Next Article in Journal
Response of Maize (Zea mays L.) to Drought under Salinity and Boron Stress in the Atacama Desert
Previous Article in Journal
Next-Generation Sequencing (NGS) Identified Species-Specific SSR and SNP Markers, Allow the Unequivocal Identification of Strawberry Tree (Arbutus unedo L.) Germplasm Accessions and Contribute to Assess Their Genetic Relationships
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identification of the NAC Transcription Factor Family during Early Seed Development in Akebia trifoliata (Thunb.) Koidz

1
College of Life Sciences, Guizhou University, Guiyang 550025, China
2
Guizhou Key Laboratory of Propagation and Cultivation on Medicinal Plants, Guizhou University, Guiyang 550025, China
*
Author to whom correspondence should be addressed.
Plants 2023, 12(7), 1518; https://doi.org/10.3390/plants12071518
Submission received: 27 February 2023 / Revised: 25 March 2023 / Accepted: 27 March 2023 / Published: 31 March 2023

Abstract

:
This study aimed to gain an understanding of the possible function of NACs by examining their physicochemical properties, structure, chromosomal location, and expression. Being a family of plant-specific transcription factors, NAC (petunia no apical meristem and Arabidopsis thaliana ATAF1, ATAF2, and CUC2) is involved in plant growth and development. None of the NAC genes has been reported in Akebia trifoliata (Thunb.) Koidz (A. trifoliata). In this study, we identified 101 NAC proteins (AktNACs) in the A. trifoliata genome by bioinformatic analysis. One hundred one AktNACs were classified into the following twelve categories based on the phylogenetic analysis of NAC protein: NAC-a, NAC-b, NAC-c, NAC-d, NAC-e, NAC-f, NAC-g, NAC-h, NAC-i, NAC-j, NAC-k, and NAC-l. The accuracy of the clustering results was demonstrated based on the gene structure and conserved motif analysis of AktNACs. In addition, we identified 44 pairs of duplication genes, confirming the importance of purifying selection in the evolution of AktNACs. The morphology and microstructure of early A. trifoliata seed development showed that it mainly underwent rapid cell division, seed enlargement, embryo formation and endosperm development. We constructed AktNACs co-expression network and metabolite correlation network based on transcriptomic and metabolomic data of A. trifoliata seeds. The results of the co-expression network showed that 25 AtNAC genes were co-expressed with 233 transcription factors. Metabolite correlation analysis showed that 23 AktNACs were highly correlated with 28 upregulated metabolites. Additionally, 25 AktNACs and 235 transcription factors formed co-expression networks with 141 metabolites, based on correlation analysis involving AktNACs, transcription factors, and metabolites. Notably, AktNAC095 participates in the synthesis of 35 distinct metabolites. Eight of these metabolites, strongly correlated with AktNAC095, were upregulated during early seed development. These studies may provide insight into the evolution, possible function, and expression of AktNACs genes.

1. Introduction

The NAC petunia no apical meristem (NAM) and Arabidopsis thaliana ATAF1, ATAF2, and CUC2 transcription factors are one of the largest families found in plants [1,2]. They exhibit a highly conserved NAC subdomain at the N-terminal region consisting of approximately 150 amino acid residues, which contain five conserved regions (A–E) [3,4]. NAC proteins are homo- and hetero-dimerized, and the ability to dimerize is localized to the NAC domain [4,5]. They can form functional NAC protein dimers through salt bridging, with rich positive charges on one side of the surface, which may be related to DNA binding [6,7]. The C-terminal region is a highly differentiated and transcriptional regulatory regulator, characterized by the frequent occurrence of simple amino acid repeats and serine and threonine, proline and glutamine or acidic residues [8,9,10,11].
NAC transcription factors have a wide range of biological functions and play a vital role in plant growth and development, metabolic regulation, hormone response, stress resistance and crop quality improvement [12,13,14]. The first NAC transcription factor, NAM, was found to be associated with embryo growth and development in Petunia [1]. Petunia embryos carrying the no apical meristem (NAM) mutation failed to develop apical meristems, resulting in seedlings lacking roots and leaves. Both GmNAC11 and GmNAC20 are induced by biotic stress and phytohormones in soybean. Overexpression of the GmNAC20 not only enhanced the tolerance of plants to salt and low temperature but also promoted the formation of lateral roots [15]. SlNAM1, a salt-sensitive NAC gene in tomato (Solanum lycopersicum), is highly expressed in flowers and ripe fruit. Furthermore, SlNAC1 and SlNAM1 in tomato cultivars can be induced by salt stress, suggesting a possible involvement in tomato resistance responses to adversity stress [16]. Notably, NAC genes are capable of regulating aspects of plant embryo development, seed size and seed development. The BcNAC2 in turnips (Brassica campestris) was most abundantly expressed in young horn fruits at 8 days after pollination [17]. In situ hybridization in tissues showed that BcNAC2 was highly expressed in the embryo sac, indicating that the NAC transcription factor is associated with seed or embryo development. NARS1 and NARS2 (also known as NAC2 and NAM) in Arabidopsis thaliana control embryogenesis by regulating the growth and development and degeneration of the integument at the torpedo-shaped embryo stage [18]. NARS1 and NARS2 are not expressed at the torpedo-shaped embryo stage, but they are expressed in the outer integument. The double knockout mutant nars1/nars2 exhibits an abnormal seed shape and delayed degeneration of the integument. Pollination of wild-type plants with the double mutant nars1/nars2 produced normal seeds and, conversely, abnormal seeds. In rice, ONAC020, ONAC026 and ONAC023 showed a particularly high horizontal expression during seed development [19]. Knockout of watermelon CLNAC68 using the CRISPR-CAS9 system leads to delayed seed development and germination inhibition in a watermelon mutant [14].
After double fertilization is completed in plants, the ovary begins to expand, accompanied by seed development. Seed development is a highly complex process involving changes in seed morphology, size, and metabolites, and it is regulated by numerous key genes. Metabolites during seed development are divided into primary and secondary metabolites. Primary metabolites include sugars, lipids, etc., which are important indicators of seed quality and provide essential nutrients and energy for seed development. Secondary metabolism is a normal physiological activity that has evolved over time and interacts with both biotic and abiotic environments, playing a crucial role in plant growth and development. NAC transcription factors are involved not only in regulating plant life processes such as seed embryo development, leaf senescence, flower formation, and root development but also in regulating primary and secondary metabolic processes. For instance, plants overexpressing OsNAC23 in rice exhibited higher photosynthetic rates, sugar translocation, and larger organ sizes, resulting in sustained increases in rice yield [20]. StNAC103 gene silencing in potato epidermis was associated with increased wax and wax loading, particularly in alkanes, ω-hydroxy acids, diacids, ferulic acid, and primary alcohols [21]. Transcript levels of several genes associated with flavonoid biosynthesis and anthocyanin levels were significantly higher in Arabidopsis plants overexpressing ANAC078 (Ox-ANAC078) and significantly lower in knockout ANAC078 plants (KO-ANAC078) [22]. Additionally, NAC transcription factors can regulate key genes in lignin synthesis, affecting lignin synthesis and cell wall structure. The Eucalyptus NAC transcription factor EgNAC141 positively regulates lignin biosynthesis and increases lignin deposition [23].
At present, more data on NAC transcription factors of plants have been collected from studies based on plant genomics [24,25] and bioinformatics [26]. Moreover, NAC transcription factors have promising applications in plant molecular breeding. Akebia trifoliata is a woody vine plant belonging to genus Akebia (Lardizabalaceae), geographically distributed in eastern Asia (China, Japan, and Korea) [27], which is used as a traditional Chinese herb consisting of many bioactive compounds such as triterpenoid saponins and flavonoid [28]. The dried fruit alias “Akebiae Fructus”, comprises seeds, peel and pulp, which have anti-inflammatory, antibacterial, antioxidant and anticancer properties [29,30,31]. Interestingly, A. trifoliata can also be eaten as a fruit, as it has well-developed placental tissue and a sweet taste. It is rich in nutrients, including soluble sugars, vitamin C and many amino acids. However, its seed content holds 70% of the entire edible part of the fruit, affecting its delicacy. Furthermore, the severely impacted natural environment has led to a decline in wild resources, and there is an urgent need for artificial cultivation for the exploitation and conservation of resources.
The research team in this study completed the sequencing of the whole genome of A. trifoliata and obtained the first information on the NAC gene family. In this study, we aim to gain an understanding of the possible function of NACs by analysing physicochemical properties, structure, chromosomal location and expression. The results will provide a theoretical basis for genetic breeding and improvement work on A. trifoliata fruits.

