Next Article in Journal
Curcumin: A Novel Way to Improve Quality of Life for Colorectal Cancer Patients?
Previous Article in Journal
The SARS-CoV-2 Delta-Omicron Recombinant Lineage (XD) Exhibits Immune-Escape Properties Similar to the Omicron (BA.1) Variant
Previous Article in Special Issue
Chromosome Number and Genome Size Evolution in Brasolia and Sobralia (Sobralieae, Orchidaceae)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Whole Genome Analysis of SLs Pathway Genes and Functional Characterization of DlSMXL6 in Longan Early Somatic Embryo Development

Institute of Horticultural Biotechnology, Fujian Agriculture and Forestry University, Fuzhou 350002, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(22), 14047; https://doi.org/10.3390/ijms232214047
Submission received: 30 September 2022 / Revised: 10 November 2022 / Accepted: 11 November 2022 / Published: 14 November 2022
(This article belongs to the Special Issue Plant Genetic Diversity and Genomics)

Abstract

:
Strigolactones (SLs), a new class of plant hormones, are implicated in the regulation of various biological processes. However, the related family members and functions are not identified in longan (Dimocarpus longan Lour.). In this study, 23 genes in the CCD, D27, and SMXL family were identified in the longan genome. The phylogenetic relationships, gene structure, conserved motifs, promoter elements, and transcription factor-binding site predictions were comprehensively analysed. The expression profiles indicated that these genes may play important roles in longan organ development and abiotic stress responses, especially during early somatic embryogenesis (SE). Furthermore, GR24 (synthetic SL analogue) and Tis108 (SL biosynthesis inhibitor) could affect longan early SE by regulating the levels of endogenous IAA (indole-3-acetic acid), JA (jasmonic acid), GA (gibberellin), and ABA (abscisic acid). Overexpression of SMXL6 resulted in inhibition of longan SE by regulating the synthesis of SLs, carotenoids, and IAA levels. This study establishes a foundation for further investigation of SL genes and provides novel insights into their biological functions.

1. Introduction

Strigolactones (SLs) are defined as carotenoid-derived plant hormones, which are divided into oxygen-free carotenes, such as b-carotene, and xanthophylls that contain oxygen, such as lutein [1,2]. SLs were originally identified in root parasitic plants as root communication signals between parasites and symbionts and were rediscovered as phytohormones involved in plant development in recent years [3,4,5,6]. SLs play important roles in plant growth and stress responses, such as seed germination, internode elongation, anthocyanin accumulation, secondary growth, shoot branching, leaf development, and adventitious rooting [7,8,9,10,11,12,13,14,15,16]. β-carotene isomerase DWARF27 (D27), carotenoid cleavage dioxygenases, CCD7 (MAX3, RMS5 D17/HTD1, DAD3), and CCD8 (MAX4, RMS1, D10, DAD1) are involved in the biosynthesis of SLs [17,18,19]. Reportedly, D27-catalysed isomerisation of all-trans-β-carotene is the first step of carlactone formation [20], and D27 was demonstrated to be responsible for SL biosynthesis in plastids [19,21,22]). D27 is a component of the MAX/RMS/D pathway and plays an essential role in biosynthesizing strigolactones, D27 acts upstream of MAX1 in the SL biosynthetic pathway in Arabidopsis [22]. The mutation in D27 leads to a deficiency in strigolactone biosynthesis [23,24]. In rice, the levels of D27 mRNAs are highly elevated in roots under P deficiency and contribute to the high levels of SLs [25]. D27 also plays an important role in adaptation to sulfur deficiency in rice [26]. CCD1, CCD4, and CCD7 have broad substrate specificity, while CCD8 may be specific for SL synthesis to produce several forms of SLs [18,27,28,29,30]. Endogenous SL levels are very low in ccd7 and ccd8 mutants [5,6]. Studies demonstrated that mutants ccd7 and ccd8 are deficient in SL and exogenous application of GR24, a synthetic analogue to SL, and rescued the excess branching phenotype of ccd7 and ccd8 [31].
It was demonstrated that SLs employ SMXL (SUPPRESSOR OF MAX2-LIKE) family members for signal transduction in Arabidopsis [32,33], and SMXL2 plays critical roles in seedling photomorphogenesis and downstream gene expression [34]. SMXL can affect shoot and root architecture and leaf shape as a transcriptional repressor [35,36,37,38,39]. Meanwhile, SMXL6 can function as a transcription factor in SL signalling, which can directly bind to the promoters of SMXL6, 7 and 8 in Arabidopsis [40].
Plant somatic embryogenesis (SE) is defined as a general cultivation technique in vitro, where embryos develop from somatic cells without gamete formation and fertilisation [41,42]. Recently, SE became a model system to understand embryonic reproductive programs and epigenetic reprogramming in plants [43]. Plant hormones play important roles in plant SE [44]. Auxin biosynthesis and polar transport are foundational for somatic embryo formation [45,46]. The functional mechanism of SL and auxin biosynthesis, transport, and signalling was extensively studied [47,48]. SL could interfere with auxin polar transport by modifying the PIN-FORMED (PIN) auxin transporter phenotypic output, thereby regulating growth and developmental responses in pea and Arabidopsis [28,49]. In Arabidopsis, an exogenous SL analogue (GR24) inhibits auxin transport, and SLs act downstream of auxin biosynthesis [49]. The results of a previous study demonstrate that SLs can improve the embryogenic process in Arabidopsis, as well as GR24 and Tis108 (an SL biosynthetic inhibitor) on SE correlated with changes in the expression of AUXIN RESPONSIVE FACTORs, which are required for the production of embryogenic tissue [50]. These studies suggest that the interaction between SL and auxin plays an important role in plant SE.
The interactions between SLs and other plant hormones were also reported. In tomato, the level of SLs decreased after treatment with specific ABA inhibitors [51]. Conversely, chemical inhibition of SL biosynthetic enzymes CCD7 and CCD8 did not change root ABA contents [51], indicating a complex regulatory mechanism between ABA and SL signalling pathways. Red to far red (R/FR) light sensing positively influences the arbuscular mycorrhizal (AM) symbiosis of a legume and a non-legume through JA and SL signalling [52]. Previous studies showed that gibberellin (GA) plays an important role in the induction of SE [53,54,55]. Reportedly, the GA signalling pathway can regulate the biosynthesis of SLs [56]. However, the interactions between SLs and other plant hormones during early SE are not uncovered, especially in perennial woody plants.
Dimocarpus longan Lour is cultivated widely for tropical/subtropical evergreen fruit trees. Since the establishment of the longan SE system by Lai [57,58], which provided ideal experimental materials for studying SE in woody angiosperms, SL genes were found to participate in a wide range of plant developmental processes, including shoot architecture, root growth, photomorphogenesis, leaf senescence, and flower development [11]. However, SL signalling and biosynthesis members and functions are not identified in longan SE. Therefore, based on whole-genome identification and bioinformatics, analysis of the genes involved in the SL pathway was carried out. GR24 (1μM) and Tis108 (3 μM) treatments on longan ECs were carried out (embryogenic callus) to analyse their effects on the transformation of longan ECs into GEs (globular embryos). Microscopic observations showed that GR24 promoted the differentiation of longan ECs, while Tis108 delayed the process, and IAA (indole-3-acetic acid), JA (jasmonic acid), GA (gibberellin), and ABA (abscisic acid) participated in the regulation. Meanwhile, exogenous hormones also regulated the expression levels of SL-related family members and the content of endogenous SLs during longan SE. Expression pattern analysis revealed that these genes were involved in the development of different organs and responded to heat and light. Overexpression of longan SMXL6 (DlSMXL6) inhibited the differentiation of longan somatic embryos, which may be related to the synthesis of IAA by upregulating the expression of longan YUC5 (DlYUC5) and longan YUC10 (DlYUC10). These findings provide a valuable resource for better understanding the biological roles of SL biosynthesis and response genes in longan SE.

2. Results

2.1. Identification and Chromosomal Localisation Analysis of Genes Involved in the SL Pathway in Longan

In this study, a total of 23 genes were identified in the SL biosynthesis and signalling pathway, including 11 members in the CCD gene family (DlCCD), eight members in the SMXL gene family (DlSMXL), and four members in the D27 gene family (DlD27). To gain insight into the biological functions of DlCCD, DlSMXL, and DlD27 family members, the protein physicochemical properties of the three gene families were further analysed (Table S2). DlCCD, DlSMXL, and DlD27 encoded proteins with lengths of 526~625 aa, 153~1134 aa, and 142~280 aa, respectively. The MWs of these 23 proteins ranged from 15.4 to 123.80 kDa. The pI ranged from 4.8 to 10.59. The grand average of hydropathicity (GRAVY) values of all members were negative, indicating that they were hydrophilic proteins.
Most family members were predicted to localise in the chloroplast. DlCCD6 and DlCCD7 were found in peroxisomes. DlSMXL6 was predicted to localise in the extracellular matrix. Four members were found primarily in the nucleus, including DlSMXL2, DlSMXL3, DlSMXL7, and DlD27c. In addition, no signal peptides exist for all full members of the three gene families, suggesting that none of the DlCCD, DlSMXL, or DlD27 family members were secreted proteins.
As shown in Figure 1A, 23 putative DlCCD, DlSMXL, and DlD27 family members were unevenly distributed on 10 of the 15 chromosomes of longan. The highest number of genes was observed on chromosome 1 (Chr1), including DlCCD1, DlCCD2, DlCCD3, DlSMXL,1 and DlSMXL2. DlSMXL family genes were located on Chr1, Chr5, Chr7, Chr12, and Chr14. The DlD27 genes were located on Chr5, Chr7, and Chr15. Additionally, Chr4, Chr10, Chr12, and Chr15 only contained one number.
Gene duplication is an important pathway for the amplification in family gene numbers and the source of diverse phenotypes [59]. A previous study showed that 7.59% of tandem repeats were identified in the longan genome [60]. Among the 23 genes in the longan genome, only two pairs in the DlSMXL family had a collinearity relationship (Figure 1A), including DlSMXL3-DlSMXL7 and DlSMXL4-DlSMXL8, which were both classified in subgroup VI of the phylogenetic tree (Figure 2). All paralogous homologous genes were derived from segmental duplication, and no tandem duplication events occurred. The Ka/Ks values of the homologous pairs DlSMXL3-DlSMXL7 and DlSMXL4-DlSMXL8 were 0.24 and 0.16, respectively, suggesting that these two gene pairs were subject to strong purifying selection and were functionally stable. Furthermore, colinear analysis of longan and Arabidopsis genomes suggests that 13 pairs of collinear genes exist between longan and Arabidopsis. Among them, DlCCD8 (Chr10) and DlSMXL7 (Chr12) corresponded to two genes in Arabidopsis (Figure 1B).

2.2. Phylogenetic Analysis and Classification of CCD, SMXL, and D27 Gene Families

To gain insight into the evolutionary relationships and predict the functions of the plant genes involved in SL biosynthesis and signalling pathways, we next compared the full length of these protein sequences as found in Dimocarpus longan Lour, Arabidopsis, Oryza sativa, Citrus sinensis, Zea mays, Glycine max, Solanum lycopersicum, and Triticum aestivum. We constructed phylogenetic trees (Figure 2). Based on the amino acid sequence similarity and topological structure of the phylogenetic tree with MEGA 8 software, these genes were divided into six subfamilies, including Group I to Group Ⅵ. In our study, phylogenetic analysis showed that Group I only belonged to DlCCD5, which may be due to the unique genetic structure. SMLX family genes were highly conserved, as well as homologous in longan and Arabidopsis, which mostly belonged to Group III. DlCCDs were distributed in Group I~Group VI, indicating diversity during evolution.

