Next Article in Journal
IMPDH Inhibition Decreases TERT Expression and Synergizes the Cytotoxic Effect of Chemotherapeutic Agents in Glioblastoma Cells
Previous Article in Journal
High-Yield α-Synuclein Purification and Ionic Strength Modification Pivotal to Seed Amplification Assay Performance and Reproducibility
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Natural Variation in the Promoter of GmSPL9d Affects Branch Number in Soybean

1
School of Agronomy, Anhui Agricultural University, Hefei 230036, China
2
The National Key Facility for Crop Gene Resources and Genetic Improvement (NFCRI), Key Laboratory of Crop Gene Resource and Germplasm Enhancement (MOA), Institute of Crop Sciences, Chinese Academy of Agricultural Sciences, Beijing 100081, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2024, 25(11), 5991; https://doi.org/10.3390/ijms25115991
Submission received: 23 April 2024 / Revised: 17 May 2024 / Accepted: 25 May 2024 / Published: 30 May 2024
(This article belongs to the Section Molecular Plant Sciences)

Abstract

:
The branch number is a crucial factor that influences density tolerance and is closely associated with the yield of soybean. However, its molecular regulation mechanisms remain poorly understood. This study cloned a candidate gene GmSPL9d for regulating the soybean branch number based on the rice OsSPL14 homologous gene. Meanwhile, the genetic diversity of the GmSPL9d was analyzed using 3599 resequencing data and identified 55 SNP/InDel variations, which were categorized into seven haplotypes. Evolutionary analysis classified these haplotypes into two groups: GmSPL9d H-I and GmSPL9d H-II. Soybean varieties carrying the GmSPL9d H-II haplotype exhibited a significantly lower branch number compared with those carrying the GmSPL9d H-I haplotype. Association analysis between the variation sites and branch number phenotypes revealed a significant correlation between the promoter variations and the branch number. Promoter activity assays demonstrated that the GmSPL9d H-II promoter displayed significantly higher activity than the GmSPL9d H-I promoter. Transgenic experiments confirmed that the plants that carried the GmSPL9d H-II promoter exhibited a significantly lower branch number compared with those that carried the GmSPL9d H-I promoter. These findings indicate that the variation in the GmSPL9d promoter affected its transcription level, leading to differences in the soybean branch number. This study provides valuable molecular targets for improving the soybean plant structure.

1. Introduction

Soybean is a crucial economic crop and a primary source of plant-based protein and oil for human consumption [1,2]. To meet the growing demand for soybean, it is essential to increase its yield. The yield of soybean is influenced by the plant architecture and planting density, with varieties that have a lower branch number typically demonstrating higher tolerance to dense planting [3,4]. Therefore, optimizing the plant structure is crucial for enhancing the soybean yield [5,6]. Only a few genes regulating soybean branch number have been identified, and genome-wide association analysis suggests that GmBRC1 is a candidate gene for regulating the soybean branch number [7]. Natural variation in the Dt2 promoter affects the transcription level of GmAP1s, significantly impacting the soybean branch number [8]. Thus, discovering genes that regulate soybean branching is vital for improving the density and yield.
SQUAMOSA promoter-binding protein-like (SPL) transcription factors form a major family of plant-specific transcription factors related to flower and vegetative development [9]. IPA1 is a key regulatory factor for rice tiller formation and yield [10,11], and belongs to the SPL gene family. The SPL gene family was first identified in Arabidopsis thaliana, which possesses a highly conserved SBP domain containing 76 amino acids [12]. The SPL gene family plays a significant role in plant growth and development, including regulating branch formation, leaf primordium development, floral organ formation, and fruit ripening [13,14]. Among these, the SPL gene family has been extensively studied in the process of plant branch formation and development, such as the overexpression of SPL10 weakening the apical dominance of Arabidopsis and promoting a higher branch number [15]. SPL9 binds to DELLA to facilitate axillary bud formation in Arabidopsis [16], and SPL15 directly regulates the transcription level of BRC1, inhibiting Arabidopsis branch production [17]. The miR156-SPL module is an essential pathway that regulates soybean branch development, and miR156b regulates branch formation and development by cutting the transcript of GmSPL9d [18]. In recent years, the function of GmSPL9d in regulating the plant branch number has been reported [18,19], but its natural variation remains unexplored. Therefore, investigating the natural variation of the GmSPL9d gene is of significant importance.
This study utilized 3,599 soybean resequencing data to analyze the genetic diversity of the GmSPL9d gene. It was found that natural variations in the promoter region affected the soybean branch number. Transient transcriptional activity assays and GUS tissue staining experiments confirmed the impact of natural variations on promoter activity. Furthermore, transgenic Arabidopsis experiments demonstrated that natural variations in the promoter region influenced the branch number by affecting its transcriptional levels. The study also investigated the selection effects of different alleles of the GmSPL9d gene in different latitudinal regions, providing molecular targets for soybean plant structure improvement.

