Skip to main content

Differential gene expression of salt-tolerant alfalfa in response to salinity and inoculation by Ensifer meliloti

Abstract

Background

Alfalfa (Medicago sativa L.) experiences many negative effects under salinity stress, which may be mediated by recurrent selection. Salt-tolerant alfalfa may display unique adaptations in association with rhizobium under salt stress.

Results

To elucidate inoculation effects on salt-tolerant alfalfa under salt stress, this study leveraged a salt-tolerant alfalfa population selected through two cycles of recurrent selection under high salt stress. After experiencing 120-day salt stress, mRNA was extracted from 8 random genotypes either grown in 0 or 8 dS/m salt stress with or without inoculation by Ensifer meliloti. Results showed 320 and 176 differentially expressed genes (DEGs) modulated in response to salinity stress or inoculation x salinity stress, respectively. Notable results in plants under 8 dS/m stress included upregulation of a key gene involved in the Target of Rapamycin (TOR) signaling pathway with a concomitant decrease in expression of the SNrK pathway. Inoculation of salt-stressed plants stimulated increased transcription of a sulfate-uptake gene as well as upregulation of the Lysine-27-trimethyltransferase (EZH2), Histone 3 (H3), and argonaute (AGO, a component of miRISC silencing complexes) genes related to epigenetic and post-transcriptional gene control.

Conclusions

Salt-tolerant alfalfa may benefit from improved activity of TOR and decreased activity of SNrK1 in salt stress, while inoculation by rhizobiumstimulates production of sulfate uptake- and other unique genes.

Peer Review reports

Background

One of the greatest impacts of salinity stress on legumes, including alfalfa (Medicago sativa L.), is the abortion of nodulation by beneficial nitrogen fixing bacteria (i.e., Ensifer meliloti). Alfalfa associates with rhizobium such as E. meliloti in a mutualistic relationship, which involves the exchange of photosynthetically-derived sugars from the plant for biologically-fixed nitrogen from the bacteroid. Under non-stressful growth conditions, this exchange is mutually beneficial. However, in periods of environmental stress, such as salinity stress, the bacteroids may transit from being net-assets to net-parasites as they drain the plant of carbohydrates [1]. Under these conditions, the host plant will commonly terminate the relationship as it attempts to save valuable energy resources for its own usage, rather than exchanging them for fixed nitrogen [1,2,3].

A number of studies have been conducted to increase the salinity tolerance of the alfalfa-rhizobium complex by inoculating alfalfa with salt-tolerant bacteria [4,5,6]. While inoculation with salt-tolerant bacteria often provides benefits such as increased yield and vigorous plant growth, the superior salt tolerance of E. meliloti relative to alfalfa [7] suggests that it may be a yet-more effective approach to inoculate salt-tolerant alfalfa with E. meliloti to promote plant fitness under saline conditions.

The findings of previous transcriptomic studies over the past decade have revealed many stress-related genes in alfalfa. RNA-seq studies examining salt-tolerant vs susceptible cultivars have revealed that tolerant genotypes can display increased transcription for genes related to ROS scavenging, Ca2+ signaling (including the Salt Overly Sensitive (SOS1, SOS2, & SOS3) series of pathways), as well as High affinity Potassium Transporters (HKT) genes [8,9,10,11,12]. The complexity of these effects would most likely be compounded by inoculation and breeding, as inoculants themselves can stimulate up- or down regulation of genes [13], and genetic improvement can cause allelic shifts. Despite these interactions, few studies have examined the combination effects of selection and inoculation in salt-stressed alfalfa.

To address this knowledge gap, we conducted an experiment whereby a salt tolerant population of alfalfa was grown in non-saline (0 dS/m E.C.) or moderately (8 dS/m E.C.) saline conditions, with or without inoculation by E. meliloti. We predicted that inoculation would stimulate transcription of genes related to nitrogen metabolism, potentially related to bacteroid sourced nitrogen, as well as stress-related genes responsible for producing ROS scavengers, osmoprotectants, calcium signaling, etc. The objective of this project was thus to determine the gene expression profile of salt-tolerant alfalfa in response to inoculation by the rhizobium E. meliloti under 8 dS/m E.C. salinity stress.

Methods

Experimental design, plant & bacterial materials

The experiment, and breeding history of alfalfa populations, are described in our previous publication [14]. Briefly, a split-plot experiment with three treatments (rhizobium, three generations of alfalfa populations, and three salinity levels) was implemented. The bacteria used were Ensifer meliloti strain rm1021. Three alfalfa populations were developed through three cycles (Generations) of recurrent selection for salt tolerance. This experiment extracted RNA from Generation 2 (G2) population (the best performing alfalfa generation based on biomass, chlorophyll content, and osmoprotectant production), which originated from the 2nd cycle of selection at 20 dS/m E.C. salinity stress. Leaf samples were collected at 120 days from 0 and 8 dS/m treated G2 plants with or without rhizobium (E. meliloti) inoculation. Each sample taken was replicated twice for a total of 8 samples (1 alfalfa population × 2 rhizobium treatments × 2 salinity levels × 2 replicates).

RNA isolation, library preparation and sequencing

Samples were lyophilized in liquid nitrogen before being stored in a -80 °C freezer until analysis. A Qiagen RNeasy Plant mini kit (Qiagen Gmbh, Hilden, Germany) was used to extract RNA as per the manufacturer’s instructions. Briefly, plant tissues were ground in liquid nitrogen and homogenized with a QIAshredder during centrifugation. Ethanol was added to bind with RNA. The resulting solid was washed 3 × in the buffers provided, and the RNA was eluted in RNA free water. Quality of samples was determined through 260/280 nm readings on a Thermofisher NanoDrop 8000 (Waltham, MA, USA). After collection, 40 uL of resulting suspended mRNA from each sample was sent on ice to Novogene (California, USA) for library preparation and sequencing on an Illumina Novaseq 6000 platform (Illumina, San Diego CA, USA).

Mapping and differential expression analysis

Raw reads were cleaned of adapters using Trimmomatic v.0.36 software [15], with threads = 4, phred33, and default quality cut-off values. The quality of resulting files was checked with FASTQC software [16]. After checking file quality, reads were moved through the bioinformatics pipeline in the Digital Research alliance of Canada for most initial steps, with differential expression being calculated in R. Once prepared, trimmed reads were matched to the reference genome prepared by a previous work [17] (available at https://doi.org/10.6084/m9.figshare.12327602.v3) [18] using STAR (v2.6.1a)[19] from within the quantification software ‘RSEM’ with the –star flag (parameters from https://github.com/ENCODE-DCC/long-rna-seq-pipeline/blob/master/DAC/STAR_RSEM.sh) [20]. From the same code, read counts were conducted using the software ‘RSEM’ [21]. ‘isoform.results’ files were passed on to the differential expression analysis software ‘DESeq2’ [22] in R version 3.6.3 [23]. ‘DESeq2’ was deemed suitable for the dataset as it is capable of defining differentially expressed genes from experiments with 2–3 replicates by accounting for the high ‘noise’ of background expression levels through shrinking the dispersion values to an average fitted value [22]. Through this conservative approach, genes with very low dispersion values are fitted with a higher value to prevent false-positives, while values with dispersion values > 2 standard deviations above the fitted mean are treated ‘as-is’ also to prevent false positives. To further reduce the risk of false positives and discover significant DEGs, we identified DEGs using a Log fold change (Log2FC) > 2 and a stringent false discovery rate (FDR) of P < 0.001, an FDR which either matched or exceeded the stringency of many previous studies [8,9,10,11,12, 24,25,26]. P values were adjusted in DESeq2 using the Benjamini and Hochberg method. Significant DEGs were calculated separately for both salinity (8 vs 0 dS/m) and inoculant (inoculated vs non-inoculated) treatments as well as an interaction term (inoculated plants in 8 dS/m vs non-inoculated plants in 8 dS/m).

