Skip to main content

Functional analysis of PagNAC045 transcription factor that improves salt and ABA tolerance in transgenic tobacco

Abstract

Background

Salt stress causes inhibition of plant growth and development, and always leads to an increasing threat to plant agriculture. Transcription factors regulate the expression of various genes for stress response and adaptation. It’s crucial to reveal the regulatory mechanisms of transcription factors in the response to salt stress.

Results

A salt-inducible NAC transcription factor gene PagNAC045 was isolated from Populus alba×P. glandulosa. The PagNAC045 had a high sequence similarity with NAC045 (Potri.007G099400.1) in P. trichocarpa, and they both contained the same conserved motifs 1 and 2, which constitute the highly conserved NAM domain at the N-terminus. Protein-protein interaction (PPI) prediction showed that PagNAC045 potentially interacts with many proteins involved in plant hormone signaling, DNA-binding and transcriptional regulation. The results of subcellular localization and transient expression in tobacco leaves confirmed the nuclear localization of PagNAC045. Yeast two-hybrid revealed that PagNAC045 protein exhibits transcriptional activation property and the activation domain located in its C-terminus. In addition, the 1063 bp promoter of PagNAC045 was able to drive GUS gene expression in the leaves and roots. In poplar leaves and roots, PagNAC045 expression increased significantly by salt and ABA treatments. Tobacco seedlings overexpressing PagNAC045 exhibited enhanced tolerance to NaCl and ABA compared to the wild-type (WT). Yeast one-hybrid assay demonstrated that a bHLH104-like transcription factor can bind to the promoter sequence of PagNAC045.

Conclusion

The PagNAC045 functions as positive regulator in plant responses to NaCl and ABA-mediated stresses.

Peer Review reports

Background

Plants often encounter different abiotic stresses, such as extreme temperatures, ultraviolet radiation, water deficits, oxidative stress, heavy metal toxicity, drought, and salinity [1]. These stresses pose a severe threat to plant growth and development, crop productivity and ecosystem balance worldwide [2]. Among them, salt stress is considered as the most significant environmental challenge, causing osmotic and toxic effect, oxidative damage, physiological water deficit, nutritional imbalance, and metabolic perturbation during plant growth and development [3,4,5,6,7]. In addition, soil salinization has been a primary global problem for a long time because of natural and human-induced actions [8]. However, the knowledge about the mechanisms of salt response in woody plants is scarce. P. alba×P. glandulosa (84 K) is a male P. alba×P. tremula var. glandulosa interspecific hybrid, and has been a representative model woody plant in plant stress biology research [9,10,11]. It has the advantages of easy rooting, short seedling period and strong tolerance to abiotic stresses [12], including drought [13], salt [14], cadmium [15], and light stress [16]. Additionally, 84 K poplar is widely used for timber, firewood, and pulp production [11].

Abscisic acid (ABA) is one of important phytohormones in plant stress response (e.g., to drought) [17]. Its induced biosynthesis appears to be one of the fastest responses to abiotic stresses [18]. Both ABA-dependent and ABA-independent pathways are involved in the regulation of plant stress response and osmotic stress-responsive gene expression [19, 20]. During the response process, ABA acts as a critical regulator in plant cells to cause significant physiological changes [21], e.g., activating ABA-induced gene expression and stomatal closure to reduce water loss [22]. Salt stress triggers osmotic stress signaling and ABA pathways [23]. However, the relationship between ABA and salt response in plants still needs to be uncovered.

Transcription factors (TFs) are key regulatory proteins that play fundamental roles in various biological processes and regulate different metabolic, developmental and stress response pathways by binding to specific cis-elements to control gene expression [24]. To date, 58 TF families were recorded including AP2, ERF, HD-ZIP, HSF, and NAC on PlantTFDB (http://planttfdb.gao-lab.org/index.php). Here we focus on NAC (NAM, ATAF1,ATAF2, CUC2) TFs [25], which contain a highly conserved N-terminal NAC domain for DNA-binding and dimerization capability, and a highly variable C-terminal regulatory domain for transcriptional activation or repression function [26, 27]. According to the PlantTFDB, 138, 228, 280, 155, 289, 189, and 145 NACs have been identified in Arabidopsis thaliana, Nicotiana benthamiana, N. tabacum, P. euphratica, P. trichocarpa, Zea mays, and Citrus sinensis, respectively. NACs have been reported to be involved in various plant processes, such as cell division [28], wood formation [29], and plant senescence [30]. Besides, NACs also play a crucial role in response to different environmental stresses [31,32,33]. For example, a NAC TF SlJUB1 gene induced by various abiotic stresses can control the expression of SlDREB1, SlDRED2, and SlDELLA to enhance tomato drought tolerance [34]. Overexpression of NAC13 can improve the salt stress tolerance in both transgenic poplar [35] and tobacco plants [36]. Overexpression of an OsNAC066 in rice improved drought and oxidative stress tolerance and increased ABA sensitivity [37]. In addition, 289 NAC TFs were retrieved and quantified based on RNA-seq datasets in our previous studies [38,39,40]. Based on log2 fold change (FC) ≥1 and false discovery rate (FDR) ≤0.05, 37 genes were significantly up-regulated under salt stress [40] and these genes could be vertically clustered into three groups. The PagNAC045 TF displayed increased expression under salt stress and was classified into the same group with PagNAC13 (Potri.001G404100.1) [35] and PagNAC036 (Potri.011G123300.1) [40].

In this study, we focused on the PagNAC045 TF gene from the 84 K poplar. The expression of PagNAC045 was significantly induced by salt stress [39, 40]. To better understand the characteristics of PagNAC045, we cloned the gene from the 84 K poplar leaves based on the sequence of a homolog NAC045 from P. trichocarpa. Next, we amplified the promoter sequence of the PagNAC045 to identify the upstream regulatory elements using the yeast one-hybrid system. Additionally, to reveal the function of PagNAC045 in response to salinity and ABA, we obtained PagNAC045-overexpressing transgenic tobacco lines by Agrobacterium-mediated transformation, and treated them with 200 mM NaCl and 50 μM ABA. Physiological parameters and the expression of downstream stress-related genes were analyzed. These results provide insight into the function of PagNAC045 in plant response to salt stress.

