Next Article in Journal
Polyphenol Composition and Antioxidant Activity of Tapirira guianensis Aubl. (Anarcadiaceae) Leaves
Next Article in Special Issue
Flowering Mechanisms and Environmental Stimuli for Flower Transition: Bases for Production Scheduling in Greenhouse Floriculture
Previous Article in Journal
Evaluation of Pseudomonas fulva PS9.1 and Bacillus velezensis NWUMFkBS10.5 as Candidate Plant Growth Promoters during Maize-Fusarium Interaction
Previous Article in Special Issue
SNP- and Haplotype-Based GWAS of Flowering-Related Traits in Brassica napus
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identification and Functional Characterization of FLOWERING LOCUS T in Platycodon grandiflorus

1
Department of Industrial Plant Science and Technology, Chungbuk National University, Cheongju 28644, Korea
2
Plant Molecular Biology and Biotechnology Research Center, Gyeongsang National University, Jinju 52828, Korea
*
Authors to whom correspondence should be addressed.
Plants 2022, 11(3), 325; https://doi.org/10.3390/plants11030325
Submission received: 8 October 2021 / Revised: 21 January 2022 / Accepted: 21 January 2022 / Published: 26 January 2022
(This article belongs to the Special Issue Control of Flowering Time and Their Environmental Regulation in Crops)

Abstract

:
Platycodon grandiflorus roots have been used as a foodstuff and traditional medicine for thousands of years in East Asia. In order to increase the root development of P. grandiflorus, cultivators removed the inflorescences, suggesting the possible negative effect of flowering on root development. This indicates that the genetic improvement of P. grandiflorus by late flowering is a potential approach to increase productivity. However, nothing is known about key genes integrating multiple flowering pathways in P. grandiflorus. In order to fill this gap, we identified potential homologs of the FLOWERING LOCUS T (FT) gene in P. grandiflorus. The alignment with other FT members and phylogenetic analysis revealed that the P. grandiflorus FT (PlgFT) protein contains highly conserved functional domains and belongs to the FT-like clade. The expression analysis revealed spatial variations in the transcription of PlgFT in different organs. In addition, the expression level of PlgFT was increased by high temperature but not by photoperiodic light input signals, presumably due to lacking the CONSTANS binding motif in its promoter region. Furthermore, PlgFT induced early flowering upon its overexpression in P. grandiflorus, suggesting the functional role of PlgFT in flowering. Taken together, we functionally characterized PlgFT as a master regulator of P. grandiflorus flowering under inductive high temperature, which will serve as an important target gene for improving the root productivity.

1. Introduction

The transition from vegetative to the reproductive stage is a critical event in plant development. The floral transition is controlled by diverse endogenous and environmental cues such as plant age, autonomous pathway, gibberellic acid (GA), photoperiod, vernalization, and ambient temperature change [1,2]. In Arabidopsis, the master regulator of floral transition, FLOWERING LOCUS T (FT) protein, is synthesized in the phloem companion cell of the leaf, translocated into a sieve tube system, and moves to shoot apical meristem (SAM) to systematically reprogram the floral gene network under the inductive photoperiod [3,4]. At the Arabidopsis SAM, FT physically interacts with basic leucine-zipper (bZIP)-type transcription factor FLOWERING LOCUS D (FD) and 14-3-3 proteins. These interactors form florigen activation complex (FAC) in Arabidopsis to promote the floral meristem identity genes such as APETALA1 (AP1) [5] and SUPPRESSOR OF OVEREXPRESSION OF CONSTANS1 (SOC1) [6] either directly or indirectly. In addition, CONSTANS (CO) functions as the master regulator of photoperiodic induction of FT expression under long-day (LD) and short-day (SD) conditions [3,7,8,9], whereas the flowering of day-neutral plants is not affected by the CO-FT module [3,10,11]. Currently, ambient temperature-responsive mechanisms have also been reported as regulating FT-mediated flower induction [12]. In Arabidopsis, PHYTOCHROME INTERACTING FACTOR4 (PIF4) directly binds the FT promoter to activate FT at high temperature, resulting in the early flowering under short-day conditions [13]. In addition, the Arabidopsis FT homolog FT-like (Narcissus FLOWERING LOCUS T1) NFT1 gene was up-regulated by heat treatment (30 °C) and promoted flower initiation [14]. These findings indicate the importance of ambient temperature to induce flowering.
Platycodon grandiflorus, a day-neutral and herbaceous perennial plant, is mainly cultivated for its storage roots. The roots of P. grandiflorus were used as traditional herbal medicine and functional foodstuff in East Asia due to their considerable bioactive components, such as triterpenoid saponin [15]. In P. grandiflorus, the transition from vegetative to reproductive growth has been known to have an inhibitory effect on storage root development, presumably due to the consumption of accumulated energy source (starch) from their storage root and delayed energy distribution to the bidirectional source to sink flow for flowering [16,17]. This indicates that delayed flowering time should affect root development via the availability of more time for vegetative growth. It was shown that the flowering timing of P. grandiflorus was not affected by photoperiod and gibberellin together but was significantly accelerated by high temperature [18]. Similarly, the day-natural tomato flowering is regulated by the FT-homologue SFT/SP3D gene, which is regulated independently of CO and day length [11,19]. In plants, the flowering transition and inflorescence architecture are modulated by eukaryotic phosphatidylethanolamine-binding protein family, which contains TWIN SISTER OF FT (TSF), TERMINAL FLOWER 1 (TFL1), ARABIDOPSIS THALIANA CENTRORADIALIS (ATC), BROTHER OF FT AND TFL1 (BFT), and MOTHER OF FT AND TFL1 (MFT) [20]. Among them, FT functions as the mobile flower-promoting signal, which promotes the transition to reproductive development and flowering, whereas TFL1 plays as a repressor of this transition [21]. Although MFT also functions as a floral inducer [22], the core determinant gene for floral transition is assumed to be FT in higher plants [23]. However, in P. grandiflorus, the information of multiple flowering pathways is completely unknown due to the missing genetic information about floral inducers such as FT.
Therefore, in this study, we identified P. grandiflorus FT (PlgFT) and analyzed the expression pattern of PlgFT in response to gibberellin and high temperature (30 °C) to investigate its role in the flowering mechanism of P. grandiflorus. Overexpression of PlgFT resulted in the induction of early flowering in P. grandiflorus, which suggests that PlgFT is potentially a gene resource for improving storage root development using the plant molecular breeding approach in this day-neutral, economically important medicinal plant.

