extent than in WL363 (Fig. [170]8). These observations suggest that WL363 accumulated more H₂O₂ under atrazine stress, thereby activating the MEKK1 and MPK4 signaling pathways. At the downstream level, MPK4 may also suppress the expression of the GST-1 gene, which encodes a key detoxification enzyme crucial for eliminating harmful exogenous substances in plants. A kinase induced by oxidative signals, particularly hydrogen peroxide, OXI1 plays an essential role in transmitting hydrogen peroxide signals downstream [[171]77, [172]78]. Stress conditions have been shown to induce the gene encoding this kinase, potentially resulting in leaf cell death [[173]79, [174]80]. The WRKY transcription factor family plays critical roles in plant growth, development, metabolism, and responses to environmental stresses, with WRKY22 being a validated participant in tolerance mechanisms across multiple plant species [[175]81–[176]83]. For example, partial knockout of the CsWRKY22 gene increased citrus tolerance to canker disease, while overexpression of LlWRKY22 enhanced heat tolerance in lilies, and silencing it produced the opposite effect [[177]84, [178]85]. In this experiment, under atrazine stress, the expression of the H₂O₂-induced OXI1 gene in alfalfa showed an upregulated trend. Meanwhile, the WRKY22 gene at the end of the pathway exhibited clear varietal specificity. Of the four differential genes encoding WRKY22, two were downregulated in the shoot parts and roots of JN5010, whereas no expression differences were evident in the shoot parts of WL363, which exhibited marked upregulation in the roots (Table. S3). These findings indicate that atrazine stress stimulated the signaling pathway from OXI1 to WRKY22, with WL363 showing more pronounced changes. The atrazine induced increase in H₂O₂ in alfalfa plants activated related signaling pathways, including WRKY22-mediated cell death and the inhibition of GST-1 gene expression by MPK4. These mechanisms may explain why WL363 displays greater toxicity under atrazine stress. Abscisic acid (ABA) is a vital plant hormone involved in various biological processes, playing a key role in plant growth and development [[179]86–[180]88]. Extensive research on the mechanisms of the ABA signaling pathway under stress conditions has demonstrated that ABA enhances plant stress tolerance by inducing the expression of specific stress response genes during adverse conditions [[181]89–[182]91]. At the core of this signaling pathway lies the PYL-PP2C-SnRK2 complex, which operates by interacting with PYL family proteins to inhibit the phosphatase activity of PP2Cs in the presence of ABA. Conversely, in the absence of ABA, PP2Cs interact with SnRK2s through dephosphorylation, leading to the inactivation of SnRK2s, which prevents them from phosphorylating downstream response factors. This study investigated the expression changes of ABA-related genes under atrazine stress, since the plants are more likely to produce ABA under stress. The findings revealed that in the shoot parts and roots of WL363, the ABA-activated PYL-encoding genes (e.g., MS.gene027877, MS.gene031922, MS.gene054122, MS.gene057108, MS.gene29692, MS.gene86009) were significantly upregulated, possibly leading to ABA insensitivity. In contrast, for JN5010, only MS.gene027877, MS.gene031922, and MS.gene29692 were significantly upregulated in the shoot parts, while other genes showed an upregulation trend but did not reach statistically significant levels (Table. S3). The lower upregulation of these PYL genes may result in greater ABA sensitivity. Furthermore, most PP2C-encoding genes exhibited a downregulation trend under atrazine stress, with the exception of MS.gene64771, which showed significant upregulation in WL363, suggesting a possible negative feedback mechanism to counteract the effects of ABA. This downregulation of PP2C allowed for the significant upregulation of the SnRK2 gene in the roots of WL363, resulting in a marked increase in the expression of associated genes, and a greater reaction to the ABA. Interestingly, the gene encoding MPK1_2 (MS.gene69463) was significantly downregulated in both varieties, with statistically significant reductions observed in the shoot parts and roots of WL363, which may mean that it plays an important role in stress responses. Additionally, the katE-encoding gene (MS.gene064838) displayed significant upregulation in both varieties across the shoot parts and roots (Table. S3), likely to combat stress. In summary, atrazine stress stimulates ABA production, with the upregulated PYL genes negatively regulating PP2C, thereby activating downstream response factors (Fig. [183]8). These findings align with previous studies and offer valuable insights into the regulatory mechanisms of the ABA signaling pathway. In the roots of the highly tolerant variety JN5010, the GST gene (MS.gene21351) was upregulated and acted synergistically with the ABCB1 transporter (e.g., MS.gene07361), which was upregulated 4.1-fold, enabling efficient extrusion of atrazine into the extracellular space or vacuolar compartments. Therefore, this combined detoxification mechanism offers two targets, the GSTC domain and the ABCB1 coding region, suitable for both marker assisted selection and CRISPR based improvement. Photosynthetic carbon fixation and sugar metabolism Photosynthesis is the process whereby plants convert carbon dioxide and water into organic matter using light energy while releasing oxygen. This process is vital for the survival of living organisms and plays a key role in the carbon and oxygen cycles on Earth [[184]92, [185]93]. Abiotic stress often disrupts photosynthesis by reducing photochemical electron transport rates, deactivating photosystem (PS) centers, degrading pigments and proteins, and interfering with other biological pathways, ultimately affecting normal plant growth [[186]94, [187]95]. In this experiment, atrazine stress led to a decrease in the expression levels of most genes associated with photosynthesis (Table. S3). Specifically, for genes related to photosystem I (PS I), MS.gene018588 and MS.gene32425 were downregulated in the shoot parts of WL363, while MS.gene044778 and MS.gene88194 were downregulated in JN5010. Regarding PS II genes, MS.gene039151, MS.gene036154, MS.gene040618, and MS.gene065561 in WL363 exhibited expression changes under atrazine stress, while no significant expression differences were observed in the shoot parts of JN5010. For electron transport-related genes, atrazine suppressed the expression of MS.gene66754, MS.gene003300, MS.gene02098, and MS.gene014378 in WL363, whereas only MS.gene002059 was upregulated in JN5010 (Table.S3). These findings demonstrate that atrazine stress suppressed the expression of most photosynthesis-related genes, with expression variations in PS II and electron transport genes being more closely linked to atrazine tolerance. Gene expression changes were significant in WL363, while only a limited number of genes were affected in the highly tolerant variety JN5010 (Fig. [188]8). This indicates that tolerance differences among alfalfa varieties primarily lie within the PS II and electron transport processes, consistent with previous research on atrazine toxicity mechanisms [[189]8, [190]96]. The Calvin cycle is a crucial component of photosynthesis, acting as an essential pathway for plants to fix carbon from the environment. However, its processes can be disrupted by factors such as reduced carbon dioxide concentration or fluctuations in light intensity. Under abiotic stress, plants modulate the expression of key enzymes and intermediates via multiple pathways, leading to alterations in the Calvin cycle. The inhibition of these enzymes is often responsible for the stunted growth observed in plants subjected to such stressful conditions [[191]97–[192]101]. Ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco), a pivotal enzyme in the Calvin cycle, plays a central role in carbon fixation. Its catalytic activity depends on the cooperative function of large (L) and small (S) subunits, which collectively enhance its efficiency [[193]102, [194]103]. Transcriptomic analyses have revealed that under atrazine-induced stress, the expression levels of rbcS-encoding genes in the shoot parts of both studied plant varieties were upregulated. Conversely, the expression of key enzymes-glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and phosphoglycerate kinase (PGK)-responsible for the conversion of 3-phosphoglycerate into glyceraldehyde-3-phosphate, was suppressed (Fig. [195]8). The carbon fixation product, glyceraldehyde-3-phosphate, is a precursor for the synthesis of starch or fructose-1,6-bisphosphate through gluconeogenesis and glycolysis pathways. Notably, the genes encoding aldolase (ALDO) and triose phosphate isomerase (TPI), which are integral to these subsequent steps, also exhibited a downregulation trend. Numerous studies have demonstrated that the Calvin cycle serves as a source of organic matter and chemical energy, underpinning high crop yields and stress tolerance in plants. However, environmental changes can impair certain enzymes and reduce carbon fixation efficiency [[196]104–[197]106]. The findings from this experiment indicate that atrazine stress disrupted the carbon fixation process in alfalfa, ultimately influencing plant growth and development. Sugars, as central molecules in biological systems, play a vital role in gene expression, metabolism, the cell cycle, and the overall development of both eukaryotes and prokaryotes [[198]107]. Glycolysis, an essential carbohydrate metabolism pathway, oxidizes sugars to produce ATP, pyruvate, and NADH in plants. Under abiotic stress, plants regulate enzymatic processes within glycolysis, which not only serve as signaling mechanisms but also function as key centers for managing energy during stress tolerance [[199]108, [200]109]. Within the glycolytic pathway, hexokinase acts as a critical rate-limiting enzyme in glycolytic metabolism and plant respiration. It catalyzes the phosphorylation of fructose to generate fructose-6-phosphate, marking the initial step of glycolysis, while simultaneously influencing sugar metabolism and storage. Furthermore, hexokinase plays an integral role in sugar signaling pathways, with research showing its expression is induced under conditions of salt, osmotic, and cold stress [[201]109–[202]111]. Fructose-6-phosphate kinase and pyruvate kinase also play significant roles in glycolysis by consuming ATP to catalyze the conversion of fructose-6-phosphate into fructose-1,6-bisphosphate and phosphoenolpyruvate into pyruvate, respectively. These reactions are both irreversible and represent key regulatory nodes within the glycolytic pathway. Experimental results indicated that the expression levels of genes encoding these three enzymes were consistently downregulated under atrazine stress. However, notable differences were observed between the two varieties studied: while hexokinase and fructose-6-phosphate kinase showed reduced expression levels in WL363, their downregulation was less pronounced in JN5010. On the other hand, pyruvate kinase exhibited a sharper decline in expression in JN5010 compared to WL363 (Fig. [203]8). These findings highlight that atrazine stress leads to changes in the expression patterns of key enzyme-encoding genes in the glycolytic pathways of the shoot parts and roots of purple flowering alfalfa, with distinct variations between the two plant varieties. Trehalose (Tre) is a non-reducing disaccharide composed of two glucose molecules. It is typically found in low concentrations across most plants but plays a crucial role in various physiological processes, including carbohydrate storage, transport, and signaling regulation. Numerous studies have demonstrated that trehalose and its precursor, trehalose-6-phosphate, serve as vital signaling molecules, regulating plant growth and development while contributing to metabolic control and gene expression. Furthermore, trehalose is strongly associated with plant responses to abiotic stress [[204]112–[205]115], playing a key role in supporting normal growth and development under unfavorable environmental conditions. Research has shown that the exogenous application of trehalose can effectively alleviate high-temperature stress during the grain-filling stage in wheat, primarily by decreasing reactive oxygen species levels and enhancing antioxidant enzyme activity. Additionally, trehalose can mitigate drought stress in sweet sorghum, improving the plant’s seedling drought tolerance [[206]98, [207]115, [208]116]. Trehalose-6-phosphate synthase (TPS), the enzyme critical to trehalose synthesis, has been observed to exhibit upregulated gene expression under atrazine stress. Experimental results indicated that the gene encoding TPS showed higher levels of upregulation in JN5010 compared to WL363 in the aboveground parts of the plant (Fig. [209]8). Similarly, most TPS-related genes demonstrated more pronounced upregulation in JN5010 roots (Table. S3). These findings suggest that both plant varieties deploy trehalose synthesis pathways as a response to atrazine stress. However, the atrazine-tolerant variety JN5010 exhibited a greater degree of TPS gene expression upregulation, potentially driving increased trehalose synthesis and enhancing its tolerance to atrazine stress (Fig. [210]8). Detoxification pathways mediated by ABC transporters and flavonoids The plant ABC transporter family comprises eight subtypes of membrane transport proteins essential for detoxification processes, facilitating the transport of diverse substrates [[211]117]. These proteins utilize ATP to expel harmful substances, such as heavy metal ions, organic pollutants, secondary metabolites, and xenobiotics, either to the extracellular space or vacuoles, thereby mitigating cellular toxicity [[212]28, [213]118, [214]119]. Research has revealed the functional diversity within the ABC transporter family. The ABCA subfamily is primarily responsible for lipid transport, maintenance of membrane structure, and stress response [[215]120–[216]122]. For instance, the soybean protein GmABCA7, localized in peroxisomes, promotes fatty acid β-oxidation and facilitates seed germination in Arabidopsis experiments [[217]123]. The ABCB subfamily predominantly regulates and transports hormonal substances. In tomatoes, SIABCB demonstrates high expression in roots, suggesting its involvement in ion and heavy metal transport [[218]123]. Likewise, the ABCC subfamily has been identified to transport metabolites, such as anthocyanins and phytic acid, alongside detoxification functions [[219]124–[220]126]. Studies in Arabidopsis have shown that AtABCC1 and AtABCC2 contribute to arsenic, cadmium, and mercury tolerance. Under cadmium stress, the expression