Results

Bioinformatic characterization of PagNAC045 transcription factor

The PagNAC045 gene of 84 K poplar is 915 bp in length to encode a protein of 305 amino acids (aa), which contains a highly conserved NAM domain at its N-terminus (Fig. 1A). The protein consists of 16.78% alpha helix, 11.18% extended strand, 3.62% beta turn and 68.42% random coil (Fig. 1B). According to NCBI protein blast result, the PagNAC045 amino acid sequence had 97.04% similarity to NAC045 (XP_002309945.1, Potri.007G099400.1) of P. trichocarpa, and they share six highly conserved motifs. Motifs 1 and 2 consist of the conserved NAM domain at N-terminus (Fig. 1C). Seven highly homologous proteins were found from the NCBI database. They include P. trichocarpa (XP_002309945.1), P. euphratica (XP_011022862.1), Salix brachista (KAB5548219.1), P. trichocarpa (XP_002306280.1), Hevea brasiliensis, Manihot esculenta and Durio zibethinus, sharing 97.04, 96.38, 90.46, 88.52, 80.13, 78.98 and 75.08% sequence similarity with the PagNAC045, respectively (Fig. 1D). Based on the results of STRING analysis, 40 proteins were predicted to form a network (Fig. 1E and Supplementary Table S1), with an average node degree of 10.8, a local clustering coefficient of 0.845, and a PPI enrichment p-value < 1.0e-16. Among these proteins, some are involved in the process of plant hormone signal transduction (red ball) including bZIP TF 6 family protein POPTR_0002s12710.1 and POPTR_0014s02810.1, ABA-insensitive 5-like protein 7 POPTR_0009s10400.1, ABA responsive elements-binding protein 2 ABF2-1, jasmonate (JA) zim domain-containing protein POPTR_0003s16350.1, JA-amido synthetase JAR1 GH3-12 and GH3-1, and JA zim domain-containing protein POPTR_0001s13240.1. Besides, some proteins play crucial roles in DNA binding (green ball), in nucleus (purple ball), and functioning as transcription regulators (yellow ball). Additionally, 10 proteins connected with black lines were predicted to be co-expressed with the PagNAC045. These include zinc finger CCCH domain-containing protein 30 POPTR_0001s27370.1 and POPTR_0009s06580.1, Tau class glutathione transferase GSTU45 POPTR_0016s10120.1, WRKY TF 42 family protein POPTR_0021s00280.1, syringolide-induced protein 1-3-1B POPTR_0008s01730.1, probable WRKY TF 48 POPTR_0010s15750.1 and its isoform POPTR_0008s10280.1, WRKY TF 6 family protein POPTR_0004s00890.1, TF salt-related MYB1 (srm1) POPTR_0010s24710.1 and POPTR_0001s22660.1.

Fig. 1
figure 1

Bioinformatic analysis of the poplar PagNAC045 gene and protein. A Structure of the PagNAC045 protein. B Conserved motifs composition of the PagNAC045 and homologous proteins in Populus trichocarpa. C PagNAC045 protein structure prediction. D Phylogenetic tree analysis of the PagNAC045 by MEGA7 with the Neighbor Joining method. E Functional protein association networks of the PagNAC045 by STRING

Subcellular localization of PagNAC045 protein in tobacco

According to the CELLO2GO prediction (Supplementary Fig. S1), the PagNAC045 was localized in the nucleus. To confirm the result in vivo, 35S::NAC045-GFP was constructed (Fig. 2A). The result of transient transformation of tobacco leaves showed that the fluorescence of NAC045-GFP was exclusively localized in the nucleus, while the positive control 35S::GFP was expressed throughout the whole cell (Fig. 2B). These results clearly showed that the PagNAC045 was a nuclear-localized protein.

Fig. 2
figure 2

Subcellular localization of the PagNAC045 protein. A Schematic map of the T-DNA inserted in the 35S::GFP binary vector. B The 35S::NAC045-GFP fusion construct and the positive control 35S::GFP plasmid were introduced into tobacco epidermal cells. GFP fluorescence was observed by confocal laser scanning microscopy. (a) and (d) were fluorescence images observed in dark field (green), (b) and (e) were light images observed in bright field, (c) and (f) were merged images of dark field and bright field. Scale bar = 50 μm

Transactivation activity analysis of PagNAC045 protein

To test the transactivation activity of poplar PagNAC045 TF, the full length and truncated sequences of PagNAC045 were cloned into pGBKT7 vector, named pGBKT7-NAC045 (1-305 aa), pGBKT7-NAC045a (1-136 aa) and pGBKT7-NAC045b (137-305 aa) (Fig. 3A). Transactivation assay showed that all the transformants including positive control pGBKT7-53/pGADT7-T and negative control pGBKT7 can grow well on SD/−Trp selection medium, while only positive control, pGBKT7-NAC045 and pGBKT7-NAC045b can grow and turn blue on SD/−Trp/−His/X-ɑ-Gal medium (Fig. 3B). The results showed that the PagNAC045 had transactivation activity and the amino acid residues at the C-terminal part are essential for the activity.

Fig. 3
figure 3

Transactivation analysis of the PagNAC045 protein. A Schematic map of pGBKT7-NAC045 (1-305 aa), pGBKT7-NAC045a (1-136 aa) and pGBKT7-NAC045b (137-305 aa) constructs. B Yeast colony assay showing transactivation activities of pGBKT7-NAC045 (1-305 aa) and pGBKT7-NAC045b (137-305 aa)

PagNAC045 gene promoter analysis and its upstream regulator

To explore the function of the PagNAC045 promoter, 1063 bp promoter sequence was ligated to the vector pBI121 (Fig. 4A) to drive the GUS gene expression, and the construct was used for tobacco transformation. GUS histochemical staining showed that the PagNAC045 promoter can drive GUS gene expression in the plant leaves and roots (Fig. 4B).

Fig. 4
figure 4

PagNAC045 promoter analyses. A Schematic map of PagNAC045 promoter that replaces the vector promoter CaMV35S; B GUS expression driven by the PagNAC045 promoter in tobacco seedlings. C Schematic map of bait vector and prey vector. D Yeast one-hybrid assay to determine that the PagNAC045 promoter was specially bound by a bHLH104-like transcription factor

To discover the upstream regulators of PagNAC045, the promoter bait strain with PagNAC045 promoter was constructed (Fig. 4C) for cDNA library screening. A total of 48 positive colonies were selected by PCR with T7 primers (Supplementary Fig. S2). According to the blast sequence by NCBI Blast server, a TF gene bHLH104-like (XM_035069641.1) was selected. Based on the homolog sequence from P. trichocarpa, specific primers of bHLH104-like was designed (Supplementary Table S2) and the TF gene was isolated from the 84 K poplar. Then the full length of bHLH104-like gene was inserted into pGADT7 as a prey vector for further identification. The results of yeast-one hybrid assay revealed that the bHLH104-like protein can bind to the PagNAC045 promoter sequence, i.e., it functions as an upstream regulator of the PagNAC045 (Fig. 4D).

Relative expression levels of PagNAC045 in poplar

Based on in silico prediction, PagNAC045 was highly expressed in mature leaves and roots, followed by internode and young leaves (Fig. 5A). To further analyze the relative expression levels of PagNAC045 in different poplar tissues after salinity and ABA treatments, we collected roots, stems, and leaves under 150 mM NaCl or 50 μM ABA at 0, 3, 6, 12, 24 and 48 h, respectively. The results of RT-qPCR showed that the PagNAC045 was significantly induced by salt and ABA in the roots and leaves of poplar, but not in stems when treated with salt. After treatment for 12 h, the expression level of PagNAC045 reached the highest point, especially in roots under salt stress (Fig. 5B) and in leaves after ABA treatment (Fig. 5C).

Fig. 5
figure 5

Tissue-specific expression pattern of PagNAC045 in poplar. A Expression pattern of PagNAC045 based on in silico prediction; B Relative expression levels of PagNAC045 in root, stem, and leaf tissues under 150 mM NaCl treatment for 0, 3, 6, 12, 24 and 48 h; C after 50 μM ABA treatment for 3, 6,12,24 and 48 h. The 0 h time point was used as control. Student’s t-test: t: *, P < 0.05; **, P < 0.01. Error bars indicate mean ± SD

Identification of transgenic tobacco lines

Using an Agrobacterium-mediated leaf disc method, the transgenic plants with PagNAC045-pBI121 recombinant construct were generated and selected on the selective 1/2 MS medium containing 100 mg/L kanamycin. According to PCR with the specific primers PagNAC045F1 and PagNAC045R1 (Supplementary Table S2) and semi-quantitative RT-PCR confirmation with PagNAC045F4 and PagNAC045R4, six transgenic lines had the same length of band as that in the positive control, but the band was absent in the WT plants (Fig. 6A, Supplementary Fig. S3). In this study, transgenic lines T1, T4, and T5 were selected for further analysis.

