Skip to main content

Overexpression of four MiTFL1 genes from mango delays the flowering time in transgenic Arabidopsis

Abstract

Background

TERMINAL FLOWER 1 (TFL1) belongs to the phosphatidylethanolamine-binding protein (PEBP) family, which is involved in inflorescence meristem development and represses flowering in several plant species. In the present study, four TFL1 genes were cloned from the mango (Mangifera indica L.) variety ‘SiJiMi’ and named MiTFL1-1, MiTFL1-2, MiTFL1-3 and MiTFL1-4.

Results

Sequence analysis showed that the encoded MiTFL1 proteins contained a conserved PEBP domain and belonged to the TFL1 group. Expression analysis showed that the MiTFL1 genes were expressed in not only vegetative organs but also reproductive organs and that the expression levels were related to floral development. Overexpression of the four MiTFL1 genes delayed flowering in transgenic Arabidopsis. Additionally, MiTFL1-1 and MiTFL1-3 changed the flower morphology in some transgenic plants. Yeast two-hybrid (Y2H) analysis showed that several stress-related proteins interacted with MiTFL1 proteins.

Conclusions

The four MiTFL1 genes exhibited a similar expression pattern, and overexpression in Arabidopsis resulted in delayed flowering. Additionally, MiTFL1-1 and MiTFL1-3 overexpression affected floral organ development. Furthermore, the MiTFL1 proteins could interact with bHLH and 14-3-3 proteins. These results indicate that the MiTFL1 genes may play an important role in the flowering process in mango.

Peer Review reports

Background

Floral transition is an important stage in the life cycle of higher plants. The process underlying the floral transition from the vegetative to reproductive phase is regulated by complex internal signals and external environmental factors [1, 2]. Various flowering response pathways, such as the photoperiod, vernalization, gibberellin, autonomous, ambient temperature, and age-related pathways, have been identified in the model plant Arabidopsis thaliana [3]. Numerous genes play essential roles in these processes, and the involved interactions specify the meristem fate [4]. CONSTANS (CO) contains a zinc finger structure and CCT domain, which activates the transcription of the FLOWERING LOCUS T (FT) gene by binding to the FT promoter region, and the FT protein moves from the leaf tissue to the stem apex to initiate the transition of the plant from vegetative to reproductive growth [5, 6]. The FLOWER LOCUS C (FLC) gene plays a central role in vernalization—the induction of plant flowering [7, 8]—and inhibits flowering by binding to FT, FLOWERING LOCUS D (FD) and SUPPRESSOR OF OVER-EXPRESSION OF CONSTANS1 (SOC1) and inhibiting the expression of these genes [9]. The flower meristem-specific genes LEAFY (LFY) and APETALA1 (AP1) directly induce shoot apical meristem differentiation, which promotes the entry of plants into the flowering stage, and then, these genes are activated by FT or SOC1 [10].

TERMINAL FLOWER 1 (TFL1), which belongs to the phosphatidylethanolamine-binding protein (PEBP) family, was first identified in Arabidopsis [11]. TFL1 and FT are genes with highly homologous sequences but opposite functions. TFL1 encodes proteins with conserved His88 and Asp144 residues and the typical amino acid triad modules EYD, YFG, and END [12]. The FT protein does not have this structure, a critical reason for the opposite functions of these two genes [13]. TFL1 genes, as flowering repressors, determine the timing of the transition of the apical meristem into an inflorescence meristem and the branching pattern of the inflorescence [14]. In Arabidopsis, the AtTFL1 gene not only maintains the infinite growth of the stem apical meristem and inflorescence meristem but is also involved in flower formation [15]. In most fruit trees, the function of TFL1 homologous genes is to delay flowering, similar to the function of AtTFL1. For example, the function of PmTFL1 (from Prunus mume) is to delay flowering, as shown in transformed Arabidopsis [16]. The antagonistic effect between FT and TFL1 exhibits a certain relationship with competition with FD [17, 18]. The TFL1 gene inhibits the expression of LFY and AP1, which are downstream of the FT gene, by binding to FD and inhibiting flowering [19]. Additionally, LFY and AP1 regulate the expression of the TFL1 gene in an opposite manner: LFY serves as an activator, AP1 is a suppressor, and these two genes form an unclear feedback loop. The flowering of plants depends on the ratio of TFL1 to LFY gene expression. A high ratio maintains the plants in a flowering inhibition period, whereas plants with a low ratio are in the early flowering period [20].

Compared with annual plants, woody fruit trees have a longer juvenile period, which severely affects breeding. Several studies have shown that the overexpression or silencing of flowering-related genes can shorten the juvenile period and promote flowering. For example, the BpMADS4 gene from silver birch is constitutively overexpressed in apple, and transgenic plants exhibit markedly shortened juvenile and flowering periods [21]. PcTFL1-1 and PcTFL1-2 were silenced in European pear using RNAi technology, and the plants showed early flowering traits and a shortened juvenile period [22].

Mango (Mangifera indica L.) is a world-famous woody fruit tree that is widely grown in tropical and subtropical areas. Several environmental factors affect mango flowering, including low temperature, water stress, and carbohydrates. Exogenous spraying of potassium nitrate, paclobutrazol and ethephon can promote flowering, whereas spraying of gibberellin inhibits flowering [23]. The publication of the mango genome has provided valuable information regarding the mining of flowering genes [24]. In recent years, several flowering-regulating genes have been isolated and functionally identified in mango, including the flowering-promoting gene MiSOC1 [25], two MiAP1 genes [26], three MiFT genes [27], and a flowering-suppressing gene (MiCO) [28]. The function of MiTFL1 has not been identified in this plant. In the present study, four MiTFL1 homologous genes, namely, MiTFT1-1, MiTFT1-2, MiTFT1-3 and MiTFT1-4, were cloned from M. indica L. cv. ‘SiJiMi’. The expression patterns of the four MiTFL1 genes in different tissues and at different flowering development stages were evaluated. MiTFL1 gene overexpression vectors were constructed, and the functions of these genes were determined by transformation in Arabidopsis. Proteins interacting with MiTFL1 proteins were screened through yeast two-hybrid (Y2H) assays. The results suggest that these four MiTFL1 genes inhibit mango flowering.

Results

Isolation and sequence analysis of MiTFL1 genes

Four TFL1 homologous genes were identified from the transcriptome and genomic data of ‘SiJiMi’ mango (unpublished data). We further verified the sequences by RT-PCR and showed that these sequences were consistent with those obtained from the transcriptome data. The four genes were named MiTFL1-1, MiTFL1-2, MiTFL1-3 and MiTFL1-4, and their DNA sequence lengths were 1175 bp, 1054 bp, 962 bp and 1299 bp, respectively. All four MiTFL1 genes contained four exons and three introns (Fig. 1a). Sequence alignment analysis showed that the length of the first exon differed among the four MiTFL1 genes, while the second and third exons had the same length; the fourth exon was 215 bp in length in all the genes except MiTFL1-2, in which the length was 218 bp. The full coding sequences of the four MiTFL1 genes were 516 bp, 525 bp, 519 bp and 510 bp and encoded 172 aa, 175 aa, 173 aa and 170 aa, respectively. Amino acid sequence alignment analysis showed that the similarity between MiTFL1-3 and MiTFL1-4 was 90% higher than that between MiTFL1-1 and MiTFL1-2 at 70%. The amino acid sequences of the MiTFL1-1, MiTFL1-2, MiTFL1-3, and MiTFL1-4 proteins exhibited 68.9, 69.1, 62.9 and 60.1% similarity with the sequence of AtTFL1 (NP_196004.1) of Arabidopsis, respectively. Additionally, all the MiTFL1 proteins were identified as TFL1 proteins containing the crucial conserved amino acid residues of TFL1-like proteins (Fig. 1b), namely, histidine at position 85 (H85) and aspartic acid at position 140 (D140).