2. Results

2.1. Identification and Analysis of NAC Genes in A. trifoliata

Based on the study of protein homology, 101 NAC genes were identified: AktNAC001–AktNAC101 (Table S1). The AktNAC gene family encodes proteins with the length ranging from 96aa (AktNAC021) to 691aa (AktNAC024), with predicted molecular weights of 11,463.13 (AktNAC021) to 77,893.7 Da (AktNAC076). Most of the NAC proteins (73.2%) had an isoelectric point (pI) < 7 and were acidic in nature, suggesting that AktNAC proteins were rich in acidic amino acids. The hydrophobicity index was <0, indicating that all NAC proteins were hydrophilic. The instability index ranged from 20.76 (AktNAC013) to 64.37 (AktNAC063), while the aliphatic index ranged from 51.79 (AktNAC049) to 83.36 (AktNAC014). The subcellular localization of proteins suggested that AktNACs were distributed in the cytoplasm, nucleus, plasma membrane, and chloroplast. Moreover, the number of potential phosphorylation sites ranged from 17 (AktNAC016) to 93 (AktNAC036).
We aligned 428 proteins to construct a phylogenetic tree to further examine the evolutionary relationship between NAC proteins of A. trifoliata, grape, rice, and Arabidopsis thaliana (Figure 1). The NAC proteins were classified into eight subfamilies (NAC-a, NAC-b, NAC-c, NAC-d, NAC-e, NAC-f, NAC-g, and NAC-h, NAC-i, NAC-j, NAC-k) via the phylogenetic analysis. The NAC-j and NAC-e groups contain the most AktNACs, 17 and 15, respectively, followed by the NAC-g (12), NAC-a (10), NAC-b (10) and NAC-d (10) groups, with NAC-l constituting the least number of AktNAC (1). The close phylogenetic relationship between A. trifoliata and Arabidopsis thaliana NAC proteins suggested that they might have similar structures or biological functions. For example, ANAC072 showed a significantly increased tolerance to drought [32] and the novel abscisic acid (ABA)–dependent stress-signaling pathway [33], indicating that AktNAC053 and AktNAC084 might be involved in drought tolerance and in ABA response.
We performed a multiple sequence alignment and phylogenetic analysis of the protein sequences from the NAC gene family in A. trifoliata, constructing an evolutionary tree (Figure S1). The tree classified the AktNAC gene family into 11 subclasses, including OsNAC8, TIP, NAC2, ANACO11, NAM, OsNAC7, OsNAC022, ATAF/NAP, ONAC079, ANAC001, and ONAC003/ANAC063. Furthermore, we conducted a conserved motif analysis of the AktNAC protein sequences. The analysis revealed that the majority of AktNAC proteins contain subdomains A–E at the N-terminus. Motifs 3, 4, and 1 were designated as NAM sub-domains A, B, and C, respectively. Additionally, motif 6 and motif 2 were identified as subdomains D and E, respectively. AktNAC proteins within the same subclasses exhibit broadly similar structures. For instance, OsNAC8 and TIP, belonging to the NAC-e group, have the highest number of motifs. NAM and OsNAC7 belong to the NAC-a group and are similar in length, both containing six identical motifs. NAC2 and ANAC011 subclasses, belonging to the NAC-d group, possess a similar arrangement and position of motifs. ANAC063 subclass proteins contain motifs 1–6, and their structural diversity may be associated with their distinct functions.
We analyzed the gene structure of members within the NAC gene family, presenting information on exons and introns. This analysis aids in understanding the structural conservation and variability of the gene family during evolution. Introns contribute to proteome complexity by producing different combinations of exons, which in turn translate into various proteins through alternative splicing [34,35]. The arrangement of exons and introns offers valuable insights into the evolution of gene families [36]. With the exception of the NAC-a, NAC-e, and NAC-d subclasses, other AktNAC subclasses possess three exons, a pattern also observed in Arabidopsis thaliana [4]. Additionally, 33 genes contained 5′-UTRs, and 31 genes contained 3′-UTRs. We further analyzed the conserved structural domains of selected AktNACs (Figure 2), demonstrating that substructural domains A, C, and D of AktNACs are highly conserved. This high conservation suggests that domains A, C, and D play crucial roles in the function of NAC family genes, a situation similar to that in Arabidopsis thaliana [37]. Previous research has reported that subdomain C may be involved in DNA binding [38]. The sequence diversity in subdomains B and E could imply distinct roles for these sequences within the NAC domain.

2.2. Chromosomal Localization and Duplication of NAC Genes in A. trifoliata

We identified the chromosomal location of 101 AktNACs (Figure 3), scattered unevenly across 16 chromosomes. Chr01 constituted the most AktNAC genes (twenty), while chromosome 12 comprised only one.
Gene duplication events have a significant effect on the gene family abundance and genome complexity in Eukaryotes [39]. We analyzed the duplication events of this family to understand the possible relationships between the potential gene duplications of AktNACs (Figure S2). Gene families form as species diverge and their evolution can be assessed using PAL2NAL to calculate synonymous (Ks) and nonsynonymous (Ka) values, thus revealing the history of gene evolution [40]. Therefore, Ks and Ka values determined whether the selection acted on protein-coding genes [41]. We estimated Ka/Ks values for 44 pairs of duplicated genes (Table S2). We identified 13 tandem replication gene pairs and 31 segmental ones. The Ka/Ks ratio was <1 suggested that purifying selection acted on these duplicated gene pairs [42].
To further investigate the phylogenetic mechanism of the NAC gene family in A. trifoliata, a collinearity analysis was performed between A. trifoliata (A), Arabidopsis thaliana and the published genome of the A. trifoliata (B) (Figure S3). The number of co-linear gene pairs between A. trifoliata (A) and Arabidopsis thaliana was 2. The number of co-linear gene pairs between A. trifoliata (A) and A. trifoliata (B) was 76, distributed on 15 chromosomes, with no co-linear gene pairs present on chromosome 2. This indicates that the A. trifoliata genome (A) is closely related to the published A. trifoliata genome(B) and has a high degree of NAC gene homology.

2.3. Analysis of Promoter Cis-Acting Elements and Protein Interaction Networks

Gene function is found to be linked to differences in promoter regions [43]. This difference is primarily due to the type and number of cis-acting elements involved in the regulation of gene expression. Hence, we identified 52 cis-acting elements associated with hormone response, light responsiveness, stress responsiveness, and tissue-specific expression, distributed in the 2000-bp upstream region of 101 AktNAC genes (Figure S4). In terms of hormone response, ABA-responsive elements (ABREs), CGTCA motif (MEJA response), TCA-element (salicylic acid response), and TGACG motif (MEJA response) were present in most AktNAC genes, with AktNAC084 comprising 25 ABREs. The light-responsive elements have the largest number of types (31), of which six elements, Box 4, GATA motif, G-box, GT1 motif, I-box, and TCT motif, are present in most AktNAC genes. Elements such as 3-AF1 binding site, AAAC motif, AT1 motif, CAG motif, Gap box, ACA-motif and Sp1 are present in fewer genes, while GATA-motif, chs-Unit 1 m1, and ACA-motif were present only in AktNAC057, AktNAC064, AktNAC092. Tissue-specific expression-related elements, such as CAT box, GCN4 motif, MBSI, O2 site, and RY element, were also identified, which were associated with phloem expression, endosperm expression, flavonoid biosynthesis, and nitrogen-responsive and seed-specific expression, respectively.

2.4. Transcriptomic Analysis of the Early Developmental Process in the Seeds of A. trifoliata

Seed development can have three stages: early embryogenesis, maturity, and post-embryogenesis [44]. During the early-embryogenesis stage, seed weight and lipid content are low. At the maturity stage, the seeds increase rapidly in terms of dry weight and accumulate large amounts of storage oil and storage protein. In the post-embryogenesis stage, the dry weight of the seeds remains constant while severe water loss occurs. The seed structure of A. trifoliata showed that 0DAF–70DAF (days after flower) is at the early stage of embryogenesis when seed endosperm cells divide rapidly and actively and the seeds increase rapidly in size (Figure 4). Moreover, we analysed the expression of AktNACs genes during the early stages of seed development using transcriptomic data. Based on FPKM > 1, we plotted a heatmap of AktNACs (Figure 5). Among them, 63 genes had expression, and 38 genes had no detectable expression (FPKM < 1). We filtered differential genes based on the value of |log2FC > 1| and FDR < 0.01. Among the genes analyzed, 63 exhibited expression, while 38 had no detectable expression (FPKM < 1). We filtered for differentially expressed genes using the criteria |log2FC > 1| and FDR < 0.01. Comparing 30 days after flowering (DAF) versus 70 DAF, we identified 30 differentially expressed AktNAC genes, with 10 genes (AktNAC022, AktNAC021, AktNAC60, AktNAC85, AktNAC95, AktNAC064, AktNAC034, AkNAC047, AktNAC062, AktNAC071) showing significant upregulation during seed development, and 17 AktNACs (AktNAC030, AktNAC053, AktNAC035, AktNAC031, AktNAC036, AktNAC067, AktNAC032, AktNAC037, AktNAC098, AktNAC072, AktNAC059, AktNAC088, AktNAC033, AktNAC049, AktNAC007, AktNAC011, AktNAC058) exhibiting significant downregulation. These results suggest that NAC genes play an important role in early seed development. We analyzed the expression of seven AktNAC genes during the early seed development stage using quantitative real-time polymerase chain reaction (qRT-PCR; Figure S5). Gene names and primer sequences are presented in Table S3.

2.5. Co-Expression Network and Enrichment Analysis of Differential Genes during Early Seed Development of A. trifoliata

NAC genes are usually involved in the regulation of plant seed growth and development together with other transcription factors. To explore the regulatory network between AktNACs genes and other transcription factors during early seed development, we performed transcription factor identification of transcriptome data during early seed development. A total of 1647 transcription factors were identified and classified into 58 types. The first five families of transcription factors with the highest numbers were bHLH (155); MYB (126); NAC (109); ERF (102); and C2H2 (102). Based on Pearson’s correlation coefficient ≥ 0.9, a co-expression network analysis was performed using a Python script to explore AktNACs co-expression transcription factors, and the results revealed (Figure 6) that the network had 258 nodes and 606 network pairs, with a total of 25 AktNAC genes co-expressed with 233 transcription factors (r > 0.9 or r < −0.9), with the top five transcription factors in highest number being bHLH (19); GRAS (14); HD-ZIP (13); ARF (12); ERF (12); MYB (12).
In parallel, we annotated and performed pathway analysis of 25 AktNACs and 233 transcription factors in the co-expression network (Figure S6). Seed development undergoes cell division and differentiation and presents in structural and morphological changes, with the embryo maintaining mitotic activity through the continuous accumulation of starch converted to glucose. Sucrose upregulates the expression of genes associated with storage proteins and initiates the cellular differentiation process [44]. Plant hormones are also involved in the regulation of seed development. During seed development genes related to transport and energy are highly transcribed and energy metabolism is constantly changing. These genes are annotated in biological processes mainly related to the cellular process, metabolic process, biological regulation, and development process [44]. In biological process, genes were involved in positive regulation of transcription, DNA-templated, positive regulation of RNA biosynthetic process, positive regulation of nucleic acid-templated transcription, response to gibberellin, regulation of flavonoid biosynthetic process, floral organ development, etc. In cellular component category, genes were involved in transcription factor complex, endoplasmic reticulum membrane, nuclear outer membrane-endoplasmic reticulum membrane network, endoplasmic reticulum subcompartment, endoplasmic reticulum part and others. In molecular function, genes were involved in regulatory region nucleic acid binding, transcription regulatory region DNA binding, protein dimerization activity, transcription factor binding and others.
We further performed the GO and KEGG enrichment analyses (Figure S6). GO enrichment analysis indicated that genes were involved in co-expressed network regulatory region nucleic acid binding (53), transcription regulatory region DNA binding (53), positive regulation of transcription, DNA-templated (52), positive regulation of RNA biosynthetic process (53), positive regulation of nucleic acid-templated transcription (53) and others. KEGG enrichment analysis indicated that genes were involved in plant hormone signal transduction (22), circadian rhythm—plant (4), MAPK signaling pathway (5).