2. Results and Discussion

2.1. Identification of FT Gene in P. grandiflorus

FT is well known as a master flowering regulator, and the amino acid sequences are highly conserved in many flowering plants [23]. In order to investigate the FT-like gene integrating multiple flowering pathways in P. grandiflorus, an FT-like gene was identified in the P. grandiflorus genome data [24] using the sequences of FTs from various species and was cloned for functional characterization. The PlgFT gene was found to have an open reading frame of 555 bp with a stop codon, encoding a protein of 184 amino acids (Figure 1). The PlgFT protein conserved the PEBP domain (DPDxP and GxHR) and key amino acid residues (Gln139 (Q139) and Tyr84 (Y84)), similar to Arabidopsis FT [25]. In addition, segment A (LGRQTVYAPGWRQN), which forms an external loop and is required for FT proteins to act as a floral activator [26], and LYN, which is required for FT full functionality with segment A [25], were found in PlgFT protein. In flowering pathways, FT and TFL1 act as flowering regulators with homology to phosphatidylethanolamine-binding proteins [20,21]. Although Arabidopsis FT exhibits protein sequence similarity of ~60% with Arabidopsis TFL1, they function in an opposite manner [21,27]. Based on the sequence alignment of FT-like proteins, flowering inducer and repressor functions in FT-like proteins were distinguished by two conserved sites at positions 134 and 138 (based on Arabidopsis FT protein sequence) [21]. As shown in Figure 1, PlgFT protein contains tyrosine at position 134 and tryptophan at position 138, similar to all flowering inducer FTs, indicating that putative PlgFT protein belongs to the flowering inducer FT.
In order to explore the phylogenic relationships between the FT-like proteins from various species, a phylogenetic tree was constructed using the neighbor-joining method. As shown in Figure S1, PlgFT protein belongs to the FT clade but is not grouped into TFL clades. In addition, PlgFT is located genetically close to GtFT and StSP3D, which are well-known flowering inducers [28,29], collectively suggesting that PlgFT could be a functional ortholog of FT, which acts as a flowering inducer.
The analysis of protein subcellular localization is important for further understanding its function. In order to investigate the subcellular localization of PlgFT in plants, we tested a transient expression system using an Agro-infiltration approach. The green fluorescence protein (GFP) fused to PlgFT was expressed under the control of the 35S promoter. As shown in Figure 2A, PlgFT-GFP was localized in both the nucleus and cytoplasm. This result was similar to the subcellular localization pattern of previously studied FT proteins in various plants, including switchgrass, sweet cherry, alfalfa, carrot, tobacco, and Eriobotrya deflexa [30,31,32,33,34,35].
FT and FD are known to be interdependent partners in protein interaction for their transcriptional regulation, then FT-FD complex binds and promotes the expression of floral identity genes in SAM [5]. Therefore, we performed a yeast two-hybrid assay to test whether PlgFT can interact with FD. Arabidopsis FD (AtFD) and PlgFT were used as prey and bait, respectively. As shown in Figure 2B, the transformed yeast expressing PlgFT and AtFD grew in SD medium without Leu, Trp, and His, indicating that PlgFT directly interacts with AtFD.

2.2. Expression Pattern of PlgFT in Various Organs

The analysis of organ-specific transcription can be used to determine the potential role of a gene of interest. Although the expression of Arabidopsis FT was detected at the vascular tissue (phloem) of cotyledons and leaves [36], the transcript levels of FTs identified from various plants were detected in various tissues, including leaf, flower, fruit, and root [33,37,38]. Therefore, to evaluate the expression profile of PlgFT, the transcripts of PlgFT were monitored by RT-qPCR in various organs of P. grandiflorus, including the leaves, stems, flowers, and roots. As shown in Figure 3A, PlgFT was highly expressed in unexpanded young leaves compared to fully expanded leaves and senescent leaves. In four-month-old plants, PlgFT was highly expressed in the root, followed by unexpanded young leaves and stems. In onion (Allium cepa), FT genes are not only involved in flowering but also in the regulation of onion bulb formation [26,39]. In addition, flowering and tuberization are controlled by StSP6A (FT ortholog of potato) in potato [29], indicating that FT also acts as a key mobile signal controlling different organ developments. Therefore, based on the expression pattern of PlgFT, it can be hypothesized that PlgFT has a distinct role in storage root development, which requires further investigation.
In the flower development stages, PlgFT was highly expressed in the flower bud compared with the flower and wilted flower. In addition, Eriobotrya deflexa FT [33], Jatropha curcas FT [40], and grapevine FT [41] are primarily expressed in the reproductive organs rather than leaves, similar to PlgFT. Taken together, the expression pattern of PlgFT indicates that PlgFT might play a role in both flower and storage root.