Fig. 1
figure 1

Multiple sequence alignment and gene structure of TFL1 genes. a Gene structures of MiTFL1 genes. b Amino acid sequence alignments of TFL1 proteins from different fruit trees and Arabidopsis thaliana. The following species were included in the analysis (the GenBank accession numbers are shown in parentheses): Arabidopsis thaliana (AtTFL1, NP_196004.1), Citrus sinensis (CsTFL1, NP_001275848), Dimocarpus longan (DlTFL1, AHY24028.1), Juglans regia (JrTFL1, XP_018811176.1), Malus domestica (MdTFL1, NP_001280887.1), Pyrus x bretschneideri (PpTFL1, NP_001289244.1), and Vitis vinifera (VvTFL1A, NP_001267929.1). The black color indicates that the sequences are exactly the same. The red color indicates ≥75% similarity. The blue color indicates ≥50% similarity. The red box indicates the key residues His85 and Asp140 of TFL1-like proteins

The PEBP gene family is divided into the TFL1, FT and MFT subfamilies. According to phylogenetic tree analysis (Fig. 2), the MiTFL1-1, MiTFL1-2, MiTFL1-3 and MiTFL1-4 proteins were clustered with the TFL1 proteins of other species. Among the investigated proteins, the MiTFL1-1 protein is closely related to the TFL1 proteins of apple, pear, apricot, plum, walnut, jujube and other fruit trees of Rutaceae and rose plants. The MiTFL1-2 protein is closely related to the TFL1 proteins of longan and grape and the CEN proteins of apple and cocoa. The MiTFL1-3 and MiTFL1-4 proteins are clustered together and closely related to the PvCEN protein of pistachio.

Fig. 2
figure 2

Phylogenetic tree analysis of PEBP proteins. The following species were included in the analysis (the GenBank accession numbers are shown in parentheses): Arabidopsis thaliana (AtTFL1, NP_196004.1), Citrus sinensis (CsTFL1, NP_001275848), Dimocarpus longan (DlTFL1, AHY24028.1), Juglans regia (JrTFL1, XP_018811176.1), Malus domestica (MdTFL1, NP_001280887.1), Pyrus x bretschneideri (PpTFL1, NP_001289244.1), Malus domestica (MdCEN, NP_001280940.1), Pistacia vera (PvCEN protein 1, XP_031269481.1), Theobroma cacao (TcCEN, XP_017973069.1), Hydrangea macrophylla (HmTFL1, MF374628.1), Glycine max (GmTFL1.3, FJ573238.1), Ficus carica (FcTFL1-2, AB746842.1), Citrus clementina (CcTFL1, XP_006430226.1), Citrus limon (ClTFL1, AWW25018.1), Ficus carica (FcTFL1, BAX00857.1), Nicotiana sylvestris (NsTFL1, XP_009766168.1), Prunus mume (PmTFL1, AEO72021.1), Prunus yedoensis (PyTFL1, AEO72023.1), Vitis vinifera (VvTFL1A, NP_001267929.1), Vitis vinifera (VvTFL1C, NP_001267933.1), Ziziphus jujube (ZjTFL1, XP_015898753.1), Actinidia chinensis (AcFT, AJA40932.1), Arabidopsis thaliana (AtFT, BAA77838.1), Eriobotrya deflexa (EdFT, AMB72867.1), Mangifera indica (MiFT, AGA19021.1), Vitis vinifera (VvFT, NP_001267907.1), Arabidopsis thaliana (AtMFT, OAP13671.1), Citrus limon (ClMFT, AWW25016.1), Dimocarpus longan (DlMFT, AUG98253.1), and Vitis vinifera (VvMFT, NP_001267935.1)

Expression analysis of MiTFL1 genes

The expression pattern of MiTFL1 genes in different mango tissues, including flowers, leaves, and stems, was determined by qRT-PCR (Fig. 3A). MiTFL1-1, MiTFL1-2, MiTFL1-3 and MiTFL1-4 were expressed in flowers, leaves, and stems. MiTFL1-1 and MiTFL1-4 exhibited higher expression levels in leaves than in other tissues. MiTFL1-2 was highly expressed in flowers, while MiTFL1-3 was highly expressed in stems.

Fig. 3
figure 3

qRT-PCR analysis of the expression profiles of MiTFL1 genes in mango. A: Expression patterns in different tissues. B: Expression patterns in different flowering developmental stages. The significance of the differences among the samples was assessed by Student’s t-test (P < 0.05). A Expression pattern of MiTFL1 genes in various tissues (flowers, leaves, and stems). B Expression pattern of MiTFL1 genes in mature leaves over time

To explore the expression patterns of the MiTFL1 genes at different flowering development stages of mango, mature leaves of M. indica L. cv. ‘SiJiMi’ were collected from the vegetative growth period to the flowering and fruiting period (November 2020–May 2021), and the results from the analysis of these leaves are shown in Fig. 3B. The expression patterns of the MiTFL1-1 and MiTFL1-2 genes were similar. The expression levels of these two genes were significantly higher at the floral induction stage than at other stages. The expression level of the MiTFL1-3 gene was high during vegetative growth, decreased significantly during the floral induction period, and increased significantly during the floral differentiation period. The lowest expression level was found at the flowering stage, and a relatively high expression level in leaves was observed during fruit development. The expression level of MiTFL1-4 first decreased during the floral induction period and then increased, peaking during the floral differentiation period, which was followed by a decrease and eventually stabilization during the fruit development period.

Subcellular localization of MiTFL1 proteins

To examine the subcellular localization of MiTFL1 proteins, 35S::GFP-MiTFL1-1, 35S::GFP-MiTFL1-2, 35S::GFP-MiTFL1-3, 35S::GFP- MiTFL1-4 and 35S::GFP-P1300 were separately transformed into onion epidermal cells (Fig. 4). The fluorescence signal of the empty vector 35S::GFP-P1300 was observed in the entire cell. The 35S::GFP-MiTFL1-1, 35S::GFP-MiTFL1-2, 35S::GFP-MiTFL1-3, and 35S::GFP-MiTFL1-4 fusion proteins were visible in only the nucleus and were stained with DAPI.

Fig. 4
figure 4

Subcellular localization analysis of MiTFL1 proteins. 35S::GFP-P1300 and 35S::GFP- MiTFL1s were localized in onion epidermal cells

Phenotypic analysis of MiTFL1 overexpression in Arabidopsis thaliana

MiTFL1 genes delay flowering in Arabidopsis thaliana

To explore the function of MiTFL1-1, MiTFL1-2, MiTFL1-3 and MiTFL1-4 in the flowering process of mango, individual overexpression vectors of pBI121-MiTFL1 were constructed and transferred separately into WT A. thaliana. Phenotypic observations of T3-generation homozygous plants were conducted, and WT and pBI121 empty vector-expressing Arabidopsis served as controls.