of AtABCC6 is upregulated, further validating its role in heavy metal detoxification [[221]119, [222]127, [223]128]. The ABCG subfamily, the largest branch of ABC transporters, is integral to plant detoxification and defense mechanisms [[224]121, [225]129]. Research indicates that this subfamily enhances tolerance to biotic and abiotic stresses by mediating the secretion of secondary metabolites and facilitating the efflux of harmful substances. For example, transport proteins such as NpPDR1/NpABC1 and NtPDR1 have been shown to promote the secretion of antifungal terpenoids and antimicrobial metabolites, playing a pivotal role in plants’ chemical defense systems [[226]130, [227]131]. Further supporting the role of ABC transporters in stress response, transcriptome analysis suggested an upregulation of the GST gene (MS.gene21351) in the roots of the highly tolerant variety JN5010 under atrazine stress. Subsequent qRT-PCR analysis confirmed a significant upregulation of the ABCB1 transporter (MS.gene005430) under the same conditions (P < 0.05). The tolerant variety JN5010 appears to establish a robust atrazine tolerance mechanism through the coordinated action of the ABCA3 gene in the shoot parts and the ABCG2 gene in the underground parts, while the sensitive variety WL363 primarily depends on the ABCB1 gene in the shoot parts. Notably, the CHS gene, responsible for synthesizing flavonoids and anthocyanins through phenylpropanoid metabolism, was upregulated by 3.5-fold in both root and shoot tissues of WL363 under atrazine stress (Fig. [228]7E). However, its MDA content still increased significantly (from 70.61 to 84.76 nmol/g, Fig. [229]1F), suggesting that while flavonoid (especially anthocyanin) accumulation may partially alleviate membrane lipid peroxidation, it fails to completely inhibit oxidative damage. Given that MYBL2 acts as a negative regulator of flavonoid biosynthesis, its CRISPR-mediated knockout could further amplify this endogenous antioxidant response in WL363 to enhance stress tolerance. For the selection and improvement of highly atrazine-tolerant alfalfa lines like JN5010, we propose focusing on detoxification pathways. Although the GST gene (MS.gene21351) was significantly upregulated in JN5010 roots under atrazine stress, the ABCB1 transporter (MS.gene005430) showed a slight yet significant increase in expression (Fig. [230]7G). Thus, the ABCB1 locus could still be considered as a potential molecular marker for preliminary screening. Thereafter, CRISPR-Cas9 multiplex editing could be employed to simultaneously target ABCA3 and ABCG2 to enhance detoxification pathways in tolerant lines, and CHS and MYBL2 to boost antioxidant defenses specifically in sensitive lines like WL363. Conclusion This study revealed the differences in gene expression between the tolerant variety JN5010 and the sensitive variety WL363 under atrazine stress, as well as their potential molecular regulatory mechanisms. The results showed that JN5010 exhibits significantly higher gene expression stability in both root and shoot tissues compared to WL363, indicating that JN5010 displayed stronger atrazine tolerance. Further analysis demonstrated that the DEGs in JN5010 were primarily enriched in proline metabolism and antioxidant defense pathways, whereas DEGs involved in cell cycle regulation and phenylpropanoid biosynthesis pathways were enriched in WL363. Despite photosynthesis and sugar metabolism were affected under atrazine stress in both alfalfa varieties, WL363 showed significantly greater suppression of the expression of genes related to photosystem II and electron transport compared to JN5010. Moreover, plant hormone signaling and MAPK signaling pathways played crucial roles in the response to atrazine stress, further supporting the advantages of JN5010 in detoxification and metabolism. Notably, the ABC transport proteins in JN5010 were significantly upregulated under stress, suggesting their key roles in detoxification and substance transport. Differential gene expression contributed to JN5010’s functional stability under stress, demonstrating better adaptability. These findings enhance our understanding of how alfalfa tolerates atrazine stress and provide important insights for developing atrazine-tolerant varieties. Nevertheless, our future research still needs to further validate specific pathways, especially through metabolomic analysis of the dynamic changes of proline and flavonoid substances, as well as through proteomic methods to verify the activity and content changes of key enzymes. Supplementary Information Below is the link to the electronic supplementary material. [231]Supplementary Material 1^ (12.4KB, xlsx) [232]Supplementary Material 2^ (17.3KB, docx) [233]Supplementary Material 3^ (1,004.9KB, docx) [234]Supplementary Material 4^ (44.5KB, xls) [235]Supplementary Material 5^ (459.9KB, docx) [236]Supplementary Material 6^ (275.8KB, docx) Acknowledgements