2.3. The Diurnal Expression Pattern of PlgFT

During photoperiodic flowering, CO integrates multiple upstream signals to respond to the photoperiod for the optimal flowering time. In both Arabidopsis and rice, the expression of CO and CO ortholog Hd1 have distinct diurnal expression profiles [3,8,9], indicating that expression of FT is controlled by the circadian clock. In order to investigate the diurnal expression pattern of PlgFT, the transcription level of PlgFT was investigated over a 24 h period. PlgFT exhibited similar expression levels throughout the LD or SD-diurnal time course (Figure 4), indicating that PlgFT is not diurnally regulated.
The circadian clock regulation of CO transcription is crucial for the expression of FT [4]. In the Arabidopsis FT promoter region, two CO-responsive elements (CORE1 and CORE1) with consensus sequence TGTG(N2-3)ATG were identified [42]. The mutation in the CORE1 reduced the expression of FT and resulted in later flowering. In addition, the late-flowering phenotype of ft-10 plants was not recovered by FT under the control of the native promoter lacking the COREs [7], which indicates that CO-mediated FT expression is mediated by COREs. However, as shown in Figure S2, the PlgFT promoter region did not contain a consensus sequence corresponding to COREs [7]. Taken together, the absence of the CO binding motif in the promoter region led to the non-responsiveness of PlgFT expression under light conditions.

2.4. Effects of High Temperature and GA3 on PlgFT Expression

As described above, the floral transition is controlled not only by the photoperiod but also by endogenous and environmental cues, including ambient temperature change and phytohormone GA. In P. grandiflorus, flowering is accelerated at elevated temperatures, whereas the treatment of GA3 does not affect flowering [18,43], indicating the importance of ambient temperature to P. grandiflorus flowering. In order to further support the effect of high temperature on P. grandiflorus flowering, variation in PlgFT expression was determined after the treatment with high temperature or GA3. As shown in Figure 5, the GA3 application did not change the expression level of PlgFT, whereas the high-temperature treatment (30 °C) strongly induced PlgFT expression (2.5-fold) compared to the control plant (22 °C). The increment of PlgFT expression continued for seven days by high-temperature treatment, indicating that the high-temperature flowering response is mediated by the induction of PlgFT in P. grandiflorus. In higher plants, the levels of FT transcription correlated with temperature effects on floral initiation. The expression of FaFT3 (Fragaria × ananassa FT3) increased by low temperature, in accordance with the result that these treatments promoted floral initiation, whereas high temperature inhibited the expression of FaFT3 and floral initiation [44]. However, contrary to vernalization, high temperature provided a flowering signal by increasing NtFT (Narcissus FT gene homolog) expression, whereas the down-regulation of NtFT by low temperature inhibited flower induction [45]. In addition, high-temperature promoted soybean flowering through the induction of soybean floral activators (soybean FT homologs GmFT2a and GmFT5a) and the suppression of floral repressors [46]. Under high-temperature conditions, the histone variant H2A.Z plays a pivotal role in temperature-dependent FT transcription by PHYTOCHROME INTERACTING FACTOR 4 (PIF4) and key thermosensory bHLH transcription factor by improving the accessibility of PIF4 to the FT promoter [4]. Our data revealed that high-temperature functions as an activator of PlgFT expression, but further studies are remained to identify and characterize the high temperature-mediated H2A.Z-PIF4 regulatory module in P. grandiflorus.

2.5. Ectopic Overexpression of PlgFT in P. grandiflorus Accelerates flowering

In order to explore the biological function of PlgFT on flowering in P. grandiflorus, we generated transgenic P. grandiflorus overexpressing PlgFT under the control of the 35S promoter. We obtained several independent transgenic lines and checked the presence of PlgFT-GFP transgene by reverse transcription (RT)-PCR (Figure 6A). In addition, we selected regenerated plants that survived on hygromycin selection medium but showed no transcription level of PlgFT-GFP as transgenic controls (TCs). Overexpression of PlgFT-GFP induced extreme early flowering compared to the TC in both the medium and soil growth condition (Figure 6B,C). In addition, PlgFT-GFP overexpressing plants exhibited dwarfism and the early leaf senescence phenotype; this should be mediated by the consequence of growth deficiency due to the short vegetative growth period and premature transition to the reproductive phase. In storage root plants, including cassava, photoassimilate translocation from leaves to roots is very important for the bulking of storage roots [47]. This indicates that the short vegetative growth period by PlgFT-GFP overexpression should affect the bulking of storage roots. Unlike PlgFT-GFP overexpressing plants, the overexpression of StSP6A induces the early flowering and tuberization without early leaf senescence phenotype [29]. Therefore, the inducible promoter would be required for analyzing the function of PlgFT in the development of storage roots. Based on the extreme early flowering phenotype caused by the ectopic expression of PlgFT-GFP, we conclude that PlgFT is a functional ortholog of Arabidopsis FT and acts as a flowering-inducer in P. grandiflorus.