Four independent lines with MiTFL1-1 overexpression (OE-1#13, OE-1#22, OE-1#25 and OE-1#29) and three independent lines with MiTFL1-2 overexpression (OE-2#24, OE-2#45 and OE-2#55) were selected for functional analysis. Semiquantitative RT-PCR analysis showed that MiTFL1-1 and MiTFL1-2 were abundantly expressed in the transgenic lines but absent in the empty vector-expressing transgenic or WT plants (Fig. 5A1 and B1). All the independent lines of MiTFL1-1 and MiTFL1-2 showed delayed bolting and flowering: these processes occurred at 28.7–32.5 and 33.3–42.2 days, respectively, in these lines and at 24.9–25.3 and 28.5–28.8 days, respectively, in the control plants (Fig. 5A and B and Table 1). All the transformant lines with MiTFL1-1 and MiTFL1-2 showed normal bolting similar to that observed in the WT plants. Additionally, compared with the heights of the WT plants, those of the MiTFL1-1 and MiTFL1-2 plants were significantly increased, and the rosette leaves were not significantly affected except in OE-1#13, OE-1#29 and OE-2#45 (Table 1).

Fig. 5
figure 5

Phenotype of transgenic Arabidopsis lines and expression profiles of the transgenes. A Phenotype of MiTFL1-1-overexpressing transgenic lines showing delayed flowering (right) and that of the WT and pBI121 transgenic lines as controls (left) under LD conditions. A1 RT-PCR analysis of MiTFL1-1 in the control and MiTFL1-1 overexpression transgenic lines. B Phenotype of MiTFL1-2 overexpression transgenic lines showing delayed flowering (right) and that of the WT and pBI121 transgenic lines as controls (left) under LD conditions. B1 RT-PCR analysis of MiTFL1-2 in the control and MiTFL1-2-overexpressing transgenic lines. The original data can be viewed in Fig. S1a-b

Table 1 Flowering phenotype analysis of WT, pBI121, MiTFL1-1-overexpressing (OE-1) and MiTFL1-2-overexpressing (OE-2) transgenic plants

Three independent MiTFL1-3 overexpression (OE-3#19, OE-3#23 and OE-3#42) and MiTFL1-4 overexpression (OE-4#24, OE-4#45 and OE-4#55) lines were selected for functional analysis. Semiquantitative RT-PCR analysis demonstrated that MiTFL1-3 and MiTFL1-4 were abundantly expressed in the transgenic lines but absent in WT and pBI121 transgenic Arabidopsis plants (Fig. 6A1 and B1). The bolting time of the MiTFL1-3-overexpressing and MiTFL1-4-overexpressing transgenic plants was significantly delayed compared with that of the WT and pBI121 lines under long-day (LD) conditions (Fig. 6a and b and Table 2). The inhibitory effect of MiTFL1-4 on flowering was weaker than those of the other three MiTFL1 genes. The plant heights of some MiTFL1-3 and MiTFL1-4 transgenic lines showed significant differences, but the heights of some of the plants did not significantly differ from those of the control lines. The rosette leaves were not significantly affected in any of the plants (Table 2).

Fig. 6
figure 6

Phenotype of transgenic Arabidopsis lines and expression profiles of the transgenes. A Phenotype of MiTFL1-3-overexpressing transgenic lines showing delayed flowering (right) and that of WT and pBI121 transgenic lines as controls (left) under LD conditions. A1 RT-PCR analysis of MiTFL1-3 in the control and MiTFL1-3-overexpressing transgenic lines. B Phenotype of the MiTFL1-4-overexpressing transgenic lines, showing delayed flowering (right), and that of the WT and pBI121 transgenic lines as controls (left) under LD conditions. B1 RT-PCR analysis of MiTFL1-4 in the control and MiTFL1-4-overexpressing transgenic lines. The original data can be viewed in Fig. S1c-d

Table 2 Flowering phenotype analysis of WT, pBI121, MiTFL1-3-overexpressing (OE-3) and MiTFL1-4-overexpressing (OE-4) plants

MiTFL1-1 and MiTFL1-3 affect the flower phenotype in Arabidopsis

The MiTFL1-1-overexpressing (Fig. 7b) and MiTFL1-3-overexpressing transgenic lines (Fig. 7c) exhibited similar abnormal phenotypes to those of the WT lines (Fig. 7a). In the transgenic plants, some carpels developed into new inflorescences (Fig. 7B-a and C-a), and some flower structures lacked petals (Fig. 7B-b and C-b), in contrast with the results obtained for the WT plants (Fig. 7A-a and A-b). Two types of silique variations were found in the transgenic plants compared with the WT plants (Fig. 7A-b, A-c): in some siliques, the fruit stalk continued to lengthen from the flower during formation (Fig. 7B-c and C-c); in other siliques, the fruit stalk exhibited curved growth, and the siliques were shorter (Fig. 7B-c and C-c) than those of the WT plants. Additionally, the inflorescences of the transgenic plants were significantly different from those of the WT plants because of the variations in flower morphology (Fig. 7A-d, B-e and C-e). Furthermore, whorled leaves grew on the lateral branches of transgenic Arabidopsis thaliana but not in the control plants (Fig. 7A-e, B-f and C-f).

Fig. 7
figure 7

Photographs of the plant architecture, inflorescence, and floral phenotypes of WT and MiTFL1-1 and MiTFL1-3 transgenic Arabidopsis lines. A Phenotypes of WT Arabidopsis: (a, b) flowers, (c) silique, (d) inflorescence, and (f) stem. B Phenotypes of the MiTFL1-1-overexpressing line OE-1#13: (a) altered flower organs and no petals or carpels becoming a new inflorescence; (b) flower lacking petals; (c) a longer pod stalk formed in flowers; (d) pods with curved growth; (e) abnormal inflorescence; and (f) whorled leaves growing on the stem. C Phenotypes of the MiTFL1-3-overexpressing line OE-3#19: (a) altered flower organs and no petals or carpels becoming a new inflorescence; (b) flower lacking petals; (c) a longer pod stalk formed in flowers; (d) pods with curved growth; (e) abnormal inflorescence; and (f) whorled leaves growing on the stem

Expression patterns of endogenous genes in transgenic Arabidopsis expressing MiTFL1 genes

To determine whether MiTFL1 gene overexpression in transgenic Arabidopsis changed the expression of some flowering-related genes, such as the AtFT, AtFD, and AtAP1 homologs in Arabidopsis, the entire aboveground portion of T3 generation homozygous transgenic A. thaliana was collected 30 days after planting and subjected to qRT-PCR analysis (Fig. 8). AtACTIN2 was used as the internal reference gene. A similar expression pattern was found for the AtFT, AtFD, and AtAP1 transcripts in Arabidopsis after overexpressing each of the four MiTFL1 genes (Fig. 8a-d). The expression levels of the AtFT and AtAP1 genes were significantly lower in all the MiTFL1-overexpressing transgenic lines than in the WT plants. However, the expression of the AtFD gene was significantly increased in many transgenic lines but not in the MiTFL1-2-overexpressing line OE-2#45.

Fig. 8
figure 8

Expression analyses of flowering-related genes in transgenic Arabidopsis. a-d qRT-PCR analysis of endogenous flowering-related genes, including Arabidopsis AtAP1, AtFT and AtFD, in the MiTFL1-overexpressing and WT lines. The expression level was normalized to that of Arabidopsis AtACTIN2. The data are shown as the means ± SEs from three biological replicates. The significance of the differences among the samples was determined by Duncan’s test (P < 0.05)

Proteins that interact with MiTFL1 proteins

The Y2H system was used to screen the proteins interacting with MiTFL1 proteins and verify their interactions. The bait vector pGBKT7-MiTFL1 was constructed by double enzyme digestion, and no autoactivation or toxicity was detected (shown in Fig. S2). Yeast cells with bait plasmids were combined with the cDNA homogenization library of ‘SiJiMi’ to screen for positive clones. After Y2H assays, we obtained 9 candidate proteins associated with the MiTFL1-1 protein, 7 with MiTFL1-2, 6 with MiTFL1-3, and 7 with MiTFL1-4 (Table S1). Three stress- or flowering-related proteins were selected from the candidate proteins for further point-to-point verification on DDO/X and QDO/X/A media. The three proteins were basic helix-loop-helix protein 13 (bHLH13), bHLH162 and 14-3-3D, as shown in Fig. 9. The cells with the candidate protein bHLH13 in the pGADT7 recombinant vector turned blue and exhibited normal growth on QDO/X/A solid medium, indicating that the protein interacts with MiTFL1-2 and MiTFL1-3 proteins. bHLH162 interacted with MiTFL1-1, MiTFL1-2 and MiTFL1-4, whereas 14-3-3D interacted with only MiTFL1-1 and MiTFL1-2.