Fig. 6
figure 6

Identification and stress treatments of PagNAC045 over-expressing transgenic tobacco lines. A RT-PCR validation of over-expressing transgenic lines with specific primers PagNAC045F1 and PagNAC045R1 (Supplementary Table S2); M, 2000 DNA marker; WT, wild type; T1-T6, transgenic lines; P, positive plasmid. B Biochemical analyses of superoxide dismutase (SOD), peroxidase (POD), malondialdehyde (MDA) and proline contents in the leaves of transgenic lines and WT. C Images of nitroblue tertazolium (NBT) staining. D Images of 3,3′-diaminobenzidine (DAB) staining. E Four-week-old tobacco plants in soil before the treatments (upper panels), and plants treated with 200 mM NaCl or 50 μM ABA for one week (bottom panels). Student’s t-test: t: *, P < 0.05; **, P < 0.01. Error bars indicate mean ± SD

Salt and ABA tolerance of the transgenic tobacco plants

To investigate the functions of PagNAC045, three-week-old tobacco seedlings including WT and three transgenic lines T1, T4 and T5 were treated with 200 mM NaCl or 50 μM ABA for 12 h. Those treated with water were used as the control. The leaves were collected for SOD, POD, MDA, and proline measurements (Fig. 6B). The results showed that under control conditions, there were no significant changes of SOD, POD, MDA and proline contents between the WT and transgenic lines. However, after treatment with salt, the activities of SOD and POD, and proline content of the transgenic lines were 1.26 ± 0.06, 1.39 ± 0.04, and 1.35 ± 0.59 times of those of WT. In contrast, the MDA content of WT was 2.17 ± 0.31 times of the transgenic lines (Fig. 6B). Similarly, after ABA treatment, the activities of SOD and POD and proline content of transgenic lines were also increased to 1.48 ± 0.05, 1.38 ± 0.02 and 1.75 ± 0.76 times of those of WT. The MDA content of WT had a 1.74 ± 0.19 times increase compared to the transgenic lines (Fig. 6B).

Histochemical staining with DAB and NBT is often used to detect the accumulation of hydrogen peroxide (H2O2) and superoxide anion (O2), respectively [41]. Three-week-old tobacco seedlings were irrigated with water, 200 mM NaCl or 50 μM ABA for 12 h. The whole seedlings were immersed in DAB and NBT staining solutions at 37 °C overnight. As shown in Fig. 6C and D, under control conditions, there was no significant difference between WT and transgenic lines, while after treatments with salt and ABA, WT showed more intense DAB- and NBT-stainings than the transgenic plants.

Based on the morphological phenotype, before the salt and ABA treatments the WT and transgenic lines grow indistinguishably. After treatments for 1 week, the transgenic lines grew obviously better than the WT (Fig. 6E).

Expression of stress-related genes in the PagNAC045 transgenic plants

The relative expression levels of 16 stress-related genes were analyzed after the tobacco seedlings were exposed to 50 μM ABA or 200 mM NaCl for 12 h. As shown in Fig. 7, under normal conditions, the expression levels of NtPOD, NtSOD, NtPPO NtSOS and NtCAT showed no differences between the WT and transgenic lines. However, under salt or ABA treatments, these five genes were significantly increased in transcripts. The upregulated expression levels of these genes under salt stress were much higher than those under ABA treatment. In addition, NtDREB3, NtNCED1, NtP5CS, NtLEA5, NtERD10A/C and NtHKT521/586 were significantly upregulated in the transgenic lines compared to WT under normal conditions. Under salt stress, the relative expression levels of NtDREB3, NtNCED1, NtP5CS and NtHKT521/555/586 were increased 105.48, 8.27, 5.42, 3.30, 41.10, and 46.75 folds, respectively, compared to WT, respectively. Under ABA treatment, the relative expression levels of these genes were increased 10.77, 6.18, 2.47, 7.87, 13.36 and 21.08 folds, respectively. Clearly, the stress-related genes displayed different expression patterns when responding to different stresses, and these differentially expressed genes may play vital roles in ABA and NaCl responses.

Fig. 7
figure 7

RT-qPCR-based gene expression profiling of stress-responsive genes under NaCl or ABA treatments of WT and PagNAC045 transgenic seedlings. The 16 stress-related genes include NtPOD, NtSOD, NtPPO, NtSOS, NtDREB3, NtNCED1, NtP5CS, NtLEA5, NtERD10A/B/C/D, NtCAT and NtHKT521/555/586. Student’s t-test: t: *, P < 0.05; **, P < 0.01; ***, P < 0.001. Error bars indicate mean ± SD

Discussion

In poplar, 289 NAC genes were identified and can be clustered in to 15 distinct subgroups based on their deduced protein sequences [42]. Among these genes, 37 were significantly upregulated by salt stress according to the RNA-seq data [40]. Here we studied one of salt-induced NAC genes, PagNAC045 from the 84 K poplar. It was expressed in different tissues of the poplar seedlings, and was significantly upregulated by NaCl and ABA in poplar roots and leaves, but not in stems. According to Yao et al [42], PagNAC045 belongs to subgroup I of NAC TF family, and this gene has the highly conserved domain NAM at its N-terminus and transcriptional activation domain at its C-terminus. PPInetwork showed that PagNAC045 may interact with eight candidate proteins and take a part in plant hormone signal transduction (Fig. 1). Four of them are related to ABA, including ABA-insensitive 5-like protein 7 (POPTR_0009s10400.1), ABA responsive elements-binding protein 2 (POPTR_0004s14790.1, ABF2-1), ABA-insensitive 5-like protein 5 isoform × 2 (POPTR_0014s02810.1) and bZIP transcription factor 6 family protein (POPTR_0002s12710.1). Thus, PagNAC045 may be involved in ABA signaling pathway. The homologous gene of PagNAC045 in Arabidopsis is ATAF1 (AT1G01720.1). ATAF1 was reported to be induced by drought, high salinity, ABA, methyl jasmonate, mechanical and wounding [43]. The results from this study showed that the expression level of PagNAC045 in poplar was induced by salt and ABA, leading us to hypothesize the PagNAC045 may have similar functions as ATAF1 in Arabidopsis in the response of salt and ABA treatments. How PagNAC045 is activated by salt and ABA, and what its targets are in the signaling pathways are not known.

It was reported that TFs as molecular switches can drive specific temporal and spatial gene expression to regulate diverse plant processes like stress signaling by binding to cis-acting elements in the promoter region [44, 45]. For the identification of upstream interacting factors, yeast one-hybrid was used to screen prey protein interactions according to the bait DNA promoter sequence [46]. For example, a potential TF which annotated as CRM-domain containing factor CFM3 and a sulfite oxidase-like protein in Nicotiana tabacum were identified to bind to the promoter of nitrate reductase gene [47]. A VASCULAR-RELATED NAC DOMAIN7 can bind to HD-ZIPIII TFs REV and PHB promoters in Arabidopsis [48]. The bZIP TF TGA9 can upregulate the expression of autophagy-related ATG8B and ATG8E by binding the special sites TGA in their promoters [49]. In our study, among the 48 positive colonies selected, a basis helix-loop-helix (bHLH) 104 TF was found to act as a positive regulator in ABA signaling pathway [50] and confers response to salinity stress [51]. Our results showed that bHLH104 TF can bind to the upstream of PagNAC045 that would function as a positive regulator to activate the expression of PagNAC045, especially under stress treatment.