3. Materials and Methods

3.1. Plant Materials and Experimental Treatment

P. grandiflorus were sown and grown in soil under long-day condition (LD, 16 h light/8 h dark) or short-day condition (SD, 8 h light/16 h dark) at 22 °C. In order to analyze the organ-specific expression pattern of PlgFT, various organs of P. grandiflorus, including the leaves, stems, and roots, were harvested from two- or four-month-old plants, which were grown under LD. The bud, flowers, and wilted flowers were harvested from four- to five-month-old plants. Plants entered into the adult vegetative phase were used for analyzing the diurnal expression pattern of FTs [48,49]. Thus, the diurnal expression patterns of PlgFT were investigated over a 24 h period using three-month-old plants. For GA3 treatment, GA3 solution was prepared using 1% EtOH and 0.02% tween-20. Three-month-old plants (LD condition) were sprayed with 1 mM GA3 or with a solution containing 1% EtOH and 0.02% tween-20 as a mock control. In addition, plants were transferred to a growth chamber (30 °C, LD conditions) to investigate the effect of temperature. The unexpanded young leaves were collected at different time points after treatment (0 d, 1 d, 3 d, and 7 d), immediately frozen in liquid nitrogen, and stored at −80 °C prior to analysis.

3.2. Identification FT Gene in P. grandiflorus

In order to identify FT in P. grandiflorus, the P. grandiflorus genomic sequence data [24] were searched using FT protein sequences obtained from different plant species as direct queries with BLAT-search. The protein sequence of putative PlgFT was aligned with FT and TFL proteins in different plant species, and a phylogenetic tree using the condensed alignment in MEGA7.0 with the neighbor-joining algorithm was constructed. Table S1 lists the accession number of proteins used for the analyses of multiple-sequence alignment and phylogenetic tree.

3.3. Plasmid Construction

The full-length PlgFT gene was amplified using gene-specific primers (Table S2) designed using P. grandiflorus genomic sequence data. The full-length coding sequence of PlgFT was cloned into the pENTR/D-TOPO vector and sub-cloned into a gateway binary vector pGWB505 or pGBKT7 (for yeast two-hybrid analysis). The sequence of PlgFT (NS-3080) was deposited in the National Agricultural Biotechnology Information Center (NABIC, http://nabic.rda.go.kr).

3.4. RNA Extraction and Quantitative Real-Time PCR

Total RNA was extracted from P. grandiflorus samples using the FavorPrep Plant Total RNA Purification Mini Kit (Favogen, Pingtung, Taiwan), and 500 ng of total RNA were reverse-transcribed using a cDNA synthesis kit (TOYOBO Co., Ltd., Osaka, Japan). In order to determine the expression levels of the PlgFT gene, RT-qPCR was performed using the CFX96TM Real-time system (Bio-Rad, Hercules, CA, USA). The expression levels of PlgFT were normalized to the constitutive expression level of actin using the 2−∆∆CT method. Table S2 lists the primer sequences.

3.5. Subcellular Localization of PlgFT in Tobacco Leaves

For transient PlgFT-GFP expression, a single colony of Agrobacterium tumefaciens GV3101 containing 35S:PlgFT-GFP (PlgFT-GFP) construct was inoculated into 3 mL LB medium with antibiotics at 28 °C. After culturing overnight, the bacteria were harvested and resuspended in MMA (10 mM MES and 10 mM MgCl2 containing 200 mM acetosyringone) to an OD600 of 0.3 and used for syringe agro-infiltration into Nicotiana benthamiana. GFP fluorescence was visualized 48 h post-infiltration using a confocal laser scanning microscope.

3.6. Yeast Two-Hybrid Assay

The pGBKT7-PlgFT and pGADT7-AtFD (Arabidopsis FD) constructs were co-transformed into yeast strain AH109 according to Matchmaker™ GAL4 Two-hybrid System 3 protocol. The transformed yeast cells were selected on SD medium without Leu and Trp, and then protein–protein interactions were confirmed on SD medium without Leu, Trp, and His. The empty pGBKT7 and pGADT7 vectors were used as negative controls.

3.7. Agrobacterium Mediated Transformation

P. grandiflorus seeds were germinated on a half-strength MS medium under LD, and leaves of three-week-old P. grandiflorus grown in vitro were used for the Agrobacterium-mediated transformation approach. Leaf explants were co-cultivated with the cell suspension of GV3101 containing 35S: PlgFT-GFP (OD600 = 1.0) for three days under dark condition and then placed on the shoot-induced medium (MS medium containing 1 mg/L 6-benzyladenine, 0.5 mg/L 1-naphthaleneacetic acid, 300 mg/L cefotaxime, 30 mg/L hygromycin, 30 g/L sucrose, and 7.5 g/L plant agar). After shoot formation from explant cultures, shoots were transferred to a BA-free medium for the induction of roots. The transgenic lines showing hygromycin resistance were transplanted in soil, and lines with a high level of PlgFT-GFP protein were selected by RT-PCR. In addition, we selected regenerated plants, which were survived on the selection medium but exhibited no expression of PlgFT-GFP, as transgenic controls.

3.8. Statistical Analysis

All experiments were conducted with independent biological replicates and presented as the mean ± standard division. Statistically significant differences were determined using Duncan’s multiple range tests, and the significance was set at p < 0.05.

4. Conclusions

Here, we first characterized the FT gene in P. grandiflorus, an economically important medicinal plant in East Asia. PlgFT contains highly conserved functional motifs and key amino acid residues necessary for FT function as a floral inducer. PlgFT was not diurnally regulated because it lacks CORE motifs in its promoter region. In addition, we suggested that the thermoresponsive flowering of P. grandiflorus is mediated by the induction of PlgFT. Furthermore, ectopic expression of PlgFT in P. grandiflorus clearly suggests that PlgFT acts as a flowering inducer. Our results provide an important starting point for future efforts to understand the molecular mechanisms of flowering in P. grandiflorus, and will be helpful for further studies to improve storage root development by the control of flowering times.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/plants11030325/s1, Figure S1: Phylogenetic tree analysis of the FT/TFL1 proteins in angiosperms, Figure S2: Sequence alignment of conserved FT promoter region, Table S1: The accession number of proteins used for the analyses of multiple-sequence alignment and phylogenetic tree, Table S2: Primer list used for experiments.