Fig. 9
figure 9

Identification of interactions between MiTFL1 proteins and other proteins in yeast. The interacting proteins were obtained through library screening; pGBKT7-53 and pGADT7-T served as positive controls, and pGBKT7-Lam and pGADT7-T were used as negative controls. The strains were cultured on DDO/−Trp/−Leu/X-α-gal (200 ng/ml) and QDO/−Trp/−Leu/−His/−Ade/X-α-gal (200 ng/ml)/AbA (500 μg/ml) media

Discussion

FT and TFL1 encode a pair of flowering regulators belonging to the PEBP family and play critical roles during the switch from vegetative growth to reproductive development [19]. FT and TFL1 share high homology, as demonstrated by their high nucleotide and amino acid sequence identities, but have opposite functions. FT promotes the transition to reproductive development and flowering, whereas TFL1 represses flowering [19, 29]. Our previous study identified three FT genes from mango and further confirmed that all FT genes significantly promoted flowering in the transgenic plants [27]. In the present study, four TFL1 genes were obtained from transcriptomic and genomic data analysis of ‘SiJiMi’ mango (unpublished data) and were named MiTFL1-1, MiTFL1-2, MiTFL1-3 and MiTFL1-4. Sequence analysis showed that the four TFL1 genes of ‘SiJiMi’ mango were highly consistent with the nucleotide sequences of ‘Alphonso’ (Fig. S3).

Multiple copies of TFL1 homologous genes are also found in other plants, such as two in soybean [30], moso bamboo [13] and loquat [31], three in petunia [32], four in cotton [33] and five in Hevea brasiliensis [34]. In a previous study, only two TFL1 genes, namely, MiTFL1 and MiTFL1a, were found in another mango variety, ‘Alphonso’, corresponding to MiTFL1-2 and MiTFL1-1, respectively, in the present study [35]. MiTFL1 genes contain four exons and three introns. The nucleotide lengths of the second and third exons were highly consistent among different MiTFL1 genes (Fig. 1a). The critical amino acids His88 and Asp144 of TFL1 proteins were also found in these four MiTFL1 proteins (Fig. 1b). The constructed phylogenetic tree indicated that the four MiTFL1 proteins belong to the TFL1 protein branch. In the tree, MiTFL1-1 and MiTFL1-2 were close to each other, whereas MiTFL1-3 and MiTFL1-4 were also located close to each other (Fig. 2).

The TFL1 gene expression pattern is related to floral development in most fruit trees. Citrus CsTFL1 is expressed only in floral organs [36]. P. mume PmTFL1 is expressed in the leaves, stems, and roots in the juvenile phase, whereas PmTFL1 has been detected only in leaf buds and young leaves at the mature stage [16]. The PpTFL1 gene of peach is mainly expressed in mature young leaves but is not found in mature leaves and flower organs [37]. EjTFL1-1 is mainly expressed in roots and leaf buds but is expressed at low levels in shoots, flower buds and flowers. EjTFL1-2 is mainly expressed in leaf buds, flowers, and fruits and is not expressed in other tissues [31]. In the present study, MiTFL1 genes showed different expression patterns compared to those reported above. Tissue expression analysis showed that MiTFL1-1, MiTFL1-2, MiTFL1-3 and MiTFL1-4 were expressed in both vegetative and reproductive tissues but differed in expression levels. Our results were similar to those found in other mango varieties, namely, ‘Alphonso’ and ‘Ratna’ [35]. These results indicated that the tissue expression pattern of the TFL1 gene was significantly different among different species.

Some studies have shown that the expression pattern of the TFL1 gene is different at different stages of floral development. For example, in pear, apple, and quince, the TFL1-1 and TFL1-2 genes are highly expressed in buds before floral differentiation, and their expression appears to decrease after floral differentiation [38]. In H. brasiliensis, HbTFL1-1, HbTFL1-2 and HbTFL1-3 expression increases gradually during inflorescence development, but the expression of HbCEN1 and HbCEN2 continuously decreases over this period [39]. The expression levels of EjTFL1-1 and EjTFL1-2 gradually decrease before floral bud differentiation and start to increase again during the flower-opening period [31]. In the mango variety ‘Alphonso’, MiTFL1 expression increases during the flowering induction period and subsequently decreases, whereas MiTFL1a expression remains low during the flowering period [35]. In this study, the expression of the MiTFL1 gene in mango was similar to that reported above. MiTFL1-1 and MiTFL1-2 were highly expressed in leaves at the floral induction stage, while MiTFL1-3 and MiTFL1-4 were highly expressed in leaves at the floral differentiation stage.

TFL1 homologous genes have similar functions in many species, and these functions include delaying the flowering time and maintaining the inflorescence meristem through suppression of AP1 and LFY expression [15]. Overexpression of apple MdTFL1 in Arabidopsis delays flowering time, and inhibition of the expression of the MdTFL1 gene by RNAi technology results in early flowering traits [34, 40]. The Japanese apricot PmTFL1 [16] and five rubber TFL1 genes showed the same function in transgenic Arabidopsis [39]. In the present study, four mango MiTFL1 genes had the same function of delaying flowering time in transgenic Arabidopsis. Moreover, the MiTFL1-1 and MiTFL1-3 transgenic lines exhibited abnormal flower organ phenotypes, such as missing petals, carpel development into a new inflorescence, curved pod growth and seed abortion. These results suggest that MiTFL1-1 and MiTFL1-3 are involved in flower organ development. Similar phenotypes were also found in chrysanthemum CmTFL1c- and Prunus PsTFL1-overexpressing transgenic Arabidopsis lines [41, 42].

Introduction of the exogenous TFL1 gene significantly downregulated the expression levels of the endogenous genes AtFT and AtAP1 in transgenic Arabidopsis, leading to delayed flowering. For example, overexpressing the HkTFL1 gene in Hemerocallis delayed flowering, and the expression levels of AtFT and AtAP1 in transgenic Arabidopsis were decreased compared with those in WT Arabidopsis [43]. The Chrysanthemum morifolium CmTFL1c gene negatively regulates flowering by inhibiting AtFT, AtLFY and AtAP1 expression [42]. Cucumber CsTFL1b also delays flowering in Arabidopsis and decreases and increases the expression levels of AtFT and AtFD, respectively [4]. In the present study, we found that four MiTFL1 genes downregulated AtFT and AtAP1 expression but upregulated AtFD expression in transgenic Arabidopsis.

In rice, the TFL1-like protein RICE CENTRORADIALIS (RCN) can directly interact with the 14-3-3 protein [44]. In this study, we also screened some proteins interacting with MiTFL1 using Y2H assays (Table S1). For example, the 14-3-3D protein was found to interact with MiTFL1-1 and MiTFL1-2, and this result was similar to that found in rice. Two other bHLH proteins were also found to interact with different MiTFL1 proteins and play a role in several processes, including growth, development, and the response to various stresses [45]. Additionally, other proteins that may interact with MiTFL1 proteins have been screened, and they are also involved in the stress response, plant growth and development. However, these interacting proteins must be further tested.