Most stress adaptive mechanisms in plants are accompanied by certain morphological and physiological changes [52], especially, the ROS accumulation in various forms like hydrogen peroxide (H2O2), superoxide anions, hydroxyl radical (OH) and singlet oxygen (1O2) [53]. SOD and POD are key antioxidant enzymes for ROS-scavenging and protecting different cellular structures from damage under stress conditions [54]. MDA is commonly used to determine the level of oxidative stress in plants [55]. NBT and DAB are used for measuring superoxide anion and H2O2, respectively [56]. Free proline is another useful index to monitor plant physiological status and evaluate the stress tolerance [57]. In this study, tobacco seedlings with PagNAC045 overexpression displayed higher activities of POD and SOD, and higher proline content than WT when challenged with NaCl and ABA. Together with the lower MDA and ROS levels in the transgenic plants than WT, these results indicate that PagNAC045 plays a positive role in salt and ABA response through activating cellular antioxidant system to scavenge excess ROS.

Salt stress can alter the expression of many stress-responsive genes [23]. It was reported that salt stress induced 932 genes and repressed 367 genes in Arabidopsis based on transcriptomics [57]. Similarity, ABA activated the expression of many stress-related genes in plants [58]. Here we selected 16 stress-related genes and quantified their relative expression levels under salt or ABA treatment. Among them, two ROS scavenging NtSOD and NtPOD expression levels were significantly increased under salt and ABA treatments. The results were consistent with the transcriptional regulation of the enzymatic activities (Fig. 6B). Polyphenol oxidase (PPO) is a ubiquitous enzyme for catalyzing the oxidation of phenols to highly reactive quinones. It has been confirmed to be induced by wounding, ABA, and methyl jasmonate (MeJA) in tobacco plants [59]. In the PagNAC045 overexpressing transgenic tobacco, NtPPO was significantly induced by ABA and salt stress. This is a strong indication that PagNAC045 could regulate the expression of NtPPO to decrease the stress damage. The salt overly sensitive (SOS) pathway is the key pathway for regulating Na+/K+ ion homeostasis to decrease ionic damage in response to salt stress [60]. In the transgenic lines, no difference in NtSOS expression was observed under normal conditions. However, after treatment with salt and ABA, NtSOS expression was significantly increased (Fig. 7). This result suggests that PagNAC045 may function to activate NtSOS under stress conditions. NtDREB3 is C-repeat-binding-factor/dehydration-responsive element (CBF/DRE) TF, NtERD10A/B are downstream CBF/DRE-binding regulon, and NtERD10C/D are early dehydration-responsive genes [61]. The levels of these transcripts were significantly lower in WT than those in the PagNAC045 transgenic plants, indicating these genes may be targets of the PagNAC045 TF. NtP5CS encodes an enzyme to catalyze the first two steps in proline biosynthesis [62]. The elevated contents of proline in the transgenic lines under salt and ABA treatments may be attributed to the upregulation of NtP5CS in the transgenic tobacco (Fig. 7). NtLEA5 gene encodes a group 5 late embryogenesis abundant (LEA) protein, which functions in stabilizing labile enzymes and protects the structure of macromolecule and membranes [63]. The high affinity K+ transporter (HKT) works in limiting Na+ accumulation and minimizing the osmotic imbalance [64]. It was demonstrated that HKT can improve the salt tolerance in Populus [65]. All these genes were significantly induced by salt and ABA in the transgenic tobacco lines. These results suggest that overexpression of PagNAC045 facilitates the tolerance to salt and ABA by regulating the expression of downstream stress-related genes. Further analyses are needed to elucidate the molecular actions of PagNAC045 in regulating downstream genes when plants experience abiotic stresses (Fig. 8).

Fig. 8
figure 8

Schematic presentation of PagNAC045 in plant salt and ABA tolerance. SOD, superoxide dismutase; POD, peroxidase; MDA, malondialdehyde; SOS, plasmalemma Na+/H+antiporter; PPO, polyphenol oxidase; CAT, ascorbate peroxidase; DREB, dehydration responsive element binding protein; NCED, 9-cis-epoxycarotenoid dioxygenase1; P5CS, 1-pyrroline-5-carhoxylate synthetase; LEA5, late-embryogenesis-abundant protein5; ERD, early responsive to dehydration; HKT, high-affinity potassium transporter

Conclusions

In this study, we isolated the PagNAC045 gene from 84 K poplar, which was induced by salt stress and ABA treatment. The characteristics of its protein sequence and structure were analyzed. The PagNAC045 was a nucleus-localized protein and had transactivation domain in its C-terminus. In addition, the promoter of PagNAC045 can drive GUS gene expression and a bHLH104 TF was upstream of PagNAC045 function. Furthermore, the PagNAC045 overexpressing tobacco plants were generated. Biochemical analyses of ROS antioxidant enzymes SOD and POD, and the contents of proline and MDA of the transgenic tobacco seedlings showed that overexpression of PagNAC045 significantly increased the ROS scavenging ability and thereby reduced cellular membrane damage. Moreover, the PagNAC045 differentially regulated the expression of stress-related genes in the transgenic tobacco plants when treated with NaCl and ABA. Our results provide important insight into the molecular mechanisms underlying the PagNAC045 functions in plant response to salt and ABA (Fig. 8).

Materials and methods

Plant materials

P. alba×P. glandulosa (84 K poplar) seedlings were cultured on 1/2 MS (Murashige and Skoog medium) plant medium (pH 5.7) supplemented with 0.1 mg/mL indole-3-butytric acid (IBA) and 0.01 mg/mL 1-naphthaleneacetic acid (NAA) and placed in a growth chamber with a photoperiod of 16/8-h light/dark cycle and an average temperature of 25 °C [35]. For tobacco transformation, Nicotiana benthamiana was used. The seeds were sterilized with 20% bleach for 15 min and washed with sterile water for 5 times, and then grown on the MS medium plates for 7 days under the growth chamber conditions. The seedlings with four true leaves were transferred to transparent glass bottle with 100 mL MS medium for a month. The fully expanded leaves were cut into 1 × 1 cm squares for leaf dish transformation. The materials poplar and tobacco used in this study were planted in the experimental field of Northeast Forestry University, Harbin, China.

Cloning and sequence analysis of PagNAC045

RNA was extracted from the 84 K poplar leaves and then reverse transcripted into cDNA following instructions included in a Prime Script RT reagent kit (Takara, China). According to the sequence of a highly homologous gene NAC045 (Potri.007G099400.1) of Populus trichocarpa from the Plant Transcription Factor Database v5.0 (http://planttfdb.gao-lab.org/), specific primes PagNAC045F1 and PagNAC045R1 (Supplementary Table S2) were designed to amplify the gene PagNAC045 from cDNA of 84 K poplar by polymerase chain reaction (PCR). Based on the translated amino acid sequence, the conserved domain and motifs of PagNAC045 were analyzed using NCBI conserved domain search (https://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi) and MEME Suite 5.3.0 (http://meme-suite.org/index.html). PagNAC045 protein structure was predicted by SWISS-MODEL (https://swissmodel.expasy.org/interactive/NHZXxy/models/). Also, we found some homologous proteins by NCBI blast and constructed a phylogenetic tree using MEGA7.0 with a Neighbor-Joining method. In addition, through STRING database, we analyzed the potential interactions of proteins, and constructed the protein-protein interactions (PPI) network.