Author Contributions

Conceptualization, G.K., H.C. and T.K.H.; Investigation, G.K. and Y.R.; writing—original draft preparation, G.K., H.C. and T.K.H.; writing—review and editing, H.C. and T.K.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation of Korea(NRF) grant funded by the Korean government (MSIT) (No. NRF-2021R1A4A1032888).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to reasons of privacy.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fornara, F.; de Montaigu, A.; Coupland, G. SnapShot: Control of flowering in Arabidopsis. Cell 2010, 141, 550. [Google Scholar] [CrossRef] [Green Version]
  2. Sang, N.; Cai, D.; Li, C.; Sun, Y.; Huang, X. Characterization and activity analyses of the FLOWERING LOCUS T promoter in Gossypium Hirsutum. Int. J. Mol. Sci. 2019, 20, 4769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Turck, F.; Fornara, F.; Coupland, G. Regulation and identity of florigen: FLOWERING LOCUS T moves center stage. Annu. Rev. Plant Biol. 2008, 59, 573–594. [Google Scholar] [CrossRef] [Green Version]
  4. Song, Y.H.; Ito, S.; Imaizumi, T. Flowering time regulation: Photoperiod-and temperature-sensing in leaves. Trends Plant Sci. 2013, 18, 575–583. [Google Scholar] [CrossRef] [Green Version]
  5. Abe, M.; Kobayashi, Y.; Yamamoto, S.; Daimon, Y.; Yamaguchi, A.; Ikeda, Y.; Ichinoki, H.; Notaguchi, M.; Goto, K.; Araki, T. FD, a bZIP protein mediating signals from the floral pathway integrator FT at the shoot apex. Science 2005, 309, 1052–1056. [Google Scholar] [CrossRef]
  6. Collani, S.; Neumann, M.; Yant, L.; Schmid, M. FT Modulates genome-wide DNA-binding of the bZIP transcription factor FD. Plant Physiol. 2019, 180, 367–380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Cao, S.; Kumimoto, R.W.; Gnesutta, N.; Calogero, A.M.; Mantovani, R.; Holt III, B.F. A distal CCAAT/NUCLEAR FACTOR Y complex promotes chromatin looping at the FLOWERING LOCUS T promoter and regulates the timing of flowering in Arabidopsis. Plant Cell 2014, 26, 1009–1017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Hayama, R.; Yokoi, S.; Tamaki, S.; Yano, M.; Shimamoto, K. Adaptation of photoperiodic control pathways produces short-day flowering in rice. Nature 2003, 422, 719–722. [Google Scholar] [CrossRef]
  9. Hayama, R.; Agashe, B.; Luley, E.; King, R.; Coupland, G. A circadian rhythm set by dusk determines the expression of FT homologs and the short-day photoperiodic flowering response in Pharbitis. Plant Cell 2007, 19, 2988–3000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Samach, A.; Lotan, H. The transition to flowering in tomato. Plant Biotechnol. J. 2007, 24, 71–82. [Google Scholar] [CrossRef] [Green Version]
  11. Lifschitz, E.; Eviatar, T.; Rozman, A.; Shalit, A.; Goldshmidt, A.; Amsellem, Z.; Alvarez, J.P.; Eshed, Y. The tomato FT ortholog triggers systemic signals that regulate growth and flowering and substitute for diverse environmental stimuli. Proc. Natl. Acad. Sci. USA 2006, 103, 6398–6403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. McClung, C.; Lou, P.; Hermand, V.; Kim, J.A. The importance of ambient temperature to growth and the induction of flowering. Front. Plant Sci. 2016, 7, 1266. [Google Scholar] [CrossRef]
  13. Kumar, S.V.; Lucyshyn, D.; Jaeger, K.E.; Alós, E.; Alvey, E.; Harberd, N.P.; Wigge, P.A. Transcription factor PIF4 controls the thermosensory activation of flowering. Nature 2012, 484, 242–245. [Google Scholar] [CrossRef]
  14. Li, X.F.; Jia, L.Y.; Xu, J.; Deng, X.J.; Wang, Y.; Zhang, W.; Zhang, X.P.; Fang, Q.; Zhang, D.M.; Sun, Y.; et al. FT-like NFT1 gene may play a role in flower transition induced by heat accumulation in Narcissus tazetta var. chinensis. Plant Cell Physiol. 2013, 54, 270–281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Tamura, K.; Teranishi, Y.; Ueda, S.; Suzuki, H.; Kawano, N.; Yoshimatsu, K.; Saito, K.; Kawahara, N.; Muranaka, T.; Seki, H. Cytochrome P450 monooxygenase CYP716A141 is a unique β-amyrin C-16β oxidase involved in triterpenoid saponin biosynthesis in Platycodon grandiflorus. Plant Cell Physiol. 2017, 58, 874–884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Yoon Sun, H.; Hee Doo, L.; Joung Kwan, L.; Bo Goo, K.; Ki Yeol, L. Effect of pinching time and position on growth and flowering of Platycodon grandiflorum var. duplex Makino in cut flower cultivation. Flower Res. J. 2015, 23, 25–30. [Google Scholar]
  17. Cho, J.T. Physiological and ecological studies of Chinese bellflower Platycodon grandiflorum DC. J. Korean Soc. Hortic. Sci. 1984, 23, 194–200. [Google Scholar]
