Skip to main content

Identification and expression analysis of expansin gene family in Salvia miltiorrhiza

Abstract

Background

Expansins (EXP) are important enzymes that are involved in the extension of plant cells and regulation of root configurations, which play important roles in resisting various stresses. As a model medicinal plant, Salvia miltiorrhiza is well recognized for treating coronary heart disease, myocardial infection, and other cardiovascular and cerebrovascular diseases; however, the SmEXP gene family has not yet been analyzed.

Methods

The SmEXP family was systematically analyzed using bioinformatics. Quantitative real-time PCR was employed to analyze the tissue expression patterns of the SmEXP family, as well as its expression under abscisic acid (ABA) treatment and abiotic stress. Subcellular localization assay revealed the localization of SmEXLA1, SmEXLB1, and SmEXPA2.

Results

This study identified 29 SmEXP that belonged to four different subfamilies. SmEXP promoter analysis suggested that it may be involved in the growth, development, and stress adaptation of S. miltiorrhiza. An analysis of the expression patterns of SmEXP revealed that ABA, Cu2+, and NaCl had regulatory effects on its expression. A subcellular localization assay showed that SmEXLA1 and SmEXLB1 were located on the nucleus and cell membrane, while SmEXPA2 was located on the cell wall.

Conclusion

For this study, the SmEXP family was systematically analyzed for the first time, which lays a foundation for further elucidating its physiological and biological functionality.

Introduction

Salvia miltiorrhiza Bunge (S. miltiorrhiza) is a well-recognized model medicinal plant, the dried roots of which are used for the treatment of cardiovascular and cerebrovascular diseases [1,2,3,4]. Plant cell walls are dynamic and complex structures that determine the shapes and sizes of plant cells by controlling the degree and direction of cell elongation [5]. Through a combination of cell wall loosening and cell expansion, plant cells have the capacity grow to more than 100 times the size of their initial meristem [6].

EXPANSIN (EXP) is a key regulator of cell wall elongation during plant growth [7, 8], which can be divided into four subfamilies: EXPA (Expansin A), EXPB (Expansin B), EXLA (Expansin-like A), and EXLB (Expansin-like B) [9]. EXP are a multi-gene family that are widely distributed in angiosperms, gymnosperms, ferns, and bryophytes. To date, the genome-wide identification and functional verification of EXP has been conducted for many plants, including Arabidopsis thaliana, Oryza sativa [10], Nicotiana tabacum [11], Solanum lycopersicum [12], Populus [13], Triticum aestivum [14], Glycine max [15], and Malus domestica [16].

EXP was initially discovered in Cucumis sativus in 1992. EXP can induce heat deactivated cell walls to stretch again; however, the stretching process is influenced by pH with the highest activity of this expansive protein occurring in the range of from pH 4.5–6. For example, AtEXPA1 requires low pH activation to exert the relaxation of cell wall activities [17]. There is ample evidence that EXP can relax and stretch the cell wall, albeit the mechanism is not clear. Currently, there are two hypotheses regarding the kinetics of EXP, with one being the acid growth theory and the other, the mechanistic non-enzymatic action model [10, 18].

The primary function of EXP is to cleave the non-covalent bonds between cellulose and other polysaccharides in the cell wall, such that it becomes relaxed and increasingly flexible to regulate plant growth and development [19, 20]. In addition to relaxing cellular walls, EXP also plays an important role in seed germination [21], leaf and stem development [22, 23], root elongation [24], fruit ripening [25], and plant stress resistance [26, 27].

It was found that the overexpression of NtEXPA1 in tobacco resulted in increased leaf and stem sizes [28]. OsEXPB2 affects root structures and plant heights by inhibiting cell growth [29]. Interference with OsEXPA8 in rice results in shorter primary fibrous roots and fewer lateral roots [30]. Other root hair EXPAs (OsEXPA30 and AtEXPA7) can complement the OsEXPA17 mutant to restore root hair elongation [31]. GmEXPB2 can adapt soybean to low phosphorus stress by modifying lateral root development and it configurations [32]. HvEXPB7 has also been shown to be associated with root hair formation in Hordeum vulgare, which was significantly inhibited after HvEXPB7 silencing [33]. Another study showed that SlEXP1 expression was closely related to the degree of fruit softening during fruit ripening [34]. ZmEXPB15, ZmNAC11, and ZmNAC29 impact the early development of corn kernels and regulates the grain size and weight by promoting the elimination of nucellar tissue [35]. Further, GhEXPA3-1 promotes fibrocyte elongation through the brassinolide signaling pathway [36].

Studies have revealed that EXP plays an important role in plant responses to abiotic stress. Compared with wild-type tobacco, NtEXPA11 overexpressed strains had strong roots, significantly increased leaf and internode length, greater cell wall flexibility, and improved tolerance to drought and salt stress [37]. The overexpression of EXPA4 in tobacco confers greater tolerance to salt and drought stress, which results in less cell damage and a higher fresh weight [38]. The overexpression of TaEXPA2 increased drought stress tolerance in transgenic tobacco [39] as well as wheat [40]. CqEXPA50 plays an important role in the exposure of quinoa seedlings to salt stress [41]. The overexpression of AtEXP3 and AtEXPB1 can enhance the sensitivity of Arabidopsis to salt stress [42]. OsEXPA7 enhances salt stress tolerance by coordinating sodium transport, ROS clearance, and cell wall loosening [43]. Further, plant cell walls can prevent heavy metals from entering its interior [44, 45].

Multiple studies have indicated that EXP can affect almost every plant growth stage. Meanwhile, S. miltiorrhiza is affected by several abiotic stress factors such as salt, drought, and heavy metals in the soil during its growth process. As the first barrier against external environmental stress, cell walls play a critical role in cell responses to external stress. Thus, it is essential to identify and analyze SmEXP to establish a foundation for the further study of its functionality.