2. Results

2.1. Identification of Soybean SPL Gene Family Members

To identify members of the SPL (SQUAMOSA Promoter-Binding Protein-Like) gene family, soybean genomic data were obtained from the Phytozome database. Utilizing the Pfam (PF03110) domain, 61 candidate SPL genes were initially identified. These candidates were further analyzed using the CDD, SMART, and Pfam databases to confirm the presence of the SBP domain, eliminate sequences lacking a complete SBP domain, and exclude distantly related sequences. Ultimately, 41 sequences with typical SBP domains were obtained. A phylogenetic tree was constructed using the selected SPL genes and previously reported rice and Arabidopsis SPL genes (Figure 1a and Figure S1a). The results show that GmSPL9s had the highest homology with the IPA1 (OsSPL14) gene. Notably, IPA1 plays a vital role in regulating plant architecture and yield, particularly in controlling the tiller number during the vegetative growth stage of rice [10,11]. Furthermore, the expression pattern of the soybean SPL family was analyzed, and it was found that GmSPL9d exhibited a higher expression in apical meristems (Figure 1b). Meanwhile, the expression pattern of GmSPL9s was further analyzed, and the results show that GmSPL9d was highly expressed in SAM due to soybean branches developing from SAM (Figure S1b). Therefore, GmSPL9d may play a crucial role in regulating the soybean branch number.

2.2. GmSPL9d Expression Pattern and Subcellular Localization

Validation analysis of the expression pattern of GmSPL9d was conducted using qRT-PCR. The findings revealed that GmSPL9d exhibited a high expression in the shoot apical meristem (SAM) tissue (Figure 2a). To determine the subcellular localization of GmSPL9d, a GFP fusion protein was created by fusing GmSPL9d with GFP and subsequently expressed in Nicotiana benthamiana leaves. The results indicated that GmSPL9d was localized in the nucleus, which aligned with the characteristic localization of the SPL gene family transcription factors (Figure 2b).

2.3. Haplotype Analysis of GmSPL9d

The genetic diversity of the GmSPL9d gene analysis results show that a total of 55 SNP/InDel variation sites were identified based on 3599 re-sequencing data (MAF > 0.05). These included 45 variations in the promoter region, 1 in the 5′ UTR, 1 in the 3′ UTR, 3 synonymous mutations, 5 variations in the intron, and a total of 7 haplotypes (Figure 3). Evolutionary analysis categorized the haplotypes into two types: GmSPL9d H-I (including Hap 1, Hap 2, Hap 3, Hap 4, and Hap 5) and GmSPL9d H-II (including Hap 6 and Hap 7) (Figure 4a). Through correlation analysis between the variation sites and branch number phenotype, it was observed that a significant number of variations in the promoter region were associated with the branch number (Table S2). Further analysis of the branch number phenotype of 561 materials revealed that the average branch numbers in 2017 and 2018 for the GmSPL9d H-I types were 3.72 and 3.71, respectively, whereas for the GmSPL9d H-II types, the average branch numbers in 2017 and 2018 were 1.95 and 2.25, respectively, representing a 43.5% decrease compared with GmSPL9d H-I (Figure 4b,c). Meanwhile, we analyzed the frequency and geographical distribution of two haplotypes using resequencing data from 1037 landraces and 2267 elites. In the landraces, the main haplotype was GmSPL9d H-I (93.5%), with a small percentage of GmSPL9d H-II (6.5%). However, in the elites, the proportion of the GmSPL9d H-II haplotype increased to 14.8% (Figure S2). Further analysis of the soybean varieties in China revealed a higher proportion of the GmSPL9d H-II haplotype in high-latitude ecological regions (Figure S3). Previous studies demonstrated that miR156b regulates the branch number by targeting GmSPL9d (Sun et al., 2019); however, no variation was detected in the 20 bp base sequence at the target site (Figure S4). In conclusion, natural variations in the promoter region of GmSPL9d may represent an essential regulatory site for controlling the branch number in soybean.

2.4. The Activity of the GmSPL9d Promoter Affected the Branch Number in Soybean

To investigate the mechanism by which GmSPL9d promoter variation influences the branch number, we initially performed a correlation analysis between the expression level of GmSPL9d in 82 natural germplasm AM tissues and the branch number phenotype. The results revealed a significant negative correlation between the expression level of GmSPL9d and the branch number (Figure S5). In addition, most GmSPL9d H-I had a low expression and a high branch number, while most GmSPL9d H-II had a high expression and a low branch number (Figure 5a). Thus, it can be inferred that natural variation in the GmSPL9d promoter might regulate the branch number by impacting the gene’s expression level. To validate this hypothesis, we assessed the promoter activity of two haplotypes: GmSPL9d H-I and GmSPL9d H-II. The results demonstrated that the promoter activity of GmSPL9d H-II was significantly higher than that of GmSPL9d H-I (Figure 5b,c). Furthermore, a GUS histochemical staining experiment confirmed that the GUS enzyme activity in the SAM and AM tissues of Arabidopsis thaliana with the GmSPL9d H-II promoter type was notably higher than that with the GmSPL9d H-I promoter type (Figure 5d–f). These findings indicate that natural variation in the GmSPL9d promoter impacted the promoter activity, thereby influencing the branch number.