  18. Halevy, A.H.; Shlomo, E.; Ziv, O. Improving cut flower production of balloon flower. HortScience 2002, 37, 759–761. [Google Scholar] [CrossRef] [Green Version]
  19. Ben-Naim, O.; Eshed, R.; Parnis, A.; Teper-Bamnolker, P.; Shalit, A.; Coupland, G.; Samach, A.; Lifschitz, E. The CCAAT binding factor can mediate interactions between CONSTANS-like proteins and DNA. Plant J. 2006, 46, 462–476. [Google Scholar] [CrossRef] [Green Version]
  20. Bennett, T.; Dixon, L.E. Asymmetric expansions of FT and TFL1 lineages characterize differential evolution of the EuPEBP family in the major angiosperm lineages. BMC Biol. 2021, 19, 181. [Google Scholar] [CrossRef]
  21. Wickland, D.P.; Hanzawa, Y. The FLOWERING LOCUS T/TERMINAL FLOWER 1 gene family: Functional evolution and molecular mechanisms. Mol. Plant 2015, 8, 983–997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Yoo, S.Y.; Kardailsky, I.; Lee, J.S.; Weigel, D.; Ahn, J.H. Acceleration of flowering by overexpression of MFT (MOTHER OF FT AND TFL1). Mol. Cells 2004, 17, 95–101. [Google Scholar]
  23. Ho, W.W.H.; Weigel, D. Structural features determining flower-promoting activity of Arabidopsis FLOWERING LOCUS T. Plant Cell 2014, 26, 552–564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Kim, J.; Kang, S.H.; Park, S.G.; Yang, T.J.; Lee, Y.; Kim, O.T.; Chung, O.; Lee, J.; Choi, J.P.; Kwon, S.J.; et al. Whole-genome, transcriptome, and methylome analyses provide insights into the evolution of platycoside biosynthesis in Platycodon grandiflorus, a medicinal plant. Hortic. Res. 2020, 7, 112. [Google Scholar] [CrossRef] [PubMed]
  25. Ahn, J.H.; Miller, D.; Winter, V.J.; Banfield, M.J.; Lee, J.H.; Yoo, S.Y.; Henz, S.R.; Brady, R.L.; Weigel, D. A divergent external loop confers antagonistic activity on floral regulators FT and TFL1. EMBO J. 2006, 25, 605–614. [Google Scholar] [CrossRef]
  26. Manoharan, R.K.; Han, J.S.H.; Vijayakumar, H.; Subramani, B.; Thamilarasan, S.K.; Park, J.-I.; Nou, I.-S. Molecular and functional characterization of FLOWERING LOCUS T homologs in Allium cepa. Molecules 2016, 21, 217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Hanzawa, Y.; Money, T.; Bradley, D. A single amino acid converts a repressor to an activator of flowering. Proc. Natl. Acad. Sci. USA 2005, 102, 7748–7753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Imamura, T.; Nakatsuka, T.; Higuchi, A.; Nishihara, M.; Takahashi, H. The gentian orthologs of the FT/TFL1 gene family control floral initiation in Gentiana. Plant Cell Physiol. 2011, 52, 1031–1041. [Google Scholar] [CrossRef] [Green Version]
  29. Navarro, C.; Abelenda, J.A.; Cruz-Oró, E.; Cuellar, C.A.; Tamaki, S.; Silva, J.; Shimamoto, K.; Prat, S. Control of flowering and storage organ formation in potato by FLOWERING LOCUS T. Nature 2011, 478, 119–122. [Google Scholar] [CrossRef]
  30. Niu, L.; Fu, C.; Lin, H.; Wolabu, T.W.; Wu, Y.; Wang, Z.Y.; Tadege, M. Control of floral transition in the bioenergy crop switchgrass. Plant Cell Environ. 2016, 39, 2158–2171. [Google Scholar] [CrossRef]
  31. Yarur, A.; Soto, E.; León, G.; Almeida, A.M. The sweet cherry (Prunus avium) FLOWERING LOCUS T gene is expressed during floral bud determination and can promote flowering in a winter-annual Arabidopsis accession. Plant Reprod. 2016, 29, 311–322. [Google Scholar] [CrossRef]
  32. Kang, J.; Zhang, T.; Guo, T.; Ding, W.; Long, R.; Yang, Q.; Wang, Z. Isolation and functional characterization of MsFTa, a FLOWERING LOCUS T homolog from alfalfa (Medicago sativa). Int. J. Mol. Sci. 2019, 20, 1968. [Google Scholar] [CrossRef] [Green Version]
  33. Zhang, L.; Yu, H.; Lin, S.; Gao, Y. Molecular characterization of FT and FD homologs from Eriobotrya deflexa Nakai forma koshunensis. Front. Plant Sci. 2016, 7, 8. [Google Scholar] [CrossRef] [Green Version]
  34. Zhan, Z.; Zhang, C.; Zhang, H.; Li, X.; Wen, C.; Liang, Y. Molecular cloning, expression analysis, and subcellular localization of FLOWERING LOCUS T (FT) in carrot (Daucus carota L.). Mol. Breed. 2017, 37, 1–9. [Google Scholar] [CrossRef]
  35. Harig, L.; Beinecke, F.A.; Oltmanns, J.; Muth, J.; Müller, O.; Rüping, B.; Twyman, R.M.; Fischer, R.; Prüfer, D.; Noll, G.A. Proteins from the FLOWERING LOCUS T-like subclade of the PEBP family act antagonistically to regulate floral initiation in tobacco. Plant J. 2012, 72, 908–921. [Google Scholar] [CrossRef] [PubMed]