Functional annotation and analysis

Functional annotations of genes were retrieved for significant DEGs from the data frames prepared in association with the reference genome provided by a previous work [27]. These included Gene Ontology (GO) terms, KEGG Orthology (KO) terms, and a BLASTp search against NCBIs Non-Redundant (NR) protein database. KAAS analysis was also conducted with DEG associated protein sequences (also supplied by the previous work [27]) to elucidate relationships between DEGs in KEGG metabolic pathways. Furthermore, gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) enrichment tests were conducted. GO enrichment analyses were conducted using the R package ‘topgo’ [28]; GO terms were tested using Fishers exact test which utilizes a hypergeometric distribution, and were considered enriched with a P = 0.05. KEGG enrichment was also done in R with a hypergeometric calculation and with significance determined at P = 0.05.

Results

High throughput sequencing and assembly

Illumina sequencing generated total 173,023,874 paired end (PE) raw reads for the 8 samples sequenced. Of the reads, 171,165,017 (98.9%) remained after trimming, and 94.91% of trimmed reads were mapped to the alfalfa reference genome using STAR (v2.7.9). All samples showed a high percentage of mapping to the alfalfa reference genome, ranging from 91.96% to 95.97% of reads mapped. There were a high percentage of multi-mapped reads. For information regarding raw and trimmed read counts and lengths, refer to Table 1.

Table 1 Summary of Illumina sequencing data and mapped read data for the assayed alfalfa samples. Samples are given in Replicate-bacterial treatment (Control vs Ensifer meliloti). Salinity level (0 vs 8 dS/m)

Identification of DEGs

Between all treatments, there were many up- and down-regulated DEGs identified after mapping to the alfalfa reference genome. In total, there were 495 significant genes including 364 rhizobia-related genes, 320 salinity-related genes, and 179 interaction- (salinity x rhizobium) related genes. Specifically, rhizobium inoculated plants displayed 173 (47.5%) upregulated and 191 (52.5%) downregulated genes relative to non-inoculated plants. Salinity treated plants displayed 158 (49.4%) upregulated and 162 (50.6%) down-regulated genes compared to control plants. Inoculated, salt-stressed plants displayed 76 (42.5%) up-regulated and 103 (57.5%) down-regulated genes relative to non-inoculated, salt-stressed plants (Fig. 1).

Fig. 1
figure 1

Venn diagram for number of overlapping DEGs in leaves of salt tolerant alfalfa affected by rhizobium inoculation, salinity, and salinity x rhizobium inoculation

Functional annotation of DEGs

GO enrichment

Of the total 495 significant DEGs, 251 salinity, 260 rhizobium, and 136 of their interaction were annotated into the three GO categories: Molecular Function (MF), Biological Process (BP), and Cellular Component (CC). In addition to this information, the proportion of total DEGs assigned to a GO category are given in Table 2. Significant candidate genes annotated to GO terms are also given in Table 3. There were a total of 12 enriched GO terms across BP and CC categories. There were no enriched terms within the MF category. Within BP, there were two significant GO terms; GO:0001505 (regulation of neurotransmitters) and GO:0000103 (sulfate assimilation, refer to Fig. 2A). Within salinity-related DEGs, GO:0001505 was annotated to two significant DEGs (MS.gene21474.t1, MS.gene012657.t1), while within the salinity x rhizobium interaction, GO:0000103 was annotated to 1 gene (MS.gene006381, see Table 3). Of salinity-related DEG in the CC category, there were 10 enriched GO terms (see Table 4, refer to Fig. 2B).

Table 2 Total number of DEGs as well as the number and percent annotated with GO categories
Table 3 List of candidate genes associated with the enriched GO terms, as well as KEGG or Nr annotations (if present)
Fig. 2
figure 2

Gene ontology (GO) analysis of differentially expressed genes (DEGs) in leaf tissues of salt tolerant alfalfa: Biological Process GO graph of enriched terms annotated to inoculation x salinity DEGs; Cellular Component GO graph annotated to salinity DEGs. P values and the number of significant/total annotations given in each box. Color of boxes represents most to least significant from Red–orange-yellow. Enrichment based on Fishers exact test

Table 4 List of significantly enriched GO terms in the Cellular Component category within the salinity pool of differentially expressed genes

KEGG enrichment

A total of 154 unique KEGG pathways were assigned to the 495 total DEGs (Table S1). Of the 320 salinity-related DEGs, 85 (26.6%) were assigned KO terms. Of the 364 rhizobia-related DEGs, 90 (24.7%) were assigned KO terms. Of the 179 DEGs expressed in the salinity x rhizobium interaction pool, 46 (25.7%) were assigned KO terms. Hypergeometric enrichment tests for salinity, rhizobia, and interaction DEGs returned no significantly enriched KEGG Orthology terms. For KEGG Orthology terms and the associated number of DEGs in a given DEG pool refer to Table S1.

The most common KEGG term within salinity-related DEGs was K17606 (annotated to 3 DEGs), coding for the TAP46 protein. Also, of note was K05016, coding for chloride channel 7 (2 DEGs), K09489 coding for the heat shock 70 kDA protein 4 (2 DEGs), and K09489 coding for the enhanced disease susceptibility 1 (EDS1) protein (2 DEGs). The most common KEGG terms among rhizobia-related DEGs (2 DEGs each) were K18875 (EDS1, enhanced disease susceptibility protein 1), K10523 (SPOP, speckled-type POZ protein), K17302 (COPB2/SEC27 protein coatomer subunit beta), K00966 (GMPP protein, mannose-1-phosphate guanylytransferase), K11253 (H3, histone H3), and K22520 (LQY1, protein disulfide isomerase).

There were no KEGG terms annotated to more than 1 DEG in the salinity x rhizobium interaction gene pool (refer to Table S1). K00830 (annotated to 1 DEG in the interaction pool) is associated with the glycine, serine, and threonine metabolism, similar to K02437. There were a total of 15 DEGs related to transcriptional or translational machinery that were significantly changed in rhizobium inoculated alfalfa at 8 dS/m salinity stress (Table 5).