2.5. GmSPL9d H-II Haplotype Inhibited Arabidopsis Branching

To further confirm the role of the two haplotypes in regulating the branch number, we transformed the CDS encoding region of GmSPL9d driven by 35S, pGmSPL9d H-I, and pGmSPL9d H-II promoters into Col-0 Arabidopsis (Figure 6a and Figure S6). The expression level of GmSPL9d H-II transgenic lines in the axillary meristem was significantly higher than that of GmSPL9d H-I, and both haplotypes exhibited significantly lower expression levels compared with the over-expression plants (Figure 6b). Consistently, the branch number in GmSPL9d H-II was significantly lower than in GmSPL9d H-I, and both haplotypes had a significantly higher branch number compared with the over-expression plants (Figure 6c,d). Notably, there was no significant difference in the branch numbers between the GmSPL9d H-I and Col-0 plants. These results indicated that GmSPL9d could only inhibit Arabidopsis branching when expressed at a high level, and furthermore, the GmSPL9d H-II haplotype exhibited a higher expression level, thereby suppressing Arabidopsis branching.

3. Discussion

Transcription factors known as SPLs have been extensively researched in plants and are known to play crucial roles in plant growth and development [13,20,21,22,23,24]. SPLs form a small gene family, with 16 SPL genes in Arabidopsis and 19 in rice [25]. In this study, we identified 41 SPL genes in soybean. Compared with Arabidopsis and rice, the number of soybean SPL genes is higher. The genome-wide duplication (WGD) events in soybean evolutionary history may explain this phenomenon. WGD is very common in plants, leading to double gene copies in the genome. The functional divergence of duplicate gene pairs is the source of new genes. Soybean has been reported to experience at least two WGD events, approximately 59 and 13 million years ago, resulting in a highly duplicated genome, with nearly 75% of the genes present in multiple copies [26]. Therefore, a greater expansion of SPL genes may have occurred in the soybean genome than in other species. It was further discovered that GmSPL9s (GmSPL9a, GmSPL9b, GmSPL9c, GmSPL9d) exhibits the highest homology with IPA1 (OsSPL14). IPA1, also known as a “star gene” in rice, is a key regulator of tillering, stalk strength, and panicle traits (Jiao et al., 2010; Miura et al., 2010). Therefore, GmSPL9s may play an important role in the regulation of soybean plant phenotypes. As a member of the soybean SPL gene family, GmSPL9d is highly expressed in the shoot apical meristem (SAM), where the shoot apical meristem (SAM) contains undifferentiated stem cells that are responsible for the initiation of aboveground organs [27]. Thus, GmSPL9d may play a critical role in multiple traits. As expected, previous studies do validate this. For instance, Arbuscular mycorrhizal fungus Rhizophagus irregularis alleviates drought stress in soybean by overexpressing the GmSPL9d gene by promoting the photosynthetic apparatus and regulating the antioxidant system [28]. The miR156b-GmSPL9d module modulates nodulation by targeting multiple core nodulation genes in soybean [29]. In addition, the miR156b-GmSPL9d module also regulates the branch number in soybean [18]. Despite numerous studies that demonstrated the regulatory role of GmSPL9d in the soybean branch number, the impact of natural variation on gene expression and function remains unexplored [18,19]. Therefore, investigating the natural variation in GmSPL9d contributes to the expansion and enhancement of our understanding of the molecular mechanisms underlying the soybean branch number.
By conducting a genetic diversity analysis of GmSPL9d, two haplotypes, namely, GmSPL9d H-I and GmSPL9d H-II, were identified that exhibited significant differences in the branch number. In rice, a natural variation in the IPA1 coding region (C-T) resulted in the inability of miR156 to recognize and target it, leading to differences in the branch number [10,11]. In this study, no natural variation was found in the miR156b targeting sequence, and no non-synonymous mutations were detected in the coding region of GmSPL9d. However, a considerable number of SNP/InDel variation sites were discovered in the promoter region, which were significantly associated with the branch number. Increasing evidence suggests that the natural variations present in promoter regions also play critical roles in altering agronomic traits by regulating the gene expression levels [30,31]. For example, segmental deletion in the promoter region of qSW5/GW5 alters its expression level and results in slender grains [32,33]. The insertion of a 148/150 bp fragment in the GmCHX1 promoter leads to differences in salt tolerance [34]. This suggests that natural variation in the promoter region of GmSPL9d may be the primary factor influencing the difference in the branch number between haplotypes.
The different haplotypes of GmSPL9d exhibited significant differences in the promoter activity. GmSPL9d H-II showed a significantly higher promoter activity than GmSPL9d H-I in both the transient and stable transformation assays. Importantly, significant differences in GUS enzyme activity driven by GmSPL9d H-II and GmSPL9d H-I promoters were observed in the shoot apical meristem (SAM) and axillary meristem (AM) of transgenic Arabidopsis. SAM and AM are crucial tissues involved in plant branching, as extensively reported [35,36,37,38]. Furthermore, when the GmSPL9d gene was expressed in Arabidopsis using different promoters, the overexpression of GmSPL9d H-II resulted in a significantly lower branch number compared with GmSPL9d H-I plants, indicating a synergistic relationship between the phenotype and expression level. Previous studies on salt tolerance traits in soybean demonstrated that the transcript levels of ProHap2: GsERD15B were significantly higher than 35S: GsERD15B and ProHap1: GsERD15B, indicating a stronger salt tolerance [39]. Similarly, in this study, the GmSPL9d H-II promoter activity was significantly higher than that of GmSPL9d H-I, and plants carrying the GmSPL9d H-II promoter type exhibited significantly lower branch numbers compared with those carrying the GmSPL9d H-I promoter type. Interestingly, the expression level of GmSPL9d H-I was significantly higher than that of Col-0, but no difference in the branch number was observed. This is consistent with previous studies, where a significant reduction in the branch number in Col-0 plants was only observed at high expression levels of GmSPL9d [18]. It is possible that GmSPL9d is targeted by miR156s in Arabidopsis, and only higher expression levels can exert the corresponding gene function. These results suggest that natural variation in the promoter region of GmSPL9d affects the soybean branch number by influencing its expression level. We have not disclosed the causal genetic variations responsible for the functional divergence of the two haplotypes. A further investigation of the upstream regulatory genes may help us to determine which polymorphisms are essential for the transcription of GmSPL9d, which will make the regulatory network more complete.