  36. Farrona, S.; Thorpe, F.L.; Engelhorn, J.; Adrian, J.; Dong, X.; Sarid-Krebs, L.; Goodrich, J.; Turck, F. Tissue-specific expression of FLOWERING LOCUS T in Arabidopsis is maintained independently of polycomb group protein repression. Plant Cell 2011, 23, 3204–3214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Kotoda, N.; Hayashi, H.; Suzuki, M.; Igarashi, M.; Hatsuyama, Y.; Kidou, S.; Igasaki, T.; Nishiguchi, M.; Yano, K.; Shimizu, T.; et al. Molecular characterization of FLOWERING LOCUS T-like genes of apple (Malus x domestica Borkh.). Plant Cell Physiol. 2010, 51, 561–575. [Google Scholar] [CrossRef] [Green Version]
  38. Varkonyi-Gasic, E.; Moss, S.M.A.; Voogd, C.; Wang, T.; Putterill, J.; Hellens, R.P. Homologs of FT, CEN and FD respond to developmental and environmental signals affecting growth and flowering in the perennial vine kiwifruit. New Phytol. 2013, 198, 732–746. [Google Scholar] [CrossRef]
  39. Lee, R.; Baldwin, S.; Kenel, F.; McCallum, J.; Macknight, R. FLOWERING LOCUS T genes control onion bulb formation and flowering. Nat. Commun. 2013, 4, 1–9. [Google Scholar] [CrossRef] [Green Version]
  40. Li, C.; Luo, L.; Fu, Q.; Niu, L.; Xu, Z.-F. Isolation and functional characterization of JcFT, a FLOWERING LOCUS T (FT) homologous gene from the biofuel plant Jatropha curcas. BMC Plant Biol. 2014, 14, 1–12. [Google Scholar] [CrossRef] [Green Version]
  41. Sreekantan, L.; Thomas, M.R. VvFT and VvMADS8, the grapevine homologues of the floral integrators FT and SOC1, have unique expression patterns in grapevine and hasten flowering in Arabidopsis. Funct. Plant Biol. 2006, 33, 1129–1139. [Google Scholar] [CrossRef] [PubMed]
  42. Tiwari, S.B.; Shen, Y.; Chang, H.C.; Hou, Y.; Harris, A.; Ma, S.F.; McPartland, M.; Hymus, G.J.; Adam, L.; Marion, C. The flowering time regulator CONSTANS is recruited to the FLOWERING LOCUS T promoter via a unique cis-element. New Phytol. 2010, 187, 57–66. [Google Scholar] [CrossRef]
  43. Park, B.H.; Oliveira, N.; Pearson, S. Temperature affects growth and flowering of the balloon flower [Platycodon grandiflorus (Jacq.) A. DC. cv. Astra Blue]. HortScience 1998, 33, 233–236. [Google Scholar]
  44. Nakano, Y.; Higuchi, Y.; Yoshida, Y.; Hisamatsu, T. Environmental responses of the FT/TFL1 gene family and their involvement in flower induction in Fragaria × ananassa. J. Plant Physiol. 2015, 177, 60–66. [Google Scholar] [CrossRef]
  45. Noy-Porat, T.; Cohen, D.; Mathew, D.; Eshel, A.; Kamenetsky, R.; Flaishman, M.A. Turned on by heat: Differential expression of FT and LFY-like genes in Narcissus tazetta during floral transition. J. Exp. Bot. 2013, 64, 3273–3284. [Google Scholar] [CrossRef] [Green Version]
  46. No, D.H.; Baek, D.; Lee, S.H.; Cheong, M.S.; Chun, H.J.; Park, M.S.; Cho, H.M.; Jin, B.J.; Lim, L.H.; Lee, Y.B.; et al. High-temperature conditions promote soybean flowering through the transcriptional reprograming of flowering genes in the photoperiod pathway. Int. J. Mol. Sci. 2021, 22, 1314. [Google Scholar] [CrossRef]
  47. Tavora, F.J.A.F.; Melo, F.I.O.; de Pinho, J.L.N.; de Queiroz, G.M. Yield, crop growth rate and assimilatory characteristics of cassava at the coastal area of Ceara. Rev. Bras. Fisiol. Veg. 1995, 7, 81–88. [Google Scholar]
  48. Chaurasia, A.K.; Patil, H.B.; Krishna, B.; Subramaniam, V.R.; Sane, P.V.; Sane, A.P. Flowering time in banana (Musa spp.), a day neutral plant, is controlled by at least three FLOWERING LOCUS T homologues. Sci. Rep. 2017, 7, 5935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Kang, M.Y.; Kwon, H.Y.; Kim, N.Y.; Sakuraba, Y.; Paek, N.C. CONSTITUTIVE PHOTOMORPHOGENIC 10 (COP10) contributes to floral repression under non-inductive short days in Arabidopsis. Int. J. Mol. Sci. 2015, 16, 26493–26505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Sequence alignment of PlgFT protein and other reported FT proteins. Red boxes indicate the conserved motifs and segments, including DPDxP, GXHR, segment A, and LYN. Blue boxes indicate key amino acids for flowering activator.
Figure 1. Sequence alignment of PlgFT protein and other reported FT proteins. Red boxes indicate the conserved motifs and segments, including DPDxP, GXHR, segment A, and LYN. Blue boxes indicate key amino acids for flowering activator.
Plants 11 00325 g001
Figure 2. Subcellular localization and yeast two-hybrid assay for P. grandiflorus FT. (A) Subcellular localization of PlgFT in Nicotiana benthamiana leaves. PlgFT recombinant proteins transiently expressed in N. benthamiana leaves through agro-infiltration. Scale bar = 20 μm. (B) Analysis of protein–protein interaction between PlgFT and Arabidopsis FD (AtAD) using yeast two-hybrid assay. Yeast cells (AH109) were grown on SD agar medium lacking either Trp and Leu (−LT) or Trp, Leu, and His (−LTH).