Table 5 Transcriptional/translational related KEGG Ontology terms with gene identifiers, log2foldchange values, adjusted p-values, and brief protein descriptions

Discussion

This experiment was designed to determine differential gene expression of salt-tolerant alfalfa stimulated by the rhizobium inoculant (E. meliloti) with or without salinity stress. This study detected a total of 495 DEGs related to salinity stress, inoculation, or inoculation x salinity stress. Many DEGs were related both to either salinity stress (320 DEGs) or rhizobium inoculation (364 DEGs), however fewer for the interaction between the two (179 DEGs). Of these DEGs, 78.4% of salinity-related DEGs, 71.4% of inoculant-related DEGs, and 76% of rhizobium x salinity interaction-related DEGs matched to GO terms, respectively. Of the 12 significantly enriched GO terms, 10 were associated with CC between salt-treated and non-salt treated plants; these GO terms all referenced membranes, vacuoles, and/or proton pumping. The remaining 2 were associated with BP (one related to salt vs non-salt treated plants and one related to the rhizobium x salinity interaction). The salinity-related term referenced the GO term “regulation of neurotransmitters” (NR/KEGG annotation “Glycine Cleavage System H protein”), while the interaction-related term referenced “sulfate assimilation”.

Ion transport and storage related DEGs under salt stress

There were 11 enriched GO terms within salinity-related DEGs (Table 1). All of the enriched CC GO terms were associated with two genes; MS.gene 067465.t1 and MS.gene50737.t1. Furthermore, all of the 10 CC GO terms referenced cell membranes, intracellular organelles (including vacuoles), or membrane transporters, which highlights the importance of ion transport and storage in the salt tolerant alfalfa.

MS.gene50737.t1 was annotated to the GO term GO:0000220 (V-type ATPase) and the KEGG term K02154 (V-type H+ transporting ATPase sub-unit a). The V-type (V = vacuolar) transporting ATPase is an ATP powered H+ transporter which creates a proton gradient across membranes [29, 30]. V-type ATPases are important for the stress response of the plant [29], whereby they are capable aiding in the compartmentalization of toxic ions [31]. Various ATPases have been found to be associated with salinity tolerance of alfalfa in many previous transcriptomic studies [8, 10, 11, 24].

MS.gene067465.t1 was associated with GO:0043231 (intracellular membrane bounded organelle), which had no associated KEGG term. However, this gene was associated with the NR accession KEH39627, which codes for a SNAP receptor complex protein ( SNARE protein). SNARE proteins can be divided into v-type (vesicular) or t-type (target) and are responsible for facilitating vesicular transport between intracellular bodies [32]. Specific SNAREs have been shown to be active in transporting intracellular Na+ ions to the vacuole for compartmentalization [33], and have been identified as being upregulated in a previous transcriptomic study of wild rice (Oryza australiensis) when experiencing ionic stress [34]. Taken together, these results show an important role of vacuolar ion storage in the salinity tolerance of alfalfa.

In addition to vacuolar storage, 3 DEGs coding for ion channels were also upregulated in salt-stressed alfalfa. Two of these DEGs (MS.gene64059.t1, MS.gene74439.t1) coded for chloride channels (CLCN7, K05016), which are anion transporters responsible for transporting Cl and NO3 [35], and one (MS.gene20294.t1) coded for a voltage dependant anion channel (VDAC2, K15040). An NR protein database cross-search returned separate terms for the CLC proteins: CLCN7 = CLC-a & CLC-c. CLC-a is a nitrate selective transporter while CLC-c is more chloride specific [36, 37]. CLCs are important for maintaining a healthy ratio of Cl:NO3, as excess Cl can impede nitrate assimilation [37, 38]. Both of the DEGs associated with CLCN7 were upregulated under 8 dS/m salinity stress, indicating that the salt-tolerant population used in this study had an intrinsic mechanism to attenuate chloride stress, thereby possibly increasing nitrate uptake. In our previous paper [14] we reported that nitrogen appeared to be a key factor in mediating salinity stress in salt-adapted alfalfa.

VDACs[1–5] are the major transport protein of the mitochondria and serve diverse functions, including transporting anions, accepting tRNA, ATP, and Ca+ ions, and mediating ROS gradients [39,40,41]. In this study, we found the transcripts of VDAC2 to increase in salt stressed plants. This is in contrast to recent work in A. thaliana and Tritichum aestivum L. where short term salt stress decreased transcription [42, 43]. The results of the current experiment suggest that after the potential decrease in expression noted by the previous studies, expression of VDAC2 may increase in alfalfa in response to a prolonged period of salinity stress.

TOR, SNrK, and general stress response-related DEGs

Salt-adapted plants displayed upregulation of three transcripts related to the TAP42/TAP46 protein (MS.gene06912.t1, MS.gene06913.t1, and MS.gene49210.t1) as well as upregulation of a SNF1 (a.k.a SNrK, MS.gene75239.t1) protein. TAP42 is an important protein, which responds downstream to the TOR signalling pathway [44]. The TOR pathway, when active, promotes translation, anabolism, as well as nutrient uptake and metabolism of nitrogen. However, when TOR is deactivated by rapamycin, K+ deficiencies may be caused in addition to programmed cell death (PCD), and nutrient recycling [44, 45]. Research has shown Tap42- mutants to deactivate the TOR pathway and cause major downstream effects including reduced activity of nitrogen reductase (Nr) and other nitrogen metabolic genes [44, 45]. TOR’s close counterpart is the Sucrose Non-feRmenting Kinase (SNrK) pathway, controlled by the SNrk gene. The SNrK pathway is generally stimulated by the same factors which inhibit TOR [46]. The SNrK gene was downregulated in salt-stressed alfalfa. In contrast to salt-stressed plants, rhizobium-inoculated salt-stressed plants displayed upregulation of the SNrK1 gene and no change in the TAP42 gene, which suggests that growth of salt-tolerant alfalfa was inhibited by inoculation by E. meliloti under salt stress.

Interestingly, the AuTophaGy 1 gene (ATG1, MS.gene92581.t1), a downstream effector proteins of the TOR pathway, was also downregulated in salt-stressed plants. This protein is responsible for autophagy, and the ATG1 gene is modulated by the TOR complex, which in turn is stimulated in times of nitrogen starvation [47]. The fact that ATG1 was downregulated in salt-stressed vs control plants reinforces that the TOR complex was stimulated in salt-tolerant alfalfa.

There were 3 DEGs annotated to heat shock proteins, including HSP90 (K09487, MS.gene64023.t1) and HSPA4 a.k.a HSP70 (K09489, MS.gene33363.t1, MS.gene33365.t1). Both HSP90 and one of the HSP70 proteins were downregulated under salt stress, however this is congruent with the diverse nature of heat shock proteins. Heat shock proteins are not only metabolized in response to heat stress but are also produced or repressed in response to multiple other forms of abiotic stress including drought, cold, and salinity stress [48], and are reported by a previous transcriptomic study examining salt stress in alfalfa [10]. Previous work [48] reports that the upregulated HSP70 protein found in this study responds both to salinity and heat stress in plants. Interestingly, HSP70 may be transcriptionally regulated by a heat shock factor HSFA4, which also interacts with the SOS1 pathway to induce ion homeostasis under salt-stress [49].