4. Materials and Methods

4.1. Plant Material and Growth Conditions

A total of 561 soybean natural germplasm resources were employed as experimental materials. The experiment was conducted using a randomized block design with a row length of 2 m, row spacing of 40 cm, plant spacing of 10 cm, and comprising two row blocks. The soybean plants were sown at the experimental base of Anhui Agricultural University (Table S1). Col-0 Arabidopsis thaliana served as the transgenic receptor, while Nicotiana benthamiana was used as the transient transformation receptor. These plants were grown in a greenhouse under conditions of 22 °C temperature, 16 h of light, and 8 h of darkness. These planting conditions were selected to provide an appropriate environment that ensured the accuracy and reproducibility of the experiment.

4.2. RNA Extraction and qRT-PCR

Total RNA was isolated from soybean tissues using an RNA prep Pure Plant Kit (TaKaRa, Otsu, Japan), and first-strand cDNA was synthesized using a PrimeScript™ 1st Strand cDNA Synthesis Kit (TaKaRa). qRT-PCR analysis was performed with a CFX96TM Real-time System (BIO-RAD, Hercules, CA, USA) with 10 μL of SYBR Mix (VAZYME, Nanjing, China), 6 μL of ddH2O, 1 μL of each primer, and 2 μL of cDNA template to a final volume of 20 μL. The implemented reaction procedure was 95 °C for 60 s, followed by 40 cycles of 95 °C for 30 s, 60 °C for 30 s, and 72 °C for 30 s. The relative gene expression levels were calculated using the 2−ΔΔCT method (Livak and Schmittgen. 2001) by normalization to GmACTIN (Glyma.19G147900) [40] or AtACTIN7 (AT5G09810) in soybean and Arabidopsis, respectively [41]. All the primers used in this experiment are listed in Table S3.

4.3. Identification and Bioinformatics Analysis of GmSPL9d

The latest reference genome of soybean and the hidden Markov model (HMM) profile of SPL (PF03110) were downloaded from the Phytozome soybean genome and PFAM databases. The hmmsearch program in HMMER software [42] was used to search the SPL HMM profile against the soybean genome, and the reliable results were screened based on an E-value less than or equal to 1 × 10−10. The soybean-specific SPL HMM profile was constructed using the hmmbuild program, and then the soybean reference genome was searched again with an E-value threshold 1 × 10−10. To avoid missing other members of the SPL gene family, the SPL-type protein sequences from soybean reported on the NCBI website were used as query sequences for BLASTP alignment with a E-value threshold of 1 × 10−5. We combined the genes predicted using the two methods and removed the duplicates. If multiple transcripts corresponding to a single SPL gene were obtained, only the most reliable one was retained. The protein sequences of the putative genes were submitted to the NCBI–CDD3 and PFAM online databases for verification, and the putative genes without SBP domains were excluded. The genes with an E-value greater than or equal to 1 × 10−20 predicted by the Pfam database were also excluded. Finally, members of the soybean SPL gene family were obtained with high confidence. For the multiple-alignment analysis, the amino acid sequences of SPL proteins from soybean, Arabidopsis, and rice were downloaded from the NCBI website. The phylogenetic tree was constructed with MEGA7.0 software using the NJ method [43]. The bootstrap tests were conducted with 1000 replicates, and other parameters were set to default.

4.4. Subcellular Localization

The coding sequence (CDS) of GmSPL9d was amplified from the soybean variety Williams82 using specific primers (Table S3). The CDS of GmSPL9d was then cloned into pCAMBIA1305-GFP using the homology recombination kit (VAZYME, Nanjing, China). The resulting construct was transformed into Agrobacterium GV3101 and subsequently injected into the leaves of Nicotiana benthamiana. After a culture period of 36–48 h, the subcellular localization of GmSPL9d within the epidermal cells of the inoculated tobacco leaves was observed using a laser scanning confocal microscope (Zeiss LSM880, Carl Zeiss, Jena, Germany).