2.6. AkNACs Involved in Metabolic Regulation in Early Seed Development

By integrating the transcriptome and metabolome data, we performed an association analysis of the differential NAC and all upregulated metabolites in the transcriptome found in the NAC and metabolite association analysis (Figure 7). There were 51 nodes with 175 network pairs (r > 0.95) in the co-expression network, including 23 AktNACs and 28 metabolites, of which 122 network pairs are negatively correlated and 53 network pairs are positively correlated. 7 genes with Degree ≥ 10 were AktNAC050 (17), AktNAC095 (16), AktNAC058 (13), AktNAC089 (11), AktNAC035 (10), AktNAC031 (10), AktNAC060 (10). Six metabolites with Degree > 10 were Procyanidin B1(15), L-Methionine (14), N6-Succinyl Adenosine (14), Cytidine (14). Methylmalonic acid (13), Kaempferol 7-O-glucosdie (12). AktNAC095 is a significantly up-regulated gene in seed development in Trifolium. By correlation analysis with metabolites (Table 1), we found that eight metabolites strongly associated with AktNAC095, Kinic acid, Trifolin, Histamine, γ-Aminobutyric, Cytosine, Uridine 5′-diphospho-d-glucose, Kaempferol 7-O-glucosdie, d-Glucose 6-phosphate expression was significantly up-regulated during seed development (Figure 8).

2.7. Co-Expression Analysis of AktNAC, TFs and Metabolites

We performed co-expression network analysis of AktNACs, transcription factor and relative peak areas of metabolites (Figure 9). The results showed that there were 401 nodes in the co-expression network, containing 25 AktNACs, 235 transcription factors and 141 metabolites. The positive correlation network was 596 pairs (r > 0.9), and the negative correlation network was 263 pairs (r < −0.9). The degree of correlation between each node was expressed as Pearson’s correlation coefficient (r), where the top 5 AktNACs were AktNAC030 (133), AktNAC032 (106), AktNAC089 (83), AktNAC037 (79), AktNAC093 (65), with degrees between 10 and 60 containing 12 AktNACs. In the co-expression process, there were multiple transcription factors associated with AktNACs, and the transcription factors with degree > 5 contained 14 transcription factors, namely mikado.Chr08G960 (TALE; degree = 7), mikado.Chr06G1013 (BES1, degree = 7), mikado.Chr03G263 (ERF, degree = 6), mikado.Chr04G2651 (HD-ZIP, degree = 6), mikado.Chr13G492 (HD-ZIP, degree = 6), mikado.Chr12G1156 (bHLH, degree = 6), mikado.Chr11G1290 (GRAS, degree = 6), mikado.Chr09G896 (MIKC_MADS, degree = 6), mikado.Chr07G2034 (CO-like, degree = 6), mikado.Chr06G1656 (CO-like, degree = 6), mikado.Chr06G1417 (HD-ZIP, degree = 6), mikado.Chr06G1350 (GRAS, degree = 6), mikado.Chr02G2324 (WRKY, degree = 6), mikado.Chr01G2513 (LBD, degree = 6).
It is hypothesized that AktNACs synergistically regulated the biosynthesis of metabolites with transcription factors, the most typical of which are AktNAC030 and AktNAC32 through synergy with multiple transcription factors, with AktNAC030 co-expressing with 96 transcription factors and AktNAC32 co-expressing with 63 transcription factors, resulting in multiple transcription factors synergistically regulating NAC transcription factors. Among the transcription factors, AktNAC032 and AktNAC095 were the most typical ones, with AktNAC032 regulating the synthesis of 41 metabolites (r > |0.9|) and AktNAC095 regulating the synthesis of 35 metabolites (r > |0.9|). Integration analysis of AktNACs, transcription factors and metabolites, we identified the most likely transcription factors that synergistically regulate metabolites with AktNAC060, AktNAC095. This indicates that NAC plays an important role in the synthesis of compounds and that transcription factors are co-expressed in concert with NAC in order to achieve regulation of metabolite biosynthesis (Table 2).

2.8. Protein Interaction Networks of AktNAC

Protein interaction networks consist of interacting proteins that participate in various aspects of biological processes, such as signaling and regulation of gene expression. Systematic analysis of a large number of protein interactions is crucial for understanding the functional connections between proteins. To further comprehend the functions of 25 AktNACs within the co-expression network of AktNACs, transcription factors (TFs), and associated metabolites, we constructed a NAC protein interaction network based on Arabidopsis homologs (Figure 10). Among these, 12 genes from M. tridentata were involved in the network. Five AktNACs were directly linked to NTL (homolog of AktNAC038), and NTL was linked to SOG1 (homolog of AktNAC058), each acting within their respective clusters. ATAF1 (homolog of AktNAC006) serves as the nodal gene connecting the two clusters. Gene Ontology (GO) predictions of gene functions within the network include nucleus, DNA binding, organic cyclic compound binding, regulation of cellular metabolic processes, regulation of macromolecule metabolic processes, regulation of transcription (DNA-templated), regulation of nucleobase-containing compound metabolism, cell cycle, and DNA repair (Table S9).

3. Discussion

In this study, we identified, for the first time, members of the NAC gene family at the genomic level using the transcriptome and metabonomics to characterize expression during the early seed development stage, which provided a valuable resource for further studies on the function of A. trifoliata NAC genes.
We identified 101 AktNAC genes encoding these protein families, a number close to Arabidopsis thaliana (105) [37] and pepper (104) [45], but less than rice (151) [9] and maize (152) [46], and more than grape (74) [47] and mung bean (81) [48]. These results suggest that the number of NAC gene families varies across species, which may be due to the evolutionary expansion of NAC gene families. Exons and introns are essential features of gene evolution. The AktNAC gene exons identified in this study ranged from one to nine, with most AktNAC genes having three exons, similar to species such as walnut [49], cucumber [50], and Chenopodium quinoa [51]. This may be related to the functional differentiation and structural diversity of the AktNAC gene family. There are differences in the arrangement of conserved AktNAC motifs located in different groups, but most AktNAC genes have a similar motif composition, which corresponds to the NAC gene structure and function.
In this study, 101 AktNACs were classified into 11 subgroups based on the results of phylogenetic tree analysis of A. trifoliata and Arabidopsis, grape, and rice genes. NAC-a (NAM subclass) is mainly involved in embryo, stem tip, flower, and root growth and development [1,2], while NAC-g and NAC-k (ATAF/NAP subclass) usually regulate different types of stresses [52]. This suggests that AktNAC genes have multiple biological functions in plant growth and development and stress response.
Gene duplication can provide an evolutionary impetus for species to generate new functions [53]. Genome-wide duplication, which occurs in genetic lines of several domesticated crops including wheat (Triticum aestivum), cotton (Gossypium hirsutum), and soybean (Glycine max), contributes to significant agronomic traits, such as grain quality, fruit shape, and flowering time [39]. Segmental duplications and tandem duplications are the main types of eukaryotic duplication events that occur [54]. We identified 44 pairs of duplicated genes, with the earliest duplication occurring at 94.629 MYA (million years ago) and the most recent duplication occurring at 3.578 MYA (Table S2). In this study, the Ka/Ks ratio was mostly <1, indicating that the duplicated AktNAC genes were subjected to strong purifying selection pressure. The results suggested that segmental duplication contributed significantly to the expansion and evolution of the NAC gene family in A. trifoliata.
The morphology and microstructure of A. trifoliata seeds showed that before 70DAF the seeds mainly underwent cell division and differentiation, a process involving rapid increase in volume, with the endosperm gradually forming to cover the entire seed and the embryo developing into a globular embryo. Approximately 70DAF–150DAF seeds showed a less pronounced increase in volume, and the embryo gradually developed into a rod-shaped structure. The transcriptome data revealed the expression pattern of A. trifoliata seeds during the early developmental stage. In this study, we found that six differentially expressed class I NAC genes were significantly upregulated at 70 DAF, including four in the NAM subclass (AktNAC060, AktNAC064, AktNAC047, AktNAC071), one in the OsNAC8 subclass (AktNAC34), and one in the ATAF/NAP subclass (AktNAC085). Additionally, there were four class II NAC genes (AktNAC62, AktNAC95, AktNAC22, AktNAC21) significantly upregulated at 70 DAF. Thirteen differentially expressed class I NAC genes showed significant downregulation, including seven in the OsNAC8 subclass and four in the ATAF/NAP subclass. Furthermore, four class II NAC genes were observed. In Arabidopsis thaliana, class I NACs exhibit greater conservation in subdomains B and E compared to class II NACs, a pattern also observed in A. trifoliata. The differential expression of class I and class II AktNAC genes in seeds may be related to their structural differences in the B and E subdomains. AktNAC039, AktNAC040, AktNAC085, and AktNAC100 were homologous to Arabidopsis thaliana NAC056, which when combined with NAC018/NARS2 regulated embryogenesis by expressing the development and degeneration of the ovule integument, a process related to intertissue communication between the embryo and the maternal ovule integument [18]. In this study, AktNAC085 was upregulated, while embryo formation started after 50 DAF, indicating that AktNAC085 is highly likely involved in the regulation of embryonic development. The interaction of embryonic development gene VvNAC26 (GSVIVT01019952001), homologous to AktNAC059 and VvMADS9 affects the development of grape seeds and fruits [55]. Expression of the VvNAC26 gene was more pronounced in the ovules of seedless grape varieties than in the ovules of seeded grape varieties. Moreover, overexpression of VvNAC26 in tomato resulted in smaller fruits and seeds of transgenic tomato VvNAC26, and significantly fewer seeds at the fruit ripening stage than nontransgenic tomato and transgenic control containing only empty vector. In this study, AktNAC059 expression was higher at 30 DAF than at 50 DAF and 70 DAF, indicating that AktNAC059 may be expressed earlier in seed development to influence seed size and number.
Plant growth and development are regulated by the expression of a large number of genes, including many specifically expressed transcription factors. Studies have shown that NAC transcription factors are involved in fruit development and ripening, embryonic development, stress response, and phytohormone signaling. FtNAC70 in buckwheat is highly expressed under exogenous PEG, NaCl, cold, MeJA, ABA and GA induction. Tartary Buckwheat plants overexpressing FtNAC70 showed greater salt and drought tolerance resistance [56]. Arabidopsis thaliana ATAF2 promotes ethylene biosynthesis and response through the activation of related genes [57]. Plant secondary metabolites are nonessential small molecule compounds for the proper functioning of plant growth and development, which are usually not directly involved in growth and development and have an important role in plant ecological adaptation [58]. The NAC transcription factors can synergistically regulate the biosynthesis of flavonoids, caffeine and theanine with MYB, bHLH and other transcription factors [59]. CsNAC7 positively regulates the caffeine synthase gene yhNAMT1, whose overexpression increases caffeine accumulation in transgenic tea healing tissues [60]. Watermelon ClNAC68 inhibits the expression of IAA-amino acid synthase (ClNAC3.6) in the IAA signaling pathway, positively regulates IAA accumulation in fruit and promotes normal seed development [14]. Overexpression of Norway spruce PaNAC03 leads to downregulation of CHS, F3′H and PaLAR3 expression in the flavonoid synthesis pathway, resulting in abnormal embryo development [61]. In this study, transcriptome and metabolome association analysis of seed development in A. trifoliata identified strong correlations (r > 0.95) between AktNACs and many metabolites, which mainly include organic acids, alkaloids, flavonoids, lignans, coumarins, terpenoid and lipids. There are also numerous transcription factor families that are synergistically regulated with AktNACs, including MYB, bHLH, WRKY, C3H, B3, AP2 and others. This indicates that AktNACs have an important role in the synthesis of metabolites during the development of A. trifoliata seeds, which is consistent with literature reports. However, further studies are needed for their specific regulatory relationships and functions.
Protein interactions have also demonstrated that five AktNACs (AktNAC042, AktNAC071, AktNAC089, AktNAC050, and AktNAC038) are linked to NTL proteins, which are NAC with TM motif1-like (NTL) proteins, similar to tail-anchored (TA) protein structures. NTL11 has been previously shown to upregulate the expression of genes associated with flavonoid biosynthesis, leading to the accumulation of anthocyanins in response to intense light stress [22]. NTL9 is a negative regulator of formative layer activity and plays a repressive role in the developmental transition of the stem vascular system to the secondary growth stage [62]. Consequently, AktNAC042 and AktNAC038 may be involved in the regulation of flavonoid biosynthesis and secondary vascular development. SOG1 is thought to be implicated in the DNA damage response of plant p53 [63]. Moreover, other related NAC factors, such as NAC044, have also been shown to be involved in the DNA damage response [64]. This suggests that their homologs, AktNAC058 and AktNAC011, may also mediate the DNA damage response in A. trifoliata. NAC083, also known as VNI2, exhibits specificity in regulating xylem cells [65]. RD26 regulates the ABA and the brassinosteroid signaling pathway [33,66]. Therefore, AktNAC095 and AktNAC084 may mediate xylem differentiation and regulation of ABA signaling, while AktNAC071 (a homolog of Arabidopsis NAC100) may bind to the promoter regions of genes involved in chlorophyll catabolic processes. In summary, AktNAC071 and AktNAC095 are important candidates for photosynthesis, flavonoid biosynthesis, and stress response in A. trifoliata seed development, and their functions require further validation.