Disease resistance DEGs under salt stress

Similar to a previous study [11], this study found several immune-related genes modulated in response to salinity stress in alfalfa, including the relA, SPOP, SIAH1, and EDS1 genes. GTP pyrophosphokinase (relA, K00951, MS.gene26785.t1) was upregulated in salt stressed plants. This gene is also conserved in the chloroplasts of plants, and plays diverse roles in modulating transcription and translation to many stresses including pathogens and salinity [50]. Interestingly, the relA gene was upregulated under salinity stress, but downregulated in rhizobium inoculated, salt-stressed plants. The relA gene of the bacteroid, or its product ppGpp, have previously been shown to be important for the successful mutualism between legume and rhizobium [51,52,53]. It would be interesting for future studies to examine the root transcriptomics in rhizobium inoculated, salt-stressed alfalfa to determine the effects on relA transcription.

A speckle-type POZ protein (SPOP, K10523, MS.gene63029.t1), an E3-ubiquitin-protein ligase (SIAH1, K04506, MS.gene85969.t1), and two enhanced disease susceptibility protein (EDS1, K04506, MS.gene38404.t1, MS.gene38408.t1) genes were also upregulated in salt-stressed vs control plants. The SPOP gene codes for production of the Broad complex-Tramway-Bric-a-brac (BTB)-POZ protein, which gives specificity to the Cul3 Cullin-Ring E3 ligases (the E3-ligase upregulated in this study) when ubiquinating cellular compounds for disposal [54,55,56,57]. Interestingly, E3-ligases were also implicated in the salt-stress response of alfalfa in previous transcriptomic studies examining salt-stressed alfalfa [8, 26, 27]. One of the functions of this ligase is to mediate levels of the EDS1 protein, upregulated in this study. EDS1 is a central protein to much of plants’ immune-response [58, 59], however research has also suggested that it is required for the functioning of specific drought-response genes in Arabidopsis thaliana [60]. The upregulation of the various components of the EDS1 pathway, in addition to corroboration by a previous transcriptomic study, suggests an important role of the pathway in mediation of salt-stress in alfalfa.

Effect of rhizobium inoculation on salt-stressed plants

The results of the differential expression analysis showed that there were 179 genes which were differentially expressed in rhizobium inoculated, salt-stressed plants. Of these 179 genes, 76 were upregulated and 103 were downregulated. Inoculation of salt-stressed alfalfa induced 46 genes with known KEGG proteins. Several of these were directly or indirectly associated with salinity tolerance in relation to epigenetics, nutrient uptake, osmoprotectants, or photorespiration.

Fifteen DEGs related to transcriptional, translational, or other forms of genetic infrastructure in the host plants were modulated in inoculated, salt-stressed plants (see Table 5). These included a DEAD box RNA helicase (DDX5, K12823, MS.gene00082.t1), an argonaute protein (AGO, K11593, MS.gene05149.t1), and several genes related to epigenetic control (EZH2, K11430, MS.gene07294.t1 & H3, K11253, MS.gene04748.t1). Of these proteins, DEAD box helicases have previously been shown to mediate salt stress in A. thaliana by increasing activity of ROS scavengers and stimulating production of the osmoprotectant proline [61], which was a key finding in our original work [14]. An upregulated RNA helicase gene has also been previously reported in salt-tolerant alfalfa [8], and a previous review [62] reports many studies in which chloroplastic/mitochondrial DEAD box helicases mediate various abiotic stresses.

The argonaute (AGO) gene was also upregulated in inoculated/salt-stressed plants. argonaute is the housing protein into which micro-RNAs (miRNA) are housed to create an miRNA-Induced-Silencing Complex miRISC; miRISCs have been implicated in plant responses to many stresses whereby they silence translation of specific mRNA transcripts [63], and non-coding RNAs (of which miRNAs are a subset) are reported in previous transcriptomic work with salt-stressed alfalfa [24]. The fact that AGO was upregulated in inoculated/salt-stressed plants indicates that post-transcriptional control was taking place to modify the expression of potentially stress-related genes. Interestingly, assembly of the miRISC complex requires HSP90 [63], another DEG found in this study.

Perhaps the most interesting effects were the upregulation of both the histone H3 (H3, MS.gene04748.t1) as well as the lysine 27 N-trimethyltransferase (EZH2 a.k.a H3K27me3, MS.gene07294.t1), both related to epigenetics. Five variants of histones work together to form a nucleosome in combination with DNA [64]: H2A, H2B, H3, and H4 (with H5 serving as a link histone [65]). DNA is wrapped around these histones for storage but unwound for transcription. Histones can be methylated, acetylated, or phosphorylated at their N-terminus (tail end) to promote or preclude transcription, causing epigenetic control of genes [66]. Recent work with Castor bean (Ricinus communis) under salinity stress [67] has shown that the EZH2 (a.k.a H3K27me3) enzyme exerts epigenetic control over salt-stress related genes. The authors report that the addition of H3K27me3 to histones was correlated with DEG repression, while its loss was correlated with normal transcription. Another study [68] showed that removal of H3K27me3 from the High Affinity Potassium transporter 1 (HKT1) gene in A. thaliana induced an improved ability to mediate salt stress in future events. Previous transcriptomic studies examining salt-stressed alfalfa have also reported epigenetic-related methyltransferases/demethylases [10, 11]. Very interestingly in this study, the EZH2 DEG was significantly downregulated in inoculated plants whereas inoculated/salt-stressed plants experienced a large-scale increase in transcription of EZH2. This result matches another recent transcriptomic study which also found upregulation of a H3-lysine methyltransferase in salt-tolerant alfalfa [8].

The significant upregulation of H3K27me3 under inoculated-saline conditions in this study clearly indicates that E. meliloti is associated with inhibition of gene transcription for stress-related genes in alfalfa. As the addition/removal of H3K27me3 can prime plants for future stress, understanding the relationship between this bacteria and plant transcription could serve to protect plants from future stress events. Additionally, the overlap between breeding and epigenetics is beginning to receive attention (i.e., [69]), and future study may benefit from examining the effects of mutualistic-bacterial inoculation on this phenomenon, as bacteria could serve as a mechanism to repress/de-repress transcription of important genes.

DEGs for nutrient uptake under salt stress

The Adenosine-5-phosphosulfate reductase (APS reductase, K13811, MS.gene006381.t1) and Fe-S Cluster Assembly ATP-binding protein (sufC, K09013, MS.gene55865.t1) genes related to sulfur metabolism were upregulated in rhizobium inoculated, salt-stressed plants. The APS reductase gene was annotated by the enriched GO term GO:0000103 (“sulfate assimilation”). APS reductase is one of the key enzymes controlling the assimilation of sulfate, and the production of the plants first sulfur containing product: cysteine [70]. The APS reductase enzyme requires an Fe-S cluster as a cofactor [71], and the sufC gene also upregulated in this study, is at least partially responsible for synthesizing Fe-S clusters [72]. Sulfur metabolism, with cysteine being the organic building block, is vital for the stress response of plants as the important stress-signalling hormones abscisic acid, salicylate, and jasmonate in addition to the ROS scavenger glutathione are all sulfur based [73, 74]. However, we did not discover significant DEGs related to N update under stress.