4.5. Histochemical Analysis of GUS Activity

The 2000 bp sequence of the GmSPL9d promoter was downloaded from Phytozome v10 (https://phytozome.jgi.doe.gov/pz/portal.html, accessed on 13 December 2022 ). DNA fragments of the Baimao (GmSPL9d H-I) and Heihe 37 (GmSPL9d H-II) varieties were amplified and cloned into the phzm33 vector, followed by transformation into the GV3101 strain (Shanghai Weidi Biological, Shanghai, China). Transgenic Arabidopsis was obtained using the floral dip method. X-gluc (5-bromo-4-chloro-3-indolyl β-d-glucuronide) and GUS staining buffer (Jefferson et al., 1987) containing 1 mM X-gluc enzyme (Gold BioTechnology, St. Louis, MO, USA), 100 mM sodium phosphate (pH 7.5), 0.5 mM potassium ferrocyanide, 0.5 mM potassium ferricyanide, 10 mM EDTA, and 0.1% (v/v) Triton X-100 were used to stain the transgenic Arabidopsis seedling SAM and AM. All treatments were incubated at 37 °C for 4 h and cleared with 70% (v/v) ethanol. The GUS activity was determined using the GUS reporter gene quantitative analysis kit (Coolaber, Beijing, China).

4.6. Transient Transcription Activity Assay

To generate the constructs pGmSPL9d H-I:LUC and pGmSPL9d H-II:LUC, a 2000 bp promoter fragment was amplified from the GmSPL9d gene and ligated upstream of the pGreen0800-LUC reporter vector. These plasmids were then transformed into the GV3101 (P19) strain (Shanghai Weidi Biological). Subsequently, the plasmids were injected into Nicotiana benthamiana leaves. For the luciferase imaging, the Dual-luciferase assay reagent (Promega, Madison, WI, USA, VPE1910) was used, with Renilla luciferase serving as an internal control.

4.7. Arabidopsis Genetic Transformation

The constructs 35S: GmSPL9d, pGmSPL9d H-I: GmSPL9d, and pGmSPL9d H-II: GmSPL9d and the empty vector pCAMBIA1305 were individually transformed into Agrobacterium tumefaciens GV3101. These transformed bacteria were subsequently used for the floral dip transformation of Arabidopsis to generate transgenic plants [44].

4.8. Genetic Diversity Analysis

Variant data for GmSPL9d were retrieved from the public database SoyGVD (https://yanglab.hzau.edu.cn/SoyGVD/#/variation, accessed on 11 July 2023) and utilized for conducting genetic diversity analysis (Yang et al., 2023b).

4.9. Statistical Analysis

SPSS (https://www.ibm.com/spss, accessed on 11 July 2023) was employed to perform the statistical analysis. Student’s t-test or the Wilcoxon test were used to compare the differences between two groups, whereas Duncan’s multiple range test was used to compare the differences between multiple groups.

Supplementary Materials

The following supporting information can be downloaded from https://www.mdpi.com/article/10.3390/ijms25115991/s1.

Author Contributions

D.Z., H.Z., J.L., L.Q. and X.W. conceived and designed the contents. D.Z., H.Z., J.L., M.W., K.S., W.W., X.H. and Y.H. conducted the experiment; D.Z., H.Z., J.L., L.Q. and X.W. wrote and revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Key Research and Development Program of China (2021YFD1201605), the Natural Science Foundation of Anhui Province (2208085MC61), the Natural Science Research Project of Colleges and Universities in Anhui Province (KJ2021A0200), and the Special Fund for Anhui Agriculture Research System.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data supporting the findings of this study are available within the paper and within its supplementary data published online.