Materials and methods

Identification and characterization of SmEXPs

The AtEXPs sequence was obtained from the TAIR database (https://www.arabidopsis.org/) [46], while the OsEXPs sequence was acquired through the Rice Genome Annotation Project (http:/ /rice.plantbiology.msu.edu/). The S. miltiorrhiza genome file was obtained from the China Traditional Chinese Medicine Data Center (https://ngdc.cncb.ac.cn/). BioEdit software (Borland, Scotts Valley, CA, USA) was used to compare AtEXPs with the S. miltiorrhiza protein database to obtain candidate proteins. The candidate SmEXPs protein sequences were submitted to InterProScan (https://www.ebi.ac.uk/interpro/), and a CD search (https://www.ncbi.nlm.nih.gov/cdd/) was conducted to further confirm whether Pollen_allerg_1 (PF01357) and DPBB_1 (PF03330) domains existed in the retrieved candidate SmEXPs protein sequences. It was confirmed that there were 29 SmEXPs in S. miltiorrhiza.

Phylogenetic analysis and cis-elements analysis

The phylogenetic tree of AtEXPs, OsEXPs, and SmEXPs proteins was constructed using MEGA X [47], which was beautified with the online website Evolview (http://www.evolgenius.info/evolview/).

Promoter sequences of 29 SmEXPs (− 1000 bp) were used to predict the cis-elements of the SmEXPs promoter region using PlantCARE (http://bioinformatics.psb.ugent.be/webtools/plantcare/htmL/) [48], whereas Tbtools v1.089 software (Chengjie Chen et al., China) [49] was used for visualization.

Physicochemical properties, gene structure, and Ka/Ks analysis

The SmEXP sequences were submitted to the ProtParam tool website (https://web.expasy.org/protparam/) [50] for an analysis of their amino acid numbers (AA), molecular weights (Mw), protein instability indices, isoelectric points (pH), and grand average of hydropathicity (GRAVY) and aliphatic indices. Plant-mPLoc (http://www.csbio.sjtu.edu.cn/bioinf/plant-multi/) was used for subcellular localization analysis. The gene structures and motifs of the SmEXPs were studied by Tbtools, and the Ka and Ks values were calculated using TBtools.

GO analysis and prediction of protein interactions

GO and KEGG analysis was performed using OmicShare tools, a free online platform for data analysis (https://www.omicshare.com/tools). The Orthovenn2 website (https://orthovenn2.bioinfotoolkits.net/home) was employed to search for the homologous EXP genes of A. thaliana and S. miltiorrhiza, while the sequence consistencies of AtEXP and SmEXP were calculated using DNAMAN (LynnonBiosoft, USA). The homologous genes of Arabidopsis that corresponded to SmEXPs were uploaded to the String10 website (http://string-db.org/) to visualize the interaction network.

Plant materials and treatments

Salvia miltiorrhiza plants were grown in a field nursery. The four tissue components of two-year-old S. miltiorrhiza plants (roots, stems, leavess, and flowers) were obtained from a resource nursery. The 2-month-old S. miltiorrhiza tissue culture seedlings were employed for hormone and stress treatments. The 2-month-old S. miltiorrhiza plants were sprayed with 100 μM ABA, 100 mm NaCl, and 200 μM Cu2+, respectively, and then sampled according to selected time points.

RNA extraction and quantitative real-time PCR (qRT-PCR) analysis

For qRT-PCR analysis, the total RNA was extracted from the S. miltiorrhiza plants using Plant Total RNA isolation kits (Aidlab, Beijing, China). Reverse transcriptase was employed to reverse transcribe the whole RNA to cDNA (Invitrogen, Waltham, MA, USA).

All qPCR assays were performed with the CFX96 real-time system (Bio-Rad, USA) using S. miltiorrhiza β-actin (DQ243702) [51] as the internal control, and quantification was performed by SYBR (Takara, Japan). Each reaction had three biological and technical replicates, using 40-fold diluted cDNA as a template. The expression of SmEXPs was calculated via the comparative CT method (2−ΔΔCT) [52]. One-way ANOVA was employed to examine the statistical significance of the data. The primer sequences used in this study are listed in Additional file 1: Table S1.

Subcellular localization of SmEXPA2, SmEXLA1, and SmEXLB1

The SmEXPA2, SmEXLA1, and SmEXLB1 without stop codons were cloned into the pEarleyGate103 (35-GFP) vector to obtain the recombinant vector GFP-SmEXPA2/SmEXLA1/SmEXLB1. The recombinant vector GFP-SmEXPA2/SmEXLA1/SmEXLB1 and 35-GFP were transformed to Agrobacterium EHA105. The bacterial solution carrying GFP-SmEXPA2/SmEXLA1/SmEXLB1 and 35-GFP were impregnated with onion epidermis and cultured for 2 days, respectively. After soaking, the onion skin was cultured in the dark on MS medium for 2 days, and then observed with a fluorescence microscope (Leica DM6000B, Wetzlar, Germany).

Result

Identification of SmEXP family

A total of 29 SmEXPs in the S. miltiorrhiza genome were found to contain both Pollen_allerg_1(PF01357) and DPBB_1(PF03330) domains. The sequences of each SmEXPs are shown in Additional file 2. The physicochemical properties of SmEXPs were then further analyzed (Table 1). The number of amino acids encoded by SmEXPs ranged from 241 to 328 aa, whereas the molecular weights of the proteins ranged from 26.05 to 36.19 kDa. The SmEXPs had a pI that ranged from between 4.83 and 9.64.

Table 1 Physicochemical properties of SmEXPs. Nomenclature, peptide lengths, molecular weights (MW), theoretical isoelectric points (pI), instability indices, aliphatic indices, Grand Average of Hydropathicity (gravy)

Furthermore, the GRAVY of SmEXPs was predicted to determine the hydrophilic and hydrophobic properties of the SmEXPs. The analysis revealed that 26 SmEXPs belonged to hydrophilic proteins, and three SmEXPs were hydrophobic proteins. SmEXPA21 exhibited the strongest hydrophilicity, while SmEXPB2 showed the greatest hydrophobicity. There were 24 SmEXPs with a low Instability index (Instability index < 40), indicating that most SmEXPs were stable, which is necessary for their functionality and the maintenance of normal plant growth. Subcellular localization prediction showed that all SmEXPs were localized to the cell wall.