2.3. Gene Structure and Motif Composition of the CCD, SMXL, and D27 Gene Families in Longan

Gene structural diversity and conserved motif divergence are possible mechanisms for the evolution of multigene families [61]. To further explore the diversity and functional differentiation of the DlCCD, DlSMXL, and DlD27 gene family members, gene structure and protein-conserved motif analyses were performed (Figure S1). Gene structure analysis showed that DlD27 gene family members were scattered both in exons and introns, and the first intron fell within the mature coding sequence, which facilitated the neat separation of the signal sequence from the coding sequence. DlCCD1, DlCCD7, DlCCD9, DlSMXL1, DlSMXL2, DlSMXL4, and DlSMXL7 contained the 5′-UTR, CDS and 3′-UTR, respectively, with a good gene structure. However, none of the DlD27 family members had a 5′-UTR or 3′-UTR, which may affect gene transcription and translation. Analysis of the conserved motif showed that DlCCD family members contained 7 or 10 motifs, and motif 1, motif 2, motif 7, motif 8, motif 9, and motif 10 were present in all CCD family members, indicating a high conservation in longan evolution. DlSMXL5 and DlSMXL6 only contained motif 15 and motif 12, respectively. DlD27 family genes all only contained one motif, and DlD27a, DlD27c, and DlD27d all contained motif 16.

2.4. Cis-Acting Elements in Longan CCD, DlSMXL, and D27 Family Genes in Longan

To further analyse the potential function of the DlCCD, DlSMXL, and DlD27 genes, a 2000-bp regulatory region upstream of the ATG (promoter) was searched for cis-acting element analysis (Figure 3). Among the promoter cis-acting elements of the three gene family members, the largest number of photoresponse elements were present in all members of the three gene families. In addition, responsive cell cycle elements were found only in DlCCD3, which may play an important role in longan cell cycle progression. Most of the members responded to ABA, MeJA, GA, SA, and auxin. Three gene family members also had elements in response to low temperature, drought, stress, seed-specific regulation, endosperm expression, and meristem expression, indicating that these genes may play important roles in reverse resistance, cold tolerance, seed growth, and embryonic mature development.
The cis-acting elements of the promoter could combine with specific transcription factors and then form a transcription initiation complex to initiate gene-specific expression in various biological processes [62]. To investigate the regulation of transcription factors (TFs) on the expression of DlCCD, Dl27, and DlSMXL family genes, transcription factor-binding sites (TFBSs) on the promoter were predicted using the PlantTFDB online tool. In total, 44, 42, and 42 TF families were identified in the DlCCD, DlD27, and DlSMXL family gene promoters, respectively. With the exception of GRF and LFY on DlCCD family gene promoters, most TFBSs were identified on DlCCD, DlD27, and DlSMXL family gene promoters, such as AP2, ARF, BBR-BPC, bZIP, bHLH, Dof, ERF, NAC, WOX, WRKY, MYB, and TCP; this indicates a potential common upstream regulatory mechanism of these genes in longan growth and development and various stress responses, especially in regulating longan early somatic embryogenesis.

2.5. RNA-Seq Revealed the Expression Profiling of Longan CCD, SMXL, and D27 Family Genes in Different Tissues and Treatments

We further analysed the expression patterns of DlCCD, DlSMXL, and DlD27 family genes in different tissues and treatments by transcriptome data (Figure 4). The results show that most genes were highly expressed in stems (Figure 4A), which might be related to the widespread involvement of SL in plant shoot branching. DlCCD2, DlCCD3, DlCCD11, and DlSMXL2 were highly expressed in the flower. DlCCD2, DlCCD4, and DlSMXL7 were highly expressed in the leaf. DlD27a, DlSMXL6, and DlSMXL7 were highly expressed in the roots. DlCCD1, DlCCD3, DlSMXL1, DlSMXL2, DlSMXL6, DlSMXL7, DlD27a, and DlD27d were highly expressed in the seed and might be involved in longan seed dormancy and germination. DlSMXL2, DlCCD5, DlCCD6, and DlCCD7 were highly expressed in the young fruit, indicating that these genes might be involved in fruit development and maturation.
A large number of photoresponsive elements were identified in the three gene family member promoters (Figure 3), and transcriptome data show that DlCCD2, DlCCD6, DlCCD7, DlCCD11, DlSMXL3, and DlSMXL6 were highly expressed under blue light. DlD27a, DlD27d, DlCCD5, DlCCD8, DlCCD10, DlSMXL5, and DlSMXL7 were highly expressed under white light (Figure 4B). In addition, the three gene family members showed different expression patterns under different temperature treatments. Moreover, with the exception of DlSMXL2, DlSMXL3, DlCCD5, and DlCCD8, all three gene family members were upregulated in response to high-temperature stress (35 °C), indicating that these genes might be involved in cell self-repair under high temperature (Figure 4C).

2.6. Genome Expression Profiling of Longan CCD, SMXL and D27 Family Genes and SL Content Determination during Early SE in LONGAN

To further investigate the potential functions of members of the DlCCD, DlSMXL, and DlD27 gene families in longan SE, a heatmap was drawn according to the FPKM values of the genes from the longan transcriptome database (Figure 5B). The results show that DlCCD3, DlCCD5, DlCCD7, DlCCD9, and DlCCD10 were highly expressed in the ECs stage, while DlCCD6 was highly expressed in both the ECs and ICpEC stages. The expression of DlCCD11 was significantly higher in the ICpEC stage than in the EC sand GEs stages, while DlCCD1, DlCCD2, DlCCD4, and DlCCD8 were highly expressed in the GE stage. In the DlSMXL gene family, DlSMXL4 and DlSMXL8 were both expressed at low levels in the three stages. DlSMXL7 was highly expressed in the ECs stage, DlSMXL5 and DlSMXL6 were highly expressed in the ICpEC stage, and DlSMXL1, DlSMXL2, and DlSMXL3 were highly expressed in the GEs stage. DlD27b and DlD27c were expressed at low levels in all three stages of longan somatic embryo development, while DlD27a was highly expressed in the GE stage, and DlD27d was expressed at high levels in both ECs and GEs. The qRT–PCR results show that the genes had the same expression trends as those obtained by RNA-seq (Figure 5C). The SL content increased significantly during longan SE, with the highest content in GEs, approximately 1.26 times that in ECs (Figure 5D). Taken together, the results indicate that the SL pathway may be involved in longan early SE.

2.7. Effects of Exogenous Hormones on CCD, SMXL and D27 Family Gene Expression and SL Content in Longan SE

According to the cis-acting element analysis of the DlCCD, DlSMXL and DlD27 families, we chose 10 validated genes by qRT–PCR to confirm whether the expression patterns were influenced by hormone treatments (IAA, ABA, GA, and JA) for 24 h. We first determined the endogenous SL content of longan ECs under different hormone treatments. The results show that exogenous IAA and GA significantly increased the SL content compared to the control, while ABA and JA decreased the SL content in longan ECs (Figure 6B).
Furthermore, we found that the expression patterns of DlCCD, DlSMXL and DlD27 family genes under different hormone treatments differed (Figure 6A, Table S3). The results show that the expression of DlCCD4, DlCCD8 and DlD27a was upregulated, while DlSMXL2 and DlSMXL6 were downregulated by IAA treatment. DlCCD5, DlCCD8 and DlD27d were upregulated, and DlSMXL2 was downregulated by GA treatment. ABA inhibited the expression of DlCCD5, DlCCD7, and DlD27d, and treatment downregulated the expression of D27a and upregulated the expression of DlSMXL6. These results indicate that these genes could respond to hormones and participate in SL biosynthesis and the accumulation of longan ECs.

2.8. Effects of Exogenous GR24 and Tis108 on Longan SE

To further investigate the function of SL in longan early SE, the synthetic strigolactone analogue GR24 (1 μM) and SL biosynthetic inhibitor Tis108 (3 μM) were applied in this study (Figure 7). The microscopic examination results show that EC cells were already at the ICpEC stage under GR24 treatment, while the EC cells were still loose under Tis108 treatment at 6 days. A few globular embryos were already appeared at 8 d under the GR24 treatment. The typical GE structure appeared in both the control group and GR24 treatment, however, the more numbers and higher compactness of the GEs were observed under GR24 treatment, while materials had not reached the GEs stage under Tis108 treatment at 10 days (Figure 7A).
To further explore the interaction relationship between SLs and other hormones in longan, endogenous hormone contents were determined with the SL synthesis analogue GR24 and the synthesis inhibitor Tis108. The results show that GR24 decreased the IAA levels of longan somatic embryos compared with the control, while Tis108 increased the IAA levels at 6 and 8 days (Figure 7C). These results further suggest that SLs played important roles in longan early SE, which was accompanied by regulating the synthesis of IAA. Further results show that the content of endogenous ABA was globally decreased upon Tis108 treatment at 8 and 10 days (Figure 7D). The content of GA was significantly decreased at 6 and 10 days upon GR24 and Tis108 treatment (Figure 7D). GR24 significantly increased the JA levels at 6 and 8 days, while Tis108 significantly decreased the JA levels at 10 days (Figure 7E). These results show that GR24 and Tis108 may regulate longan early SE by changing the content of endogenous hormones.

2.9. Overexpression of SMXL6 Inhibited Differentiation of the Early Somatic Embryogenesis in Longan

The above results show that DlSMXL6 expression decreased gradually during longan SE and was downregulated by IAA treatment, indicating that it might play a role in somatic embryo development. To verify this prediction, the full-length DlSMXL6 CDS was cloned into the pCAMBIA-1301-35S overexpression vector and stably transformed into longan ECs. We selected the OE-1, OE-2, and OE-3 cell lines with high expression levels for further analyses (Figure 8A). Phenotypic observation results show that the SMXL6-OE lines could not differentiate into GEs at 12 days; however, the WT and 1301-GUS lines exhibited typical GEs, revealing that overexpression of DlSMXL6 inhibited the differentiation of longan somatic embryos (Figure 8B–F). Compared with the WT and 1301-GUS cell lines, the contents of SLs were significantly decreased (Figure 8H). We further found that the precursors for SL biosynthesis and carotenoids were significantly increased in the DlSMXL6-overexpressing cell lines, which may be due to negative feedback regulation (Figure 8G). Furthermore, the levels of IAA were significantly increased in the DlSMXL6-overexpressing cell lines (Figure 8I). To further investigate the regulatory mechanism of IAA synthesis in DlSMXL6-overexpressing cell lines, we also analysed the expression of IAA biosynthetic genes, including DlYUC3, DlYUC5, and DlYUC10, which were differentially expressed during longan early SE stages based on transcriptome data. Compared with WT and 1301-GUS, DlYUC5 and DlYUC10 were obviously upregulated in the DlSMXL6-overexpressing cell lines, which may play important roles in the interaction of auxin and SLs during longan early SE (Figure 8J–L).