Osmoprotectants and photorespiration related DEGs

The Xaa-Pro peptidase (pepP, K01262, MS.gene89849.t1) gene related to the osmoprotectant proline, alpha-galactosidase (galA, K07407, MS.gene87685.t1) gene responsible for galactinol/myo-inositol conversions, and a gene related to glycine (AGXT, K00830, MS.gene92042.t1) production were all downregulated in inoculated/salt-stressed plants. pepP proteins can act to cleaving off any amino acid off of the N-terminal of a peptide that is attached to a proline element [75], which can prepare proteins with an N-terminal proline to be acted upon by a proline aminopeptidase (PAP), which have been shown to increase proline levels in response to abiotic stress [76]. Proline acts both to combat ROS accumulation and to stabilize osmotic potential across membranes [77, 78]. In contrast to proline, the galA gene is responsible for myo-inositol/galactinol interconversions; myo-inositol is an important osmoprotectant while galactinol may serve as an ROS scavenger [79, 80]. Modulation of beta-galactosidase and inositol transporter genes have been implicated by previous transcriptomic studies of salt-stressed alfalfa [10, 11]. Another interesting result was the downregulation of the AGXT gene. AGXT directly governs the reversible reaction creating glycine (a vital componenet of the osmoprotectant glycine betaine) from serine glyoxylate [81]. As these enzymes catalyzed reversible reactions, the effect on the pool of osmoprotectants is unknown.

Another interesting result was the down regulation of the glycine cleavage proteins in inoculated/salt-stressed plants, which also acts upon the compound glycine. The glycine cleavage proteins are part of the glyoxylate decarboxylase complex (GDC) which is composed of glycine cleavage proteins P, L, T, and H. gcvH (K02437, MS.gene21474.t1) was strongly downregulated in inoculated/salt-stressed plants. The GDC complex is responsible for creating serine from glycine in a two-step reaction within the mitochondria [82], an important step in the photorespiratory pathway in which phosphoglycolate is recycled into 3-phosphoglycerate after oxygen is mistaken for CO2 by the rubisco enzyme. Salt-stressed vs control plants in this trial exhibited elevated expression of the gcvH complex, however expression in inoculated/salt-stressed plants was depressed showing that bacteria were associated with a strong effect on transcription of this gene. Recent studies [83, 84] clearly demonstrate that mutualistic bacteria can reduce the rates of photorespiration in salt-stressed crops by overexpressing photorespiratory enzymes, including elements of the glycine cleavage system.

Conclusion

Altogether, results of this study show that salt-tolerant alfalfa responded to 8 dS/m salinity by increasing production of ion transporters and by repressing PCD and nitrogen catabolism through a promoted TOR and inhibited SNrK pathway. E. meliloti caused both advantageous and deleterious changes to gene expression levels of salt-stressed alfalfa; these included stimulating sulfate assimilation through upregulation of APS reductase, but also downregulating genes responsible for creating more free proline (pepP) and effective photorespiration (gcvH). An ambiguous result which may deserve future research was the differential expression of genes related to epigenetics and post-transcriptional control of genes, including the EZH2 and AGO genes seen in inoculated/salt-stressed plants. The effect on these two genes offers a tantalizing glimpse at the complex genetic relationship between salt stress, legumes, and rhizobium.

Availability of data and materials

NCBI SRA Database, Accession PRJNA1071613.

Abbreviations

KEGG:

Kyoto Encyclopedia of Genes and Genomes

RNA-seq:

RNA Sequencing

GO:

Gene Ontology

NR:

NCBI Non-Redundant Protein Database

TOR:

Target of Rapamycin

SNrK:

Sucrose Non Fermenting Kinase

EZH2:

Histone H3 Lysine-27-Trimethyltransferase

AGO:

Argonaute

CLC:

Chloride Channel

VDAC:

Voltage Dependent Anion Channel

TAP42:

Type 2A Phosphatase Associated Protein 42Kd

PP2A:

Protein Phosphatase 2A

ATG1:

Autophagy 1

ppGpp:

Guanosine Tetraphosphate

SPOP:

Speckle type POZ Protein

EDS1:

Enhanceded Disease Susceptibility 1

HSP:

Heat Shock Protein

pepP:

Proline Peptidase

AGXT:

Alanine Glyoxylate Transaminase

gcvH:

Glycine Cleavage system H protein

GDC:

Glyoxylate Decarboxylase Complex

References

  1. Ferguson BJ, Mens C, Hastwell AH, Zhang M, Su H, Jones CH, et al. Legume nodulation: the host controls the party. Plant Cell Environ. 2019;42(1):41–51.

    Article  CAS  PubMed  Google Scholar 

  2. Prell J, White JP, Bourdes A, Bunnewell S, Bongaerts RJ, Poole PS. Legumes regulate Rhizobium bacteroid development and persistence by the supply of branched-chain amino acids. Proc Natl Acad Sci U S A. 2009;106(30):12477–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Chakraborty S, Harris JM. At the crossroads of salinity and rhizobium-legume symbiosis. Mol Plant-Microbe Interact. 2022;35(7):540–53.

    Article  CAS  PubMed  Google Scholar 

  4. Martínez R, Espejo A, Sierra M, Ortiz-Bernad I, Correa D, Bedmar E, et al. Co-inoculation of halomonas maura and ensifer meliloti to improve alfalfa yield in saline soils. Appl Soil Ecol. 2015;87:81–6 Available from: https://doi.org/10.1016/j.apsoil.2014.11.013 .

    Article  Google Scholar 

  5. Noori F, Etesami H, Najafi Zarini H, Khoshkholgh-Sima NA, Hosseini Salekdeh G, Alishahi F. Mining alfalfa (Medicago sativa L.) nodules for salinity tolerant non-rhizobial bacteria to improve growth of alfalfa under salinity stress. Ecotoxicol Environ Saf. 2018;162(June):129–38.

    Article  CAS  PubMed  Google Scholar 

  6. Bertrand A, Gatzke C, Bipfubusa M, Lévesque V, Chalifour FP, Claessens A, et al. Physiological and biochemical responses to salt stress of alfalfa populations selected for salinity tolerance and grown in symbiosis with salt-tolerant rhizobium. Agronomy. 2020;10(4):569. https://doi.org/10.3390/agronomy10040569.

    Article  CAS  Google Scholar 

  7. Talebi MB, Bahar M, Saeidi G, Mengoni A, Bazzicalupo M. Diversity of sinorhizobium strains nodulating medicago sativa from different iranian regions. FEMS Microbiol Lett. 2008;288(1):40–6.

    Article  CAS  PubMed  Google Scholar 

  8. Bhattarai S, Fu YB, Coulman B, Tanino K, Karunakaran C, Biligetu B. Transcriptomic analysis of differentially expressed genes in leaves and roots of two alfalfa (Medicago sativa L.) cultivars with different salt tolerance. BMC Plant Biol. 2021;21(1):1–16.

    Article  Google Scholar 

  9. Lei Y, Xu Y, Hettenhausen C, Lu C, Shen G, Zhang C, et al. Comparative analysis of alfalfa (Medicago sativa L.) seedling transcriptomes reveals genotype-specific drought tolerance mechanisms. BMC Plant Biol. 2018;18(35). https://doi.org/10.1186/s12870-018-1250-4.