Phylogenetic and Ka/Ks analyses

To explore the phylogenetic relationships between the SmEXPs, AtEXPs, and OsEXPs, 136 protein sequences were compared and phylogenetic trees were constructed (Fig. 1). The results revealed that the EXP members of S. miltiorrhiza, A. thaliana, and O. sativa could be divided into EXPA, EXPB, EXLA, and EXLB subfamilies. The SmEXPA subfamily, with the highest number, included 21 members. Next, the SmEXPB subfamily had five members, the SmEXLA subfamily had one member, and the SmEXLB subfamily had two members.

Fig. 1
figure 1

Phylogenetic relationships between EXP sequences. The full-length amino acid sequences of SmEXP, AtEXP, and OsEXP were used to construct the phylogenetic tree using MEGA X via the neighbor-joining (NJ) method. The value at the branch is the confidence level determined by performing 1000 bootstrap tests

According to the sequence similarity of the SmEXP family, 11 homologous gene pairs were selected for prediction (Table 2). It was predicted that the Ka/Ks ratio of all gene pairs were less than 1. This indicated a strong purifying selection in the evolution of the SmEXP family that tended to be more stable, which was conducive to maintaining the conservation of gene family functionality. The Ka/Ks ratio of the SmEXPA12&SmEXPA13 group was highest, which indicated that the evolution rate of this group was faster. Conversely, the Ka/Ks ratio of the SmEXPA6&SmEXPA7 group was lowest, which implied that the amino acid sites in this group were more conserved and not easily changed.

Table 2 Ka/Ks values of homologous SmEXP gene pairs

Motif and gene structure analyses

The gene structures and conserved motifs within each subfamily were almost consistent (Fig. 2A). Motif 5 was present in all subfamilies, while motif 6 was present in all SmEXPA subfamilies (except SmEXPA21). Motif 7 was present in all SmEXPB proteins (except for SmEXPB4) (Fig. 2B). The motif analysis revealed that the proteins translated from genes in the same subfamily had similar amino acid structures, which may have been related to tandem repetition, random repetition, and insertion during the genome evolution process. Most SmEXPs contained one-four introns, SmEXPA tended to have two introns, and SmEXPB tended to have three introns (Fig. 2C).

Fig. 2
figure 2

Evolutionary tree, conserved motifs, and gene structures of SmEXP family. A Phylogenetic relationship between SmEXPs. B Conserved motifs of SmEXPs. C Exon–intron structure of SmEXP. Yellow rectangles refer to UTR, and green rectangles refer to CDS

Cis-element SmEXP analysis

A total of 12 cis-elements were predicted, while the GATA-motif was associated with light response. The ABRE, TGACG-motif, P-box, and TCA-element were associated with hormone responses, and the WUN-motif was associated with wound response. Circadian was associated with circadian control, whereas CAT-box and CCGTCC-box were associated with tissue-specific and developmental associations, whereas MBS, Myb-binding sites, and LTR were correlated with abiotic stress (Fig. 3). The analysis results suggested that SmEXPs may have played important roles in the photoresponse, hormonal response, abiotic stress response-related, tissue-specific, and developmental responses.

Fig. 3
figure 3

Predicted cis-elements in the SmEXPs prompters (− 1000 bp). Different cis-elements are represented by different colors

GO annotation and prediction of protein interactions

The GO annotation was divided into three levels (molecular function (MF), biological process (BP), and cellular component (CC)). The SmEXPs had only biological process and cellular component annotations and GO annotation assisted with understanding the functions of proteins at the molecular level. At the CC level, the SmEXPs were primarily enriched in the extracellular region, while at the BP level the SmEXPs were primarily enriched in cellular processes, cellular component organization, or biogenesis (Fig. 4A).

Fig. 4
figure 4

Functional analysis. A GO terms enriched with SmEXPs. B Protein–protein interaction networks of SmEXPs. The strength of the protein interaction is shown by the line's thickness

Orthologous genes often have similar functions [53]; thus, the Arabidopsis interaction network to was used to map the potential interactive SmEXP network. OrthoVenn2 was employed to screen the orthologous genes of SmEXP in the Arabidopsis database (Table 3). The sequence alignment results of orthologous genes are shown in Additional file 3: Fig. S1. The results indicated that the sequence consistency of these 13 pairs of orthologous genes was > 50%. The results showed that AtEXPA4 could interact with AtEXPB3, and AtEXPA8 could interact with AtEXPA11. The interactive network of AtEXP18 was rich, and AtEXP18 could interact with another 10 genes to facilitate its functions (Fig. 4B).

Table 3 Orthologous SmEXP and AtEXP genes

Analysis of SmEXP tissue expression patterns

To explore the tissue expression patterns of SmEXPs, qRT-PCR was used to detect the expression levels of 18 SmEXPs members in four organs (roots, stems, leaves, and flowers) of S. miltiorrhiza (Fig. 5). The results showed that although SmEXP members were expressed in roots, stems, leaves, and flowers, their expression levels generally varied. The same genes had significant differences in their expressions in the four organs, and most SmEXPs had low expression levels in roots and high expression levels in flowers.

Fig. 5
figure 5

Expression levels of SmEXP in roots, stems, leaves, and flowers. The transcript levels were analyzed using the qRT-PCR. Data are shown as means ± SD, n = 3. Asterisks indicate significance differences from roots, as determined by the one-way ANOVA Dunnett’s test (*P < 0.05; **P < 0.01)

The SmEXPA5 expression level was highest in the leaves and flowers, which indicated that it played important roles in their growth. The expression levels of SmEXPA2 and SmEXLA1 were highest in the root, implying their important roles in this organ.

Analysis of SmEXP expression patterns under ABA treatments

The analysis of cis-acting elements based on SmEXP promoters revealed that SmEXP members had a large number of ABA responsive elements. S. miltiorrhiza was treated with 100 μM of ABA and sampled at 0 h, 10 min, 30 min, 1 h, 3 h, 6 h, 12 h, and 24 h. The expression levels of 21 SmEXPs genes were detected by qRT-PCR (Fig. 6).