4. Materials and Methods

4.1. Plant Material

In this study, wild germplasm was transplanted in the teaching farm of Guizhou university 3 years ago, whose plant origin is Akebia trifoliata (Thunb.) Koidz, belonging to Lardizabalaceae, genus Akebia Decne., identified by the plant identification center of Guizhou University. In this study, different stages in the same tree were harvested and then classified according to the seed development stage, with different developmental stages used as different samples. The tree was labeled on the day it fully blossomed (20 April 2019). Different fruits were randomly collected every 20 days during early seed development (20 May, 10 June and 1 July 2019), including 30 DAF (30 days after flower), 50 DAF (50 days after flower), 70 DAF (70 days after flower). Pulp and peel were quickly removed, and a total of 3 biological replicates were collected in each stage. The tender leaves were also collected from the tree. Samples were frozen in liquid nitrogen and stored at −80 °C with leaves for genomic analysis, and seeds for transcriptome and metabolome analysis.

4.2. Total RNA Extraction, cDNA Library Construction, and RNA-Seq

Total RNA was extracted from A. trifoliata seeds using the Trizol kit (Invitrogen, CA, USA). RNA degradation was assessed using an Agilent 2100 Bioanalyzer (Agilent Technologies, CA, USA) and RNase-free agarose gel electrophoresis. Total RNA was removed from rRNA and reverse transcribed into cDNA using random primers. cDNA libraries were constructed and sequenced using Illumina HiSeqTM 4000 (Gene Denovo Biotechnology Co., Guangzhou, China).
Raw data were run through fastp [67] for quality control to remove reads containing adapters, those containing >10% N, and those of low quality (number of bases with a quality value of Q ≤ 20 accounting for more than 50% of the entire read). The filtered data were mapped to the A. trifoliata reference genome using HISAT2 software (version 2.1.0) [68]. Transcripts were reconstructed using StringTie (version 1.3.4) [69], and the expression of genes in each sample was calculated with RSEM using the fragments per kilobase per million mapped reads (FPKM) method. We calculated the differentially expressed genes between the three different samples according to the FPKM method. A model with a negative binomial distribution was used for the three samples using DESeq2 [70]. The p-values obtained by DESeq were corrected using the Benjamini–Hochberg program, and the corrected p-value was used to detect the false discovery rate (FDR). We identified the differential genes by log2Fold Change <1 & >1, and FDR < 0.01.

4.3. Identification of NAC Genes in A. trifoliata

The genome, coding sequence, protein sequence, and Gff annotation file of A. trifoliata were assembled by sequencing at Genedenovo Biotechnology Co., Ltd. (Guangzhou, China). The Arabidopsis NAC protein sequence was downloaded from the Arabidopsis Information Resource (TAIR, https://www.arabidopsis.org/ accessed on 18 April 2022), and the grape and rice were downloaded from plantTFDB (http://planttfdb.gao-lab.org/index.php accessed on 18 April 2022). The Arabidopsis NAC protein was used as the query sequence, and the predicted protein sequence of A. trifoliata genome was used as the database for identification using the BLAST alignment method with TBtools (v. 1.0987663) [71]. The obtained AktNAC genes were identified using Pfam (http://pfam.xfam.org/family accessed on 18 April 2022), NCBI-CDD (https://www.ncbi.nlm.nih.gov/cdd/ accessed on 18 April 2022), and SMART (http://smart.embl-heidelberg.de/ accessed on 18 April 2022) online tools for NAC protein conserved structural domain prediction, excluding false-positive genes.

4.4. Basic Properties and Phylogenetic Analysis of AktNACs

The basic properties of NAC family proteins were analyzed using ExPaSy’s online Protparam software (https://web.expasy.org/protparam/ accessed on 22 April 2022). The hypotheses related to subcellular localization were analyzed using PSORTb version 2.0 [72] (https://www.psort.org/psortb/index.html accessed on 22 April 2022). The phosphorylation sites of proteins were predicted using NetPhos (v. 3.1) (https://services.healthtech.dtu.dk/service.php?NetPhos-3.1 accessed on 22 April 2022). The Muscle in MEGA 7.0 [73] was used to perform multiple sequence comparisons of the identified white mullein and Arabidopsis NAC proteins. We adopted the neighbor-joining method to construct the phylogenetic tree, with parameters set to the Poisson correction, P-distance, and bootstrap value of 1000 replicates, and other parameters by default. The phylogenetic trees were displayed using the online software iTOL (v. 6) [74] (https://itol.embl.de/ accessed on 22 April 2022).

4.5. Gene Structure, Conserved Motif, and Cis-Element Analyses of AktNACs

Based on the information from the Gff annotation file of A. trifoliata, the gene structure was analyzed using TBtools (v. 1.0987663) to characterize and map the gene structure. Conserved sequences of AktNACs were analyzed using the MEME version 5.0.5 [75] (http://meme-suite.org/tools/meme accessed on 26 April 2022), with the number of searches set to 10 for motif and other parameters by default plotted using TBtools (v. 1.0987663).
The 2000-bp sequence upstream of the transcriptional start site of AktNACs was extracted using TBtools (v. 1.0987663), and then submitted to PlantCARE (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/ accessed on 26 April 2022) [76] to retrieve and analyze the cis-acting elements of the promoter region. The protein interaction network map was constructed using the STRING database (https://cn.string-db.org/ accessed on 26 April 2022) [77]. Moreover, the GO enrichment analysis was performed on the protein interaction network map node genes and plotted using the R package.

4.6. Chromosomal Location and Gene Duplication

The distribution location of the AktNAC on the chromosome was determined from the gene annotation data and was drafted using the R package. Duplication events of genes were analyzed using the Multiple Collinearity Scan toolkit (MCScanX) [78] with the default settings. The differentiation rate of replicated genes was calculated using the Ka/Ks ratio of the nonsynonymous substitution rate (Ka) to the synonymous substitution rate (Ks). Genes with Ks >2.0 were discarded to avoid substitution saturation. The divergence time was calculated using the following formula: T (T = Ks/(2r) × 10−8), where r is 1.5 × 10−8 synonymous substitutions per site per year for dicotyledonous plants [79].

4.7. Paraffin-Embedded (FPE) Tissue Sections of A. trifoliata Seeds

Seeds at 30 DAF, 50 DAF, 70 DAF, and 130 DAF were taken, photographed, and recorded; their appendages were removed, washed in distilled water, and blotted on filter paper; seeds were cut with a scalpel; the taken material was placed in an anatomical flask with FAA fix solution, pumped for more than 30 min until the material sank to the bottom of the vessel, and fixed with FAA fix solution for at least 24 h. (1) Paraffin sectioning: the sections were sequentially placed in xylene I for 20 min, xylene II for 5 min, anhydrous ethanol II for 5 min, 75% alcohol for 5 min, and then washed using tap water. (2) Safranin O staining: slices were stained in safranin O staining solution for 1–2 h and slightly washed using tap water to remove excess dye. (3) Decolorization: each slice was placed in 50%, 70%, and 80% gradient alcohol for 3–8 s. (4) Fast green staining: slices were placed in a fast green staining solution for 30–60 s, stained for 30–60 s, and dehydrated using anhydrous ethanol. (5) Transparent sealing: slices were placed in fresh xylene for 5 min and subjected to the neutral resin sealing. (6) Finally, microscopic examination, image acquisition, and analysis were conducted.