  10. Gruber MY, Xia J, Yu M, Steppuhn H, Wall K, Messer D, et al. Transcript analysis in two alfalfa salt tolerance selected breeding populations relative to a non-tolerant population. Genome. 2017;60(2):104–27.

    Article  CAS  PubMed  Google Scholar 

  11. Postnikova OA, Shao J, Nemchinov LG. Analysis of the alfalfa root transcriptome in response to salinity stress. Plant Cell Physiol. 2013;54(7):1041–55.

    Article  CAS  PubMed  Google Scholar 

  12. Li J, Ma M, Sun Y, Lu P, Shi H, Guo Z, et al. Comparative Physiological and Transcriptome Profiles Uncover Salt Tolerance Mechanisms in Alfalfa. Front Plant Sci. 2022;13(June):1–12.

    Google Scholar 

  13. Defez R, Andreozzi A, Dickinson M, Charlton A, Tadini L, Pesaresi P, et al. Improved drought stress response in alfalfa plants nodulated by an IAA over-producing Rhizobium Strain. Front Microbiol. 2017;8(Dec):1–13.

    Google Scholar 

  14. Lundell S, Biligetu B. Effects of soil-mutualistic bacterial inoculation on osmolyte production and growth in three generations of alfalfa (Medicago sativa) populations selected for salt tolerance. Plant Soil. 2024. https://doi.org/10.1007/s11104-024-06603-z.

  15. Bolger AM, Lohse M, Usadel B. Genome analysis Trimmomatic : a flexible trimmer for Illumina sequence data. 2014;30(15):2114–20.

    CAS  Google Scholar 

  16. Andrews S. FastQC: A quality control tool for high throughput sequence data. Online. 2023. Available from: https://www.bioinformatics.babraham.ac.uk/projects/fastqc/

  17. Chen H, Zeng Y, Yang Y, Huang L, Tang B, Zhang H, et al. Allele-aware chromosome-level genome assembly and efficient transgene-free genome editing for the autotetraploid cultivated alfalfa. Nat Commun. 2020;11(1). Available from: https://doi.org/10.1038/s41467-020-16338-x

  18. Zeng Y. genome fasta sequence and annotation files. Figshare. 2020. https://figshare.com/projects/whole_genome_sequencing_and_assembly_of_Medicago_sativa/66380.

  19. Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, et al. Sequence analysis. 2013;29(1):15–21.

    CAS  Google Scholar 

  20. Dobin A. ENCODE-DCC/long-rna-seq-pipeline. Github. 2015. p. 2. Available from: https://github.com/ENCODE-DCC/long-rna-seq-pipeline/blob/master/DAC/STAR_RSEM.sh

  21. Li B, Dewey CN. RSEM: accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinform. 2011;12:323. https://doi.org/10.1186/1471-2105-12-323.

    Article  CAS  Google Scholar 

  22. Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014;15(12):1–21.

    Article  Google Scholar 

  23. R Core Development Team. A language and environment for statistical computing. R Foundation for Statistical Computing. Vienna; 2023. Available from: https://www.r-project.org/

  24. Medina CA, Samac DA, Yu LX. Pan-transcriptome identifying master genes and regulation network in response to drought and salt stresses in Alfalfa (Medicago sativa L.). Sci Rep. 2021;11(1):1–16 Available from: https://doi.org/10.1038/s41598-021-96712-x .

    Article  Google Scholar 

  25. Luo D, Zhou Q, Wu Y, Chai X, Liu W, Wang Y, et al. Full-length transcript sequencing and comparative transcriptomic analysis to evaluate the contribution of osmotic and ionic stress components towards salinity tolerance in the roots of cultivated alfalfa (Medicago sativa L.). BMC Plant Biol. 2019;19(1):1–20.

    Article  Google Scholar 

  26. Kaundal R, Duhan N, Acharya BR, Pudussery MV, Ferreira JFS, Suarez DL, et al. Transcriptional profiling of two contrasting genotypes uncovers molecular mechanisms underlying salt tolerance in alfalfa. Sci Rep. 2021;11(1):1–15 Available from: https://doi.org/10.1038/s41598-021-84461-w .

    Article  Google Scholar 

  27. Pan L, Yu X, Shao J, Gao T, Zheng Y, et al. Transcriptomic profiling and analysis of differentially expressed genes in asparagus bean (Vigna unguiculata ssp. sesquipedalis) under salt stress. PLoS One. 2019;14(7):1–23.

    Article  Google Scholar 

  28. Alexa A, Rahnenfuhrer J. topGO: Enrichment analysis for Gene Ontology. Bioconductor. 2009:2–5. https://bioconductor.org/packages/release/bioc/vignettes/topGO/inst/doc/topGO.pdf.

  29. Dietz KJ, Tavakoli N, Kluge C, Mimura T, Sharma SS, Harris GC, et al. Significance of the V-type ATPase for the adaptation to stressful growth conditions and its regulation on the molecular and biochemical level. J Exp Bot. 2001;52(363):1969–80.

    Article  CAS  PubMed  Google Scholar 

  30. Kluge C, Lamkemeyer P, Tavakoli N, Golldack D, Kandlbinder A, Dietz KJ. cDNA cloning of 12 subunits of the V-type ATPase from Mesembryanthemum crystallinum and their expression under stress. Mol Membr Biol. 2003;20(2):171–83.

    Article  CAS  PubMed  Google Scholar 

  31. Koyro HW, Huchzermeyer B, Zörb C. Effects of hyperosmotic salinity on protein patterns and enzyme activities. Handb Plant Crop Physiol Third Ed. 2014;(i):487–507.

  32. Hong W. SNAREs and traffic. Biochim Biophys Acta. 2005;1744(3):493–517.

    PubMed  Google Scholar 

  33. Tarte VN, Seok HY, Woo DH, Le DH, Tran HT, Baik JW, et al. Arabidopsis Qc-SNARE gene AtSFT12 is involved in salt and osmotic stress responses and Na+ accumulation in vacuoles. Plant Cell Rep. 2015;34(7):1127–38 Available from: https://doi.org/10.1007/s00299-015-1771-3 .

    Article  CAS  PubMed  Google Scholar 

  34. Yichie Y, Hasan MT, Tobias PA, Pascovici D, Goold HD, Van Sluyter SC, et al. Salt-treated roots of oryza australiensis seedlings are enriched with proteins involved in energetics and transport. Proteomics. 2019;19(19):1–12.

    Article  Google Scholar 

  35. He J, Wang M, Li S, Chen L, Zhang K, She J. Cryo-EM structure of the plant nitrate transporter AtCLCa reveals characteristics of the anion-binding site and the ATP-binding pocket. J Biol Chem. 2023;299(2):102833 Available from: https://doi.org/10.1016/j.jbc.2022.102833 .

    Article  CAS  PubMed  Google Scholar 

  36. Hechenberger M, Schwappach B, Fischer WN, Frommer WB, Jentsch TJ, Steinmeyer K. A family of putative chloride channels from Arabidopsis and functional complementation of a yeast strain with a CLC gene disruption. J Biol Chem. 1996;271(52):33632–8 Available from: https://doi.org/10.1074/jbc.271.52.33632 .