Fig. 6
figure 6

Expression of SmEXP treated with 100 μM ABA for 0 h, 10 min, 30 min, 1 h, 3 h, 6 h, 12 h, and 24 h, as determined by RT-qPCR. Data are shown as means ± SD, n = 3. Asterisks indicate significance differences from 0 min, as determined by a one-way ANOVA Dunnett’s test (*P < 0.05; **P < 0.01)

The results revealed that all 21 SmEXPs responded to ABA treatments, and ABA could negatively regulate the expressions of genes such as SmEXPA1, SmEXPA3, SmEXPA5, SmEXPA7, SmEXPA8, SmEXPA10, SmEXPA11, SmEXPA14, and SmEXPB5, etc. After 6 h of ABA treatment, the expression level of SmEXPA5 decreased ~ 1341 times, while after 1 h of ABA treatment the expression level of SmEXLA1 increased ~ 32 times. These results indicated that SmEXPA5 and SmEXLA1 played a more important role in responding to ABA treatments in contrast to other SmEXPs.

Analysis of SmEXP expression patterns under NaCl treatments

After S. miltiorrhiza was treated with 100 mM NaCl for 0 h, 10 min, 1 h, 3 h, 6 h, 12 h, and 24 h, and the gene expression levels of 21 SmEXPs members were detected by qRT-PCR (Fig. 7). The results indicated that after the 100 mM NaCl treatment (except for SmEXPB5) the expression levels of the other SmEXP were not significantly altered, as were most of the upregulated genes. The expression levels of SmEXPA2, SmEXPA4, and SmEXPA21 were upregulated ~ 149-fold, 246-fold, and 170-fold respectively at 24 h, whereas the expression level of SmEXPA3 was downregulated by ~ 36-fold at 30 min. These results indicated that the SmEXPA2, SmEXPA4, SmEXPA21, and SmEXPA3 played more important roles in the responses of the SmEXP family under salt stress.

Fig. 7
figure 7

Expression levels of SmEXP treated with 100 mM NaCl for 0 h, 10 min, 1 h, 3 h, 6 h, 12 h, and 24 h, as determined by RT-qPCR. Data are shown as means ± SD, n = 3. Asterisks indicate significant differences from 0 h, as determined by one-way ANOVA Dunnett’s test (*P < 0.05; **P < 0.01)

Analysis of SmEXPs expression patterns under Cu2+ treatments

Following the treatment of S. miltiorrhiza with 200 μM Cu2+ for 0 h, 10 min, 30 min, 1 h, 3 h, 12 h and 24 h, the gene expression levels of 19 SmEXPs members were detected by qRT-PCR (Fig. 8). Following the exogenous application of Cu2+ (except for SmEXPA15) the expression levels of most SmEXPs exhibited significant changes. The expression level of SmEXPA3 decreased ~ 42 times after 1 h of treatment with Cu2+, while the expression level of SmEXLB1 increased 22 times after 24 h of treatment. These results indicated that SmEXPA3 and SmEXLB1 played important roles in the SmEXP family when S. miltiorrhiza was contaminated with heavy metals.

Fig. 8
figure 8

Expression of SmEXP treated with 200 μM Cu2+ for 0 h, 10 min, 30 min, 1 h, 3 h, 12 h, and 24 h. As determined by RT-qPCR. Data are shown as means ± SD, n = 3. An asterisk indicates significance differences from 0 h, as determined by one-way ANOVA Dunnett’s test (*P < 0.05; **P < 0.01)

Subcellular localization of SmEXLA1, SmEXLB1, and SmEXPA2

To reveal the potential functions of SmEXLA1, SmEXLB1m, and SmEXPA2 we developed GFP-SmEXLA1, GFP-SmEXLB1, and GFP-SmEXPA2 protein expression vectors, and used 35-GFP vector as positive control. These vectors were then transferred into the inner epidermis of onion and the transient expression was observed under laser confocal microscopy. The positive control 35-GFP was expressed in the cytoplasm and nucleus, which was consistent with the biological state. The results confirmed that SmEXLA1 and SmEXLB1 were localized in the nucleus and cell membrane, and SmEXPA2 was localized in the cell wall (Fig. 9).

Fig. 9
figure 9

Subcellular localization analysis of GFP-SmEXLA1, GFP-SmEXLB1, and GFP-SmEXPA2. With 35S-GFP as a positive control, the localization of GFP-SmEXLA1, GFP-SmEXLB1, and GFP-SmEXPA2 in onion was confirmed by fluorescence microscopy (Scale bar = 50 μm)

Discussion

EXP is a protein that has the capacity to relax the cell walls of plants, which is extensively found in various plants and widely regarded as a key regulator of cell wall extension during the growth of plants [8]. In this study, SmEXP was identified and analyzed at the whole genome level. A total of 29 SmEXPs were identified based on the criteria containing both Pollen_allerg_1 and DPPB_1 domains [9].

Through phylogenetic analysis, SmEXPs were found to be distributed among four subfamilies, 21 SmEXPA, 5 SmEXPB, 1 SmEXLA, and 2 SmEXLB, which is similar to the subfamily proportions of most species. However, the number of SmEXP was generally lower than that of other plants [e.g., A. thaliana (35), O. sativa (58), Moso Bamboo (82) [54], Solanum tuberosum [55] (34), and Zea mays (88)] [56]. This indicated that repetitive events or the double amplification of EXP in the genome occurred less frequently during the evolutionary process of S. miltiorrhiza, which resulted in fewer members of the gene family.

Orthologous genes typically retain similar biological functions; thus, the identification of orthologous genes is very important in terms of the functional data of model plants to predict the functionalities of non-model plant genes [57]. Through a comparative investigation, 13 pairs of orthologous genes were selected from SmEXP and AtEXP, with a sequence consistency of > 50%. SmEXPA11 and AtEXPA4 were orthologous genes with a sequence consistency of 79.69%, where AtEXPA4 was observed to affect root growth [58]. The overexpression of AtEXPA4 can enhance the elongation of primary roots, while knocking out AtEXPA4 decelerates the growth rate of the primary root, which may also affect the root growth of S. miltiorrhiza in SmEXPA11.