4.8. Metabolomics Experimental and Analytical Methods

All chemicals and reagents were of analytical grade. Methyl alcohol, acetonitrile and ethyl alcohol were purchased from Merck Company, Germany (www.merckchemicals.com accessed on 21 April 2022). Milli-Q system (Millipore Corp., Bedford, MA, USA) ultrapure water was used throughout the study. Authentic standards were purchased from BioBioPha Co., Ltd. Kunming, China (www.biobiopha.com/ accessed on 21 April 2022) and Sigma-Aldrich, St. Louis, MO, USA (www.sigmaaldrich.com/unitedstates.html accessed on 21 April 2022).
The freeze-dried samples were crushed using a mixer mill (MM 400, Retsch) with a zirconia bead for 1.5 min at 30 Hz. Then 100 mg powder was weighed and extracted overnight at 4 °C with 1.0 mL 70% aqueous methanol containing 0.1 mg/L lidocaine for internal standard. Following centrifugation at 10,000× g for 10 min, the supernatant was absorbed and filtrated (SCAA-104, 0.22-μm pore size; ANPEL, Shanghai, China, www.anpel.com.cn/ accessed on 2 May 2022) before LC–MS/MS analysis.
The compounds extracted were analyzed using an LC-ESI-MS/MS system (UPLC, Shim-pack UFLC SHIMADZU CBM30A, http://www.shimadzu.com.cn/ accessed on 16 May 2022; MS/MS (Applied Biosystems 6500 QTRAP, http://www.appliedbiosystems.com.cn/ accessed on 16 May 2022) [80]. Two μL of samples were injected onto a Waters ACQUITY UPLC HSS T3 C18 column (2.1 mm × 100 mm, 1.8 µm) operating at 40 °C and a flow rate of 0.4 mL/min. The mobile phases used were acidified water (0.04% acetic acid) (Phase A) and acidified acetonitrile (0.04% acetic acid) (Phase B). Compounds were separated using the following gradient: 95:5 Phase A/Phase B at 0 min; 5:95 Phase A/Phase B at 11.0 min; 5:95 Phase A/Phase B at 12.0 min; 95:5 Phase A/Phase B at 12.1 min; 95:5 Phase A/Phase B at 15.0 min. The effluent was connected to an ESI-triple quadrupole-linear ion trap (Q TRAP)–MS. LIT and triple quadrupole (QQQ) scans were acquired on a triple quadrupole-linear ion trap mass spectrometer (Q TRAP), AB Sciex QTRAP6500 System, equipped with an ESI-Turbo Ion-Spray interface, operating in a positive ion mode and controlled by Analyst 1.6.1 software (AB Sciex). The operation parameters were as follows: ESI source temperature 500 °C; ion spray voltage (IS) 5500 V; curtain gas (CUR) 25 psi; the collision-activated dissociation (CAD) was set highest. QQQ scans were acquired as MRM experiments with optimized declustering potential (DP) and collision energy (CE) for each individual MRM transitions. The m/z range was set between 50 and 1000.

4.9. qRT-PCR Verification of AktNACs

Seven unigenes were randomly selected to verify the reliability of the transcriptome data, and primers were designed using the Primer 5.0 software. Total RNA was extracted using cetyl trimethylammonium bromide considering RG5 (EF-α) as an internal reference gene. Approximately 16 µL of the gDNA removal system was prepared on ice using the iScript™ gDNA Clear cDNA Synthesis kit, and then this was incubated at 25 °C for 5 min and at 75 °C for 5 min to remove the gDNA. Then, 20 µL of the reverse transcription system was prepared on ice, and incubated at 25 °C for 5 min, 46 °C for 20 min, and 95 °C for 5 min. Subsequently, 20 µL of the reverse transcription system was prepared on ice, incubated at 25 °C for 5 min and at 46 °C for 20 min, denatured at 95 °C for 1 min, cooled on ice, and used immediately for qPCR. The reaction was carried out using an ABI QuantStudio 6 Flex fluorescent PCR instrument with an SYBR Premix EX Taq kit. The experiment used 10 μL of 2× qPCR Mix, 0.5 μL of each primer, 0.5 μL of cDNA, and 8.5 μL of ddH2O. The amplification program was set as follows: 95 °C 300 s followed by 40 cycles of 95 °C for 10 s and 60 °C for 30 s. The relative expression was calculated using the 2–ΔΔCT method [81], where T1 was used as a control to normalize the experimental data (T1, T2 and T3 represent 30DAF, 50DAF and 70DAF, respectively). Gene names and primer sequences are presented in Table S3.

4.10. Statistical Analysis

GO and KEGG annotation of A. trifoliata using Eggnog (http://eggnog-mapper.embl.de/ accessed on 16 October 2022). Transcription factor identification of genome-wide proteins using Planttfdb (http://planttfdb.gao-lab.org/prediction.php accessed on 16 October 2022) [82,83]. Transcriptome and metabolome correlations were calculated using Python scripts and correlations were calculated using Pearson correlation coefficients (PCCs) [84]. Cytoscape v3.7.1 was used for co-expression network graphs for visualization [85]. Graphpad prism 6.0 (GraphPad Inc., La Jolla, CA, USA) was used for t-test analysis of metabolite content and visualization of metabolic content.

5. Conclusions

In this study, we identified the NAC family from A. trifoliata genome. One hundred one AktNACs were divided into eight subgroups. These NAC genes were unevenly distributed on 16 chromosomes. Genes with similar conserved motifs and gene structures clustered phylogenetically into one group. cis-acting elements of the AktNACs gene promoters were associated with plant hormones, organ development and abiotic stresses. Transcriptome analysis showed that AktNAC060, AktNAC085, AktNAC071 and AktNAC095 showed upregulated at all three stages. Combined transcriptome and metabolome analyses indicated that AktNACs synergistically regulate the synthesis of seed metabolites with a variety of transcription factors. These candidate genes provide an important theoretical basis for future molecular breeding of AktNACs. These candidate genes provide an important theoretical foundation for future molecular breeding of AktNACs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/plants12071518/s1.

Author Contributions

Z.Z. and H.L. (Hongchang Liu) designed the study; H.L. (Huijuan Liu) and C.X. conducted the gene family bioinformatics analysis; S.C. and J.L. supervised the experimental work; M.H. and H.W. performed the experiments; H.L. (Huijuan Liu) wrote the manuscript; Z.Z. and X.W. provided funding. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Guizhou Province Science and Technology Plan Project (Grant number: No. 1061), the “Hundred” Level Innovative Talent Foundation of Guizhou Province (Grant number: No. GCC [2022]023-1), the Research on targeted knockout of fruit cracking-related genes based on CRISPR/Cas9 system” (Grant number: [2019] No.1061), and the Talent 532 Base Project of Organization Department in Guizhou Province, China (Grant number: QRLF (2013) 533 no. 15) (Grant number: QRLF (2016) no. 23) (Grant number: QRLF (2020) no. 2).

Data Availability Statement

The datasets of RNA-seq presented in this study can be found in online repositories (NCBI SRA with BioProject ID: PRJNA884501).