    Article  CAS  PubMed  Google Scholar 

  37. Wei P, Che B, Shen L, Cui Y, Wu S, Cheng C, et al. Identification and functional characterization of the chloride channel gene, GsCLC-c2 from wild soybean. BMC Plant Biol. 2019;19(1):1–15.

    Article  Google Scholar 

  38. Hu Y, Schmidhalter U. Drought and salinity: a comparison of their effects on mineral nutrition of plants. J Plant Nutr Soil Sci. 2005;168(4):541–9.

    Article  CAS  Google Scholar 

  39. Hemono M, Ubrig É, Azeredo K, Salinas-Giegé T, Drouard L, Duchêne AM. Arabidopsis voltage-dependent anion channels (VDACS): overlapping and specific functions in mitochondria. Cells. 2020;9(4):1–14.

    Article  Google Scholar 

  40. Homble F, Kmita H, Saidani H, Leonetti M. Plant VDAC Permeability: Molecular Basis and Role in Oxidative Stress. In: Molecular Basis for Mitochondrial Signaling. 2017. p. 161–83. Available from: https://link.springer.com/ https://doi.org/10.1007/978-3-319-55539-3

  41. Galganska H, Karachitos A, Wojtkowska M, Stobienia O, Budzinska M, Kmita H. Communication between mitochondria and nucleus: putative role for VDAC in reduction/oxidation mechanism. Biochim Biophys Acta - Bioenerg. 2010;1797(6–7):1276–80 Available from: https://doi.org/10.1016/j.bbabio.2010.02.004 .

    Article  CAS  Google Scholar 

  42. Yu M, Yu Y, Song T, Zhang Y, Wei F, Cheng J, et al. Characterization of the voltage-dependent anion channel (VDAC) gene family in wheat (Triticum aestivum L.) and its potential mechanism in response to drought and salinity stresses. Gene. 2022;809(October 2021):146031 Available from: https://doi.org/10.1016/j.gene.2021.146031 .

    Article  CAS  PubMed  Google Scholar 

  43. Sanyal SK, Kanwar P, Fernandes JL, Mahiwal S, Yadav AK, Samtani H, et al. Arabidopsis mitochondrial voltage-dependent anion channels are involved in maintaining reactive oxygen species homeostasis, oxidative and salt stress tolerance in yeast. Front Plant Sci. 2020;11(February):1–14.

    Google Scholar 

  44. Ahn CS, Han JA, Lee HS, Lee S, Pai HS. The PP2A regulatory subunit Tap46, a component of the TOR signaling pathway, modulates growth and metabolism in plants. Plant Cell. 2011;23(1):185–209.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Deng K, Wang W, Feng L, Yin H, Xiong F, Ren M. Target of rapamycin regulates potassium uptake in Arabidopsis and potato. Plant Physiol Biochem. 2020;155(July):357–66 Available from: https://doi.org/10.1016/j.plaphy.2020.07.044 .

    Article  CAS  PubMed  Google Scholar 

  46. Rodriguez M, Parola R, Andreola S, Pereyra C, Martínez-Noël G. TOR and SnRK1 signaling pathways in plant response to abiotic stresses: do they always act according to the “yin-yang” model? Plant Sci. 2019;288(April): 110220.

    Article  CAS  PubMed  Google Scholar 

  47. Wang Q, Hou S. The emerging roles of ATG1/ATG13 kinase complex in plants. J Plant Physiol. 2022;271(March):153653 Available from: https://doi.org/10.1016/j.jplph.2022.153653 .

    Article  CAS  PubMed  Google Scholar 

  48. Andrási N, Pettkó-Szandtner A, Szabados L. Diversity of plant heat shock factors: regulation, interactions, and functions. J Exp Bot. 2021;72(5):1558–75.

    Article  PubMed  Google Scholar 

  49. Li F, Zhang H, Zhao H, Gao T, Song A, Jiang J, et al. Chrysanthemum CmHSFA4 gene positively regulates salt stress tolerance in transgenic chrysanthemum. Plant Biotechnol J. 2018;16(7):1311–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Boniecka J, Prusińska J, Dąbrowska GB, Goc A. Within and beyond the stringent response-RSH and (p)ppGpp in plants. Planta. 2017;246(5):817–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Qiu X, Gao T, Yang J, Wang E, Liu L, Yuan H. Water-soluble humic materials modulating metabolism and triggering stress defense in sinorhizobium fredii. Microbiol Spectr. 2021;9(1): e0029321.

    Article  PubMed  Google Scholar 

  52. Moris M, Braeken K, Schoeters E, Verreth C, Beullens S, Vanderleyden J, et al. Effective symbiosis between rhizobium etli and phaseolus vulgaris requires the alarmone ppGpp. J Bacteriol. 2005;187(15):5460–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Wippel K, Long SR. Symbiotic performance of sinorhizobium meliloti lacking ppgpp depends on the medicago host species. Mol Plant-Microbe Interact. 2019;32(6):717–28.

    Article  CAS  PubMed  Google Scholar 

  54. Vierstra RD. The ubiquitin-26S proteasome system at the nexus of plant biology. Nat Rev Mol Cell Biol. 2009;10(6):385–97.

    Article  CAS  PubMed  Google Scholar 

  55. Chen L, Hellmann H. Plant E3 ligases: Flexible enzymes in a sessile world. Mol Plant. 2013;6(5):1388–404 Available from: https://doi.org/10.1093/mp/sst005 .

    Article  CAS  PubMed  Google Scholar 

  56. Song Y, Xu Y, Pan C, Yan L, Wang ZW, Zhu X. The emerging role of SPOP protein in tumorigenesis and cancer therapy. Mol Cancer. 2020;19(1):1–13.

    Article  Google Scholar 

  57. Ban Z, Estelle M. CUL3 E3 ligases in plant development and environmental response. Nat Plants. 2021;7(1):6–16 Available from: https://doi.org/10.1038/s41477-020-00833-6 .

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Wiermer M, Feys BJ, Parker JE. Plant immunity: the EDS1 regulatory node. Curr Opin Plant Biol. 2005;8(4):383–9.

    Article  CAS  PubMed  Google Scholar 

  59. Lapin D, Bhandari DD, Parker JE. Origins and immunity networking functions of eds1 family proteins. Annu Rev Phytopathol. 2020;58:253–76.

    Article  CAS  PubMed  Google Scholar 

  60. Chini A, Grant JJ, Seki M, Shinozaki K, Loake GJ. Drought tolerance established by enhanced expression of the CC-NBS-LRR gene, ADR1, requires salicylic acid, EDS1 and ABI1. Plant J. 2004;38(5):810–22.

    Article  CAS  PubMed  Google Scholar 

  61. Luo Y, Liu YB, Dong YX, Gao XQ, Zhang XS. Expression of a putative alfalfa helicase increases tolerance to abiotic stress in Arabidopsis by enhancing the capacities for ROS scavenging and osmotic adjustment. J Plant Physiol. 2009;166(4):385–94.

    Article  CAS  PubMed  Google Scholar 

  62. Nawaz G, Kang H. Chloroplast- or mitochondria-targeted DEAD-box RNA helicases play essential roles in organellar RNA metabolism and abiotic stress responses. Front Plant Sci. 2017;8(May):1–9.