AtEXPA18 is associated with root hair-specific 13 (RHS13), Peroxidase 7 (PER7), Proline-rich protein 3 (PRP3), Xyloglucan 14 (XTH14 endotransglucosylase/hydrolase protein), Pollen Ole e 1 allergen, extensin family protein (T30B22. 16), Probable 46 (PME46 pectinesterase/pectinesterase inhibitor), Peroxidase (PER73) 73, Putative pectinesterase/pectinesterase inhibitor 24 (PME24), Pollen Ole e 1 allergen, extensin family protein (MOP10), Leucine-rich repeat Extensin-like protein 10 genes (including 1 (LRX1)), which interacted with each other. These genes were associated with cell wall construction and root hair morphology. This prediction suggested that SmEXPA19 (a orthologous gene of AtEXPA18) may also play an important role in cell wall development and root hair growth.

An exploration of the expression patterns of SmEXPs in different tissues is of great significance for functional analysis. SmEXPs are expressed in the roots, stems, leaves, and flowers of S. miltiorrhiza, indicating that SmEXPs are broadly involved in its growth and development. Numerous ABA responsive elements were predicted for the SmEXPs promoter, and experiments revelaed that all 21 SmEXPs responded to ABA treatments.

The large accumulation of Na+ in soil can lead to the decreased water absorption capacities of plant roots, as well as damage to leaf cells through transpiration, which affects plant growth and, in severe cases, can lead to plant death [59,60,61]. When plant roots are subjected to salt stress, the earliest signals of this condition are received by the cell wall sensing system [62]. In response to salt stress, most SmEXP are activated, among which SmEXPA3 and SmEXPA4 have the strongest responses, and play the most important roles.

Under heavy metal stress, plants engage inherent mechanisms to reduce the absorption and accumulation of heavy metals. Excessive copper ions can alter the compositions and distribution of sugars in the cell walls of Arabidopsis roots, and copper can increase the contents of cellulose, hemicellulose, and pectin, which thicken cell walls to prevent harmful heavy metal ions from entering the cell [63]. It was found that Cu2+ can induce most SmEXP expression; however, the kinetics that drive the exogenous Cu2+ induction of SmEXP differential expressions require further in depth investigation.

Subcellular localization prediction results revealed that all SmEXPs were located on the cell wall. It was verified through subsequent experiments that SmEXPA2 was indeed located on the cell wall and belonged to its complement of proteins. However, SmEXLA1 and SmEXLB1 were not present on the cell wall as predicted, but on the nuclear membrane. Due to the diversity of EXPANSIN gene functions, all candidate SmEXPs will require further analysis to verify their functions during the growth and development of S. miltiorrhiza.

Conclusion

For this study, 29 SmEXPs of S. miltiorrhiza were identified and analyzed at the whole genome level. They were analyzed from the aspects of systematic evolution, gene structure, protein characteristics, promoter analysis, tissue expression patterns, and cell localization. This revealed the basic characteristics of the SmEXP family and predicted their potential functions in biological processes, which lays a foundation for further research on specific functions. Further, we investigated the responses of SmEXP family members under various ABA and stress treatments, which is of great significance for the breeding and large scale production of S. miltiorrhiza.

Availability of data and materials

The data analyzed during this study may be obtained from the corresponding author upon reasonable request.

References

  1. Su CY, Ming QL, Rahman K, Han T, Qin LP. Salvia miltiorrhiza: traditional medicinal uses, chemistry, and pharmacology. Chin J Nat Med. 2015;13(3):163–82. https://doi.org/10.1016/S1875-5364(15)30002-9.

    Article  CAS  PubMed  Google Scholar 

  2. Xu H, Song J, Luo H, Zhang Y, Li Q, Zhu Y, Xu J, Li Y, Song C, Wang B, Sun W, Shen G, Zhang X, Qian J, Ji A, Xu Z, Luo X, He L, Li C, Sun C, Chen S. Analysis of the genome sequence of the medicinal plant Salvia miltiorrhiza. Mol Plant. 2016;9(6):949–52. https://doi.org/10.1016/j.molp.2016.03.010.

    Article  CAS  PubMed  Google Scholar 

  3. Shu T, Liu C, Pang M, He L, Yang B, Fan L, Zhang S, Wang X, Liu B, Rong L. Salvianolic acid B promotes neural differentiation of induced pluripotent stem cells via PI3K/AKT/GSK3β/β-catenin pathway. Neurosci Lett. 2018;671:154–60. https://doi.org/10.1016/j.neulet.2018.02.007.

    Article  CAS  PubMed  Google Scholar 

  4. Tsai MK, Lin YL, Huang YT. Effects of salvianolic acids on oxidative stress and hepatic fibrosis in rats. Toxicol Appl Pharmacol. 2010;242(2):155–64. https://doi.org/10.1016/j.taap.2009.10.002.

    Article  CAS  PubMed  Google Scholar 

  5. Polko JK, Kieber JJ. The regulation of cellulose biosynthesis in plants. Plant Cell. 2019;31(2):282–96. https://doi.org/10.1105/tpc.18.00760.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Cosgrove DJ. Catalysts of plant cell wall loosening. F1000Research. 2016;5:F1000. https://doi.org/10.12688/f1000research.7180.1.

    Article  PubMed  PubMed Central  Google Scholar 

  7. Cosgrove DJ. Microbial expansins. Annu Rev Microbiol. 2017;71:479–97. https://doi.org/10.1146/annurev-micro-090816-093315.

    Article  CAS  PubMed  Google Scholar 

  8. Li Y, Jones L, McQueen-Mason S. Expansins and cell growth. Curr Opin Plant Biol. 2003;6(6):603–10. https://doi.org/10.1016/j.pbi.2003.09.003.

    Article  CAS  PubMed  Google Scholar 

  9. Kende H, Bradford K, Brummell D, Cho HT, Cosgrove D, Fleming A, Gehring C, Lee Y, McQueen-Mason S, Rose J, Voesenek LA. Nomenclature for members of the expansin superfamily of genes and proteins. Plant Mol Biol. 2004;55(3):311–4. https://doi.org/10.1007/s11103-004-0158-6.