Subcellular localization analysis of PagNAC045 protein

CELLO2GO (http://cello.life.nctu.edu.tw/cello2go/) for protein subcellular localization prediction with ontology annotation was used for the subcellular localization analysis. To experimentally test the result from CELLO2GO, the full length of PagNAC045 sequence without terminal codon was amplified with primers PagNAC045F2 and PagNAC045R2 including restriction enzyme sites of Xba I (TCTAGA) and Spe I (ACTAGT), respectively. After ligation to the vector pBI121 with green florescence protein (GFP), a construct 35S::NAC045-GFP was produced. The plasmid of 35S::NAC045-GFP was transformed into agrobacteria strain GV3101 for tobacco infiltration. The plasmid of 35S::GFP was used as positive control. Two days later, the GFP signal was observed by confocal laser scanning microscope (LSM 700, Zeiss, Germany).

Transcriptional activation assay

Some of NAC family proteins are plant-specific transcription factors [66]. To test whether PagNAC045 protein has transcriptional activation activity and the activation domain, the full length of PagNAC045 (0-305 aa), the highly conserved domain NAM (0-136 aa), the sequence at the carboxy-terminal domain (137-305 aa) were ligated to GAL4 DNA-BD vector pGBKT7, called pGBKT7-NAC045, pGBKT7-NAC045a and pGBKT7-NAC045b, respectively. All the primers for gene amplification were listed in Supplementary Table S2. These fusion plasmid vectors with different fragments were transformed into Y2H yeast strain and screened on selective medium without Trp and His plates. β-Galactosidase assays were then performed on filter lifts of the colonies to detect activation of the lacZ reporter gene. pGBKT7 and pGBKT7-53/pGBKT7-T were used as negative and positive controls, respectively.

PagNAC045 promoter analysis and yeast-one hybrid

Cis-acting elements in promoters function as binding sites of TFs and are important for transcription initiation [67]. To isolate and explore the function of PagNAC045 promoter (PagNAC045Pro), we extracted DNA from the 84 K poplar using a NuClean PlantGen DNA Kit (CWBIO, China). Specific primers for promoter cloning were designed as follows, PagNAC045F3 and PagNAC045R3 with restriction enzyme sites Cla I and Xba I, respectively (Supplementary Table S2). After PCR amplification, the 1063 bp PagNAC045 promoter sequence was ligated to a plant binary expression vector pBI121 with CaMV35S to drive β-glucuronidase (GUS) gene expression (Fig. 4A). The recombinant plasmid PagNAC045Pro and positive control CaMV35S were transformed into agrobacteria EHA105 strain and used for tobacco transient transformation. The agrobacteria were cultured in LB liquid medium which contains 50 mg/L rifampin and 50 mg/L kanamycin overnight. The agrobacteria were harvested by centrifugation for 10 min at the speed of 5000 rpm, resuspended in 1/2 liquid MS containing 150 μM acetosyringone until OD600 reached 0.2. Two-week-old tobacco seedlings were immersed in the agrobacteria suspension for 2 days at 25 °C. Then the seedlings were transferred to a X-Gluc staining solution (50 mM NaH2PO4, 50 mM Na2HPO4, 10 mM Na2EDTA, 0.1% TritonX-100, 10 mM K3[Fe (CN)6], 10 mM K4[Fe (CN)6] and 20 mM X-Gluc) and shaked at the speed of 120 rpm at 37 °C overnight. After washing the seedlings with a destaining solution (ethanol/acetic acid, V/V = 3:1) for three times, the blue coloring reflecting the GUS activity could be observed.

To further explore the interactions between PagNAC045 promoter and its TF in poplar, yeast one-hybrid assay was performed to map TF-DNA interactions [68] based on the manufacturer’s instructions of Matchmaker Gold Yeast one-hybrid Library Screening System (Clontech, USA) and Yao et al. [69]. The 1063 bp PagNAC045Pro sequence was inserted into pAbAi bait vector, and the recombinant bait vector was then transformed into Y1HGold yeast strain and selected on synthetic dextrose (SD)/−Ura solid culture medium. A poplar cDNA library was constructed as the method of Matchmaker™ Gold Yeast One-Hybrid Library Screening System (Clontech, USA). The double-stranded cDNA which was synthesized through long-distance (LD)-PCR and Sma I-linearized vector pGADT7-Rec was co-transformed into Y1HGold with bait PagNAC045Pro and selected on SD/−Leu/AbA solid medium at 30 °C for 3-5 days. Well-grown single colonies were cultured and identified by PCR with T7 primers (Supplementary Table S2). The PCR products were sequenced, and bioinformatics analyses were used for identification of positive clones. To further confirm the interaction between PagNAC045Pro and its potential TF, primers were designed and used for amplification of target TF gene from poplar based on the blasted sequence. The full length of target TF sequence was inserted into pGADT7 as the prey vector (Fig. 4C) and co-transformed into Y1HGold yeast strain with the bait vector PagNAC045Pro-pAbAi. The interaction between PagNAC045Pro and its TF was determined based on the yeast growth on the selective media.

Spatial and temporal quantitative expression analysis of PagNAC045

To analyze the expression levels of PagNAC045 in different tissues, we predicted the expression patterns in different tissues of PagNAC045 in silico using an exlmage tool of the PopGenIE V3 database (https://popgenie.org/gene?id=Potri.007G099400). To test the results, one-month-old poplar seedlings from culture medium were taken out and watered with 150 mM NaCl or 50 μM ABA for 0, 3, 6, 12, 24 and 48 h. Root, stem and leaf tissues were harvested with three biological replicates at each time point of control and the two treatments for RNA extraction and RT-qPCR. The Actin was used as the reference gene in RT-qPCR and the primers AF and AR were shown in Supplementary Table S2. The relative expression level in different samples was calculated using a 2−ΔΔCt method [70].

Generation of transgenic tobacco lines

One-month-old WT tobacco seedlings in culture medium were used for leaf discs transformation by agrobacteria-mediated transgenic approach [71, 72]. The full length of PagNAC045 sequence was cloned and ligated to plant expression vector pBI121 and transformed into agrobacteria strain EH105. The transformed agrobacteria were cultured in LB liquid medium containing 50 mg/L rifampin and 50 mg/L kanamycin overnight until OD600 = 0.6. The tobacco leaves were immersed in bacteria solution for 15 min and then screened on MS medium with 100 mg/L kanamycin. To identify the transgenic tobacco plants, the leaves from different lines were collected for RNA extraction and PCR detection. WT was used as a negative control. To further confirm the transgenic tobacco lines, semi-quantitative assay was conducted, the Ntactin was the reference primers for RT-PCR. The specific primers PagNAC045F4 and PagNAC045R4 (Supplementary Table S2) were designed, and cDNAs of each line were collected as template for semi-quantitative RT-PCR. The identified seedlings were transplanted into soil under greenhouse conditions for seeds and the third-generation seeds were used for further experiment.