3. Discussion

3.1. CCD, D27, and SMXL Members Are Involved in Early SE and Organ Development in Longan

Strigolactones play crucial roles in signalling transportation and gene expression during plant growth processes [7,63]. CCD, D27, and SMXL family genes are involved in SL signalling and biosynthesis [64,65,66,67,68,69,70]. In Arabidopsis, two SL biosynthetic genes, the CCD7 and CCD8 homologous genes MAX3 and MAX4, are upregulated during the induction phase of embryogenesis [50]. In this study, following early SE, SLs accumulated progressively. DlCCD6 was highly expressed in both ECs sand ICpEC stages, and DlCCD1, DlCCD2, DlCCD4, DlCCD8, and D27a were highly expressed in the GE stage, which may play important roles in the accumulation of SLs. The repressors of SLs signalling and biosynthesis, the repressors of SL signalling and biosynthesis, DlSMXL5, DlSMXL6, and DlSMXL7, were expressed at low levels at the GEs stage. These genes may play key roles in regulating SL levels during longan SE (Figure 5).
SLs were found to participate in a wide range of plant developmental processes, including root growth, leaf senescence, photomorphogenesis, and flower development [10,11]. In this study, DlCCD, DlD27, and DlSMXL members showed divergent expression patterns across different tissues of longan. Strigolactone application increased the number of flowers and also promoted flower and fruit size of tomato [14,71]. MAX1 homologues may be involved in the regulation of flower development [72]. CCD7 were highly expressed in the tomato fruit [73]. In this study, DlCCD2, DlCCD3, DlCCD11, and DlSMXL2 were highly expressed in the flower, and DlSMXL2, DlCCD5, DlCCD6, and DlCCD7 were highly expressed in the young fruit, which might be involved in longan floral induction and fruit development. Strigolactones can induce secondary growth in stem and root tissue [65]. In this study, most genes were highly expressed in stems, and DlD27a and DlSMXL7 were highly expressed in roots, which may participate in the regulation of longan root formation and elongation [74]. Studies show that SLs are involved in the regulation of leaf development and senescence [15,75]. In this study, DlCCD2, DlCCD4, and DlSMXL7 were highly expressed in the leaf, which might play a special role in longan leaf development. SL-deficient ΔCCD7 exudates reduced seed germination activity [76]. In this study, DlCCD1, DlCCD3, DlSMXL1, DlSMXL2, DlSMXL7, DlD27a, and DlD27d were highly expressed in the seed, which might be involved in seed dormancy and germination [77,78]. These results provide a foundation for further investigation of DlCCD, DlD27, and DlSMXL members in longan development and maturation.

3.2. CCD, D27, and SMXL Genes Respond to Abiotic Stress in Longan Somatic Embryos

The promoter regions of the DlCCD, DlD27, and DlSMXL genes contained a large number of stress-related cis-acting elements, indicating that these genes may be widely involved in longan somatic embryo stress responses (Figure 3). The analysis of transcriptome data under high temperature and blue light treatment revealed that the expression patterns of DlCCD, DlD27, and DlSMXL genes under different stresses differed (Figure 4). Light is an important factor that affects the synthesis of functional metabolites in longan ECs [79]. In Arabidopsis, light and strigolactone signalling together contribute to the regulation of branching [80]. In this study, DlCCD2, DlCCD6, DlCCD7, DlCCD11, DlSMXL3, and DlSMXL6 were induced by blue light. DlD27a, DlD27d, DlCCD5, DlCCD8, DlCCD10, DlSMXL5, and DlSMXL7 were induced by white light. Reportedly, strigolactone biosynthesis and signalling play crucial roles in the responses to heat stress in tomato, and high growth temperature significantly induced the transcription of CCD7 and CCD8 in the roots at 3 h after heat stress [81]. In this study, most members were upregulated in response to high-temperature stress (35 °C) at 24 h, indicating that these genes might play important roles in cell self-repair under high temperature.

3.3. SLs Are Involved in the Regulation of Longan Early SE by Interacting with Endogenous Hormones

Hormones play crucial roles in plant growth and development, especially in plant SE. Recently, the interactions between SLs and other hormones were uncovered in plant development [82,83]. Auxin plays a crucial role in plant SE [45,46,84,85,86]. In most cases, the transcription levels of CCD7 and CCD8 are induced by auxin [87,88,89,90]. Upregulation of SL biosynthetic genes (CCD7, CCD8, and D27) by auxin was reported [88,89,90]. In this study, we found that IAA significantly increased the SL content compared to the control by regulating the expression of DlCCD4, DlCCD8, DlD27a, DlSMXL2, and DlSMXL6, indicating that these genes may be involved in SL synthesis under IAA treatment to regulate longan SE. To further demonstrate the effects of SLs on ECs differentiation and the endogenous hormone content changes during longan SE, GR24 and Tis108 were applied to treat longan ECs. The results show that GR24 could promote the transformation of ECs into GEs, and Tis108 inhibited the process. In this study, GR24 decreased the IAA level of longan embryos, while Tis108 could increase the IAA level, indicating that GR24 and Tis108 could regulate the SE process by altering the endogenous IAA content [49,91].
ABA plays an important role in plant SE, which can accelerate the synthesis and accumulation of nutrients [92,93,94,95,96]. Both ABA and SLs originate from a carotenoid precursor, which indicated that their biosynthesis pathways might be interregulated [97]. Studies show that SL signalling can regulate ABA metabolism and sensitivity [98,99,100]. In this study, DlCCD5, DlCCD7, and DlD27d were downregulated in longan ECs by ABA treatment. GR24 did not lead to an increased content of ABA [99], and negative regulation of ABA metabolism by SLs was reported [97]. In cherry, GR24 can significantly increase the ABA level of roots, whereas Tis108 would decrease the ABA level [100]. However, in this study, the endogenous ABA levels were significantly decreased upon Tis108 treatment at 8 and 10 days, but did not respond to GR24.
GA was reported to be necessary for the induction of SE in spinach [101,102,103,104] The physiological interactions between SL and GA were reported previously [78,105,106,107]. In this study, exogenous GA significantly increased the SL content of longan ECs by regulating the expression of DlCCD5, DlCCD8, DlD27d, and DlSMXL2. GR24 could significantly increase the GA content in seeds [78], and SL can promote the production of bioactive GA [107]. In Jatropha curcas, SL acted antagonistically to gibberellin in the control of lateral bud outgrowth [105,106]. However, in this study, both GR24 and Tis108 significantly decreased the GA levels at 6 and 10 days, indicating that GA may not respond to GR24 and Tis108 during longan SE.
It was reported that JA inhibits embryo germination in angiosperms [108,109], and a dual mode of GR24 is found in regulating JA accumulation in rice roots [110]. In this study, exogenous JA significantly decreased the SL content of longan ECs by upregulating the expression of DlSMXL6 and downregulating the expression of DlD27a. The JA content significantly increased upon GR24 treatment at 6 and 8 days and significantly decreased upon Tis108 treatment, which may play an important role in longan SE upon SL regulation.
Taken together, GR24 promoted longan SE by decreasing the levels of IAA and increasing the levels of JA at the early stage, while Tis108 inhibited the process by regulating the levels of IAA at the early stage and ABA and JA at the late stage (Figure 9A). However, the regulatory mechanisms among SL and endogenous hormones in longan SE remain to be further investigated.

3.4. Function of SMXL6 in Longan Early Somatic Embryo Development

As a downstream regulatory factor of SL, SMXL6, 7, and 8 are important switches of the SL signalling pathway and play an important role in regulating the growth and development of plant branches, root architecture, and leaf shape [35,36,38,39]. Meanwhile, SMXL6 can function as a transcription factor in SL signalling, which can directly bind to the promoters of SMXL6, 7, and 8 in Arabidopsis [40]. Compared to the progress in other species, little is known about the functions of SL in longan SE. In the present study, DlSMXL6 was specifically highly expressed in ECs and IcpEC, and expressed at low levels in GE, indicating its role in somatic embryo development. The results show that SMXL6-OE cells could not differentiate into GE cells on the 12th day, while WT and 1301-GUS cells exhibited typical GEs.
It was shown that auxin biosynthesis and polar transport are foundational for somatic embryo formation [45,46]. Auxin is involved in the regulation of SLs, which may play a direct role in meristematic processes [68]. SLs can regulate shoot branching and root formation by changing the auxin optimum needed [24,111], and auxin signalling acts downstream of SLs [13,112]. In this study, the content of IAA was significantly increased in the DlSMXL6-overexpressing cell lines compared to the WT and 1301-GUS cell lines by upregulating the expression of DlYUC5 and DlYUC10. As reported, the requirement of IAA content decreased during the plant SE [113]. Overexpression of DlSMXL6 in longan ECs also promoted the accumulation of SL precursor carotenoids, which may be due to the feedback regulation of the SL pathway (Figure 9B).
In summary, we first identified SL pathway genes and revealed the expression patterns of different organ development and abiotic stress responses, especially in early SE in longan. We found that GR24 and Tis108 could affect longan early SE by regulating the levels of endogenous hormones. Functional analysis of DlSMXL6 in longan ECs showed that it was involved in the regulation of SE. However, the regulatory mechanism of DlSMXL6 in longan SE remains to be further investigated.

4. Materials and Methods

4.1. Plant Materials

The ‘Hong He Zi’ (‘HHZ’) longan ECs were obtained as previously described [58]. The experiment was conducted at the Institute of Horticultural Biotechnology at Fujian Agriculture and Forestry University from 2020 to 2021. The ECs were cultured in Murashige and Skoog (MS) liquid medium (20 g∙L−1 sucrose and 7 g∙L−1 agar, pH 5.8~6.0) supplemented with 100 μM IAA, ABA, JA, or GA for 24 h. Additionally, to investigate the effects of SL on the conversion of ECs into GEs in longan, the ECs were cultured in MS liquid medium added with GR24 (1 μM stock in acetone) and Tis108 (3 μM stock in DMSO). The same volume of acetone (Mock1) and DMSO (Mock2) were as controls at 25 °C with 120 r·min−1 shaking in the dark for 10 days. The test materials were immediately frozen in liquid nitrogen and stored at −80 °C for subsequent tests. Each assay was repeated three times.

4.2. Identification of Genes Involved in SL Biosynthesis and Signalling in Longan

The genes involved in SL biosynthesis and signalling in Arabidopsis were obtained from TAIR (http://www.arabido-psis.org (accessed on 10 May 2021). They were used as base alignment sequences with the local BLAST of TBtools [114]. The ‘HHZ’ longan genome database from the NCBI Sequence Read Archive (SRA) database (SRR17675476) was used for homologyously alignment. All of the sequences identified for SL biosynthesis and signalling genes in longan were subjected to NCBI’s Batch CD-search (https://www.ncbi.nlm.nih.gov/Structure/bwrpsb/bwrpsb.cgi (accessed on 10 May 2021) and SMART (http://smart.embl-heidelberg.de/ (accessed on 15 May 2021)) tools to verify their reliability as targets [115]. We also examined the physicochemical properties of SL biosynthesis and signalling proteins, including determining the protein length (aa), molecular weight (MW), isoelectric point (pI), and GRAVY using ExPASy (https://www.expasy.org/ (accessed on 15 May 2021). The subcellular localisation of SL biosynthesis and signalling proteins were predicted with Plant-PLoc (http://www.csbio.sjtu.edu.cn/bioinf/plant/ (accessed on 15 May 2021) [116]. The signal peptides were analysed with SignalP 4.0 (http://www.cbs.dtu.dk/services/SignalP/ (accessed on 15 May 2021). Chromosomal localisation information was extracted from the GFF file, and the results were visualised with Tbtools. The cis-acting elements were analysed with PlantCARE (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/ (accessed on 15 May 2021), which were present from 2000 bp upstream of the gene initiation codon. PlantTFDB (http://planttfdb.cbi.pku.edu.cn/ (accessed on 15 May 2021) was used to predict the transcription factor-binding sites (TFBSs) on promoters with the parameter set of p value ≤ 1 × 10−6. MEME (https://meme-suite.org/meme/tools/meme (accessed on 15 May 2021) was used to analyse the proteins’ conserved motifs. Finally, TBtools software (v1.098669, South China Agricultural University, Guangzhou, China) was used to annotate the gene structure and visualise the conserved domains and motifs. MCScanX software was used to predict the collinearity of the genes of SL biosynthesis and signalling and to mark the collinear genes [117]. KaKs_Calclator 2.0 software (https://sourceforge.net/ (accessed on 20 May 2021) was used to estimate synonymous (Ks) and nonsynonymous (Ka) substitution rates [118].