    Article  CAS  PubMed  Google Scholar 

  10. Sampedro J, Cosgrove DJ. The expansin superfamily. Genome Biol. 2005;6(12):242. https://doi.org/10.1186/gb-2005-6-12-242.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Ding A, Marowa P, Kong Y. Genome-wide identification of the expansin gene family in tobacco (Nicotiana tabacum). Mol Genet Genomics MGG. 2016;291(5):1891–907. https://doi.org/10.1007/s00438-016-1226-8.

    Article  CAS  PubMed  Google Scholar 

  12. Lu Y, Liu L, Wang X, Han Z, Ouyang B, Zhang J, Li H. Genome-wide identification and expression analysis of the expansin gene family in tomato. Mol Genet Genomics MGG. 2016;291(2):597–608. https://doi.org/10.1007/s00438-015-1133-4.

    Article  CAS  PubMed  Google Scholar 

  13. Yin Z, Zhou F, Chen Y, Wu H, Yin T. Genome-wide analysis of the expansin gene family in Populus and characterization of expression changes in response to phytohormone (abscisic acid) and abiotic (low-temperature) stresses. Int J Mol Sci. 2023;24(9):7759. https://doi.org/10.3390/ijms24097759.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Han Z, Liu Y, Deng X, Liu D, Liu Y, Hu Y, Yan Y. Genome-wide identification and expression analysis of expansin gene family in common wheat (Triticum aestivum L.). BMC Genomics. 2019;20(1):101. https://doi.org/10.1186/s12864-019-5455-1.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Zhu Y, Wu N, Song W, et al. Soybean (Glycine max) expansin gene superfamily origins: segmental and tandem duplication events followed by divergent selection among subfamilies. BMC Plant Biol. 2014;14:93. https://doi.org/10.1186/1471-2229-14-93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Zhang S, Xu R, Gao Z, Chen C, Jiang Z, Shu H. A genome-wide analysis of the expansin genes in Malus × Domestica. Mol Genet Genomics MGG. 2014;289(2):225–36. https://doi.org/10.1007/s00438-013-0796-y.

    Article  CAS  PubMed  Google Scholar 

  17. McQueen-Mason S, Durachko DM, Cosgrove DJ. Two endogenous proteins that induce cell wall extension in plants. Plant Cell. 1992;4(11):1425–33. https://doi.org/10.1105/tpc.4.11.1425.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Rayle DL, Cleland RE. The acid growth theory of auxin-induced cell elongation is alive and well. Plant Physiol. 1992;99(4):1271–4. https://doi.org/10.1104/pp.99.4.1271.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Cosgrove D. Loosening of plant cell walls by expansins. Nature. 2000;407:321–6. https://doi.org/10.1038/35030000.

    Article  ADS  CAS  PubMed  Google Scholar 

  20. Cosgrove DJ. Growth of the plant cell wall. Nat Rev Mol Cell Biol. 2005;6(11):850–61. https://doi.org/10.1038/nrm1746.

    Article  CAS  PubMed  Google Scholar 

  21. Yan A, Wu M, Yan L, Hu R, Ali I, Gan Y. AtEXP2 is involved in seed germination and abiotic stress response in Arabidopsis. PLoS ONE. 2014;9(1):e85208. https://doi.org/10.1371/journal.pone.0085208.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  22. Cho HT, Cosgrove DJ. Altered expression of expansin modulates leaf growth and pedicel abscission in Arabidopsis thaliana. Proc Natl Acad Sci USA. 2000;97(17):9783–8. https://doi.org/10.1073/pnas.160276997.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  23. Keller E, Cosgrove DJ. Expansins in growing tomato leaves. Plant J Cell Mol Biol. 1995;8(6):795–802. https://doi.org/10.1046/j.1365-313x.1995.8060795.x.

    Article  CAS  Google Scholar 

  24. Lee HW, Kim J. EXPANSINA17 up-regulated by LBD18/ASL20 promotes lateral root formation during the auxin response. Plant Cell Physiol. 2013;54(10):1600–11. https://doi.org/10.1093/pcp/pct105.

    Article  CAS  PubMed  Google Scholar 

  25. Brummell DA, Harpster MH, Civello PM, Palys JM, Bennett AB, Dunsmuir P. Modification of expansin protein abundance in tomato fruit alters softening and cell wall polymer metabolism during ripening. Plant Cell. 1999;11(11):2203–16. https://doi.org/10.1105/tpc.11.11.2203.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Kaashyap M, Ford R, Kudapa H, Jain M, Edwards D, Varshney R, Mantri N. Differential regulation of genes involved in root morphogenesis and cell wall modification is associated with salinity tolerance in chickpea. Sci Rep. 2018;8(1):4855. https://doi.org/10.1038/s41598-018-23116-9.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  27. Zhang JF, Xu YQ, Dong JM, Peng LN, Feng X, Wang X, Li F, Miao Y, Yao SK, Zhao QQ, Feng SS, Hu BZ, Li FL. Genome-wide identification of wheat (Triticum aestivum) expansins and expansin expression analysis in cold-tolerant and cold-sensitive wheat cultivars. PLoS ONE. 2018;13(3):e0195138. https://doi.org/10.1371/journal.pone.0195138.

    Article  PubMed  PubMed Central  Google Scholar 

  28. Kuluev BR, Knyazev AV, Nikonorov YM, et al. Role of the expansin genes NtEXPA1 and NtEXPA4 in the regulation of cell extension during tobacco leaf growth. Russ J Genet. 2014;50:489–97. https://doi.org/10.1134/S1022795414040061.

    Article  CAS  Google Scholar 

  29. Zou H, Wenwen Y, Zang G, et al. OsEXPB2, a β-expansin gene, is involved in rice root system architecture. Mol Breeding. 2015;35:41. https://doi.org/10.1007/s11032-015-0203-y.

    Article  CAS  Google Scholar 

  30. Wang Y, Ma N, Qiu S, et al. Regulation of the α-expansin gene OsEXPA8 expression affects root system architecture in transgenic rice plants. Mol Breeding. 2014;34:47–57. https://doi.org/10.1007/s11032-014-0016-4.