Conclusions

In the present study, four MiTFL1 genes were identified in mango. These proteins contained both the critical amino acids His88 and Asp144. Expression analysis showed that the MiTFL1 genes exhibited a similar expression pattern: the MiTFL1 genes were expressed in vegetative and reproductive tissues and were highly expressed in mature leaves during the flowering induction period and floral differentiation stage. Overexpression of the four MiTFL1 genes in Arabidopsis resulted in delayed flowering, whereas MiTFL1-1 and MiTFL1-3 overexpression affected floral organ development. Y2H analysis showed that the MiTFL1 proteins interact with bHLH and 14-3-3 proteins. These results provide preliminary evidence that MiTFL1 genes negatively regulate floral induction in mango, but their interaction mechanisms must be further validated.

Methods

Plant materials and growth conditions

The M. indica L. cv. ‘SiJiMi’ plants used in this study were planted in an orchard at Guangxi University, Nanning, Guangxi, China (22°502 N, 108°17′E). For tissue expression analysis, leaves, stems and flowers were collected from 17-year-old trees on 10 March 2021. For seasonal expression analysis, leaves (closer to the terminal bud) were collected once per month from 1 November 2020 (vegetative period) to 1 May 2021 (fruit development period). All the samples were immediately frozen in liquid nitrogen and stored at − 80 °C. The Arabidopsis ecotype Col-0 was maintained in our laboratory.

Isolation of MiTFL1 genes from mango

Total RNA was extracted from mango leaves using the RNAprep Pure Plant Kit (TianGen, Beijing, China) according to the manufacturer’s instructions. First-strand cDNA was synthesized from 1 μg of total RNA using M-MLV reverse transcriptase (TaKaRa, Dalian, China) according to the manufacturer’s instructions. Genomic DNA was extracted from mango leaves using the CTAB method. Four TFL1 genes were obtained from mango leaves and named MiTFL1-1, MiTFL1-2, MiTFL1-3 and MiTFL1-4. Specific primers (QTFL1-1u/d, QTFL1-2u/d, QTFL1-3u/d, and QTFL1-4u/d; Table S2) were designed to amplify MiTFL1 genes from genomic DNA and cDNA. The polymerase chain reaction (PCR) mixture contained 2.5 μl of 10× PCR buffer (with MgCl2+), 0.5 μl of 10 mM dNTPs (Sangon Biotech, Shanghai, China), 1 μl of each of the upstream and downstream primers (10 μM), 0.15 μl of TransTaq-T DNA polymerase (TianGen), 1 μl of genomic DNA (100 ng/μl) or cDNA (100 ng/μl), which served as the templates, and sterile water (25 μl). The PCR amplification conditions included an initial denaturation step of 4 min at 95 °C; 38 cycles of 95 °C for 40 s, 56 °C for 50 s, and 72 °C for N min (N = 1 min/kb); and a final extension at 72 °C for 10 min. The amplified fragments were cloned into the pMD18-T vector (Takara) and then sequenced.

Sequence analysis

Sequence analysis and amino acid prediction were performed using BioXM 2.6 software. IBS version 1.0 was used to generate exon-intron structures. The conserved domains were analyzed using the NCBI BLAST search engine (https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/Structure/cdd/wrpsb.cgi). The amino acid sequences of the FT1/TFL1 family were downloaded through BLAST searches of GenBank (http://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/BLAST/). Multiple sequence alignments of TFL1 proteins were analyzed using DNAMAN software. The phylogenetic tree was constructed using the neighbor-joining method in MEGA-Χ with 1000 bootstrap replicates.

Expression analysis of MiTFL1 genes

The expression of MiTFL1 genes was detected by quantitative real-time PCR. Total RNA from all the samples was extracted using the RNAprep Pure Plant Kit (TianGen) according to the manufacturer’s instructions. First-strand cDNA was synthesized and used as a template. Gene-specific primers (qTFL1-1u/d, qTFL1-2u/d, qTFL1-3u/d and qTFL1-4u/d; Fig. S4) were designed to distinguish the MiTFL1 genes. The MiActin1 gene of mango was used as the internal reference gene [46]. The PCR mixture contained 10 μl of SYBR Premix Ex Taq II (Takara), 1 μl of cDNA (100 ng/μl), 0.5 μl (10 μM) of the upstream and downstream primers, 0.8 μl of ROX Reference Dye II, and sterile water to obtain a total volume of 20 μl. The PCR amplification conditions included 30 s at 95 °C; 40 cycles of 95 °C for 5 s, 60 °C for 34 s, and 95 °C for 15 s; 60 °C for 1 min; and 95 °C for 15 s. The relative transcript abundances were estimated using the 2-ΔΔCt method [47]. The analysis of each sample included three biological replicates.

Subcellular localization of MiTFL1 proteins

The complete coding sequences of MiTFL1 genes without terminator codons were constructed into the P1300-GFP vector between the XbaI and BamHI restriction enzyme cleavage sites. The constructed vectors were transformed into Agrobacterium tumefaciens EHA105. The GFP fusion vectors and empty vector were subsequently transformed into onion epidermal cells via A. tumefaciens EHA105. 4′,6-Diamidino-2-phenylindole (DAPI) was used to visualize the nucleus. Images were captured using a high-resolution laser confocal microscope (TCS-SP8MP; Leica, Germany).

Plasmid construction and genetic transformation

The MiTFL1-overexpressing (OE) vectors were constructed by cloning the genes into the pBI121 binary vector using CaMV 35S as the promoter between the XbaI and XmaI restriction enzyme cleavage sites. The overexpression plasmids were transferred into the A. tumefaciens strain EHA105. The overexpression vectors and empty vector were subsequently transformed into WT A. thaliana using the floral-dip method [48]. The transgenic seeds were selected on 1/2 MS medium containing 50 mg/l kanamycin and confirmed by genomic PCR. The specific primers MiTFL1-1u/d, MiTFL1-2u/d, MiTFL1-3u/d and MiTFL1-4u/d were used to detect whether transformation of the MiTFL1 genes was successful. Homozygous T3 transgenic plants were used for subsequent experiments.

Phenotypic analyses

Wild-type and empty vector-transformed Arabidopsis plants were used as controls. Several phenotypic indexes, including the bolting time, flowering time, time from bolting to flowering, and rosette leaves, were measured. To detect the expression levels of the MiTFL1 genes and some flowering-related genes in transgenic and control plants, 30-day-old seedlings of both transgenic and WT Arabidopsis plants were collected for total RNA extraction. Total RNA was extracted, and first-strand cDNA was synthesized as described above. Semiquantitative RT-PCR was performed to determine the expression levels of the MiTFL1 genes in the transgenic and control lines. The PCR amplification conditions comprised an initial denaturation step of 2 min at 95 °C; 30 cycles of 95 °C for 30 s, 56 °C for 30 s, and 72 °C for 30 s; and a final extension at 72 °C for 5 min. The PCR products were electrophoretically separated on a 1.8% agarose gel. The expression levels of some endogenous flowering-related genes in transgenic and control Arabidopsis lines were detected by qRT-PCR using the reaction system and conditions described above. Arabidopsis AtACTIN2 was used as the internal reference gene for qRT-PCR analysis. All the primers used in this study are listed in Table S1.