Acknowledgments

We thank Guangzhou Genedenovo Biotechnology Co., Ltd. (Guangzhou, China) for assisting in sequencing and/or bioinformatics analysis. We are grateful to Chengjie Chen from South China Agricultural University for kindly providing the TBtools software (v. 1.0987663).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Souer, E.; van Houwelingen, A.; Kloos, D.; Mol, J.; Koes, R. The No Apical Meristem Gene of Petunia Is Required for Pattern Formation in Embryos and Flowers and Is Expressed at Meristem and Primordia Boundaries. Cell 1996, 85, 159–170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Aida, M.; Ishida, T.; Fukaki, H.; Fujisawa, H.; Tasaka, M. Genes involved in organ separation in Arabidopsis: An analysis of the cup-shaped cotyledon mutant. Plant Cell 1997, 9, 841–857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Puranik, S.; Sahu, P.P.; Srivastava, P.S.; Prasad, M. NAC proteins: Regulation and role in stress tolerance. Trends Plant Sci. 2012, 17, 369–381. [Google Scholar] [CrossRef]
  4. Duval, M.; Hsieh, T.-F.; Kim, S.Y.; Thomas, T.L. Molecular characterization of AtNAM: A member of the Arabidopsis NAC domain superfamily. Plant Mol. Biol. 2002, 50, 237–248. [Google Scholar] [CrossRef] [PubMed]
  5. Xie, Q.; Frugis, G.; Colgan, D.; Chua, N.-H. Arabidopsis NAC1 transduces auxin signal downstream of TIR1 to promote lateral root development. Genes Dev. 2000, 14, 3024–3036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Ernst, H.A.; Olsen, A.N.; Skriver, K.; Larsen, S.; Leggio, L.L. Structure of the conserved domain of ANAC, a member of the NAC family of transcription factors. EMBO Rep. 2004, 5, 297–303. [Google Scholar] [CrossRef] [PubMed]
  7. Hegedus, D.; Yu, M.; Baldwin, D.; Gruber, M.; Sharpe, A.; Parkin, I.; Whitwill, S.; Lydiate, D. Molecular characterization of Brassica napus NAC domain transcriptional activators induced in response to biotic and abiotic stress. Plant Mol. Biol. 2003, 53, 383–397. [Google Scholar] [CrossRef] [PubMed]
  8. Nakashima, K.; Takasaki, H.; Mizoi, J.; Shinozaki, K.; Yamaguchi-Shinozaki, K. NAC transcription factors in plant abiotic stress responses. Biochim. Biophys. Acta (BBA) Gene Regul. Mech. 2011, 1819, 97–103. [Google Scholar] [CrossRef]
  9. Nuruzzaman, M.; Manimekalai, R.; Sharoni, A.M.; Satoh, K.; Kondoh, H.; Ooka, H.; Kikuchi, S. Genome-wide analysis of NAC transcription factor family in rice. Gene 2010, 465, 30–44. [Google Scholar] [CrossRef]
  10. Christiansen, M.W.; Holm, P.B.; Gregersen, P.L. Characterization of barley (Hordeum vulgare L.) NAC transcription factors suggests conserved functions compared to both monocots and dicots. BMC Res. Notes 2011, 4, 302. [Google Scholar] [CrossRef] [Green Version]
  11. Liu, L.; White, M.J.; MacRae, T.H. Transcription factors and their genes in higher plants: Functional domains, evolution and regulation. Eur. J. Biochem. 1999, 262, 247–257. [Google Scholar] [CrossRef]
  12. Olsen, A.N.; Ernst, H.A.; Leggio, L.L.; Skriver, K. NAC transcription factors: Structurally distinct, functionally diverse. Trends Plant Sci. 2005, 10, 79–87. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, G.-S.; Li, H.-L.; Grierson, D.; Fu, D.-Q. NAC Transcription Factor Family Regulation of Fruit Ripening and Quality: A Review. Cells 2022, 11, 525. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, J.; Wang, Y.; Zhang, J.; Ren, Y.; Li, M.; Tian, S.; Yu, Y.; Zuo, Y.; Gong, G.; Zhang, H.; et al. The NAC transcription factor ClNAC68 positively regulates sugar content and seed development in watermelon by repressing ClINV and ClGH3.6. Hortic. Res. 2021, 8, 214. [Google Scholar] [CrossRef] [PubMed]
  15. Hao, Y.-J.; Wei, W.; Song, Q.-X.; Chen, H.-W.; Zhang, Y.-Q.; Wang, F.; Zou, H.-F.; Lei, G.; Tian, A.-G.; Zhang, W.-K.; et al. Soybean NAC transcription factors promote abiotic stress tolerance and lateral root formation in transgenic plants. Plant J. 2011, 68, 302–313. [Google Scholar] [CrossRef] [PubMed]
  16. Yang, R.; Deng, C.; Ouyang, B.; Ye, Z. Molecular analysis of two salt-responsive NAC-family genes and their expression analysis in tomato. Mol. Biol. Rep. 2010, 38, 857–863. [Google Scholar] [CrossRef] [PubMed]
  17. Zhang, H.; Wu, J.; Hu, S.; Yu, X. Isolation and expression analysis of BcNAC2, a NAC transcription factor gene in turnip. Acta Hortic. Sin. 2011, 38, 1089–1096. [Google Scholar]
  18. Kunieda, T.; Mitsuda, N.; Ohme-Takagi, M.; Takeda, S.; Aida, M.; Tasaka, M.; Kondo, M.; Nishimura, M.; Hara-Nishimura, I. NAC Family Proteins NARS1/NAC2 and NARS2/NAM in the Outer Integument Regulate Embryogenesis in Arabidopsis. Plant Cell 2008, 20, 2631–2642. [Google Scholar] [CrossRef] [Green Version]
  19. Mathew, I.E.; Das, S.; Mahto, A.; Agarwal, P. Three Rice NAC Transcription Factors Heteromerize and Are Associated with Seed Size. Front. Plant Sci. 2016, 7, 1638. [Google Scholar] [CrossRef] [Green Version]
  20. Li, Z.; Wei, X.; Tong, X.; Zhao, J.; Liu, X.; Wang, H.; Tang, L.; Shu, Y.; Li, G.; Wang, Y.; et al. The OsNAC23-Tre6P-SnRK1a feed-forward loop regulates sugar homeostasis and grain yield in rice. Mol. Plant 2022, 15, 706–722. [Google Scholar] [CrossRef]
  21. Verdaguer, R.; Soler, M.; Serra, O.; Garrote, A.; Fernández, S.; Company-Arumí, D.; Anticó, E.; Molinas, M.; Figueras, M. Silencing of the potato StNAC103 gene enhances the accumulation of suberin polyester and associated wax in tuber skin. J. Exp. Bot. 2016, 67, 5415–5427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Morishita, T.; Kojima, Y.; Maruta, T.; Nishizawa-Yokoi, A.; Yabuta, Y.; Shigeoka, S. Arabidopsis NAC Transcription Factor, ANAC078, Regulates Flavonoid Biosynthesis under High-light. Plant Cell Physiol. 2009, 50, 2210–2222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Sun, Y.; Jiang, C.; Jiang, R.; Wang, F.; Zhang, Z.; Zeng, J. A Novel NAC Transcription Factor From Eucalyptus, EgNAC141, Positively Regulates Lignin Biosynthesis and Increases Lignin Deposition. Front. Plant Sci. 2021, 12, 642090. [Google Scholar] [CrossRef] [PubMed]
  24. He, F.; Zhang, L.; Zhao, G.; Kang, J.; Long, R.; Li, M.; Yang, Q.; Chen, L. Genome-Wide Identification and Expression Analysis of the NAC Gene Family in Alfalfa Revealed Its Potential Roles in Response to Multiple Abiotic Stresses. Int. J. Mol. Sci. 2022, 23, 10015. [Google Scholar] [CrossRef] [PubMed]
  25. Song, H.; Liu, Y.; Dong, G.; Zhang, M.; Wang, Y.; Xin, J.; Su, Y.; Sun, H.; Yang, M. Genome-Wide Characterization and Comprehensive Analysis of NAC Transcription Factor Family in Nelumbo nucifera. Front. Genet. 2022, 13, 901838. [Google Scholar] [CrossRef] [PubMed]
  26. Rai, N.; Rai, K.K.; Singh, M.K.; Singh, J.; Kaushik, P. Investigating NAC Transcription Factor Role in Redox Homeostasis in Solanum lycopersicum L.: Bioinformatics, Physiological and Expression Analysis under Drought Stress. Plants 2022, 11, 2930. [Google Scholar] [CrossRef] [PubMed]
  27. Flora of China Editorial Committee. Flora of China; Science Press (Beijing): Beijing, China; MBG Press: St. Louis, MO, USA, 2018; Volume 1. [Google Scholar]
  28. Gao, H.; Wang, Z. Triterpenoid saponins and phenylethanoid glycosides from stem of Akebia trifoliata var. australis. Phytochemistry 2006, 67, 2697–2705. [Google Scholar] [CrossRef]
  29. Yu, N.; Wang, X.; Ning, F.; Jiang, C.; Li, Y.; Peng, H.; Xiong, H. Development of antibacterial pectin from Akebia trifoliata var. australis waste for accelerated wound healing. Carbohydr. Polym. 2019, 217, 58–68. [Google Scholar] [CrossRef]
  30. Wang, J.; Xu, Q.-L.; Zheng, M.-F.; Ren, H.; Lei, T.; Wu, P.; Zhou, Z.-Y.; Wei, X.-Y.; Tan, J.-W. Bioactive 30-Noroleanane Triterpenes from the Pericarps of Akebia trifoliata. Molecules 2014, 19, 4301–4312. [Google Scholar] [CrossRef] [Green Version]
  31. Wang, X.; Yu, N.; Peng, H.; Hu, Z.; Sun, Y.; Zhu, X.; Jiang, L.; Xiong, H. The profiling of bioactives in Akebia trifoliata pericarp and metabolites, bioavailability and in vivo anti-inflammatory activities in DSS-induced colitis mice. Food Funct. 2019, 10, 3977–3991. [Google Scholar] [CrossRef]
  32. Tran, L.-S.P.; Nakashima, K.; Sakuma, Y.; Simpson, S.D.; Fujita, Y.; Maruyama, K.; Fujita, M.; Seki, M.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Isolation and Functional Analysis of Arabidopsis Stress-Inducible NAC Transcription Factors That Bind to a Drought-Responsive cis-Element in the early responsive to dehydration stress 1 Promoter. Plant Cell 2004, 16, 2481–2498. [Google Scholar] [CrossRef] [Green Version]
  33. Fujita, M.; Fujita, Y.; Maruyama, K.; Seki, M.; Hiratsu, K.; Ohme-Takagi, M.; Tran, L.S.; Yamaguchi-Shinozaki, K.; Shinozaki, K. A dehydration-induced NAC protein, RD26, is involved in a novel ABA-dependent stress-signaling pathway. Plant J. 2004, 39, 863–876. [Google Scholar] [CrossRef]
  34. Kriventseva, E.V.; Koch, I.; Apweiler, R.; Vingron, M.; Bork, P.; Gelfand, M.S.; Sunyaev, S. Increase of functional diversity by alternative splicing. Trends Genet. 2003, 19, 124–128. [Google Scholar] [CrossRef]
  35. Stetefeld, J.; Ruegg, M.A. Structural and functional diversity generated by alternative mRNA splicing. Trends Biochem. Sci. 2005, 30, 515–521. [Google Scholar] [CrossRef]
  36. Mach, J. Alternative Splicing Produces a JAZ Protein That Is Not Broken Down in Response to Jasmonic Acid. Plant Cell 2009, 21, 14. [Google Scholar] [CrossRef] [Green Version]
  37. Ooka, H.; Satoh, K.; Doi, K.; Nagata, T.; Otomo, Y.; Murakami, K.; Matsubara, K.; Osato, N.; Kawai, J.; Carninci, P.; et al. Comprehensive Analysis of NAC Family Genes in Oryza sativa and Arabidopsis thaliana. DNA Res. 2003, 10, 239–247. [Google Scholar] [CrossRef]
  38. Kikuchi, K.; Ueguchi-Tanaka, M.; Yoshida, K.T.; Nagato, Y.; Matsusoka, M.; Hirano, H.Y. Molecular analysis of the NAC gene family in rice. Mol. Genet. Genom. 2000, 262, 1047–1051. [Google Scholar] [CrossRef]