4.3. Phylogenetic Analysis of the Gene Families Involved in the SL Pathway in Longan

To compare the phylogenetic relationship of the genes involved in SL biosynthesis and signalling in longan with their counterparts in other plant species, we obtained the amino acid sequences of Arabidopsis (Arabidopsis thaliana) from TAIR (http://arabidopsis.org (accessed on 20 May 2021) and rice (Oryza sativa) from hytozome v12.1 (https://phytozome.jgi.doe.gov/pz/portal.html (accessed on 20 May 2021). The amino acid sequences of Zea mays, Citrus sinensis, Glycine max, Solanum lycopersicum, and Triticum aestivum were downloaded from NCBI (https://www.ncbi.nlm.nih.gov/ (accessed on 20 May 2021). The phylogenetic trees were constructed using MEGA 5.0 software with neighbour-joining (NJ) [119].

4.4. Analysis of the FPKM Expression Profile of Genes Involved in the SL Pathway in Longan

The FPKM expression profile of genes involved in the SL pathway in longan was extracted from four transcriptome datasets, including three for the ‘Hong He Zi’ (HHZ) cultivar and one for the ‘Si Ji Mi’ (SJM) cultivar. RNA-seq data during early longan SE were available in the NCBI SRA repository under accession code SRR21979789, SRR21979788 and SRR21979787 (ECs), SRR21979786, SRR21979785 and SRR21979784 (ICpECs, Incomplete compact proembryogenic culture) and SRR21979783, and SRR21979782 and SRR21979781 (GEs). RNA-seq data of different temperature treatments on ECs are available in the NCBI SRA repository under accession code SRR21921625, SRR21921624 and SRR21921623 (15 °C), SRR21921622, SRR21921621 and SRR21921620 (25 °C), SRR21921619, and SRR21921618 and SRR21921617 (35 °C). The different light qualities (blue-light, whitelight, and dark as the control) on ECs are available in dataset [79]. The ‘SJM’ transcriptome dataset was the six-organ (root, stem, leaf, flower, young fruit, and seed) dataset [60]. The fragments per kilobase of exon model per million mapped fragment (FPKM) values of isoform transcripts were analysed with Excel and visualised with TBtool software [114].

4.5. Determination of Endogenous Hormone, Carotenoid, and SL Contents in Longan

The levels of IAA, GA, JA, ABA, carotenoid, and SL were determined with a double antibody one-step sandwich enzyme-linked immunosorbent assay (ELISA). The specific steps were as follows: the standard, sample extract, and horseradish peroxidase (HRP)-labelled detection antibody were added to the coating micropores, their corresponding antibodies were added, and the samples were incubated and washed thoroughly. The substrate 3,3′,5,5′-tetramethylbenzidine was used for colour development. TMB was transformed into blue under the catalysis of peroxidase and finally into yellow under the action of acid. The colour depth was positively correlated with the content of the substance to be measured in the sample. The absorbance was determined with a microplate reader (Infinite M200 PRO, Tecan, Männedorf, Switzerland) at a 450 nm wavelength to calculate the sample activity. Each assay was repeated three times.

4.6. RNA Extraction and Real-Time Fluorescence Quantitative Polymerase Chain Reaction (qRT–PCR)

Total RNA was extracted using the TransZol Up reagent kit (TRANS, Beijing, China) and treated with RNase-free DNase I (TRANS, Beijing, China) following the manufacturer’s instructions. The RNA quality was analysed by agarose gel electrophoresis and quantified using a Nanodrop 2000 spectrophotometer (Thermo Scientific, Wilmington, DE, USA). Only RNAs that met the criteria 260/280 ratio of 1.8–2.1 and 260/230 ratio ≥ 2.0 were used for further analyses and stored at −80 °C.
The RNA was used to synthesise first-strand cDNA using a PrimeScript RT Master Mix (Perfect Real Time) cDNA Synthesis Kit (TaKaRa, Japan). DNAMAN software was used to design the qRT–PCR primers (Table S1). The efficiency of the qPCR was verified for each primer pair using the cDNA of longan somatic embryos (Figure S2). Additionally, qRT-PCRs were performed on a Roche Lightcyler 480 instrument using SYBR Green chemistry (Hieff qPCR SYBR Green Master Mix, YEASEN, China). DlFeSOD (EU330204) was used as a reference gene for the internal control [120]. The operating parameters of the qRT–PCR were as follows: 95 °C for 30 s, followed by 40 cycles of 95 °C for 10 s and 58 °C for 30 s. The relative expression of the genes involved in SL biosynthesis and signalling in longan was calculated via the 2−ΔΔCt method [121].

4.7. Vector Construction

The SMXL6 coding sequences in longan ECs were amplified by PCR and then inserted into a separate pCAMBIA-1301-35S vector at the BamH I and Sal I restriction sites using the In-Fusion HD Cloning Kit (Takara, CA, USA). The generated pCAMBIA1301-35S-SMXL6 recombinant plasmids were used for longan ECs stable transformation.

4.8. Stable Transformation of Longan Embryogenic Calli

The pCAMBIA1301-35S-SMXL6 constructs were transformed into the Agrobacterium tumefaciens strain (GV3101) following the manufacturer’s instructions (TRANS, Beijing, China). The rapidly proliferating longan ECs were transferred into the bacterial liquid of Agrobacterium tumefaciens (OD600 = 0.7–0.8). After 30 min of infection, the infected ECs were transferred into coculture MS solid medium containing 30 g/L sucrose for three days and then transferred into MS solid medium containing 1.0 mg/L 2,4-D, 20 g/L sucrose and 300 mg/L cefotaxime and harvested after 20 days. The proliferating ECs were transferred into MS solid selection medium containing 1.0 mg/L 2,4-D, 20 g/L sucrose, 300 mg/L cefotaxime, and 20 mg/L hygromycin and subcultured every one month until resistant ECs formed. GUS staining (Huayueyang Biotechnology, Beijing, China) and PCR amplification were used to detect the transformed ECs. For the differentiation of early somatic embryos, longan wild-type and transformed ECs were cultured in MS solid medium containing 300 mg/L cefotaxime, and samples were harvested after 6 and 12 days. The morphological characteristics of transformed somatic embryos were observed by microscopy (Olympus).

4.9. Statistical Analyses

Data were analysed, and graphs were drawn using GraphPad Prism 8 (GraphPad Prism Software, Inc., San Diego, CA, USA). In all graphs, error bars indicate standard deviation, and significant differences are indicated with “*” (p < 0.05), “**” (p < 0.01), “***” (p < 0.005), or “****” (p < 0.001) using Student’s t test.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232214047/s1.

Author Contributions

X.Z., C.L., M.L. and X.X. (Xiaodong Xue) performed the experiments; X.Z., Y.C. (Yan Chen), S.Z., X.X. (Xuechen Xiao), C.L., Z.Z. and Y.C. (Yukun Chen). analysed the data; X.Z., Y.L. and Z.L. wrote the article with contributions from the other authors; all authors contributed to the design of the project. All authors have read and agreed to the published version of the manuscript.