Proteins interacting with the MiTFL1 proteins

A cDNA library from M. indica L. cv. SiJiMi leaves and flowering organs was constructed using the Yeast Two-Hybrid Library Construction Kit (Clontech, Dalian, China). Y2H assays were performed according to the Yeastmaker™ Yeast Transformation System 2 protocol (Clontech). The complete coding sequences of the MiTFL1 genes were inserted into the pGBKT7 vector between the NdeI and EcoRI restriction enzyme cleavage sites. The pGBKT7-bait plasmid was transformed into competent Y2H Gold yeast cells, which were diluted 10_1 and 10_2 and coated on SDO/−Trp, SDO/−Trp/X-alpha-Gal and SDO/−Trp/X-Alpha-Gal/AbA media. The transcriptional activity and toxicity were verified using this method.

The interacting proteins were identified by screening a DNA library on QDO/−Trp/−Leu/−His/−Ade culture medium. The plasmids of the interacting proteins were extracted and used for further verification of the actual interactions. Y2H Gold yeast cells containing the pGBKT7-bait plasmid and Y187 yeast cells containing the candidate prey were mixed and cultured in liquid medium containing 2 × YPDA at 30 °C and 200 rpm for 20–24 h. The mixture was then coated on DDO/−Trp/−Leu/X and QDO/X/A media and cultured for 3–5 days. Blue colonies on the media indicated a positive interaction. The AbA concentration was 500 μg/ml, and the X-α-gal concentration was 200 ng/ml. Y2H Gold (pGBKT7-53) and Y187 (pGADT7-T) served as positive controls, and Y2H Gold (pGBKT7-LAM) and Y187 (pGADT7-T) were used as negative controls.

Statistical analysis

SPSS 19.0 statistical software (SPSS Inc., Chicago, IL, United States) was used for the statistical analyses.

Availability of data and materials

All the data generated or analyzed during this study are included in this published article and its supplementary information files. The datasets generated in this study are available in GenBank (http://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/Genbank), and the accession numbers are as follows: AtTFL1 (NP_196004.1), CsTFL1 (NP_001275848), DlTFL1 (AHY24028.1), JrTFL1 (XP_018811176.1), MdTFL1 (NP_001280887.1), PpTFL1 (NP_001289244.1), VvTFL1A (NP_001267929.1), MdCEN (NP_001280940.1), PvCEN protein 1 (XP_031269481.1), TcCEN (XP_017973069.1), HmTFL1 (MF374628.1), GmTFL1.3 (FJ573238.1), FcTFL1-2 (AB746842.1), CcTFL1 (XP_006430226.1), ClTFL1 (AWW25018.1), FcTFL1 (BAX00857.1), NsTFL1 (XP_009766168.1), PmTFL1 (AEO72021.1), PyTFL1 (AEO72023.1), VvTFL1C (NP_001267933.1), ZjTFL1 (XP_015898753.1), AcFT (AJA40932.1), AtFT (BAA77838.1), EdFT (AMB72867.1), MiFT (AGA19021.1), VvFT (NP_001267907.1), AtMFT (OAP13671.1), ClMFT (AWW25016.1), DlMFT (AUG98253.1), and VvMFT (NP_001267935.1).

Abbreviations

CTAB:

Hexadecyltrimethylammonium bromide

qRT-PCR:

Quantitative reverse transcription-polymerase chain reaction

DAPI:

6-diamidino-2-phenylindole

X-Alpha-Gal:

5-bromo-4-chloro-3-indoxyl-α-D-galactopyranoside

AbA:

Aureobasidin A

References

  1. Larsson AS, Landberg K, Meeks-Wagner DR. The TERMINAL FLOWER2 (TFL2) gene controls the reproductive transition and meristem identity in Arabidopsis thaliana. Genetics. 1998;149(2):597–605. https://0-doi-org.brum.beds.ac.uk/10.1093/genetics/149.2.597.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Prudencio AS, Hoeberichts FA, Dicenta F, Martínez-Gómez P, Sánchez-Pérez R. Identification of early and late flowering time candidate genes in endodormant and ecodormant almond flower buds. Tree Physiol. 2020. https://0-doi-org.brum.beds.ac.uk/10.1093/treephys/tpaa151.

  3. He YH. Chromatin regulation of flowering. Trends Plant Sci. 2012;17(9):556–62. https://0-doi-org.brum.beds.ac.uk/10.1016/j.tplants.2012.05.001.

    Article  CAS  PubMed  Google Scholar 

  4. Zhao W, Gu R, Che G, Cheng Z, Zhang X. CsTFL1b may regulate the flowering time and inflorescence architecture in cucumber (Cucumis sativus L.). Biochem Biophys Res Commun. 2018;499(2):307–13.

    Article  CAS  Google Scholar 

  5. Luccioni L, Krzymuski M, Sanchezlamas M, Karayekov E, Cerdán PD, Casal JJ. CONSTANS delays Arabidopsis flowering under short days. Plant J. 2018;97(5):923–32. https://0-doi-org.brum.beds.ac.uk/10.1111/tpj.14171.

    Article  CAS  Google Scholar 

  6. Wells RS, Adal AM, Bauer L, Najafianashrafi E, Mahmoud SS. Cloning and functional characterization of a floral repressor gene from Lavandula angustifolia. Planta. 2020;251(2):41. https://0-doi-org.brum.beds.ac.uk/10.1007/s00425-019-03333-w.

    Article  CAS  PubMed  Google Scholar 

  7. Sheldon CC, Rouse DT, Finnegan EJ, Peacock WJ, Dennis ES. The molecular basis of vernalization: the central role of FLOWERING LOCUS C (FLC). Proc Natl Acad Sci U S A. 2000;97(7):3753–8. https://0-doi-org.brum.beds.ac.uk/10.1073/pnas.97.7.3753.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Won SY, Jung JA, Kim JS. Genome-wide analysis of the MADS-box gene family in Chrysanthemum. Comput Biol Chem. 2021;90:107424. https://0-doi-org.brum.beds.ac.uk/10.1016/j.compbiolchem.2020.107424.

    Article  CAS  PubMed  Google Scholar 

  9. Ream TS, Woods DP, Amasino RM. The molecular basis of vernalization in different plant groups. Proc Natl Acad Sci U S A. 2011;77:105–15.

    Google Scholar 

  10. Sharma B, Batz TA, Kaundal R, Kramer EM, Larson RB. Developmental and molecular changes underlying the vernalization-induced transition to flowering in Aquilegia coerulea (James). Genes. 2019;10(10):734. https://0-doi-org.brum.beds.ac.uk/10.3390/genes10100734.

    Article  CAS  PubMed Central  Google Scholar 

  11. Shannon S, Meeks-Wagner DR. A mutation in the Arabidopsis TFL1 gene affects inflorescence meristem development. Plant Cell. 1991;3(9):877–92. https://0-doi-org.brum.beds.ac.uk/10.2307/3869152.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Carmona MJ, Calonje M, Martínez-Zapater JM. The FT/TFL1 gene family in grapevine. Plant Mol Biol. 2007;63(5):637–50. https://0-doi-org.brum.beds.ac.uk/10.1007/s11103-006-9113-z.

    Article  CAS  PubMed  Google Scholar 

  13. Yang ZL, Chen L, Kohnen MV, Xiong B, Zhen X, Liao J, et al. Identification and characterization of the PEBP family genes in Moso Bamboo ( Phyllostachys heterocycla ). Sci Rep. 2019;9:14998.

    Article  Google Scholar 

  14. Liu X, Zhang J, Xie DY, Franks RG, Xiang QJ. Functional characterization of terminal flower 1 homolog in Cornus canadensis by genetic transformation. Plant Cell Rep. 2019;38(3):333–43. https://0-doi-org.brum.beds.ac.uk/10.1007/s00299-019-02369-2.