  39. Panchy, N.; Lehti-Shiu, M.; Shiu, S.-H. Evolution of Gene Duplication in Plants. Plant Physiol. 2016, 171, 2294–2316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Suyama, M.; Torrents, D.; Bork, P. PAL2NAL: Robust conversion of protein sequence alignments into the corresponding codon alignments. Nucleic Acids Res. 2006, 34 (Suppl. S2), W609–W612. [Google Scholar] [CrossRef] [Green Version]
  41. Hurst, L.D. The Ka/Ks ratio: Diagnosing the form of sequence evolution. Trends Genet. 2002, 18, 486–487. [Google Scholar] [CrossRef]
  42. Nekrutenko, A.; Makova, K.D.; Li, W.-H. The KA/KS Ratio Test for Assessing the Protein-Coding Potential of Genomic Regions: An Empirical and Simulation Study. Genome Res. 2001, 12, 198–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Xue, T.; Wang, D.; Zhang, S.; Ehlting, J.; Ni, F.; Jakab, S.; Zheng, C.; Zhong, Y. Genome-wide and expression analysis of protein phosphatase 2C in rice and Arabidopsis. BMC Genom. 2008, 9, 550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Baud, S.; Boutin, J.-P.; Miquel, M.; Lepiniec, L.; Rochat, C. An integrated overview of seed development in Arabidopsis thaliana ecotype WS. Plant Physiol. Biochem. 2002, 40, 151–160. [Google Scholar] [CrossRef]
  45. Diao, W.; Snyder, J.C.; Wang, S.; Liu, J.; Pan, B.; Guo, G.; Ge, W.; Dawood, M.H.S.A. Genome-Wide Analyses of the NAC Transcription Factor Gene Family in Pepper (Capsicum annuum L.): Chromosome Location, Phylogeny, Structure, Expression Patterns, Cis-Elements in the Promoter, and Interaction Network. Int. J. Mol. Sci. 2018, 19, 1028. [Google Scholar] [CrossRef] [Green Version]
  46. Shiriga, K.; Sharma, R.; Kumar, K.; Yadav, S.K.; Hossain, F.; Thirunavukkarasu, N. Genome-wide identification and expression pattern of drought-responsive members of the NAC family in maize. Meta Gene 2014, 2, 407–417. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, N.; Zheng, Y.; Xin, H.; Fang, L.; Li, S. Comprehensive analysis of NAC domain transcription factor gene family in Vitis vinifera. Plant Cell Rep. 2012, 32, 61–75. [Google Scholar] [CrossRef]
  48. Tariq, R.; Hussain, A.; Tariq, A.; Bin Khalid, M.H.; Khan, I.; Basim, H.; Ingvarsson, P.K. Genome-wide analyses of the mung bean NAC gene family reveals orthologs, co-expression networking and expression profiling under abiotic and biotic stresses. BMC Plant Biol. 2022, 22, 343. [Google Scholar] [CrossRef] [PubMed]
  49. Khan, H.; Yan, F.; Yan, Y.; Chen, P.; Xi, R.; Ullah, I.; Peng, X.; Luo, X.; Yue, M.; Zhao, P. Genome-wide analysis of evolution and expression profiles of NAC transcription factor gene family in Juglans regia L. Ann. For. Sci. 2020, 77, 82. [Google Scholar] [CrossRef]
  50. Liu, X.; Wang, T.; Bartholomew, E.; Black, K.; Dong, M.; Zhang, Y.; Yang, S.; Cai, Y.; Xue, S.; Weng, Y.; et al. Comprehensive analysis of NAC transcription factors and their expression during fruit spine development in cucumber (Cucumis sativus L.). Hortic. Res. 2018, 5, 31. [Google Scholar] [CrossRef] [PubMed]
  51. Li, F.; Guo, X.; Liu, J.; Zhou, F.; Liu, W.; Wu, J.; Zhang, H.; Cao, H.; Su, H.; Wen, R. Genome-Wide Identification, Characterization, and Expression Analysis of the NAC Transcription Factor in Chenopodium quinoa. Genes 2019, 10, 500. [Google Scholar] [CrossRef] [Green Version]
  52. Collinge, M.; Boller, T. Differential induction of two potato genes, Stprx2 and StNAC, in response to infection by Phytophthora infestans and to wounding. Plant Mol. Biol. 2001, 46, 521–529. [Google Scholar] [CrossRef]
  53. Magadum, S.; Banerjee, U.; Murugan, P.; Gangapur, D.; Ravikesavan, R. Gene duplication as a major force in evolution. J. Genet. 2013, 92, 155–161. [Google Scholar] [CrossRef] [PubMed]
  54. Cannon, S.B.; Mitra, A.; Baumgarten, A.; Young, N.D.; May, G. The roles of segmental and tandem gene duplication in the evolution of large gene families in Arabidopsis thaliana. BMC Plant Biol. 2004, 4, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Zhang, S.; Dong, R.; Wang, Y.; Li, X.; Ji, M.; Wang, X. NAC domain gene VvNAC26 interacts with VvMADS9 and influences seed and fruit development. Plant Physiol. Biochem. 2021, 164, 63–72. [Google Scholar] [CrossRef] [PubMed]
  56. Huang, J.; Ren, R.; Rong, Y.; Tang, B.; Deng, J.; Chen, Q.; Shi, T. Identification, Expression, and Functional Study of Seven NAC Transcription Factor Genes Involved in Stress Response in Tartary Buckwheat (Fagopyrum tataricum (L.) Gaertn.). Agronomy 2022, 12, 849. [Google Scholar] [CrossRef]
  57. Peng, H.; Phung, J.; Stowe, E.C.; Dhingra, A.; Neff, M.M. The NAC transcription factor ATAF2 promotes ethylene biosynthesis and response in Arabidopsis thaliana seedlings. FEBS Lett. 2022, 596, 1586–1599. [Google Scholar] [CrossRef]
  58. Bourgaud, F.; Gravot, A.; Milesi, S.; Gontier, E. Production of plant secondary metabolites: A historical perspective. Plant Sci. 2001, 161, 839–851. [Google Scholar] [CrossRef]
  59. Li, C.-F.; Zhu, Y.; Yu, Y.; Zhao, Q.-Y.; Wang, S.-J.; Wang, X.-C.; Yao, M.-Z.; Luo, D.; Li, X.; Chen, L.; et al. Global transcriptome and gene regulation network for secondary metabolite biosynthesis of tea plant (Camellia sinensis). BMC Genom. 2015, 16, 560. [Google Scholar] [CrossRef] [Green Version]
  60. Ma, W.; Kang, X.; Liu, P.; She, K.; Zhang, Y.; Lin, X.; Li, B.; Chen, Z. The NAC-like transcription factor CsNAC7 positively regulates the caffeine biosynthesis-related gene yhNMT1 in Camellia sinensis. Hortic. Res. 2022, 9, uhab046. [Google Scholar] [CrossRef]
  61. Dalman, K.; Wind, J.J.; Nemesio-Gorriz, M.; Hammerbacher, A.; Lundén, K.; Ezcurra, I.; Elfstrand, M. Overexpression of PaNAC03, a stress induced NAC gene family transcription factor in Norway spruce leads to reduced flavonol biosynthesis and aberrant embryo development. BMC Plant Biol. 2017, 17, 6. [Google Scholar] [CrossRef] [Green Version]
  62. Sugimoto, H.; Tanaka, T.; Muramoto, N.; Kitagawa-Yogo, R.; Mitsukawa, N. Transcription factor NTL9 negatively regulates Arabidopsis vascular cambium development during stem secondary growth. Plant Physiol. 2022, 190, 1731–1746. [Google Scholar] [CrossRef] [PubMed]
  63. Yoshiyama, K.; Conklin, P.A.; Huefner, N.D.; Britt, A.B. Suppressor of gamma response 1 (SOG1) encodes a putative transcription factor governing multiple responses to DNA damage. Proc. Natl. Acad. Sci. USA 2009, 106, 12843–12848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Lang, L.; Pettkó-Szandtner, A.; Elbaşı, H.T.; Takatsuka, H.; Nomoto, Y.; Zaki, A.; Dorokhov, S.; De Jaeger, G.; Eeckhout, D.; Ito, M.; et al. The DREAM complex represses growth in response to DNA damage in Arabidopsis. Life Sci. Alliance 2021, 4, e202101141. [Google Scholar] [CrossRef] [PubMed]
  65. Yamaguchi, M.; Ohtani, M.; Mitsuda, N.; Kubo, M.; Ohme-Takagi, M.; Fukuda, H.; Demura, T. VND-INTERACTING2, a NAC Domain Transcription Factor, Negatively Regulates Xylem Vessel Formation in Arabidopsis. Plant Cell 2010, 22, 1249–1263. [Google Scholar] [CrossRef] [Green Version]
  66. Ye, H.; Liu, S.; Tang, B.; Chen, J.; Xie, Z.; Nolan, T.M.; Jiang, H.; Guo, H.; Lin, H.Y.; Li, L.; et al. RD26 mediates crosstalk between drought and brassinosteroid signalling pathways. Nat. Commun. 2017, 8, 14573. [Google Scholar] [CrossRef] [Green Version]
  67. Chen, S.; Zhou, Y.; Chen, Y.; Gu, J. fastp: An ultra-fast all-in-one FASTQ preprocessor. Bioinformatics 2018, 34, i884–i890. [Google Scholar] [CrossRef]
  68. Kim, D.; Langmead, B.; Salzberg, S.L. HISAT: A fast spliced aligner with low memory requirements. Nat. Methods 2015, 12, 357–360. [Google Scholar] [CrossRef] [Green Version]
  69. Pertea, M.; Pertea, G.M.; Antonescu, C.M.; Chang, T.-C.; Mendell, J.T.; Salzberg, S.L. StringTie enables improved reconstruction of a transcriptome from RNA-seq reads. Nat. Biotechnol. 2015, 33, 290–295. [Google Scholar] [CrossRef] [Green Version]
  70. Love, M.I.; Huber, W.; Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014, 15, 550. [Google Scholar] [CrossRef] [Green Version]
  71. Chen, C.J.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.H.; Xia, R. TBtools: An Integrative Toolkit Developed for Interactive Analyses of Big Biological Data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef]
  72. Gardy, J.L.; Laird, M.R.; Chen, F.; Rey, S.; Walsh, C.J.; Ester, M.; Brinkman, F.S.L. PSORTb v.2.0: Expanded prediction of bacterial protein subcellular localization and insights gained from comparative proteome analysis. Bioinformatics 2004, 21, 617–623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Letunic, I.; Bork, P. Interactive Tree Of Life (iTOL) v5: An online tool for phylogenetic tree display and annotation. Nucleic Acids Res. 2021, 49, W293–W296. [Google Scholar] [CrossRef] [PubMed]
  75. Bailey, T.L.; Williams, N.; Misleh, C.; Li, W.W. MEME: Discovering and analyzing DNA and protein sequence motifs. Nucleic Acids Res. 2006, 34 (Suppl. S2), W369–W373. [Google Scholar] [CrossRef] [PubMed]
  76. 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, 325–327. [Google Scholar] [CrossRef]
  77. Szklarczyk, D.; Gable, A.L.; Lyon, D.; Junge, A.; Wyder, S.; Huerta-Cepas, J.; Simonovic, M.; Doncheva, N.T.; Morris, J.H.; Bork, P.; et al. STRING v11: Protein–protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids Res. 2019, 47, D607–D613. [Google Scholar] [CrossRef] [Green Version]
  78. Wang, Y.; Tang, H.; DeBarry, J.D.; Tan, X.; Li, J.; Wang, X.; Lee, T.-H.; Jin, H.; Marler, B.; Guo, H.; et al. MCScanX: A toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Res. 2012, 40, e49. [Google Scholar] [CrossRef] [Green Version]
  79. Lynch, M.; Conery, J.S. The evolutionary demography of duplicate genes. J. Struct. Funct. Genom. 2003, 3, 35–44. [Google Scholar] [CrossRef]
  80. Chen, W.; Gong, L.; Guo, Z.; Wang, W.; Zhang, H.; Liu, X.; Yu, S.; Xiong, L.; Luo, J. A Novel Integrated Method for Large-Scale Detection, Identification, and Quantification of Widely Targeted Metabolites: Application in the Study of Rice Metabolomics. Mol. Plant 2013, 6, 1769–1780. [Google Scholar] [CrossRef] [Green Version]