Salt and ABA treatment of transgenic tobacco plants

One-month-old tobacco plants in soil were irrigated with 200 mM NaCl or 50 μM ABA treatments for 1 week. Three transgenic lines (T1, T4 and T5) and WT were used. The leaves from different lines under control, salt and ABA stress were collected for physiological measurements. The physiological indexes including superoxide dismutase (SOD), peroxidase (POD), malondialdehyde (MDA) and proline contents were measured according to published methods [35]. Histochemical analyses including nitroblue tertazolium (NBT) staining and 3,3′-Diaminobenzidine (DAB) staining were conducted according to the methods in Sekulska et al. [73].

Stress-related genes analysis in transgenic plants

The expression levels of stress-related genes including peroxidase NtPOD, superoxide dismutase NtSOD, polyphenol oxidase NtPPO, plasmalemma Na+/H+antiporter NtSOS, regulatory proteins NtDREB3, 9-cis-epoxycarotenoid dioxygenase1 NtNCED1, 1-pyrroline-5-carhoxylate synthetase NtP5CS, late-embryogenesis-abundant protein5 NtLEA5, early responsive to dehydration NtERD10A/B/C/D, ascorbate peroxidase NtCAT and Na+ antiporter genes NtHKT521/555/586 were profiled using RT-qPCR quantification. Ntactin and NtUbiquitin were used as reference genes. Leaves were harvested with three biological replicates for each sample. The related primer sequences were listed in Supplementary Table S3. The relative expression level in different samples was profiled and calculated by a 2−ΔΔCt method as previously described [69].

Availability of data and materials

The datasets generated and /or analyzed during the current study are available in NCBI SRA with the accession number SRP267437.

Abbreviations

84 K:

Populus alba×P. glandulosa

TFs:

Transcription factors

MS:

Murashige and Skoog medium

IBA:

Indole-3-butytric acid

NAA:

1-naphthaleneacetic acid

PCR:

Polymerase chain reaction

PPI:

Protein-protein interactions

GFP:

Green florescence protein

GUS :

β-glucuronidase

SOD:

Superoxide dismutase

POD:

Peroxidase

MDA:

Malondialdehyde

NBT:

Nitroblue tetrazolium

DAB:

3,3′-Diaminobenzidine

References

  1. Gong Z, Xiong L, Shi H, Yang S, Herrera-Estrella LR, Xu G, et al. Plant abiotic stress response and nutrient use efficiency. Sci China Life Sci. 2020;63(5):635–74.

    Article  PubMed  Google Scholar 

  2. He M, He CQ, Ding NZ. Abiotic stresses: general defenses of land plants and chances for engineering multistress tolerance. Front Plant Sci. 2018;9:1771.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Acosta-Motos JR, Ortuño MF, Bernal-Vicente A, Diaz-Vivancos P, Sanchez-Blanco MJ, Hernandez JA. Plant responses to salt stress: adaptive mechanisms. Agronomy. 2017;7(1):18.

    Article  CAS  Google Scholar 

  4. Zeng Y, Li L, Yang R, Yi X, Zhang B. Contribution and distribution of inorganic ions and organic compounds to the osmotic adjustment in Halostachys caspica response to salt stress. Sci Rep. 2015;5(1):1–2.

    Google Scholar 

  5. Liang W, Ma X, Wan P, Liu L. Plant salt-tolerance mechanism: a review. Biochem Biophys Res Commun. 2018;495(1):286–91.

    Article  CAS  PubMed  Google Scholar 

  6. Yang H, Dai L, Wei Y, Deng Z, Li D. Melatonin enhances salt stress tolerance in rubber tree (Hevea brasiliensis) seedlings. Ind Crop Prod. 2020;145:111990.

    Article  CAS  Google Scholar 

  7. Ren Z, Zheng Z, Chinnusamy V, Zhu J, Cui X, Iida K, et al. RAS1, a quantitative trait locus for salt tolerance and ABA sensitivity in Arabidopsis. Proc Natl Acad Sci. 2010;107(12):5669–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Shahid SA, Zaman M, Heng L. Soil salinity: historical perspectives and a world overview of the problem. In guideline for salinity assessment, mitigation and adaptation using nuclear and related techniques; 2018. p. 43–53.

    Google Scholar 

  9. Wang SY, Chen QJ, Wang WL, Wang XC, Lu MZ. Salt tolerance conferred by over-expression of OsNHX1 gene in poplar 84K. Chin Sci Bull. 2005;50:225–9.

    Article  CAS  Google Scholar 

  10. Shim D, Kim S, Choi YI, Song WY, Park J, Youk ES, et al. Transgenic poplar trees expressing yeast cadmium factor 1 exhibit the characteristics necessary for the phytoremediation of mine tailing soil. Chemosphere. 2013;90(4):1478–86.

    Article  CAS  PubMed  Google Scholar 

  11. Qiu D, Bai S, Ma J, Zhang L, Shao F, Zhang K, et al. The genome of Populus alba×Populus tremula var. glandulosa clone 84K. DNA Res. 2019;26(5):423–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Wang R, Yin P, Ruixia Y, Liu X, Luo L, Xu J. Genome-wide profiling of alternative splicing genes in hybrid poplar (P. alba× P. glandulosa cv. 84K) leaves. PLoS One. 2020;15(11):e0241914.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Song J, Wang Y, Pan Y, Pang J, Zhang X, Fan J, et al. The influence of nitrogen availability on anatomical and physiological responses of Populus alba× P. glandulosa to drought stress. BMC Plant Biol. 2019;19(1):1–2.

    Article  Google Scholar 

  14. Yoon SK, Bae EK, Lee H, Choi YI, Han M, Choi H, et al. Downregulation of stress-associated protein 1 (PagSAP1) increases salt stress tolerance in poplar (Populus alba× P. glandulosa). Trees. 2018;32(3):823–33.

    Article  CAS  Google Scholar 

  15. Hu YQ, Zhang CX, Wei AZ. Comparative Studies on the Resistance to Cadmium Stress for Populus canescens and P. alba× P. glandulosa cv. 84K in vitro. J Northwest Forestry Univ. 2012;27(6):43–7.

  16. Shou-Ren ZH, Rong-Fu GA. Effects of light stress on oxygen evolution and photochemical energy storage of hybrid poplar clones determined by photoacoustic technique. J Integr Plant Biol. 2000;42(8):818–23.

    Google Scholar 

  17. Fujita Y, Fujita M, Shinozaki K, Yamaguchi-Shinozaki K. ABA-mediated transcriptional regulation in response to osmotic stress in plants. J Plant Res. 2011;124(4):509–25.

    Article  CAS  PubMed  Google Scholar 

  18. Nakashima K, Takasaki H, Mizoi J, Shinozaki K, Yamaguchi-Shinozaki K. NAC transcription factors in plant abiotic stress responses. Biochim Biophys Acta. 2012;1819(2):97–103.

    Article  CAS  PubMed  Google Scholar 

  19. Yamaguchi-Shinozaki K, Shinozaki K. Organization of cis-acting regulatory elements in osmotic-and cold-stress-responsive promoters. Trends Plant Sci. 2005;10(2):88–94.

    Article  CAS  PubMed  Google Scholar 

  20. Yamaguchi-Shinozaki K, Shinozaki K. Transcriptional regulatory networks in cellular responses and tolerance to dehydration and cold stresses. Annu Rev Plant Biol. 2006;57:781–803.