    Article  CAS  PubMed  Google Scholar 

  15. Bradley D, Ratcliffe O, Vincent C, Carpenter R, Coen E. Inflorescence commitment and architecture in Arabidopsis. Science. 1997;275(5296):80–3. https://0-doi-org.brum.beds.ac.uk/10.1126/science.275.5296.80.

    Article  CAS  PubMed  Google Scholar 

  16. Esumi T, Kitamura Y, Hagihara C, Yamane H, Tao R. Identification of a TFL1 ortholog in Japanese apricot (Prunus mume Sieb. et Zucc.). Sci Hortic. 2010;125(4):608–16. https://0-doi-org.brum.beds.ac.uk/10.1016/j.scienta.2010.05.016.

    Article  CAS  Google Scholar 

  17. Moraes TS, Dornelas MC, Martinelli AP. FT/TFL1: calibrating plant architecture. Front Plant Sci. 2019;10:97. https://0-doi-org.brum.beds.ac.uk/10.3389/fpls.2019.00097.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Périlleux C, Bouché F, Randoux M, Orman-Ligeza B. Turning meristems into fortresses. Trends Plant Sci. 2019;24(5):431–42. https://0-doi-org.brum.beds.ac.uk/10.1016/j.tplants.2019.02.004.

    Article  CAS  PubMed  Google Scholar 

  19. Wickland DP, Hanzawa Y. The FLOWERING LOCUS T/TERMINAL FLOWER 1 gene family: functional evolution and molecular mechanisms. Mol Plant. 2015;8(7):983–97. https://0-doi-org.brum.beds.ac.uk/10.1016/j.molp.2015.01.007.

    Article  CAS  PubMed  Google Scholar 

  20. Conti L, Bradley D. TERMINAL FLOWER 1 is a mobile signal controlling Arabidopsis architecture. Plant Cell. 2007;19(3):767–78. https://0-doi-org.brum.beds.ac.uk/10.1105/tpc.106.049767.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Flachowsky H, Peil A, Sopanen T, Elo A, Hanke V. Overexpression of BpMADS4 from silver birch (Betula pendula Roth.) induces early-flowering in apple (Malus aldomestica Borkh.). Plant Breed. 2007;126(2):137–45. https://0-doi-org.brum.beds.ac.uk/10.1111/j.1439-0523.2007.01344.x.

    Article  CAS  Google Scholar 

  22. Freiman A, Shlizerman L, Golobovitch S, Yablovitz Z, Korchinsky R, Cohen Y, et al. Development of a transgenic early flowering pear (Pyrus communis L.) genotype by RNAi silencing of PcTFL1-1 and PcTFL1-2. Planta. 2012;235(6):1239–51. https://0-doi-org.brum.beds.ac.uk/10.1007/s00425-011-1571-0.

    Article  CAS  PubMed  Google Scholar 

  23. Luo C, Yu HX, Fan Y, Zhang XJ, He XH. Research advance on the flowering mechanism of mango. Acta Hortic. 2019;1244(2):17–22. https://0-doi-org.brum.beds.ac.uk/10.17660/ActaHortic.2019.1244.2.

    Article  Google Scholar 

  24. Wang P, Luo YF, Huang JF, Gao SH, Zhu GP, Dang ZG, et al. The genome evolution and domestication of tropical fruit mango. Genom Biol. 2020;26(1):60. https://0-doi-org.brum.beds.ac.uk/10.1186/s13059-020-01959-8.

    Article  CAS  Google Scholar 

  25. Wei JY, Liu DB, Liu GY, Tang J, Chen YY. Molecular cloning, characterization, and expression of MiSOC1: a homolog of the flowering gene SUPPRESSOR OF OVEREXPRESSION OF CONSTANS1 from mango (Mangifera indica L). Front Plant Sci. 2016;7:1758. https://0-doi-org.brum.beds.ac.uk/10.3389/fpls.2016.01758.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Yu HX, Luo C, Fan Y, Zhang XJ, Huang F, Li M, et al. Isolation and characterization of two APETALA1-Like genes from mango (Mangifera indica L.). Sci Hortic (Amsterdam, Neth.). 2020;259:108814.

    Article  CAS  Google Scholar 

  27. Fan ZY, He XH, Fan Y, Yu HX, Wang YH, Xie XJ, et al. Isolation and functional characterization of three MiFTs genes from mango. Plant Physiol Bioch. 2020;155:169–76. https://0-doi-org.brum.beds.ac.uk/10.1016/j.plaphy.2020.07.009.

    Article  CAS  Google Scholar 

  28. Liu Y, Luo C, Zhang XJ, Lu XX, Yu HX, Xie XJ, et al. Overexpression of the mango MiCO gene delayed fowering time in transgenic Arabidopsis. Plant Cell Tissue Organ Cult. 2020;143(1):219–28. https://0-doi-org.brum.beds.ac.uk/10.1007/s11240-020-01894-3.

    Article  CAS  Google Scholar 

  29. Hanzawa Y, Money T, Bradley D. A single amino acid converts a repressor to an activator of flowering. Proc Natl Acad Sci U S A. 2005;102(21):7748–53. https://0-doi-org.brum.beds.ac.uk/10.1073/pnas.0500932102.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Liu B, Watanabe S, Uchiyama T, Kong F, Kanazawa A, Xia Z, et al. The soybean stem growth habit gene Dt1 is an ortholog of Arabidopsis TERMINAL FLOWER 1. Plant Physiol. 2010;153(1):198–210. https://0-doi-org.brum.beds.ac.uk/10.1104/pp.109.150607.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Jiang YY, Zhu YM, Zhang L, Su WB, Peng JR, Yang XH, et al. EjTFL1 genes promote growth but inhibit flower bud differentiation in loquat. Front Plant Sci. 2020;11:576. https://0-doi-org.brum.beds.ac.uk/10.3389/fpls.2020.00576.

    Article  PubMed  PubMed Central  Google Scholar 

  32. Wu L, Li F, Deng Q, Zhang S, Liu G. Identification and characterization of the FLOWERING LOCUS T/TERMINAL FLOWER 1 gene family in petunia. DNA Cell Biol. 2019;38(9):982–95. https://0-doi-org.brum.beds.ac.uk/10.1089/dna.2019.4720.

    Article  CAS  PubMed  Google Scholar 

  33. Zhang X, Wang C, Pang C, Wei H, Wang H, Song M, et al. Characterization and functional analysis of PEBP family genes in upland cotton (Gossypium hirsutum L.). PLoS ONE. 2016;11(8):e0161080.

    Article  Google Scholar 

  34. Kotoda N, Iwanami H, Takahashi S, Abe K. Antisense expression of MdTFL1, a TFL1-like gene, reduces the juvenile phase in apple. J Am Soc Hortic Sci. 2006;131(1):74–81. https://0-doi-org.brum.beds.ac.uk/10.21273/JASHS.131.1.74.

    Article  CAS  Google Scholar 

  35. Vyavahare SN, Krishna B, Joshi SS, Chaudhari RS, Sane PV. Characterization of mango flowering Locus T (FT) and TERMINAL flower 1 (TFL1) gene. J Am Soc Hortic Sci. 2017;1183(1183):113–24. https://0-doi-org.brum.beds.ac.uk/10.17660/ActaHortic.2017.1183.16.