  81. Livak, K.J.; Schmittgen, T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 2-ΔΔCT method. Methods 2001, 25, 402–408. [Google Scholar] [CrossRef]
  82. Cantalapiedra, C.P.; Hernández-Plaza, A.; Letunic, I.; Bork, P.; Huerta-Cepas, J. eggNOG-mapper v2: Functional Annotation, Orthology Assignments, and Domain Prediction at the Metagenomic Scale. Mol. Biol. Evol. 2021, 38, 5825–5829. [Google Scholar] [CrossRef] [PubMed]
  83. Tian, F.; Yang, D.-C.; Meng, Y.-Q.; Jin, J.; Gao, G. PlantRegMap: Charting functional regulatory maps in plants. Nucleic Acids Res. 2019, 48, D1104–D1113. [Google Scholar] [CrossRef] [PubMed]
  84. Zhang, T.; Song, C.; Song, L.; Shang, Z.; Yang, S.; Zhang, D.; Sun, W.; Shen, Q.; Zhao, D. RNA Sequencing and Coexpression Analysis Reveal Key Genes Involved in α-Linolenic Acid Biosynthesis in Perilla frutescens Seed. Int. J. Mol. Sci. 2017, 18, 2433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Shannon, P.; Markiel, A.; Ozier, O.; Baliga, N.S.; Wang, J.T.; Ramage, D.; Amin, N.; Schwikowski, B.; Ideker, T. Cytoscape: A software environment for integrated models of Biomolecular Interaction Networks. Genome Res. 2003, 13, 2498–2504. [Google Scholar] [CrossRef]
Figure 1. Phylogenetic tree analysis of NAC protein in A. trifoliata (101), grape (71), rice (141), and Arabidopsis thaliana (115). We used the neighbour-joining method to construct the phylogenetic tree, with parameters set to the P-distance, pairwise deletion, and bootstrap value of 1000 replicates, and other parameters by default. The blue, purple, orange, and bottle green represent the NAC proteins of A. trifoliata, Arabidopsis thaliana, grape, and rice respectively.
Figure 1. Phylogenetic tree analysis of NAC protein in A. trifoliata (101), grape (71), rice (141), and Arabidopsis thaliana (115). We used the neighbour-joining method to construct the phylogenetic tree, with parameters set to the P-distance, pairwise deletion, and bootstrap value of 1000 replicates, and other parameters by default. The blue, purple, orange, and bottle green represent the NAC proteins of A. trifoliata, Arabidopsis thaliana, grape, and rice respectively.
Plants 12 01518 g001
Figure 2. The conservation of AktNAC amino acid residues was analyzed. The depth of color is used to represent conservation. A darker color indicates a higher conservation of amino acids or nucleotides at that position in the aligned sequences. A, B, C, D, and E are the five subdomains of the NAC domain. The x-axis represents the positions of amino acids, while the y-axis indicates the frequency of each amino acid at a specific position appears in a particular position. Different colors are used to represent different types of amino acids. “N” and “C” typically represent the amino terminus (N-terminus) and carboxyl terminus (C-terminus), respectively.
Figure 2. The conservation of AktNAC amino acid residues was analyzed. The depth of color is used to represent conservation. A darker color indicates a higher conservation of amino acids or nucleotides at that position in the aligned sequences. A, B, C, D, and E are the five subdomains of the NAC domain. The x-axis represents the positions of amino acids, while the y-axis indicates the frequency of each amino acid at a specific position appears in a particular position. Different colors are used to represent different types of amino acids. “N” and “C” typically represent the amino terminus (N-terminus) and carboxyl terminus (C-terminus), respectively.
Plants 12 01518 g002
Figure 3. The chromosome locations of AktNACs genes. The chromosome numbers were indicated at the top of each chromosome. Red indicates tandem-duplicated genes. The location was drafted using Perl 5, version 18, subversion 4 (v5.18.4).
Figure 3. The chromosome locations of AktNACs genes. The chromosome numbers were indicated at the top of each chromosome. Red indicates tandem-duplicated genes. The location was drafted using Perl 5, version 18, subversion 4 (v5.18.4).
Plants 12 01518 g003
Figure 4. The seed development process of A. trifoliata. (A) Seed microstructure was observed under a laser confocal microscope. Scale bars: 200 µm. SC: Seed coat, EM: Embryo, EN: endosperm. (B) Seed development morphological changes was observed by SMZ800 stereomicroscope. Scale bars: 5 mm. 30DAF, 50DAF, 70DAF, and 150DAF represent 30 days, 50 days, 70 days, and 150 days after flowering, respectively.
Figure 4. The seed development process of A. trifoliata. (A) Seed microstructure was observed under a laser confocal microscope. Scale bars: 200 µm. SC: Seed coat, EM: Embryo, EN: endosperm. (B) Seed development morphological changes was observed by SMZ800 stereomicroscope. Scale bars: 5 mm. 30DAF, 50DAF, 70DAF, and 150DAF represent 30 days, 50 days, 70 days, and 150 days after flowering, respectively.
Plants 12 01518 g004
Figure 5. The expression levels of AktNACs in early seeds at different developmental stages. Each column in the figure represents a sample and each row represents a gene, normalized by gene expression FPKM values for the rows using the z-score method.
Figure 5. The expression levels of AktNACs in early seeds at different developmental stages. Each column in the figure represents a sample and each row represents a gene, normalized by gene expression FPKM values for the rows using the z-score method.
Plants 12 01518 g005
Figure 6. Co-expression network and enrichment analysis of AktNACs during early seed development of A. trifoliata. Co-expression network of AktNAC genes constructed based on transcriptome data during early seed development. Diamonds indicates AktNAC genes, round indicated transcription factors, and color shade indicates the number of associated genes.
Figure 6. Co-expression network and enrichment analysis of AktNACs during early seed development of A. trifoliata. Co-expression network of AktNAC genes constructed based on transcriptome data during early seed development. Diamonds indicates AktNAC genes, round indicated transcription factors, and color shade indicates the number of associated genes.
Plants 12 01518 g006
Figure 7. Co-expression network analysis of AktNACs and metabolites. Different shapes and colors represent different substances.
Figure 7. Co-expression network analysis of AktNACs and metabolites. Different shapes and colors represent different substances.
Plants 12 01518 g007
Figure 8. Metabolite content in the co-expression network of AktNACs and metabolites. Error bars denote the standard deviation of three replicates (* p < 0.05, ** p < 0.01).
Figure 8. Metabolite content in the co-expression network of AktNACs and metabolites. Error bars denote the standard deviation of three replicates (* p < 0.05, ** p < 0.01).
Plants 12 01518 g008
Figure 9. Co-expression of AktNACs, TFs and compounds.
Figure 9. Co-expression of AktNACs, TFs and compounds.
Plants 12 01518 g009
Figure 10. The interaction network of Arabidopsis thaliana NAC homologs proteins. The purple dashed oval indicated the Arabidopsis thaliana homologous genes to the A. trifoliata NAC protein, and the AktNACs were shown in purple below the Arabidopsis thaliana protein.
Figure 10. The interaction network of Arabidopsis thaliana NAC homologs proteins. The purple dashed oval indicated the Arabidopsis thaliana homologous genes to the A. trifoliata NAC protein, and the AktNACs were shown in purple below the Arabidopsis thaliana protein.
Plants 12 01518 g010
Table 1. Correlation analysis of metabolites associated with co-expression of AktNAC095.
Table 1. Correlation analysis of metabolites associated with co-expression of AktNAC095.
Gene NameCompoundMolecular FormulaPCC
AktNAC095Kinic acidC7H12O60.982401 **
AktNAC095TrifolinC21H20O110.980908 **
AktNAC095HistamineC5H9N30.980678 **
AktNAC095γ-AminobutyricC4H9NO20.973989 **
AktNAC095CytosineC4H5N3O0.969859 **
AktNAC095Uridine 5′-diphospho-d-glucoseC15H24N2O17P20.962642 **
AktNAC095Kaempferol 7-O-glucosideC21H20O110.953569 **
AktNAC095d-Glucose 6-phosphateC6H13O9P0.951671 **
** Means p-value < 0.01.
Table 2. Correlation analysis of AkNACs co-expression network-regulated metabolites.
Table 2. Correlation analysis of AkNACs co-expression network-regulated metabolites.
Node1Node2CorrelationNode3Correlation
mikado.Chr05G296 (B3)AktNAC0600.994325284 **Kinic acid0.990795 **
mikado.Chr04G2027 (MYB)AktNAC0600.984943879 **Uridine 5′-diphospho-d-glucose0.984606 **
mikado.Chr04G2027 (MYB)AktNAC0600.984943879 **Histamine0.983068 **
mikado.Chr05G50 (GRAS)AktNAC0950.982596163 **Kinic acid0.982401 **
mikado.Chr01G2874 (HD-ZIP)AktNAC0950.981437254 **Trifolin0.980908 **
mikado.Chr01G2874 (HD-ZIP)AktNAC0950.981437254 **Histamine0.980678 **
** Means p-value < 0.01, Correlation > 0.9: Positive.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, H.; Chen, S.; Wu, X.; Li, J.; Xu, C.; Huang, M.; Wang, H.; Liu, H.; Zhao, Z. Identification of the NAC Transcription Factor Family during Early Seed Development in Akebia trifoliata (Thunb.) Koidz. Plants 2023, 12, 1518. https://doi.org/10.3390/plants12071518

AMA Style

Liu H, Chen S, Wu X, Li J, Xu C, Huang M, Wang H, Liu H, Zhao Z. Identification of the NAC Transcription Factor Family during Early Seed Development in Akebia trifoliata (Thunb.) Koidz. Plants. 2023; 12(7):1518. https://doi.org/10.3390/plants12071518

Chicago/Turabian Style

Liu, Huijuan, Songshu Chen, Xiaomao Wu, Jinling Li, Cunbin Xu, Mingjin Huang, Hualei Wang, Hongchang Liu, and Zhi Zhao. 2023. "Identification of the NAC Transcription Factor Family during Early Seed Development in Akebia trifoliata (Thunb.) Koidz" Plants 12, no. 7: 1518. https://doi.org/10.3390/plants12071518

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

Article Metrics

Back to TopTop