Funding

This work was Supported by the National Natural Science Foundation of China (31672127), Natural Science Foundation of Fujian Province (2020J01543), the Constructions of Plateau Discipline of Fujian Province (102/71201801101), the Technology Innovation Fund of Fujian Agriculture and Forestry University (CXZX2019033S and CXZX2018078).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article Material, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Moise, A.R.; Al-Babili, S.; Wurtzel, E.T. Mechanistic aspects of carotenoid biosynthesis. Chem. Rev. 2014, 114, 164–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Nisar, N.; Li, L.; Lu, S.; Khin, N.C.; Pogson, B.J. Carotenoid metabolism in plants. Mol. Plant 2015, 8, 68–82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Cook, C.E.; Whichard, L.P.; Turner, B.; Wall, M.E.; Egley, G.H. Germination of witchweed (Striga lutea Lour.): Isolation and properties of a potent stimulant. Science 1966, 154, 1189–1190. [Google Scholar] [CrossRef] [PubMed]
  4. Cook, C.E.; Whichard, L.P.; Wall, M.; Egley, G.H.; Coggon, P.; Luhan, P.A.; McPhail, A.T. Germination stimulants. ll. Structure of strigol, a potent seed germination stimulantfor witchweed (Striga lutea). J. Am. Chem. Soc. 1972, 94, 6198–6199. [Google Scholar] [CrossRef]
  5. Umehara, M.; Hanada, A.; Yoshida, S.; Akiyama, K.; Arite, T.; Takeda-Kamiya, N.; Magome, H.; Kamiya, Y.; Shirasu, K.; Yoneyama, K.; et al. Inhibition of shoot branching by new terpenoid plant hormones. Nature 2008, 455, 195–200. [Google Scholar] [CrossRef] [PubMed]
  6. Gomez-Roldan, V.; Fermas, S.; Brewer, P.B.; Puech-Pages, V.; Dun, E.A.; Pillot, J.P.; Letisse, F.; Matusova, R.; Danoun, S.; Portais, J.-C.; et al. Strigolactone inhibition of shoot branching. Nature 2008, 455, 189–194. [Google Scholar] [CrossRef]
  7. Haider, I.; Andreo-Jimenez, B.; Bruno, M.; Bimbo, A.; Flokova, K.; Abuauf, H.; Ruyter-Spira, C. The interaction of strigolactones with abscisic acid during the drought response in rice. J. Exp. Bot. 2018, 69, 2403–2414. [Google Scholar] [CrossRef] [Green Version]
  8. Brewer, P.B.; Koltai, H.; Beveridge, C.A. Diverse roles of strigolactones in plant development. Mol. Plant 2013, 6, 18–28. [Google Scholar] [CrossRef] [Green Version]
  9. Kapulnik, Y.; Koltai, H. Strigolactone involvement in root development, response to abiotic stress, and interactions with the biotic soil environment. Plant Physiol. 2014, 166, 560–569. [Google Scholar] [CrossRef] [Green Version]
  10. Al-Babili, S.; Bouwmeester, H.J. Strigolactones, a novel carotenoid-derived plant hormone. Annu. Rev. Plant Biol. 2015, 66, 161–186. [Google Scholar] [CrossRef]
  11. Waters, M.T.; Gutjahr, C.; Bennett, T.; Nelson, D.C. Strigolactone signaling and evolution. Annu. Rev. Plant Biol. 2017, 68, 291–322. [Google Scholar] [CrossRef] [PubMed]
  12. Brun, G.; Braem, L.; Thoiron, S.; Gevaert, K.; Goormachtig, S.; Delavault, P. Seed germination in parasitic plants: What insights can we expect from strigolactone research. J. Exp. Bot. 2018, 69, 2265–2280. [Google Scholar] [CrossRef] [PubMed]
  13. Kapulnik, Y.; Resnick, N.; Mayzlish-Gati, E.; Kaplan, Y.; Wininger, S.; Hershenhorn, J.; Koltai, H. Strigolactones interact with ethylene and auxin in regulating root-hair elongation in Arabidopsis. J. Exp. Bot. 2011, 62, 2915–2924. [Google Scholar] [CrossRef] [PubMed]
  14. Kohlen, W.; Charnikhova, T.; Lammers, M.; Pollina, T.; Toth, P.; Haider, I.; Pozo, M.J.; de Maagd, R.A.; Ruyter-Spira, C.; Bouwmeester, H.J.; et al. The tomato CAROTENOID CLEAVAGE DIOXYGENASE8 (SlCCD8) regulates rhizosphere signaling, plant architecture and affects reproductive development through strigolactone biosynthesis. New Phytol. 2012, 196, 535–547. [Google Scholar] [CrossRef] [PubMed]
  15. Ueda, H.; Kusaba, M. Strigolactone rcgulates leaf senescence inconcert with ethylene in Arabidopsis. Plant Physiol. 2015, 169, 138–147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Madmon, O.; Mazuz, M.; Kumari, P.; Dam, A.; Ion, A.; Mayzlish-Gati, E.; Belausov, E.; Wininger, S.; Abu-Abied, M.; McErlean, C.S.; et al. Expression of MAX2 under SCARECROW promoter enhances the strigolactone/MAX2 dependent response of Arabidopsis roots to low-phosphate conditions. Planta 2016, 243, 1419–1427. [Google Scholar] [CrossRef]
  17. Beveridge, C.A.; Dun, E.A.; Rameau, C. Pea has its tendrils in branching discoveries spanning a century from auxin to strigolactones. Plant Physiol. 2009, 151, 985–990. [Google Scholar] [CrossRef] [Green Version]
  18. Drummond, R.S.M.; Martinez-Sanchez, N.M.; Janssen, B.J.; Templeton, K.R.; Simons, J.L.; Quinn, B.D.; Karunairetnam, S.; Snowden, K.C. Petunia hybrida CAROTENOID CLEAVAGE DIOXYGENASE7 is involved in the production of negative and positive branching signals in petunia. Plant Physiol. 2009, 151, 1867–1877. [Google Scholar] [CrossRef] [Green Version]
  19. Alder, A.; Jamil, M.; Marzorati, M.; Bruno, M.; Vermathen, M.; Bigler, P.; Ghisla, S.; Bouwmeester, H.; Beyer, P.; Al-Babili, S. The path from β-carotene to carlactone, a strigolactone-like plant hormone. Science 2012, 335, 1348–1351. [Google Scholar] [CrossRef] [Green Version]
  20. Bruno, M.; Al-Babili, S. On the substrate specificity of the rice strigolactone biosynthesis enzyme DWARF27. Planta 2016, 243, 1429–1440. [Google Scholar] [CrossRef]
  21. Lin, H.; Wang, R.X.; Qian, Q.; Yan, M.X.; Meng, X.B.; Fu, Z.M.; Yan, C.Y.; Jiang, B.; Su, Z.; Li, J.Y.; et al. DWARF27, an iron-containing protein required for the biosynthesis of strigolactones, regulates rice tiller bud outgrowth. Plant Cell 2009, 21, 1512–1525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Waters, M.T.; Brewer, P.B.; Bussell, J.D.; Smith, S.M.; Beveridge, C.A. The Arabidopsis ortholog of rice DWARF27 acts upstream of MAX1 in the control of plant development by strigolactones. Plant Physiol. 2012, 159, 1073–1085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Beveridge, C.A.; Symons, G.M.; Turnbull, C.G. Auxin inhibition of decapitation-induced branching is dependent on grafttransmissible signals regulated by genes Rms1 and Rms2. Plant Physiol. 2000, 123, 689–698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Bennett, T.; Sieberer, T.; Willett, B.; Booker, J.; Luschnig, C.; Leyser, O. The Arabidopsis MAX pathway controls shoot branching by regulating auxin transport. Curr. Biol. 2006, 16, 553–563. [Google Scholar] [CrossRef] [Green Version]
  25. Umehara, M.; Hanada, A.; Magome, H.; Takeda-Kamiya, N.; Yamaguchi, S. Contribution of strigolactones to the inhibition of tiller bud outgrowth under phosphate deficiency in rice. Plant Cell Physiol. 2010, 51, 1118–1126. [Google Scholar] [CrossRef] [Green Version]
  26. Shindo, M.; Shimomura, K.; Yamaguchi, S.; Umehara, M. Upregulation of DWARF27 is associated with increased strigolactone levels under sulfur deficiency in rice. Plant Direct 2018, 2, e00050. [Google Scholar] [CrossRef] [Green Version]
  27. Hou, X.; Rivers, J.; León, P.; McQuinn, R.P.; Pogson, B.J. Synthesis and function of apocarotenoid signals in plants. Trends Plant Sci. 2016, 21, 792–803. [Google Scholar] [CrossRef]
  28. Johnson, X.; Brcich, T.; Dun, E.A.; Goussot, M.; Haurogne, K.; Beveridge, C.A.; Rameau, C. Branching genes are conserved across species. Genes controlling a novel signal in pea are coregulated by other long-distance signals. Plant Physiol. 2006, 142, 1014–1026. [Google Scholar] [CrossRef] [Green Version]
  29. Snowden, K.C.; Simkin, A.J.; Janssen, B.J.; Templeton, K.R.; Loucas, H.M.; Simons, J.L.; Karunairetnam, S.; Gleave, A.P.; Clark, D.G.; Klee, H.J. The Decreased apical dominance1/Petunia hybrida CAROTENOID CLEAVAGE DIOXYGENASE8 gene affects branch production and plays a role in leaf senescence, root growth, and flower development. Plant Cell 2005, 17, 746–759. [Google Scholar] [CrossRef] [Green Version]
  30. Simons, J.L.; Napoli, C.A.; Janssen, B.J.; Plummer, K.M.; Snowden, K.C. Analysis of the DECREASED APICAL DOMINANCE genes of petunia in the control of axillary branching. Plant Physiol. 2007, 143, 697–706. [Google Scholar] [CrossRef]
  31. Van Zeijl, A.; Liu, W.; Xias, T.T.; Kohlen, W.; Yang, W.C.; Bisseling, T.; Geurts, R. The strigolactone biosynthesis gene DWARF27 is coopted in rhizobium symbiosis. BMC Plant Biol. 2015, 15, 260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Mach, J. Strigolactones regulate plant growth in Arabidopsis via degradation of the DWARF53-like proteins SMXL6, 7, and 8. Plant Cell 2015, 27, 3022–3023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Stanga, J.P.; Morffy, N.; Nelson, D.C. Functional redundancy in the control of seedling growth by the karrikin signaling pathway. Planta 2016, 243, 1397–1406. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, L.; Xu, Q.; Yu, H.; Ma, H.; Li, X.; Yang, J.; Chu, J.; Xie, Q.; Wang, Y.; Smith, S.M.; et al. Strigolactone and Karrikin Signaling Pathways Elicit Ubiquitination and Proteolysis of SMXL2 to Regulate Hypocotyl Elongation in Arabidopsis. Plant Cell 2020, 32, 2251–2270. [Google Scholar] [CrossRef]
  35. Jiang, L.; Liu, X.; Xiong, G.; Liu, H.; Chen, F.; Wang, L.; Meng, X.; Liu, G.; Yu, H.; Yuan, Y.; et al. DWARF 53 acts as a repressor of strigolactone signaling in rice. Nature 2013, 504, 401–405, Corrigendum 2014, 506, 396. [Google Scholar] [CrossRef] [Green Version]
  36. Zhou, F.; Lin, Q.; Zhu, L.; Ren, Y.; Zhou, K.; Shabek, N.; Wu, F.; Mao, H.; Dong, W.; Gan, L.; et al. D14–SCFD3-dependent degradation of D53 regulates strigolactone signaling. Nature 2013, 504, 406–410, Corrigendum 2016, 532, 402. [Google Scholar] [CrossRef] [Green Version]
  37. Soundappan, I.; Bennett, T.; Morffy, N.; Liang, Y.; Stanga, J.P.; Abbas, A.; Leyser, O.; Nelson, D.C. SMAX1-LIKE/D53 family members enable distinct MAX2-dependent responses to strigolactones and karrikins in Arabidopsis. Plant Cell 2015, 27, 3143–3159. [Google Scholar] [CrossRef]
  38. Wang, L.; Wang, B.; Jiang, L.; Liu, X.; Li, X.; Lu, Z.; Meng, X.; Wang, Y.; Smith, S.M.; Li, J. Strigolactone signaling in Arabidopsis regulates shoot development by targeting D53-like SMXL repressor proteins for ubiquitination and degradation. Plant Cell 2015, 27, 3128–3142. [Google Scholar] [CrossRef] [Green Version]