    Article  CAS  PubMed  Google Scholar 

  21. Chen K, Li GJ, Bressan RA, Song CP, Zhu JK, Zhao Y. Abscisic acid dynamics, signaling, and functions in plants. J Integr Plant Biol. 2020;62(1):25–54.

    Article  CAS  PubMed  Google Scholar 

  22. Wu H, Li H, Zhang W, Tang H, Yang L. Transcriptional regulation and functional analysis of Nicotiana tabacum under salt and ABA stress. Biochem Biophys Res Commun. 2021;570:110–6.

    Article  CAS  PubMed  Google Scholar 

  23. Yang Y, Guo Y. Elucidating the molecular mechanisms mediating plant salt-stress responses. New Phytol. 2018;217(2):523–39.

    Article  CAS  PubMed  Google Scholar 

  24. Riaño-Pachón DM, Ruzicic S, Dreyer I, Mueller-Roeber B. PlnTFDB: an integrative plant transcription factor database. BMC Bioinformatics. 2007;8(1):1–0.

    Article  CAS  Google Scholar 

  25. Olsen AN, Ernst HA, Leggio LL, Skriver K. NAC transcription factors: structurally distinct, functionally diverse. Trends Plant Sci. 2005;10(2):79–87.

    Article  CAS  PubMed  Google Scholar 

  26. Trishla VS, Kirti PB. Structure-function relationship of Gossypium hirsutum NAC transcription factor, GhNAC4 with regard to ABA and abiotic stress responses. Plant Sci. 2021;302:110718.

    Article  CAS  PubMed  Google Scholar 

  27. Hu R, Qi G, Kong Y, Kong D, Gao Q, Zhou G. Comprehensive analysis of NAC domain transcription factor gene family in Populus trichocarpa. BMC Plant Biol. 2010;10(1):1–23.

    Article  CAS  Google Scholar 

  28. Kim HJ, Nam HG, Lim PO. Regulatory network of NAC transcription factors in leaf senescence. Curr Opin Plant Biol. 2016;33:48–56.

    Article  PubMed  CAS  Google Scholar 

  29. Wang H, Zhao Q, Chen F, Wang M, Dixon RA. NAC domain function and transcriptional control of a secondary cell wall master switch. Plant J. 2011;68(6):1104–14.

    Article  CAS  PubMed  Google Scholar 

  30. Podzimska-Sroka D, O'Shea C, Gregersen PL, Skriver K. NAC transcription factors in senescence: from molecular structure to function in crops. Plants. 2015;4(3):412–48.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Pinheiro GL, Marques CS, Costa MD, Reis PA, Alves MS, Carvalho CM, et al. Complete inventory of soybean NAC transcription factors: sequence conservation and expression analysis uncover their distinct roles in stress response. Gene. 2009;444(1-2):10–23.

    Article  CAS  PubMed  Google Scholar 

  32. Sun H, Hu M, Li J, Chen L, Li M, Zhang S, et al. Comprehensive analysis of NAC transcription factors uncovers their roles during fiber development and stress response in cotton. BMC Plant Biol. 2018;18(1):1–5.

    Article  CAS  Google Scholar 

  33. Hu H, Dai M, Yao J, Xiao B, Li X, Zhang Q, et al. Overexpressing a NAM, ATAF, and CUC (NAC) transcription factor enhances drought resistance and salt tolerance in rice. Proc Natl Acad Sci. 2006;103(35):12987–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Thirumalaikumar VP, Devkar V, Mehterov N, Ali S, Ozgur R, Turkan I, et al. NAC transcription factor JUNGBRUNNEN 1 enhances drought tolerance in tomato. Plant Biotechnol J. 2018;16(2):354–66.

    Article  CAS  PubMed  Google Scholar 

  35. Zhang X, Cheng Z, Zhao K, Yao W, Sun X, Jiang T, et al. Functional characterization of poplar NAC13 gene in salt tolerance. Plant Sci. 2019;281:1–8.

    Article  CAS  PubMed  Google Scholar 

  36. Cheng Z, Zhang X, Zhao K, Zhou B, Jiang T. Ectopic expression of a poplar gene NAC13 confers enhanced tolerance to salinity stress in transgenic Nicotiana tabacum. J Plant Res. 2020;133(5):727–37.

    Article  PubMed  CAS  Google Scholar 

  37. Yuan X, Wang H, Cai J, Bi Y, Li D, Song F. Rice NAC transcription factor ONAC066 functions as a positive regulator of drought and oxidative stress response. BMC Plant Biol. 2019;19(1):1–9.

    Article  Google Scholar 

  38. Yao W, Zhang D, Zhou B, Wang J, Li R, Jiang T. Over-expression of poplar NAC15 gene enhances wood formation in transgenic tobacco. BMC Plant Biol. 2020;20(1):1–1.

    Article  CAS  Google Scholar 

  39. Yao W, Zhao K, Cheng Z, Li x, Zhou B, Jiang T. Transcriptome analysis of poplar under salt stress and over-expression of transcription factor NAC57 gene confers salt tolerance in transgenic Arabidopsis. Front Plant Sci. 2018;9:1121.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Zhang X, Cheng Z, Yao W, Zhao K, Wang X, Jiang T. Functional characterization of PsnNAC036 under salinity and high temperature stresses. Int J Mol Sci. 2021;22(5):2656.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Kumar D, Yusuf MA, Singh P, Sardar M, Sarin NB. Histochemical detection of superoxide and H2O2 accumulation in Brassica juncea seedlings. Bio Protoc. 2014;4(8):e1108.

    Article  Google Scholar 

  42. Yao W, Li C, Lin S, Wang J, Zhou B, Jiang T. Transcriptome analysis of salt-responsive and wood-associated NACs in Populus simonii×Populus nigra. BMC Plant Biol. 2020;20(1):1–2.

    Article  CAS  Google Scholar 

  43. Wu Y, Deng Z, Lai J, Zhang Y, Yang C, Yin B, et al. Dual function of Arabidopsis ATAF1 in abiotic and biotic stress responses. Cell Res. 2009;19(11):1279–90.

    Article  CAS  PubMed  Google Scholar 

  44. Lehmeyer M, Kanofsky K, Hanko E, Ahrendt S, Wehrs M, Machens F, et al. Functional dissection of a strong and specific microbe-associated molecular pattern-responsive synthetic promoter. Plant Biotechnol J. 2015;1(14):61–71.

    Google Scholar 

  45. Erpen L, Devi HS, Grosser JW, Dutt M. Potential use of the DREB/ERF, MYB, NAC and WRKY transcription factors to improve abiotic and biotic stress in transgenic plants. Plant Cell Tissue Organ Cult. 2018;132:1–25.

    Article  CAS  Google Scholar 

  46. Cao Y, Bi M, Yang P, Song M, He G, Wang J, et al. Construction of yeast one-hybrid library and screening of transcription factors regulating LhMYBSPLATTER expression in Asiatic hybrid lilies (Lilium spp.). BMC Plant Biol. 2021;21:563.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Yang H, Zhou Y, Zhang Y, Wang J, Shi H. Identification of transcription factors of nitrate reductase gene promoters and NRE2 cis-element through yeast one-hybrid screening in Nicotiana tabacum. BMC Plant Biol. 2019;19:145.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Taylor-Teeples M, Lin L, de Lucas M, Turco G, Toal TW, Gaudinier A, et al. An Arabidopsis gene regulatory network for secondary cell wall synthesis. Nature. 2015;517:571–5.