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Li, Y.H.; Zhou, G.Y.; Ma, J.X.; Jiang, W.K.; Jin, L.G.; Zhang, Z.H.; Guo, Y.; Zhang, J.B.; Sui, Y.; Zheng, L.T.; et al. De novo assembly of soybean wild relatives for pan-genome analysis of diversity and agronomic traits. Nat. Biotechnol. 2014, 32, 1045. [Google Scholar] [CrossRef] [PubMed]
  2. Miao, L.; Yang, S.N.; Zhang, K.; He, J.B.; Wu, C.H.; Ren, Y.H.; Gai, J.Y.; Li, Y. Natural variation and selection in GmSWEET39 affect soybean seed oil content. New Phytol. 2020, 225, 1651–1666. [Google Scholar] [CrossRef]
  3. Agudamu; Yoshihira, T.; Shiraiwa, T. Branch development responses to planting density and yield stability in soybean cultivars. Plant. Prod. Sci. 2016, 19, 331–339. [Google Scholar] [CrossRef]
  4. Xu, C.L.; Li, R.D.; Song, W.W.; Wu, T.T.; Sun, S.; Han, T.F.; Wu, C.X. High Density and Uniform Plant Distribution Improve Soybean Yield by Regulating Population Uniformity and Canopy Light Interception. Agronomy 2021, 11, 18. [Google Scholar] [CrossRef]
  5. Li, S.C.; Sun, Z.H.; Sang, Q.; Qin, C.; Kong, L.P.; Huang, X.; Liu, H.; Su, T.; Li, H.Y.; He, M.L.; et al. Soybean reduced internode 1 determines internode length and improves grain yield at dense planting. Nat. Commun. 2023, 14, 13. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, Z.Q.; Gao, L.; Ke, M.Y.; Gao, Z.; Tu, T.L.; Huang, L.M.; Chen, J.M.; Guan, Y.F.; Huang, X.; Chen, X. GmPIN1-mediated auxin asymmetry regulates leaf petiole angle and plant architecture in soybean. J. Integr. Plant Biol. 2022, 64, 1325–1338. [Google Scholar] [CrossRef]
  7. Shim, S.; Ha, J.; Kim, M.Y.; Choi, M.S.; Kang, S.T.; Jeong, S.C.; Moon, J.K.; Lee, S.H. GmBRC1 is a Candidate Gene for Branching in Soybean (Glycine max (L.) Merrill). Int. J. Mol. Sci. 2019, 20, 15. [Google Scholar] [CrossRef] [PubMed]
  8. Liang, Q.J.; Chen, L.Y.; Yang, X.; Yang, H.; Liu, S.L.; Kou, K.; Fan, L.; Zhang, Z.F.; Duan, Z.B.; Yuan, Y.Q.; et al. Natural variation of Dt2 determines branching in soybean. Nat. Commun. 2022, 13, 9. [Google Scholar] [CrossRef]
  9. Wang, J.W.; Czech, B.; Weigel, D. miR156-Regulated SPL Transcription Factors Define an Endogenous Flowering Pathway in Arabidopsis thaliana. Cell 2009, 138, 738–749. [Google Scholar] [CrossRef]
  10. Jiao, Y.Q.; Wang, Y.H.; Xue, D.W.; Wang, J.; Yan, M.X.; Liu, G.F.; Dong, G.J.; Zeng, D.L.; Lu, Z.F.; Zhu, X.D.; et al. Regulation of OsSPL14 by OsmiR156 defines ideal plant architecture in rice. Nature Genet. 2010, 42, 541–544. [Google Scholar] [CrossRef]
  11. Miura, K.; Ikeda, M.; Matsubara, A.; Song, X.J.; Ito, M.; Asano, K.; Matsuoka, M.; Kitano, H.; Ashikari, M. OsSPL14 promotes panicle branching and higher grain productivity in rice. Nature Genet. 2010, 42, 545–549. [Google Scholar] [CrossRef]
  12. Yamasaki, K.; Kigawa, T.; Inoue, M.; Tateno, M.; Yamasaki, T.; Yabuki, T.; Aoki, M.; Seki, E.; Matsuda, T.; Nunokawa, E.; et al. A novel zinc-binding motif revealed by solution structures of DNA-binding domains of Arabidopsis SBP-family transcription factors. J. Mol. Biol. 2004, 337, 49–63. [Google Scholar] [CrossRef]
  13. Manning, K.; Tör, M.; Poole, M.; Hong, Y.; Thompson, A.J.; King, G.J.; Giovannoni, J.J.; Seymour, G.B. A naturally occurring epigenetic mutation in a gene encoding an SBP-box transcription factor inhibits tomato fruit ripening. Nature Genet. 2006, 38, 948–952. [Google Scholar] [CrossRef]
  14. Xu, M.L.; Hu, T.Q.; Zhao, J.F.; Park, M.Y.; Earley, K.W.; Wu, G.; Yang, L.; Poethig, R.S. Developmental Functions of miR156-Regulated SQUAMOSA PROMOTER BINDING PROTEIN-LIKE (SPL) Genes in Arabidopsis thaliana. PLoS Genet. 2016, 12, 29. [Google Scholar] [CrossRef]
  15. Shikata, M.; Koyama, T.; Mitsuda, N.; Ohme-Takagi, M. Arabidopsis SBP-Box Genes SPL10, SPL11 and SPL2 Control Morphological Change in Association with Shoot Maturation in the Reproductive Phase. Plant Cell Physiol. 2009, 50, 2133–2145. [Google Scholar] [CrossRef]
  16. Zhang, Q.Q.; Wang, J.G.; Wang, L.Y.; Wang, J.F.; Wang, Q.; Yu, P.; Bai, M.Y.; Fan, M. Gibberellin repression of axillary bud formation in Arabidopsis by modulation of DELLA-SPL9 complex activity. J. Integr. Plant Biol. 2020, 62, 421–432. [Google Scholar] [CrossRef] [PubMed]
  17. Xie, Y.R.; Liu, Y.; Ma, M.D.; Zhou, Q.; Zhao, Y.P.; Zhao, B.B.; Wang, B.B.; Wei, H.B.; Wang, H.Y. Arabidopsis FHY3 and FAR1 integrate light and strigolactone signaling to regulate branching. Nat. Commun. 2020, 11, 13. [Google Scholar] [CrossRef]