  39. Song, X.; Lu, Z.; Yu, H.; Shao, G.; Xiong, J.; Meng, X.; Jing, Y.; Liu, G.; Xiong, G.; Duan, J.; et al. IPA1 functions as a downstream transcription factor repressed by D53 in strigolactone signaling in rice. Cell Res. 2017, 27, 1128–1141. [Google Scholar] [CrossRef] [Green Version]
  40. Wang, L.; Wang, B.; Yu, H.; Guo, H.; Lin, T.; Kou, L.; Wang, A.; Shao, N.; Ma, H.; Xiong, G.; et al. Transcriptional regulation of strigolactone signaling in Arabidopsis. Nature 2020, 583, 277–281. [Google Scholar] [CrossRef]
  41. Grzybkowska, D.; Morończyk, J.; Wójcikowska, B.; Gaj, M.D. Azacitidine (5-AzaC)-treatment and mutations in DNA methylase genes affect embryogenic response and expression of the genes that are involved in somatic embryogenesis in Arabidopsis. Plant Growth Regul. 2018, 85, 243–256. [Google Scholar] [CrossRef] [Green Version]
  42. Sugimoto, K.; Gordon, S.P.; Meyerowitz, E.M. Regeneration in plants and animals: Dedifferentiation, trans-differentiation, or just differentiation. Trends Cell Biol. 2011, 21, 212–218. [Google Scholar] [CrossRef] [PubMed]
  43. Lowe, K.; Wu, E.; Wang, N.; Hoerster, G.; Hastings, C.; Cho, M.J.; Scelonge, C.; Lenderts, B.; Chamberlin, M.; Cushatt, J.; et al. Morphogenic Regulators Baby boom and Wuschel Improve Monocot Transformation. Plant Cell 2016, 28, 1998–2015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Jiménez, V.M. Involvement of plant hormones and plant growth regulators on in vitro somatic embryogenesis. Plant Growth Regul. 2005, 47, 91–110. [Google Scholar] [CrossRef]
  45. Bai, B.; Su, Y.H.; Yuan, J.; Zhang, X.S. Induction of somatic embryos in Arabidopsis requires local YUCCA expression mediated by the down-regulation of ethylene biosynthesis. Mol. Plant 2013, 6, 1247–1260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Soriano, M.; Li, H.; Jacquard, C.; Angenent, G.C.; Krochko, J.; Offringa, R.; Boutilier, K. Plasticity in cell division patterns and auxin transport dependency during in vitro embryogenesis in Brassica napus. Plant Cell 2014, 26, 2568–2581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Crawford, S.; Shinohara, N.; Sieberer, T.; Williamson, L.; George, G.; Hepworth, J.; Muller, D.; Domagalska, M.A.; Leyser, O. Strigolactones enhance competition between shoot branches by dampening auxin transport. Development 2010, 137, 2905–2913. [Google Scholar] [CrossRef] [Green Version]
  48. Chatfield, S.P.; Stirnberg, P.; Forde, B.G.; Leyser, O. The hormonal regulation of axillary bud growth in Arabidopsis. Plant J. 2000, 2, 159–169. [Google Scholar] [CrossRef]
  49. Zhang, J.; Mazur, E.; Balla, J.; Gallei, M.; Kalousek, P.; Medved’ová, Z.; Li, Y.; Wang, Y.; Prát, T.; Vasileva, M.; et al. Strigolactones inhibit auxin feedback on PIN-dependent auxin transport canalization. Nat. Commun. 2020, 11, 3508. [Google Scholar] [CrossRef]
  50. Elhiti, M.; Mira, M.M.; So, K.K.Y.; Stasolla, C.; Hebelstrup, K.H. Synthetic Strigolactone GR24 Improves Arabidopsis Somatic Embryogenesis through Changes in Auxin Responses. Plants 2021, 10, 2720. [Google Scholar] [CrossRef]
  51. López-Ráez, J.A.; Kohlen, W.; Charnikhova, T.; Mulder, P.; Undas, A.K.; Sergeant, M.J.; Bouwmeester, H.J. Does abscisic acid affect strigolactone biosynthesis. New Phytol. 2010, 187, 343–354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Nagata, M.; Yamamoto, N.; Shigeyama, T.; Terasawa, Y.; Anai, T.; Sakai, T.; Inada, S.; Arima, S.; Hashiguchi, M.; Akashi, R.; et al. Red/Far Red Light Controls Arbuscular Mycorrhizal Colonization via Jasmonic Acid and Strigolactone Signaling. Plant Cell Physiol. 2015, 56, 2100–2109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Knoll, K.A.; Short, K.C.; Curtis, I.S.; Power, J.B.; Davey, J.B. Shoot regeneration from cultured root explants of spinach (Spinacia oleracea L.): A system for Agrobacterium transformation. Plant Cell Rep. 1997, 17, 96–101. [Google Scholar] [CrossRef] [PubMed]
  54. Thomas, T.D. Efect of sugars, gibberellic acid and abscisic acid on somatic embryogenesis in Tylophora indica (Burm. f.) Merrill. Chin. J. Biotechnol. 2006, 22, 465–471. [Google Scholar] [CrossRef]
  55. Rose, R.J. Somatic Embryogenesis in the Medicago truncatula Model: Cellular and Molecular Mechanisms. Front. Plant Sci. 2019, 10, 267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Ito, S.; Yamagami, D.; Umehara, M.; Hanada, A.; Yoshida, S.; Sasaki, Y.; Yajima, S.; Kyozuka, J.; Ueguchi-Tanaka, M.; Matsuoka, M.; et al. Regulation of Strigolactone Biosynthesis by Gibberellin Signaling. Plant Physiol. 2017, 174, 1250–1259. [Google Scholar] [CrossRef] [PubMed]
  57. Lai, Z.X. Establishment and maintenance of longan embryogenic cell lines. J. Fujian Agric. Univ. 1997, 2, 160–167. [Google Scholar]
  58. Lai, Z.X. Somatic embryogenesis of high frequency from longan embryogenic calli. J. Fujian Agric. Univ. 1997, 26, 271–276. [Google Scholar]
  59. Teng, L.; Han, W.; Fan, X.; Xu, D.; Zhang, X.; Dittami, S.M.; Ye, N. Evolution and expansion of the prokaryote-like lipoxygenase family in the brown alga Saccharina japonica. Front. Plant Sci. 2017, 8, 2018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Lin, Y.L.; Min, J.M.; Lai, R.L.; Wu, Z.Y.; Chen, Y.K.; Yu, L.L.; Cheng, C.Z.; Jin, Y.C.; Tian, Q.L.; Liu, Q.F.; et al. Genome-wide sequencing of longan (Dimocarpus longan Lour.) provides insights into molecular basis of its polyphenol-rich characteristics. Gigascience 2017, 6, 1–14. [Google Scholar] [CrossRef] [Green Version]
  61. Hu, R.; Qi, G.; Kong, Y.; Kong, D.; Gao, Q.; Zhou, G. Comprehensive analysis of NAC domain transcription factor gene family in Populus trichocarpa. BMC Plant Biol. 2010, 10, 145. [Google Scholar] [CrossRef] [Green Version]
  62. Lindemose, S.; O’Shea, C.; Jensen, M.K.; Skriver, K. Structure, function and networks of transcription factors involved in abiotic stress responses. Int. J. Mol. Sci. 2013, 14, 5842–5878. [Google Scholar] [CrossRef]
  63. Koltai, H. Strigolactones activate different hormonal pathways for regulation of root development in response to phosphate growth conditions. Ann. Bot. 2013, 112, 409–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Dun, E.A.; Brewer, P.B.; Beveridge, C.A. Strigolactones: Discovery of the elusive shoot branching hormone. Trends Plant Sci. 2009, 14, 364–372. [Google Scholar] [CrossRef] [PubMed]
  65. Agusti, J.; Herold, S.; Schwarz, M.; Sanchez, P.; Ljung, K.; Dun, E.A.; Brewer, P.B.; Beveridge, C.A.; Sieberer, T.; Sehr, E.M.; et al. Strigolactone signaling is required for auxin-dependent stimulation of secondary growth in plants. Proc. Natl. Acad. Sci. USA 2011, 108, 20242–20247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Beveridge, C.A.; Kyozuka, J. New genes in the strigolactone-related shoot branching pathway. Curr. Opin. Plant Biol. 2010, 13, 34–39. [Google Scholar] [CrossRef]
  67. Boyer, F.D.; Germain, A.D.; Pillot, J.P.; Pouvreau, J.B.; Chen, V.X.; Ramos, S.; Stévenin, A.; Simier, P.; Delavault, P.; Beau, J.M.; et al. Structure-activity relationship studies of strigolactonerelated molecules for branching inhibition in garden pea: Molecule design for shoot branching. Plant Physiol. 2012, 159, 1524–1544. [Google Scholar] [CrossRef] [Green Version]
  68. Brewer, P.B.; Dun, E.A.; Ferguson, B.J.; Rameau, C.; Beveridge, C.A. Strigolactone acts downstream of auxin to regulate bud outgrowth in pea and Arabidopsis. Plant Physiol. 2009, 150, 482–493. [Google Scholar] [CrossRef] [Green Version]
  69. Chevalier, F.; Pata, M.; Nacry, P.; Doumas, P.; Rossignol, M. Effects of phosphate availability on the root system architecture: Large-scale analysis of the natural variation between Arabidopsis accessions. Plant Cell Environ. 2003, 26, 1839–1850. [Google Scholar] [CrossRef] [Green Version]
  70. Moturu, T.R.; Thula, S.; Singh, R.K.; Nodzyński, T.; Vařeková, R.S.; Friml, J.; Simon, S. Molecular evolution and diversification of the SMXL gene family. J. Exp. Bot. 2018, 69, 2367–2378. [Google Scholar] [CrossRef] [Green Version]
  71. Zheng, Y.; Kumar, N.; González, P.; Etxeberria, E. Strigolactones restore vegetative and reproductive developments in Huanglongbing (HLB) affected, greenhouse-grown citrus trees by modulating carbohydrate distribution. Sci. Hortic. 2018, 237, 89–95. [Google Scholar] [CrossRef]
  72. Marzec, M.; Situmorang, A.; Brewer, P.B.; Brąszewska, A. Diverse Roles of MAX1 Homologues in Rice. Genes 2020, 11, 1348. [Google Scholar] [CrossRef] [PubMed]
  73. Vogel, J.T.; Walter, M.H.; Giavalisco, P.; Lytovchenko, A.; Kohlen, W.; Charnikhova, T.; Simkin, A.J.; Goulet, C.; Strack, D.; Bouwmeester, H.J.; et al. SICCD7 controls strigolactone biosynthesis, shoot branching and mycorrhiza-induced apocarotenoid formation in tomato. Plant J. 2010, 61, 300–311. [Google Scholar] [CrossRef] [PubMed]
  74. Koltai, H.; Dor, E.; Hershenhorn, J.; Joel, D.M.; Weininger, S.; Lekalla, S.; Shealtiel, H.; Bhattacharya, C.; Eliahu, E.; Resnick, N.; et al. Strigolactones’ effect on root growth and roothair elongation may be mediated by auxin-efflux carriers. J. Plant Growth Regul. 2010, 29, 129–136. [Google Scholar] [CrossRef]
  75. Yamada, Y.; Umehara, M. Possible roles of strigolactones during leaf senescence. Plants 2015, 4, 664–677. [Google Scholar] [CrossRef] [Green Version]
  76. Decker, E.L.; Alder, A.; Hunn, S.; Ferguson, J.; Lehtonen, M.T.; Scheler, B.; Kerres, K.L.; Wiedemann, G.; Safavi-Rizi, V.; Nordzieke, S.; et al. Strigolactone biosynthesis is evolutionarily conserved, regulated by phosphate starvation and contributes to resistance against phytopathogenic fungi in a moss, Physcomitrella patens. New Phytol. 2017, 216, 455–468. [Google Scholar] [CrossRef] [Green Version]
  77. Lechat, M.M.; Pouvreau, J.B.; Peron, T.; Gauthier, M.; Montiel, G.; Veronesi, C.; Todoroki, Y.; Le Bizec, B.; Monteau, F.; Macherel, D.; et al. PrCYP707Al, an ABA catabolic gene, is a key component of Phelipanche ramosa seed germination in response to the strigolactone analogue GR24. J. Exp. Bot. 2012, 63, 5311–5322. [Google Scholar] [CrossRef]
  78. Toh, S.; Kamiya, Y.; Kawakami, N.; Nambara, E.; Mccourt, P.; Tsuchiya, Y. Thermoinhibition uncovers a role for strigolactones in Arabidopsis seed germination. Plant Cell Physiol. 2012, 53, 107–117. [Google Scholar] [CrossRef] [Green Version]