    CAS  Google Scholar 

  63. Song X, Li Y, Cao X, Qi Y. MicroRNAs and their regulatory roles in plant-environment interactions. Annu Rev Plant Biol. 2019;70:489–525.

    Article  CAS  PubMed  Google Scholar 

  64. Scott WA, Campos EI. Interactions with histone H3 & Tools to study them. Front Cell Dev Biol. 2020;8(July):1–21.

    Google Scholar 

  65. Fan L, Roberts VA. Complex of linker histone H5 with the nucleosome and its implications for chromatin packing. Proc Natl Acad Sci U S A. 2006;103(22):8384–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Chang YN, Zhu C, Jiang J, Zhang H, Zhu JK, Duan CG. Epigenetic regulation in plant abiotic stress responses. J Integr Plant Biol. 2020;62(5):563–80.

    Article  CAS  PubMed  Google Scholar 

  67. Han B, Xu W, Ahmed N, Yu A, Wang Z, Liu A. Changes and associations of genomic transcription and histone methylation with salt stress in castor bean. Plant Cell Physiol. 2020;61(6):1120–33.

    Article  CAS  PubMed  Google Scholar 

  68. Sani E, Herzyk P, Perrella G, Colot V, Amtmann A. Hyperosmotic priming of arabidopsis seedlings establishes a long-term somatic memory accompanied by specific changes of the epigenome. Genome Biol. 2013;14:R59. https://doi.org/10.1186/gb-2013-14-6-r59.

  69. Saradadevi GP, Das D, Mangrauthia SK, Mohapatra S, Chikkaputtaiah C, Roorkiwal M, et al. Genetic, epigenetic, genomic and microbial approaches to enhance salt tolerance of plants: a comprehensive review. Biology. 2021;10(12):1255. https://doi.org/10.3390/biology10121255.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Gotor C, Laureano-Marín AM, Moreno I, Aroca Á, García I, Romero LC. Signaling in the plant cytosol: cysteine or sulfide? Amino Acids. 2015;47(10):2155–64.

    Article  CAS  PubMed  Google Scholar 

  71. Kopriva S, Büchert T, Fritz G, Suter M, Benda R, Schünemann V, et al. The presence of an iron-sulfur cluster in adenosine 5′-phosphosulfate reductase separates organisms utilizing adenosine 5′-phosphosulfate and phosphoadenosine 5′-phosphosulfate for sulfate assimilation. J Biol Chem. 2002;277(24):21786–91.

    Article  CAS  PubMed  Google Scholar 

  72. Ming XuX, Adams S, Chua NH, Møller SG. AtNAP1 represents an atypical SufB protein in Arabidopsis plastids. J Biol Chem. 2005;280(8):6648–54.

    Article  Google Scholar 

  73. Takahashi H, Kopriva S, Giordano M, Saito K, Hell R. Sulfur assimilation in photosynthetic organisms: molecular functions and regulations of transporters and assimilatory enzymes. Annu Rev Plant Biol. 2011;62:157–84.

    Article  CAS  PubMed  Google Scholar 

  74. Kwon DH, Cha HJ, Lee H, Hong SH, Park C, Park SH, et al. Protective effect of glutathione against oxidative stress-induced cytotoxicity in RAW 264.7 macrophages through activating the nuclear factor erythroid 2-related factor-2/heme oxygenase-1 pathway. Antioxidants (Basel). 2019;8(4):82. https://doi.org/10.3390/antiox8040082.

  75. Chen SJ, Kim L, Song HK, Varshavsky A. Aminopeptidases trim Xaa-Pro proteins, initiating their degradation by the Pro/N-degron pathway. Proc Natl Acad Sci U S A. 2021;118(43):1–7.

    Article  Google Scholar 

  76. Ghifari AS, Teixeira PF, Kmiec B, Singh N, Glaser E, Murcha MW. The dual-targeted prolyl aminopeptidase PAP1 is involved in proline accumulation in response to stress and during pollen development. J Exp Bot. 2022;73(1):78–93.

    Article  CAS  PubMed  Google Scholar 

  77. Bertrand A, Gatzke C, Bipfubusa M, Lévesque V, Chalifour FP, Claessens A, et al. Physiological and biochemical responses to salt stress of alfalfa populations selected for salinity tolerance and grown in symbiosis with salt-tolerant rhizobium. Agronomy. 2020;10(4):1–20.

    Article  Google Scholar 

  78. Fichman Y, Gerdes SY, Kovacs H, Szabados L, Zilberstein A, Csonka LN. Evolution of proline biosynthesis: enzymology, bioinformatics, genetics, and transcriptional regulation. Biol Rev. 2015;90:1065–99.

    Article  PubMed  Google Scholar 

  79. Nishizawa A, Yabuta Y, Shigeoka S. Galactinol and raffinose constitute a novel function to protect plants from oxidative damage. Plant Physiol. 2008;147(3):1251–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Zhang Q, Song X, Bartels D. Sugar metabolism in the desiccation tolerant grass Oropetium thomaeum in response to environmental stresses. Plant Sci. 2018;270(February):30–6 Available from: https://doi.org/10.1016/j.plantsci.2018.02.004 .

    Article  CAS  PubMed  Google Scholar 

  81. Fang Y, Coulter JA, Wu J, Liu L, Li X, Dong Y, et al. Identification of differentially expressed genes involved in amino acid and lipid accumulation of winter turnip rape (Brassica rapa L.) in response to cold stress. PLoS One. 2021;16(2 February):1–18 Available from: https://doi.org/10.1371/journal.pone.0245494 .

    Google Scholar 

  82. Maurino VG, Peterhansel C. Photorespiration: Current status and approaches for metabolic engineering. Curr Opin Plant Biol. 2010;13(3):248–55 Available from: https://doi.org/10.1016/j.pbi.2010.01.006 .

    Article  Google Scholar 

  83. Ilangumaran G, Subramanian S, Smith DL. Soybean leaf proteomic profile influenced by rhizobacteria under optimal and salt stress conditions. Front Plant Sci. 2022;13(March):1–15.

    Google Scholar 

  84. Taj Z, Challabathula D. Protection of photosynthesis by halotolerant staphylococcus sciuri ET101 in Tomato (Lycoperiscon esculentum) and Rice (Oryza sativa) plants during salinity stress: possible interplay between carboxylation and oxygenation in stress mitigation. Front Microbiol. 2021;11(January):1–24.

    Google Scholar 

Download references

Acknowledgements

We would like to thank Aabroo Ahmed for her help in RNA extraction.

Funding

Beef Cattle Research Council of Canda, NERC Discovery grant, and Agriculture Development Fund of Saskatchewan.

Author information

Authors and Affiliations

Authors

Contributions

B.B. Funding acquisition; S.L., B.B. designed experiments; S.L. conducted the experiment, analyzed the data, and wrote the original manuscript; B.B., S.L. edited drafts and final manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Bill Biligetu.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lundell, S., Biligetu, B. Differential gene expression of salt-tolerant alfalfa in response to salinity and inoculation by Ensifer meliloti. BMC Plant Biol 24, 633 (2024). https://doi.org/10.1186/s12870-024-05337-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-024-05337-5

Keywords