    Article  CAS  Google Scholar 

  31. ZhiMing Y, Bo K, XiaoWei H, ShaoLei L, YouHuang B, WoNa D, Ming C, Hyung-Taeg C, Ping W. Root hair-specific expansins modulate root hair elongation in rice. Plant J Cell Mol Biol. 2011;66(5):725–34. https://doi.org/10.1111/j.1365-313X.2011.04533.x.

    Article  CAS  Google Scholar 

  32. Zhou J, Xie J, Liao H, Wang X. Overexpression of β-expansin gene GmEXPB2 improves phosphorus efficiency in soybean. Physiol Plant. 2014;150(2):194–204. https://doi.org/10.1111/ppl.12077.

    Article  CAS  PubMed  Google Scholar 

  33. He X, Zeng J, Cao F, Ahmed IM, Zhang G, Vincze E, Wu F. HvEXPB7, a novel β-expansin gene revealed by the root hair transcriptome of Tibetan wild barley, improves root hair growth under drought stress. J Exp Bot. 2015;66(22):7405–19. https://doi.org/10.1093/jxb/erv436.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Silvia M, Adnane B, Fabien M, Christelle T, Bernard Q, Francesco C, Angelo P, Jacqueline V, Marc L, Filomena C, Abdelhafid B. Induced mutations in tomato SlExp1 alter cell wall metabolism and delay fruit softening. Plant Sci. 2016;242:195–202. https://doi.org/10.1016/j.plantsci.2015.07.001.

    Article  CAS  Google Scholar 

  35. Sun Q, Li Y, Gong D, Hu A, Zhong W, Zhao H, Ning Q, Tan Z, Liang K, Mu L, Jackson D, Zhang Z, Yang F, Qiu F. A NAC-EXPANSIN module enhances maize kernel size by controlling nucellus elimination. Nat Commun. 2022;13(1):5708. https://doi.org/10.1038/s41467-022-33513-4.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  36. Zhu L, Wang H, Zhu J, Wang X, Jiang B, Hou L, Xiao G. A conserved brassinosteroid-mediated BES1-CERP-EXPA3 signaling cascade controls plant cell elongation. Cell Rep. 2023;42(4):112301. https://doi.org/10.1016/j.celrep.2023.112301.

    Article  CAS  PubMed  Google Scholar 

  37. Marowa P, Ding A, Xu Z, Kong Y. Overexpression of NtEXPA11 modulates plant growth and development and enhances stress tolerance in tobacco. Plant Physiol Biochem PPB. 2020;151:477–85. https://doi.org/10.1016/j.plaphy.2020.03.033.

    Article  CAS  PubMed  Google Scholar 

  38. Chen LJ, Zou WS, Fei CY, Wu G, Li XY, Lin HH, Xi DH. α-Expansin EXPA4 positively regulates abiotic stress tolerance but negatively regulates pathogen resistance in Nicotiana tabacum. Plant Cell Physiol. 2018;59(11):2317–30. https://doi.org/10.1093/pcp/pcy155.

    Article  CAS  PubMed  Google Scholar 

  39. Chen Y, Han Y, Zhang M, Zhou S, Kong X, Wang W. Overexpression of the wheat expansin gene TaEXPA2 improved seed production and drought tolerance in transgenic tobacco plants. PLoS ONE. 2016;11(4):e0153494. https://doi.org/10.1371/journal.pone.0153494.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Yang J, Zhang G, An J, Li Q, Chen Y, Zhao X, Wu J, Wang Y, Hao Q, Wang W, Wang W. Expansin gene TaEXPA2 positively regulates drought tolerance in transgenic wheat (Triticum aestivum L.). Plant Sci Int J Exp Plant Biol. 2020;298:110596. https://doi.org/10.1016/j.plantsci.2020.110596.

    Article  CAS  Google Scholar 

  41. Sun W, Yao M, Wang Z, Chen Y, Zhan J, Yan J, Jiang S, Jian S, Chen H, Bu T, Tang Z, Li Q, Zhao H, Wu Q. Involvement of auxin-mediated CqEXPA50 contributes to salt tolerance in quinoa (Chenopodium quinoa) by interaction with auxin pathway genes. Int J Mol Sci. 2022;23(15):8480. https://doi.org/10.3390/ijms23158480.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Kwon YR, Lee HJ, Kim KH, Hong SW, Lee SJ, Lee H. Ectopic expression of Expansin3 or Expansinbeta1 causes enhanced hormone and salt stress sensitivity in Arabidopsis. Biotech Lett. 2008;30(7):1281–8. https://doi.org/10.1007/s10529-008-9678-5.

    Article  CAS  Google Scholar 

  43. Jadamba C, Kang K, Paek NC, Lee SI, Yoo SC. Overexpression of rice Expansin7 (Osexpa7) confers enhanced tolerance to salt stress in rice. Int J Mol Sci. 2020;21(2):454. https://doi.org/10.3390/ijms21020454.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Nishizono H, Ichikawa H, Suziki S, et al. The role of the root cell wall in the heavy metal tolerance of Athyrium yokoscense. Plant Soil. 1987;101:15–20. https://doi.org/10.1007/BF02371025.

    Article  CAS  Google Scholar 

  45. Haynes RJ. Ion exchange properties of roots and ionic interactions within the root apoplasm: their role in ion accumulation by plants. Bot Rev. 1980;46:75–99. https://doi.org/10.1007/BF02860867.

    Article  CAS  Google Scholar 

  46. Poole RL. The TAIR database. Methods Mol Biol. 2007;406:179–212. https://doi.org/10.1007/978-1-59745-535-0_8.

    Article  CAS  PubMed  Google Scholar 

  47. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol. 2018;35(6):1547–9. https://doi.org/10.1093/molbev/msy096.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Lescot M, Déhais P, Thijs G, Marchal K, Moreau Y, Van de Peer Y, Rouzé P, Rombauts S. PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res. 2002;30(1):325–7. https://doi.org/10.1093/nar/30.1.325.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Chen C, Chen H, Zhang Y, Thomas HR, Frank MH, He Y, Xia R. TBtools: an integrative toolkit developed for interactive analyses of big biological data. Mol Plant. 2020;13(8):1194–202. https://doi.org/10.1016/j.molp.2020.06.009.