  18. Sun, Z.X.; Su, C.; Yun, J.X.; Jiang, Q.; Wang, L.X.; Wang, Y.N.; Cao, D.; Zhao, F.; Zhao, Q.S.; Zhang, M.C.; et al. Genetic improvement of the shoot architecture and yield in soya bean plants via the manipulation of GmmiR156b. Plant Biotechnol. J. 2019, 17, 50–62. [Google Scholar] [CrossRef]
  19. Bao, A.L.; Chen, H.F.; Chen, L.M.; Chen, S.L.; Hao, Q.N.; Guo, W.; Qiu, D.Z.; Shan, Z.H.; Yang, Z.L.; Yuan, S.L.; et al. CRISPR/Cas9-mediated targeted mutagenesis of GmSPL9 genes alters plant architecture in soybean. Bmc Plant Biol. 2019, 19, 12. [Google Scholar] [CrossRef] [PubMed]
  20. Arazi, T.; Talmor-Neiman, M.; Stav, R.; Riese, M.; Huijser, P.; Baulcombe, D.C. Cloning and characterization of micro-RNAs from moss. Plant J. 2005, 43, 837–848. [Google Scholar] [CrossRef] [PubMed]
  21. Eriksson, M.; Moseley, J.L.; Tottey, S.; del Campo, J.A.; Quinn, J.; Kim, Y.; Merchant, S. Genetic dissection of nutritional copper signaling in chlamydomonas distinguishes regulatory and target genes. Genetics 2004, 168, 795–807. [Google Scholar] [CrossRef] [PubMed]
  22. Kropat, J.; Tottey, S.; Birkenbihl, R.P.; Depège, N.; Huijser, P.; Merchant, S. A regulator of nutritional copper signaling in Chlamydomonas is an SBP domain protein that recognizes the GTAC core of copper response element. Proc. Natl. Acad. Sci. USA 2005, 102, 18730–18735. [Google Scholar] [CrossRef] [PubMed]
  23. Lännenpää, M.; Jänönen, I.; Hölttä-Vuori, M.; Gardemeister, M.; Porali, I.; Sopanen, T. A new SBP-box gene BpSPL1 in silver birch (Betula pendula). Physiol. Plant. 2004, 120, 491–500. [Google Scholar] [CrossRef] [PubMed]
  24. Moreno, M.A.; Harper, L.C.; Krueger, R.W.; Dellaporta, S.L.; Freeling, M. liguleless1 encodes a nuclear-localized protein required for induction of ligules and auricles during maize leaf organogenesis. Genes Dev. 1997, 11, 616–628. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, X.B.; Zhang, Z.L.; Liu, D.M.; Zhang, K.; Li, A.L.; Mao, L. SQUAMOSA Promoter-Binding Protein-Like Transcription Factors: Star Players for Plant Growth and Development. J. Integr. Plant Biol. 2010, 52, 946–951. [Google Scholar] [CrossRef] [PubMed]
  26. Schmutz, J.; Cannon, S.B.; Schlueter, J.; Ma, J.X.; Mitros, T.; Nelson, W.; Hyten, D.L.; Song, Q.J.; Thelen, J.J.; Cheng, J.L.; et al. Genome sequence of the palaeopolyploid soybean (vol 463, pg 178, 2010). Nature 2010, 465, 120. [Google Scholar] [CrossRef]
  27. Haerizadeh, F.; Wong, C.E.; Singh, M.B.; Bhalla, P.L. Genome-wide analysis of gene expression in soybean shoot apical meristem. Plant Mol. Biol. 2009, 69, 711–727. [Google Scholar] [CrossRef] [PubMed]
  28. Begum, N.; Xiao, Y.T.; Wang, L.; Li, D.M.; Irshad, A.; Zhao, T.J. Arbuscular mycorrhizal fungus Rhizophagus irregularis alleviates drought stress in soybean with overexpressing the GmSPL9d gene by promoting photosynthetic apparatus and regulating the antioxidant system. Microbiol. Res. 2023, 273, 14. [Google Scholar] [CrossRef]
  29. Yun, J.X.; Sun, Z.X.; Jiang, Q.; Wang, Y.N.; Wang, C.; Luo, Y.Q.; Zhang, F.R.; Li, X. The miR156b-GmSPL9d module modulates nodulation by targeting multiple core nodulation genes in soybean. New Phytol. 2022, 233, 1881–1899. [Google Scholar] [CrossRef] [PubMed]
  30. Springer, N.; de León, N.; Grotewold, E. Challenges of Translating Gene Regulatory Information into Agronomic Improvements. Trends Plant Sci. 2019, 24, 1075–1082. [Google Scholar] [CrossRef]
  31. Swinnen, G.; Goossens, A.; Pauwels, L. Lessons from Domestication: Targeting Cis-Regulatory Elements for Crop Improvement. Trends in Plant Science 2016, 21, 506–515. [Google Scholar] [CrossRef] [PubMed]
  32. Duan, P.G.; Xu, J.S.; Zeng, D.L.; Zhang, B.L.; Geng, M.F.; Zhang, G.Z.; Huang, K.; Huang, L.J.; Xu, R.; Ge, S.; et al. Natural Variation in the Promoter of GSE5 Contributes to Grain Size Diversity in Rice. Mol. Plant 2017, 10, 685–694. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, J.F.; Chen, J.; Zheng, X.M.; Wu, F.Q.; Lin, Q.B.; Heng, Y.Q.; Tian, P.; Cheng, Z.J.; Yu, X.W.; Zhou, K.N.; et al. GW5 acts in the brassinosteroid signalling pathway to regulate grain width and weight in rice. Nat. Plants 2017, 3, 7. [Google Scholar] [CrossRef] [PubMed]