  79. Li, H.S.; Lyu, Y.M.; Chen, X.H.; Wang, C.Q.; Yao, D.H.; Ni, S.S.; Lin, Y.L.; Chen, Y.K.; Zhang, Z.H.; Lai, Z.X. Exploration of the effect of blue light on functional metabolite accumulation in longan embryonic calli via RNA sequencing. Int. J. Mol. Sci. 2019, 20, 441. [Google Scholar] [CrossRef] [Green Version]
  80. Xie, Y.; Liu, Y.; Ma, M.; Zhou, Q.; Zhao, Y.; Zhao, B.; Wang, B.; Wei, H.; Wang, H. Arabidopsis FHY3 and FAR1 integrate light and strigolactone signaling to regulate branching. Nat. Commun. 2020, 11, 1955. [Google Scholar] [CrossRef] [Green Version]
  81. Chi, C.; Xu, X.; Wang, M.; Zhang, H.; Fang, P.; Zhou, J.; Xia, X.; Shi, K.; Zhou, Y.; Yu, J. Strigolactones positively regulate abscisic acid-dependent heat and cold tolerance in tomato. Hortic. Res. 2021, 8, 237. [Google Scholar] [CrossRef] [PubMed]
  82. Ha, C.V.; Leyva-Gonzalez, M.A.; Osakabe, Y.; Tran, U.T.; Nishiyama, R.; Watanabe, Y.; Tanaka, M.; Seki, M.; Yamaguchi, S.; Dong, N.V.; et al. Positive regulatory role of strigolactone in plant responses to drought and salt stress. Proc. Natl. Acad. Sci. USA 2014, 111, 851–856. [Google Scholar] [CrossRef] [PubMed]
  83. Cai, H.; Chu, Q.; Yuan, L.; Liu, J.; Chen, X.; Chen, F.; Mi, G.; Zhang, F. Identification of quantitative trait loci for leaf area and chlorophyll content in maize (Zea mays) under low nitrogen and low phosphorus supply. Mol. Breed. 2012, 30, 251–266. [Google Scholar] [CrossRef]
  84. Ikeda-Iwai, M.; Satoh, S.; Kamada, H. Establishment of a reproducible tissue culture system for the induction of Arabidopsis somatic embryos. J. Exp. Bot. 2002, 53, 1575–1580. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Fehér, A. Somatic embryogenesis-stress-induced remodeling of plant cell fate. Biochim. Biophys. Acta 2015, 1849, 385–402. [Google Scholar] [CrossRef]
  86. Gaj, M.D. Somatic embryogenesis and plant regeneration in the culture of Arabidopsis thaliana (L.) Heynh. immature zygotic embryos. Methods Mol. Biol. 2011, 710, 257–265. [Google Scholar]
  87. Sorefan, K.; Booker, J.; Haurogne, K.; Goussot, M.; Bainbridge, K.; Foo, E.; Chatfield, S.P.; Ward, S.; Beveridge, C.A.; Rameau, C.; et al. MAX4 and RMS1 are orthologous dioxygenase-like genes that regulate shoot branching in Arabidopsis and pea. Genes Dev. 2003, 17, 1469–1474. [Google Scholar] [CrossRef] [Green Version]
  88. Foo, E.; Bullier, E.; Goussot, M.; Foucher, F.; Rameau, C.; Beveridge, C.A. The branching gene RAMOSUS1 mediates interactions among two novel signals and auxin in pea. Plant Cell 2005, 17, 464–474. [Google Scholar] [CrossRef] [Green Version]
  89. Arite, T.; Iwata, H.; Ohshima, K.; Maekawa, M.; Nakajima, M.; Kojima, M.; Sakakibara, H.; Kyozuka, J. DWARF10, an RMS1/MAX4/DAD1 ortholog, controls lateral bud outgrowth in rice. Plant J. 2007, 51, 1019–1029. [Google Scholar] [CrossRef]
  90. Hayward, A.; Stirnberg, P.; Beveridge, C.; Leyser, O. Interactions between auxin and strigolactone in shoot branching control. Plant Physiol. 2009, 151, 400–412. [Google Scholar] [CrossRef]
  91. Pinto, R.T.; Freitas, N.C.; Máximo, W.P.F.; Cardoso, T.B.; Prudente, D.O.; Paiva, L.V. Genome-wide analysis, transcription factor network approach and gene expression profile of GH3 genes over early somatic embryogenesis in Coffea spp. BMC Genom. 2019, 20, 812. [Google Scholar] [CrossRef] [PubMed]
  92. Tuteja, N. Abscisic acid and abiotic stress signaling. Plant Signal. Behav. 2007, 2, 135–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Frey, A.; Godin, B.; Bonnet, M.; Sotta, B.; Marion-Poll, A. Maternal synthesis of abscisic acid controls seed development and yield in Nicotiana plumbaginifolia. Planta 2004, 218, 958–964. [Google Scholar] [CrossRef] [PubMed]
  94. Langhansová, L.; Konrádová, H.; Vaněk, T. Polyethylene glycol and abscisic acid improve maturation and regeneration of Panax ginseng somatic embryos. Plant Cell Rep. 2004, 22, 725–730. [Google Scholar] [CrossRef] [PubMed]
  95. Zur, I.; Dubas, E.; Golemiec, E.; Szechynska-Hebda, M.; Janowiak, F.; Wedzony, M. Stress-induced changes important for effective androgenic induction in isolated microspore culture of triticale (×Triticosecale Wittm.). Plant Cell Tissue Organ Cult. 2008, 94, 319–328. [Google Scholar] [CrossRef]
  96. Akula, A.; Akula, C.; Bateson, M.F. Betaine a novel candidate for rapid induction of somatic embryogenesis in tea (Camellia sinensis (L.) O. Kuntze). Plant Growth Regul. 2000, 30, 241–246. [Google Scholar] [CrossRef]
  97. Brun, G. At the crossroads of strigolactones and abscisic acid pathways: A role for miR156. Plant Cell Environ. 2020, 43, 1609–1612. [Google Scholar] [CrossRef]
  98. Bu, Q.; Lv, T.; Shen, H.; Luong, P.; Wang, J.; Wang, Z.; Huang, Z.; Xiao, L.; Engineer, C.; Kim, H.T.; et al. Regulation of drought tolerance by the F-box protein MAX2 in arabidopsis. Plant Physiol. 2014, 164, 424–439. [Google Scholar] [CrossRef] [Green Version]
  99. Liu, J.; He, H.; Vitali, M.; Visentin, I.; Charnikhova, T.; Haider, I.; Cardinale, F. Osmotic stress represses strigolactone biosynthesis in Lotus japonicus roots: Exploring the interaction between strigolactones and ABA under abiotic stress. Planta 2015, 241, 1435–1451. [Google Scholar] [CrossRef]
  100. Jiu, S.; Xu, Y.; Xie, X.; Wang, J.; Xu, J.; Liu, X.; Sun, W.; Xu, W.; Wang, S.; Zhang, C. Strigolactones affect the root system architecture of cherry rootstock by mediating hormone signaling pathways. Environ. Exp. Bot. 2022, 193, 104667. [Google Scholar] [CrossRef]
  101. Ishizaki, T.; Komai, F.; Masuda, K. Screening for strongly regenerative genotypes of spinach in tissue culture using subcultured root explants. Plant Cell Tissue Organ Cult. 2001, 67, 251–255. [Google Scholar] [CrossRef]
  102. Nguyen, Q.V.; Sun, H.J.; Boo, K.H.; Lee, D.; Lee, J.H.; Lim, P.O.; Lee, H.Y.; Riu, K.Z.; Lee, D.S. Efect of plant growth regulator combination and culture period on in vitro regeneration of spinach (Spinacia oleracea L.). Plant Biotechnol. Rep. 2013, 7, 99–108. [Google Scholar] [CrossRef]
  103. Wang, H.; Caruso, L.V.; Downie, A.B.; Perry, S.E. The embryo MADS domain protein AGAMOUS-Like 15 directly regulates expression of a gene encoding an enzyme involved in gibberellin metabolism. Plant Cell 2004, 16, 1206–1219. [Google Scholar] [CrossRef] [PubMed]
  104. Belić, M.; Zdravković-Korać, S.; Janošević, D.; Savić, J.; Todorović, S.; Banjac, N.; Milojević, J. Gibberellins and light synergistically promote somatic embryogenesis from the in vitro apical root sections of spinach. Plant Cell Tissue Organ Cult. 2020, 142, 537–548. [Google Scholar] [CrossRef]
  105. Ni, J.; Gao, C.; Chen, M.S.; Pan, B.Z.; Ye, K.; Xu, Z.F. Gibberellin Promotes Shoot Branching in the Perennial Woody Plant Jatropha curcas. Plant Cell Physiol. 2015, 56, 1655–1666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Ni, J.; Zhao, M.L.; Chen, M.S.; Pan, B.Z.; Tao, Y.B.; Xu, Z.F. Comparative Transcriptome Analysis of Axillary Buds in Response to the Shoot Branching Regulators Gibberellin A3 and 6-Benzyladenine in Jatropha curcas. Sci. Rep. 2017, 7, 11417. [Google Scholar] [CrossRef] [Green Version]
  107. De Saint Germain, A.; Ligerot, Y.; Dun, E.A.; Pillot, J.P.; Ross, J.J.; Beveridge, C.A.; Rameau, C. Strigolactones stimulate internode elongation independently of gibberellins. Plant Physiol. 2013, 163, 1012–1025. [Google Scholar] [CrossRef] [Green Version]
  108. Preston, C.A.; Betts, H.; Baldwin, I.T. Methyl jasmonate as an allelopathic agent: Sagebrush inhibits germination of a neighboring tobacco, Nicotiana attenuata. J. Chem. Ecol. 2002, 28, 2343–2369. [Google Scholar] [CrossRef]
  109. Bialecka, B.; Kepczynsky, J. Regulation of a-amylase activity in Amaranthus caudatus seeds by methyl jasmonate, gibberellin A3, benzyladenine and ethylene. Plant Growth Regul. 2003, 39, 51–56. [Google Scholar] [CrossRef]
  110. Lahari, Z.; Ullah, C.; Kyndt, T.; Gershenzon, J.; Gheysen, G. Strigolactones enhance root-knot nematode (Meloidogyne graminicola) infection in rice by antagonizing the jasmonate pathway. New Phytol. 2019, 224, 454–465. [Google Scholar] [CrossRef] [Green Version]
  111. Ruyter-Spira, C.; Kohlen, W.; Charnikhova, T.; van Zeijl, A.; van Bezouwen, L.; de Ruijter, N.; Cardoso, C.; Lopez-Raez, J.A.; Matusova, R.; Bours, R.; et al. Physiological effects of the synthetic strigolactone analog GR24 on root system architecture in Arabidopsis: Another belowground role for strigolactones. Plant Physiol. 2011, 155, 721–734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Mayzlish-Gati, E.; De-Cuyper, C.; Goormachtig, S.; Beeckman, T.; Vuylsteke, M.; Brewer, P.B.; Beveridge, C.A.; Yermiyahu, U.; Kaplan, Y.; Enzer, Y.; et al. Strigolactones are involved in root response to low phosphate conditions in Arabidopsis. Plant Physiol. 2012, 160, 1329–1341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Ayil-Gutiérrez, B.; Galaz-Ávalos, R.M.; Peña-Cabrera, E.; Loyola-Vargas, V.M. Dynamics of the concentration of IAA and some of its conjugates during the induction of somatic embryogenesis in Coffea canephora. Plant Signal. Behav. 2013, 8, e26998. [Google Scholar] [CrossRef] [PubMed]
  114. Chen, C.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.; Xia, R. TBtools: An integrative toolkit developed for interactive analyses of big biological data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef] [PubMed]
  115. Gong, X.; Dou, F.; Cheng, X.; Zhou, J.; Zou, Y.; Ma, F. Genome-wide identification of genes involved in polyamine biosynthesis and the role of exogenous polyamines in Malus hupehensis Rehd. under alkaline stress. Gene 2018, 669, 52–62. [Google Scholar] [CrossRef]
  116. Chou, K.C.; Shen, H.B. Cell-PLoc: A package of Web servers for predicting subcellular localization of proteins in various organisms. Nat. Protoc. 2008, 3, 153–162. [Google Scholar] [CrossRef]
  117. 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]
  118. Wang, D.P.; Zhang, Y.B.; Zhang, Z.; Zhu, J.; Yu, J. KaKs Calculator 2.0: A toolkit incorporating gamma-series methods and sliding window strategies. Genom. Proteom. Bioinform. 2010, 8, 77–80. [Google Scholar] [CrossRef] [Green Version]