    Article  CAS  PubMed  Google Scholar 

  50. Artimo P, Jonnalagedda M, Arnold K, Baratin D, Csardi G, de Castro E, Duvaud S, Flegel V, Fortier A, Gasteiger E, Grosdidier A, Hernandez C, Ioannidis V, Kuznetsov D, Liechti R, Moretti S, Mostaguir K, Redaschi N, Rossier G, Xenarios I, Stockinger H. ExPASy: SIB bioinformatics resource portal. Nucleic Acids Res. 2012;40(Web Server):W597–603. https://doi.org/10.1093/nar/gks400.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Yang Y, Hou S, Cui G, Chen S, Wei J, Huang L. Characterization of reference genes for quantitative real-time PCR analysis in various tissues of Salvia miltiorrhiza. Mol Biol Rep. 2010;37(1):507–13. https://doi.org/10.1007/s11033-009-9703-3.

    Article  CAS  PubMed  Google Scholar 

  52. Schmittgen TD, Livak KJ. Analyzing real-time PCR data by the comparative C(T) method. Nat Protoc. 2008;3(6):1101–8. https://doi.org/10.1038/nprot.2008.73.

    Article  CAS  PubMed  Google Scholar 

  53. Moreira D, et al. Orthologous gene. In: Amils R, et al., editors. Encyclopedia of astrobiology. Berlin, Heidelberg: Springer; 2014. https://doi.org/10.1007/978-3-642-27833-4_1731-3.

    Chapter  Google Scholar 

  54. Jin KM, Zhuo RY, Xu D, Wang YJ, Fan HJ, Huang BY, Qiao GR. Genome-wide identification of the expansin gene family and its potential association with drought stress in moso bamboo. Int J Mol Sci. 2020;21(24):9491. https://doi.org/10.3390/ijms21249491.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Chen Y, Zhang B, Li C, Lei C, Kong C, Yang Y, Gong M. A comprehensive expression analysis of the expansin gene family in potato (Solanum tuberosum) discloses stress-responsive expansin-like B genes for drought and heat tolerances. PLoS ONE. 2019;14(7):e0219837. https://doi.org/10.1371/journal.pone.0219837.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Zhang W, Yan H, Chen W, Liu J, Jiang C, Jiang H, Zhu S, Cheng B. Genome-wide identification and characterization of maize expansin genes expressed in endosperm. Mol Genet Genomics MGG. 2014;289(6):1061–74. https://doi.org/10.1007/s00438-014-0867-8.

    Article  CAS  PubMed  Google Scholar 

  57. Remm M, Storm CE, Sonnhammer EL. Automatic clustering of orthologs and in-paralogs from pairwise species comparisons. J Mol Biol. 2001;314(5):1041–52. https://doi.org/10.1006/jmbi.2000.5197.

    Article  CAS  PubMed  Google Scholar 

  58. Liu W, Xu L, Lin H, Cao J. Two expansin genes, AtEXPA4 and AtEXPB5, are redundantly required for pollen tube growth and AtEXPA4 is involved in primary root elongation in Arabidopsis thaliana. Genes. 2021;12(2):249. https://doi.org/10.3390/genes12020249.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Zhao KF, Song J, Fan H, Zhou S, Zhao M. Growth response to ionic and osmotic stress of NaCl in salt-tolerant and salt-sensitive maize. J Integr Plant Biol. 2010;52(5):468–75. https://doi.org/10.1111/j.1744-7909.2010.00947.x.

    Article  CAS  PubMed  Google Scholar 

  60. van Zelm E, Zhang Y, Testerink C. Salt tolerance mechanisms of plants. Annu Rev Plant Biol. 2020;71:403–33. https://doi.org/10.1146/annurev-arplant-050718-100005.

    Article  CAS  PubMed  Google Scholar 

  61. Munns R. Comparative physiology of salt and water stress. Plant Cell Environ. 2002;25(2):239–50. https://doi.org/10.1046/j.0016-8025.2001.00808.x.

    Article  CAS  PubMed  Google Scholar 

  62. Voxeur A, Höfte H. Cell wall integrity signaling in plants: “to grow or not to grow that’s the question.” Glycobiology. 2016;26(9):950–60. https://doi.org/10.1093/glycob/cww029.

    Article  CAS  PubMed  Google Scholar 

  63. Gao J, Sun L, Yang X, Liu JX. Transcriptomic analysis of cadmium stress response in the heavy metal hyperaccumulator Sedum alfredii Hance. PLoS ONE. 2013;8(6):e64643. https://doi.org/10.1371/journal.pone.0064643.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We greatly appreciated the contributions from all authors.

Funding

This research was supported by the National Natural Science Foundation of China (31870276, 31900254, and 32170378), the Fundamental Research Funds for the Central Universities (GK202205002, GK202205003, GK202205004), the Basic Research Project of Shaanxi Academy of Sciences (2023k-23), the Project of Shaanxi Science and Technology Department (2020JM-628), and the Basic Research Project of Shaanxi Academy of Sciences (2023k-23).

Author information

Authors and Affiliations

Authors

Contributions

ZW, DW and YL conceived and designed the experiments; YL and YL identified the SmEXP families. YL, BL and QP constructed the vectors and analyzed the data; YL wrote the manuscript. All authors reviewed the manuscript.

Corresponding authors

Correspondence to Donghao Wang or Zhezhi Wang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

All authors agreed to publish this article.

Competing interests

The authors have declared that they have 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

Additional file 1

: Table S1. Primer sequences of qRT-PCR.

Additional file 2

: Sequence of each SmEXP.

Additional file 3

: Fig.S1. Full length amino acid sequence alignments of orthologous genes.

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

Li, Y., Li, B., Pang, Q. et al. Identification and expression analysis of expansin gene family in Salvia miltiorrhiza. Chin Med 19, 22 (2024). https://doi.org/10.1186/s13020-023-00867-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13020-023-00867-w

Keywords