    Article  CAS  PubMed  Google Scholar 

  49. Wang P, Nolan TM, Yin Y, Bassham DC. Identification of transcription factors that regulate ATG8 expression and autophagy in Arabidopsis. Autophagy. 2020;16(1):123–39.

    Article  CAS  PubMed  Google Scholar 

  50. Li H, Sun J, Xu Y, Jiang H, Wu X, Li C. The bHLH-type transcription factor AtAIB positively regulates ABA response in Arabidopsis. Plant Mol Biol. 2007;65:655–65.

    Article  CAS  PubMed  Google Scholar 

  51. Qiu JR, Huang Z, Xiang XY, Xu WX, Wang JT, Chen J, et al. MfbHLH38, a Myrothamnus flabellifolia bHLH transcription factor, confers tolerance to drought and salinity stresses in Arabidopsis. BMC Plant Biol. 2020;20:542.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Larcher W. Physiological plant ecology: ecophysiology and stress physiology of functional groups. Fourth Edition. Biologia Planetarium. 2003;47(4):500.

  53. Baxter A, Mittler R, Suzuki N. ROS as key players in plant stress signalling. J Exp Bot. 2014;65(5):1229–40.

    Article  CAS  PubMed  Google Scholar 

  54. Sahi C, Singh A, Blumwald E, Grover A. Beyond osmolytes and transporters: novel plant salt-stress tolerance-related genes from transcriptional profiling data. Physiol Plant. 2006;127(1):1–9.

    Article  CAS  Google Scholar 

  55. Sunkar R, Bartels D, Kirch HH. Overexpression of a stress-inducible aldehyde dehydrogenase gene from Arabidopsis thaliana in transgenic plants improves stress tolerance. Plant J. 2003;35(4):452–64.

    Article  CAS  PubMed  Google Scholar 

  56. Jambunathan N. Determination and detection of reactive oxygen species (ROS), lipid peroxidation, and electrolyte leakage in plants. Methods Mol Biol. 2010;639:292–8.

  57. Ábrahám E, Hourton-Cabassa C, Erdei L, Szabados L. Methods for determination of proline in plants. Methods Mol Biol. 2015;639:317–31.

  58. Cutler SR, Rodriguez PL, Finkelstein RR, Abrams SR. Abscisic acid: emergence of a core signaling network. Annu Rev Plant Biol. 2010;61:651–79.

    Article  CAS  PubMed  Google Scholar 

  59. Aziz E, Akhtar W, Ilyas M, Rehman S, Koiwa H, Shinwari ZK, et al. Response of tobacco polyphenol oxidase gene to wounding, abscisic acid (ABA) and methyl jasmonate (MEJ). Pak J Bot. 2017;49:499–502.

    CAS  Google Scholar 

  60. Sun S, Wen D, Yang W, Meng Q, Shi Q, Gong B. Overexpression of caffeic acid O-methyltransferase 1 (COMT1) increases melatonin level and salt stress tolerance in tomato plant. J Plant Growth Regul. 2020;39(3):1221–35.

    Article  CAS  Google Scholar 

  61. Guo X, Zhang L, Dong G, Xu Z, Li G, Liu N, et al. A novel cold-regulated protein isolated from Saussurea involucrata confers cold and drought tolerance in transgenic tobacco (Nicotiana tabacum). Plant Sci. 2019;289:110246.

    Article  CAS  PubMed  Google Scholar 

  62. Yue Y, Zhang M, Zhang J, Tian X, Duan L, Li Z. Overexpression of the AtLOS5 gene increased abscisic acid level and drought tolerance in transgenic cotton. J Exp Bot. 2012;63(10):3741–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Hundertmark M, Hincha DK. LEA (late embryogenesis abundant) proteins and their encoding genes in Arabidopsis thaliana. BMC Genomics. 2008;9(1):1–22.

    Article  CAS  Google Scholar 

  64. Qi Q, Yanyan D, Yuanlin L, Kunzhi L, Huini X, Xudong S. Overexpression of SlMDHAR in transgenic tobacco increased salt stress tolerance involving S-nitrosylation regulation. Plant Sci. 2020;299:110609.

    Article  CAS  PubMed  Google Scholar 

  65. Xu M, Chen C, Cai H, Wu L. Overexpression of PeHKT1;1 improves salt tolerance in populus. Genes. 2018;9(10):475.

    Article  PubMed Central  CAS  Google Scholar 

  66. Hao YJ, Song QX, Chen HW, Zou HF, Wei W, Kang XS, et al. Plant NAC-type transcription factor proteins contain a NARD domain for repression of transcriptional activation. Planta. 2010;232(5):1033–43.

    Article  CAS  PubMed  Google Scholar 

  67. Blazeck J, Alper HS. Promoter engineering: recent advances in controlling transcription at the most fundamental level. Biotechnol J. 2013;8(1):46–58.

    Article  CAS  PubMed  Google Scholar 

  68. Gaudinier A, Tang M, Bågman AM, Brady SM. Identification of protein-DNA interactions using enhanced yeast one-hybrid assays and a semiautomated approach. Methods Mol Biol. 2017;1610:187–215.

  69. Yao W, Wang S, Zhou B, Wang J, Jiang T. Characterization of ERF76 promoter cloned from Populus simonii×P. nigra. Acta Physiol Plant. 2017;39(11):1–6.

    Article  CAS  Google Scholar 

  70. Dang ZH, Zheng LL, Wang J, Gao Z, Wu SB, Qi Z, et al. Transcriptomic profiling of the salt-stress response in the wild recretohalophyte Reaumuria trigyna. BMC Genomics. 2013;14(1):1–8.

    Article  CAS  Google Scholar 

  71. Meer IMVD. Agrobacterium-mediated transformation of petunia leaf discs. Methods Mol Biol. 2006;318:265–72.

  72. Zhang HY, Liu XZ, Wei L, Zhou LY, Yang YM. Transgenic tobacco plants containing Bt and GNA genes. Biol Plant. 2007;51(4):746–8.

    Article  CAS  Google Scholar 

  73. Sekulska-Nalewajko J, Gocławski J, Chojak-Koźniewska J, Kuźniak E. Automated image analysis for quantification of reactive oxygen species in plant leaves. Methods. 2016;109:114–22.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors thank Daniel Chen from the Morsani College of Medicine at the University of South Florida for critical reading and editing of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (No.2021YFD2200203), the Innovation Project of State Key Laboratory of Tree Genetics and Breeding (Northeast Forestry University) (2015A02), and the University of Florida and China Scholarship Council (201906600017).

Author information

Authors and Affiliations

Authors

Contributions

TJ and WL designed and supervised the study with SC. XZ conducted experiments and wrote the manuscript. ZC and GF performed data analysis. WY and SC revised the manuscript. All authors read and approved the manuscript.

Corresponding authors

Correspondence to Sixue Chen or Tingbo Jiang.

Ethics declarations

Ethics approval and consent to participate

All plant materials used in this study were obtained from Northeast Forestry University, Harbin, China. All the plants complied with national guidelines and legislation, and did not involve any endangered or protected species.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no conflict of interest (include financial and non-financial 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

Zhang, X., Cheng, Z., Fan, G. et al. Functional analysis of PagNAC045 transcription factor that improves salt and ABA tolerance in transgenic tobacco. BMC Plant Biol 22, 261 (2022). https://doi.org/10.1186/s12870-022-03623-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-022-03623-8

Keywords