Figure 2. Subcellular localization and yeast two-hybrid assay for P. grandiflorus FT. (A) Subcellular localization of PlgFT in Nicotiana benthamiana leaves. PlgFT recombinant proteins transiently expressed in N. benthamiana leaves through agro-infiltration. Scale bar = 20 μm. (B) Analysis of protein–protein interaction between PlgFT and Arabidopsis FD (AtAD) using yeast two-hybrid assay. Yeast cells (AH109) were grown on SD agar medium lacking either Trp and Leu (−LT) or Trp, Leu, and His (−LTH).
Plants 11 00325 g002
Figure 3. Expression pattern of PlgFT in different organs (A) and flower development stages (B). Expression levels of PlgFT were normalized to actin. Data represent the means ± SD of three independent experiments. Different letters correspond to means that are statistically different (p < 0.05). The scale bars represent 10 cm (A) and 1 cm (B).
Figure 3. Expression pattern of PlgFT in different organs (A) and flower development stages (B). Expression levels of PlgFT were normalized to actin. Data represent the means ± SD of three independent experiments. Different letters correspond to means that are statistically different (p < 0.05). The scale bars represent 10 cm (A) and 1 cm (B).
Plants 11 00325 g003
Figure 4. Diurnal expression pattern of PlgFT under different photoperiodic conditions. The expression levels of PlgFT were measured every 4 h throughout the diurnal cycle in LD and SD conditions. Expression levels of PlgFT were normalized to actin. Values represent three independent replicates ± SD. Uppercase denotes significance (p < 0.05) between the expression levels under LD, whereas lowercase denotes significance (p < 0.05) between the expression levels under SD.
Figure 4. Diurnal expression pattern of PlgFT under different photoperiodic conditions. The expression levels of PlgFT were measured every 4 h throughout the diurnal cycle in LD and SD conditions. Expression levels of PlgFT were normalized to actin. Values represent three independent replicates ± SD. Uppercase denotes significance (p < 0.05) between the expression levels under LD, whereas lowercase denotes significance (p < 0.05) between the expression levels under SD.
Plants 11 00325 g004
Figure 5. Regulation of PlgFT expression by GA3 (A) and high-temperature conditions (B). Two-month-old plants were treated with GA3 or high temperature (30 °C). The expression levels of PlgFT were normalized to actin. Values represent three independent replicates ± SD. Uppercase denotes significance (p < 0.05) between the expression levels after treatment of GA3 or high temperature (30 °C), whereas lowercase denotes significance (p < 0.05) between the expression levels after treatment of mock or normal temperature (22 °C).
Figure 5. Regulation of PlgFT expression by GA3 (A) and high-temperature conditions (B). Two-month-old plants were treated with GA3 or high temperature (30 °C). The expression levels of PlgFT were normalized to actin. Values represent three independent replicates ± SD. Uppercase denotes significance (p < 0.05) between the expression levels after treatment of GA3 or high temperature (30 °C), whereas lowercase denotes significance (p < 0.05) between the expression levels after treatment of mock or normal temperature (22 °C).
Plants 11 00325 g005
Figure 6. Ectopic expression of PlgFT in P. grandiflorus induced early flowering. (A) The expression of PlgFT-GFP in the selected transgenic plants was confirmed by RT-PCR. Phenotypes of TC and PlgFT-GFP overexpression plants (PlgFT-OX) grown in the in vitro condition (B) and soil (C). Scale bar = 1 cm (B) or 10 cm (C).
Figure 6. Ectopic expression of PlgFT in P. grandiflorus induced early flowering. (A) The expression of PlgFT-GFP in the selected transgenic plants was confirmed by RT-PCR. Phenotypes of TC and PlgFT-GFP overexpression plants (PlgFT-OX) grown in the in vitro condition (B) and soil (C). Scale bar = 1 cm (B) or 10 cm (C).
Plants 11 00325 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, G.; Rim, Y.; Cho, H.; Hyun, T.K. Identification and Functional Characterization of FLOWERING LOCUS T in Platycodon grandiflorus. Plants 2022, 11, 325. https://doi.org/10.3390/plants11030325

AMA Style

Kim G, Rim Y, Cho H, Hyun TK. Identification and Functional Characterization of FLOWERING LOCUS T in Platycodon grandiflorus. Plants. 2022; 11(3):325. https://doi.org/10.3390/plants11030325

Chicago/Turabian Style

Kim, Gayeon, Yeonggil Rim, Hyunwoo Cho, and Tae Kyung Hyun. 2022. "Identification and Functional Characterization of FLOWERING LOCUS T in Platycodon grandiflorus" Plants 11, no. 3: 325. https://doi.org/10.3390/plants11030325

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