    Article  Google Scholar 

  36. Pillitteri LJ, Lovatt CJ, Walling LL. Isolation and characterization of a TERMINAL FLOWER homolog and its correlation with juvenility in citrus. Plant Physiol. 2004;135(3):1540–51. https://0-doi-org.brum.beds.ac.uk/10.1104/pp.103.036178.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Chen Y, Jiang P, Thammannagowda S, Liang H, Wilde HD. Characterization of peach TFL1 and comparison with FT/TFL1 gene families of the rosaceae. J Am Soc Hortic Sci. 2013;138(1):12–7. https://0-doi-org.brum.beds.ac.uk/10.21273/JASHS.138.1.12.

    Article  CAS  Google Scholar 

  38. Esumi T, Tao R, Yonemori K. Isolation of LEAFY and TERMINAL FLOWER 1 homologues from six fruit tree species in the subfamily maloideae of the rosaceae. Sex Plant Reprod. 2005;17(6):277–87. https://0-doi-org.brum.beds.ac.uk/10.1007/s00497-004-0239-3.

    Article  CAS  Google Scholar 

  39. Bi Z, Tahir AT, Huang H, Hua Y. Cloning and functional analysis of five TERMINAL FLOWER 1/CENTRORADIALIS-like genes from Hevea brasiliensis. Physiol Plant. 2019;166(2):612–27.

    Article  CAS  Google Scholar 

  40. Kotoda N, Wada M. MdTFL1, a TFL1 -like gene of apple, retards the transition from the vegetative to reproductive phase in transgenic Arabidopsis. Plant Sci. 2005;168(1):95–104. https://0-doi-org.brum.beds.ac.uk/10.1016/j.plantsci.2004.07.024.

    Article  CAS  Google Scholar 

  41. Wang Y, Pijut PM. Isolation and characterization of a TERMINAL FLOWER 1 homolog from Prunus serotina Ehrh. Tree Physiol. 2013;33(8):855–65. https://0-doi-org.brum.beds.ac.uk/10.1093/treephys/tpt051.

    Article  CAS  PubMed  Google Scholar 

  42. Gao Y, Gao Y, Wu Z, Bu X, Fan M, Zhang Q. Characterization of TEMINAL FLOWER 1 homologs CmTFL1c gene from Chrysanthemum morifolium. Plant Mol Biol. 2019;99(6):587–601. https://0-doi-org.brum.beds.ac.uk/10.1007/s11103-019-00838-6.

    Article  CAS  PubMed  Google Scholar 

  43. Liu Y, Gao Y, Yuan L, Zhang Q. Functional characterization and spatial interaction of TERMINAL FLOWER 1 in Hemerocallis. Sci Hortic. 2019;253:154–62. https://0-doi-org.brum.beds.ac.uk/10.1016/j.scienta.2019.03.036.

    Article  CAS  Google Scholar 

  44. Miho KS, Rie KI, Chiaki OT, Chojiro K, Misa NF, Izuru O, et al. TFL1-like proteins in rice antagonize rice FT-like protein in inflorescence development by competition for complex formation with 14-3-3 and FD. Plant Cell Physiol. 2018;59(3):458–68. https://0-doi-org.brum.beds.ac.uk/10.1093/pcp/pcy021.

    Article  CAS  Google Scholar 

  45. Niu X, Guan YX, Chen SK, Li HF. Genome-wide analysis of basic helix-loop-helix (bHLH) transcription factors in Brachypodium distachyon. BMC Genomics. 2017;18(1):619. https://0-doi-org.brum.beds.ac.uk/10.1186/s12864-017-4044-4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Luo C, He XH, Chen H, Hu Y, Ou SJ. Molecular cloning and expression analysis of four actin genes (MiACT) from mango. Biol Plant. 2013;57(2):238–44. https://0-doi-org.brum.beds.ac.uk/10.1007/s10535-012-0278-9.

    Article  CAS  Google Scholar 

  47. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR, and the 2-ΔΔCT method. Methods. 2001;25(4):402–8.

    Article  CAS  Google Scholar 

  48. Clough SJ, Bent AF. Floral dip: a simplifed method for Agrobacterium- mediated transformation of Arabidopsis thaliana. Plant J. 1998;16(6):735–43. https://0-doi-org.brum.beds.ac.uk/10.1046/j.1365-313x.1998.00343.x.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

Not applicable.

Funding

This research was supported by grants from the National Natural Science Foundation of China (31860541 and 31660561), Science and Technology Major Projects of Guangxi (GuiKeAA17204026 and GuiKe AA17204097), State Key Laboratory for Conservation and Utilization of Subtropical Agro-bioresources (SKLCUSA-c201901) and Innovation Team of Guangxi Mango Industry Project (nycytxgxcxtd- 07-02). The funders had no role in the design of the study; collection, analysis, and interpretation of the data; and writing of the manuscript but provided financial support.

Author information

Authors and Affiliations

Authors

Contributions

CL and XHH designed and procured funding for this study. YHW performed the experiments. HXY, XM, YF, ZYF, XJX, and YL helped with the experiments and data analysis. YHW and CL wrote the manuscript. All the authors have read and approved the manuscript.

Corresponding author

Correspondence to Cong Luo.

Ethics declarations

Ethics approval and consent to participate

All the materials of this project are given by Guangxi University (Nanning, China). The experimental research on plants performed in this study complies with institutional, national and international guidelines.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1

: Supplementary Table 1. Four TFL1 interacting protein through yeast two hybrid.

Additional file 2

: Supplementary Table 2. Primers used in this study.

Additional file 3

: Supplement Figure 1. Source data for Figs. 5 and 6. (a) The red frame indicates the source data in Fig. S1a displayed for Fig. 5a-a1. (b) The source data in Fig. S1b displayed for Fig. 5b-b1. (c) The source data in Fig. S1c displayed for Fig. 6a-a1. (d) The source data in Fig. S1d displayed for Fig. 6b-b1.

Additional file 4

: Supplement Figure 2. The autoactivation and toxicity of pGBKT7-MiTFL1s vector, concentrations of 10− 1. Yeast bait expression vectors of MiTFL1-1, MiTFL1-2, MiTFL1-3 and MiTFL1-4 were constructed by double enzyme digestion method, and transferred into Y2H Gold yeast, which were cultured in SDO/−Trp, SDO/X and SDO/X/A medium. Y2H Gold (pGBKT7-53) served as positive controls, and Y2H Gold (pGBKT7-lam) were used as negative controls. The results showed that the four pGBKT7-MiTFL1s had no autoactivation and toxicity.

Additional file 5

: Supplement Figure 3. Comparison of cDNA sequences of four MiTFL1 genes in two cultivars of ‘Alphonso’ and ‘SiJiMi’. (A) The comparative similarity between the two cultivars of MiTFL1-1 gene was 99.4%. (B) The comparative similarity between the two cultivars of MiTFL1-2 gene was 100.0%. (C) The comparative similarity between the two cultivars of MiTFL1-3 gene was 98.5%. (D) The comparative similarity between the two cultivars of MiTFL1-4 gene was 98.3%.

Additional file 6

: Supplement Figure 4. The primer design site of MiTFL1s gene was used for qRT-PCR. cDNA sequence comparison of the four genes, where the yellow shaded part represents the start codon and the red shaded part represents the stop codon. The red sequence represents the upstream primer sequence and the blue sequence represents the downstream primer sequence. The black part indicates a similarity of 100%.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, YH., He, XH., Yu, HX. et al. Overexpression of four MiTFL1 genes from mango delays the flowering time in transgenic Arabidopsis. BMC Plant Biol 21, 407 (2021). https://0-doi-org.brum.beds.ac.uk/10.1186/s12870-021-03199-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s12870-021-03199-9

Keywords