  119. Subramanian, B.; Gao, S.; Lercher, M.J.; Hu, S.; Chen, W. Evolview v3: A web server for visualization, annotation, and management of phylogenetic trees. Nucleic Acids Res. 2019, 47, 270–275. [Google Scholar] [CrossRef]
  120. Xu, X.; Chen, X.; Shen, X.; Chen, R.; Zhu, C.; Zhang, Z.; Chen, Y.; Lin, W.; Xu, X.; Lin, Y.; et al. Genome-wide identification and characterization of DEAD-box helicase family associated with early somatic embryogenesis in Dimocarpus longan Lour. J. Plant Physiol. 2021, 258–259, 153364. [Google Scholar] [CrossRef]
  121. Livak, K.J.; Schmittgen, T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods 2001, 25, 402–408. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Collinear analysis of DlCCD, DlD27, and DlSMXL family members. (A) Collinearity analysis within longan. Gray lines indicate the collinear blocks within the longan genome, the red line represent the syntenic gene pairs relating to DlCCD, DlD27, and DlSMXL family genes. (B) Colinear analysis between longan and Arabidopsis. Gray lines indicate the collinear blocks within the longan and Arabidopsis, and the red lines represent the syntenic gene pairs relating to SL biosynthesis.
Figure 1. Collinear analysis of DlCCD, DlD27, and DlSMXL family members. (A) Collinearity analysis within longan. Gray lines indicate the collinear blocks within the longan genome, the red line represent the syntenic gene pairs relating to DlCCD, DlD27, and DlSMXL family genes. (B) Colinear analysis between longan and Arabidopsis. Gray lines indicate the collinear blocks within the longan and Arabidopsis, and the red lines represent the syntenic gene pairs relating to SL biosynthesis.
Ijms 23 14047 g001
Figure 2. Phylogenetic tree of CCD, SMXL, and D27 family genes. The different-colored areas indicate different groups.
Figure 2. Phylogenetic tree of CCD, SMXL, and D27 family genes. The different-colored areas indicate different groups.
Ijms 23 14047 g002
Figure 3. CRE analysis of the promoter region of DlCCD, DlSMXL, and DlD27 genes. (A) DlCCD, DlSMXL, and DlD27 family members were arranged in the order of phylogenetic tree. The names of CREs were marked on the bottom. The heatmap shows the number of main CREs corresponding to each member. (B) CREs were classified into four categories: stress response, growth and development, light response, and phytohormone response. The bar chart counts all the number of four types of CREs in each member, and different types of CREs are indicated by different colors. (C) Network diagram of predicted transcription factor binding sites with DlCCD, DlSMXL, and DlD27 family members. Gray connecting lines indicate the presence of interactions.
Figure 3. CRE analysis of the promoter region of DlCCD, DlSMXL, and DlD27 genes. (A) DlCCD, DlSMXL, and DlD27 family members were arranged in the order of phylogenetic tree. The names of CREs were marked on the bottom. The heatmap shows the number of main CREs corresponding to each member. (B) CREs were classified into four categories: stress response, growth and development, light response, and phytohormone response. The bar chart counts all the number of four types of CREs in each member, and different types of CREs are indicated by different colors. (C) Network diagram of predicted transcription factor binding sites with DlCCD, DlSMXL, and DlD27 family members. Gray connecting lines indicate the presence of interactions.
Ijms 23 14047 g003
Figure 4. The expression profiling of DlCCD, DlSMXL, and DlD27 genes detected in different tissues and in response to different treatments. (A) Transcriptome data of DlCCD, DlSMXL, and DlD27 genes of six organs in longan. (B) Transcriptome data of DlCCD, DlSMXL, and DlD27 genes in the ECs under different light quality treatments for 25 d. (C) Transcriptome data of DlCCD, DlSMXL, and DlD27 genes in the ECs under different temperature treatments for 24 h.
Figure 4. The expression profiling of DlCCD, DlSMXL, and DlD27 genes detected in different tissues and in response to different treatments. (A) Transcriptome data of DlCCD, DlSMXL, and DlD27 genes of six organs in longan. (B) Transcriptome data of DlCCD, DlSMXL, and DlD27 genes in the ECs under different light quality treatments for 25 d. (C) Transcriptome data of DlCCD, DlSMXL, and DlD27 genes in the ECs under different temperature treatments for 24 h.
Ijms 23 14047 g004
Figure 5. (A) Roles of CCD, D27, and SMXL genes in SLs pathway. (B) FPKM values of members of DlCCD, DlSMXL, and DlD27 family in longan SE. (C) Relative expression levels of DlCCD, DlSMXL, and DlD27 family in longan SE. (D) Changes in the endogenous SL content in longan SE. “****” (p < 0.001).
Figure 5. (A) Roles of CCD, D27, and SMXL genes in SLs pathway. (B) FPKM values of members of DlCCD, DlSMXL, and DlD27 family in longan SE. (C) Relative expression levels of DlCCD, DlSMXL, and DlD27 family in longan SE. (D) Changes in the endogenous SL content in longan SE. “****” (p < 0.001).
Ijms 23 14047 g005
Figure 6. (A) The gene expression analysis and SL content changes of longan ECs treated with different hormones for 24 h. (B) Changes in the endogenous SL content of longan ECs under hormones treatment. (C) Schematic diagram of changes in the content of SLs and expression analysis of SL pathway genes respond to four hormone treatments for 24 h by qRT-PCR in longan ECs. Red and green mark indicate upregulated and downregulated genes. The arrow indicates promoting the process. The “T” line indicates inhibiting the process. Significant differences are indicated with “*” (p < 0.05), “**” (p < 0.01), “***” (p < 0.005).
Figure 6. (A) The gene expression analysis and SL content changes of longan ECs treated with different hormones for 24 h. (B) Changes in the endogenous SL content of longan ECs under hormones treatment. (C) Schematic diagram of changes in the content of SLs and expression analysis of SL pathway genes respond to four hormone treatments for 24 h by qRT-PCR in longan ECs. Red and green mark indicate upregulated and downregulated genes. The arrow indicates promoting the process. The “T” line indicates inhibiting the process. Significant differences are indicated with “*” (p < 0.05), “**” (p < 0.01), “***” (p < 0.005).
Ijms 23 14047 g006
Figure 7. (A) Morphological change in embryogenic callus under GR24 and Tis108 treatment. The appearance globular embryo is marked by red arrows. (BE) Changes in the endogenous IAA, ABA, GA, and JA content of longan somatic embryos upon GR24 and Tis108 treatment. Significant differences are indicated with “*” (p < 0.05), “**” (p < 0.01), “***” (p < 0.005), or “****” (p < 0.001).
Figure 7. (A) Morphological change in embryogenic callus under GR24 and Tis108 treatment. The appearance globular embryo is marked by red arrows. (BE) Changes in the endogenous IAA, ABA, GA, and JA content of longan somatic embryos upon GR24 and Tis108 treatment. Significant differences are indicated with “*” (p < 0.05), “**” (p < 0.01), “***” (p < 0.005), or “****” (p < 0.001).
Ijms 23 14047 g007
Figure 8. Morphological characteristics and physiological indexes determination of DlSMXL6 overexpressing cell lines. (A) The expression levels of DlSMXL6 in longan transgenic cell lines. (BF) Phenotypical comparisons of wild-type (WT), transgenic 1301-GUS, and DlSMXL6-OEs. 1301: expression vector; bar = 100 µm; the recombinant vector constructed of 1301-DlSMXL6: GUS; (GI) carotenoid, SLs, and IAA content changes of wild-type (WT), transgenic 1301-GUS and DlSMXL6-OEs; (JL) the expression levels of DlYUC3, DlYUC5 and DlYUC10 in longan transgenic cell lines. Significant differences are indicated with “*” (p < 0.05), “**” (p < 0.01), “***” (p < 0.005), or “****” (p < 0.001).
Figure 8. Morphological characteristics and physiological indexes determination of DlSMXL6 overexpressing cell lines. (A) The expression levels of DlSMXL6 in longan transgenic cell lines. (BF) Phenotypical comparisons of wild-type (WT), transgenic 1301-GUS, and DlSMXL6-OEs. 1301: expression vector; bar = 100 µm; the recombinant vector constructed of 1301-DlSMXL6: GUS; (GI) carotenoid, SLs, and IAA content changes of wild-type (WT), transgenic 1301-GUS and DlSMXL6-OEs; (JL) the expression levels of DlYUC3, DlYUC5 and DlYUC10 in longan transgenic cell lines. Significant differences are indicated with “*” (p < 0.05), “**” (p < 0.01), “***” (p < 0.005), or “****” (p < 0.001).
Ijms 23 14047 g008
Figure 9. (A) The regulation network among GR24 and Tis108 with endogenous hormones during longan somatic embryo development. (B) The regulatory network composed of DlSMXL6 during longan somatic embryo development. The arrow indicates promoting the process. The “T” line indicates inhibiting the process.
Figure 9. (A) The regulation network among GR24 and Tis108 with endogenous hormones during longan somatic embryo development. (B) The regulatory network composed of DlSMXL6 during longan somatic embryo development. The arrow indicates promoting the process. The “T” line indicates inhibiting the process.
Ijms 23 14047 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, X.; Lai, C.; Liu, M.; Xue, X.; Zhang, S.; Chen, Y.; Xiao, X.; Zhang, Z.; Chen, Y.; Lai, Z.; et al. Whole Genome Analysis of SLs Pathway Genes and Functional Characterization of DlSMXL6 in Longan Early Somatic Embryo Development. Int. J. Mol. Sci. 2022, 23, 14047. https://doi.org/10.3390/ijms232214047

AMA Style

Zhang X, Lai C, Liu M, Xue X, Zhang S, Chen Y, Xiao X, Zhang Z, Chen Y, Lai Z, et al. Whole Genome Analysis of SLs Pathway Genes and Functional Characterization of DlSMXL6 in Longan Early Somatic Embryo Development. International Journal of Molecular Sciences. 2022; 23(22):14047. https://doi.org/10.3390/ijms232214047

Chicago/Turabian Style

Zhang, Xueying, Chunwang Lai, Mengyu Liu, Xiaodong Xue, Shuting Zhang, Yan Chen, Xuechen Xiao, Zihao Zhang, Yukun Chen, Zhongxiong Lai, and et al. 2022. "Whole Genome Analysis of SLs Pathway Genes and Functional Characterization of DlSMXL6 in Longan Early Somatic Embryo Development" International Journal of Molecular Sciences 23, no. 22: 14047. https://doi.org/10.3390/ijms232214047

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