  34. Li, Y.; Ye, H.; Vuong, T.D.; Zhou, L.J.; Do, T.D.; Chhapekar, S.S.; Zhao, W.Q.; Li, B.; Jin, T.; Gu, J.B.; et al. A novel natural variation in the promoter of GmCHX1 regulates conditional gene expression to improve salt tolerance in soybean. J. Exp. Bot. 2024, 75, 1051–1062. [Google Scholar] [CrossRef] [PubMed]
  35. Greb, T.; Clarenz, O.; Schäfer, E.; Müller, D.; Herrero, R.; Schmitz, G.; Theres, K. Molecular analysis of the LATERAL SUPPRESSOR gene in Arabidopsis reveals a conserved control mechanism for axillary meristem formation. Genes Dev. 2003, 17, 1175–1187. [Google Scholar] [CrossRef] [PubMed]
  36. Shi, B.H.; Zhang, C.; Tian, C.H.; Wang, J.; Wang, Q.; Xu, T.F.; Xu, Y.; Ohno, C.; Sablowski, R.; Heisler, M.G.; et al. Two-Step Regulation of a Meristematic Cell Population Acting in Shoot Branching in Arabidopsis. PLoS Genet. 2016, 12, 20. [Google Scholar] [CrossRef] [PubMed]
  37. Tucker, M.R.; Laux, T. Connecting the paths in plant stem cell regulation. Trends Cell Biol. 2007, 17, 403–410. [Google Scholar] [CrossRef] [PubMed]
  38. Yang, Q.Q.; Yuan, C.Q.; Cong, T.C.; Zhang, Q.X. The Secrets of Meristems Initiation: Axillary Meristem Initiation and Floral Meristem Initiation. Plants-Basel 2023, 12, 18. [Google Scholar] [CrossRef] [PubMed]
  39. Jin, T.; Sun, Y.Y.; Shan, Z.; He, J.B.; Wang, N.; Gai, J.Y.; Li, Y. Natural variation in the promoter of GsERD15B affects salt tolerance in soybean. Plant Biotechnol. J. 2021, 19, 1155–1169. [Google Scholar] [CrossRef]
  40. Ahmad, M.Z.; Rehman, N.U.; Yu, S.W.; Zhou, Y.Z.; ul Haq, B.; Wang, J.J.; Li, P.H.; Zeng, Z.X.; Zhao, J. GmMAX2-D14 and -KAI interaction-mediated SL and KAR signaling play essential roles in soybean root nodulation. Plant J. 2020, 101, 334–351. [Google Scholar] [CrossRef] [PubMed]
  41. Zhang, D.; Yang, H.F.; Wang, X.C.; Qiu, Y.J.; Tian, L.H.; Qi, X.Q.; Qu, L. Cytochrome P450 family member CYP96B5 hydroxylates alkanes to primary alcohols and is involved in rice leaf cuticular wax synthesis. New Phytol. 2020, 225, 2094–2107. [Google Scholar] [CrossRef] [PubMed]
  42. Mistry, J.; Finn, R.D.; Eddy, S.R.; Bateman, A.; Punta, M. Challenges in homology search: HMMER3 and convergent evolution of coiled-coil regions. Nucleic Acids Res. 2013, 41, 10. [Google Scholar] [CrossRef] [PubMed]
  43. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef]
  44. Clough; Steven, J.; Bent; Andrew, F. Floral dip: A simplified method for Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant J. 1998, 16, 735–743. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Phylogenetic and expression analyses of GmSPLs. (a) Phylogenetic tree of SPLs from Arabidopsis, soybean, and rice. (b) Expression analysis of SPLS in soybean, with the data from the Phytozome database.
Figure 1. Phylogenetic and expression analyses of GmSPLs. (a) Phylogenetic tree of SPLs from Arabidopsis, soybean, and rice. (b) Expression analysis of SPLS in soybean, with the data from the Phytozome database.
Ijms 25 05991 g001
Figure 2. Expression patterns of GmSPL9d and the subcellular localization of GmSPL9d. (a) Gene expression of GmSPL9d in the root, nodule, stem, leaf, flower, and SAM. (b) GmSPL9d was localized to the nucleus. Scale bars, 20 μm.
Figure 2. Expression patterns of GmSPL9d and the subcellular localization of GmSPL9d. (a) Gene expression of GmSPL9d in the root, nodule, stem, leaf, flower, and SAM. (b) GmSPL9d was localized to the nucleus. Scale bars, 20 μm.
Ijms 25 05991 g002
Figure 3. Haplotype analysis for GmSPL9d locus in the soybean population. Sequence and allelic variations in GmSPL9d locus, including 2 kb promoters and the full−length gene sequence. GmSPL9d H-I (Hap 1−Hap 5) and GmSPL9d H-II (Hap 6−Hap 7) represent seven different haplotypes detected by 55 polymorphic sites. The black boxes represent exons, the lines represent introns and the grey boxes represent UTR.
Figure 3. Haplotype analysis for GmSPL9d locus in the soybean population. Sequence and allelic variations in GmSPL9d locus, including 2 kb promoters and the full−length gene sequence. GmSPL9d H-I (Hap 1−Hap 5) and GmSPL9d H-II (Hap 6−Hap 7) represent seven different haplotypes detected by 55 polymorphic sites. The black boxes represent exons, the lines represent introns and the grey boxes represent UTR.
Ijms 25 05991 g003
Figure 4. Branch number analysis of the different haplotypes in GmSPL9d. (a) Phylogenetic analysis of seven different haplotypes of GmSPL9d. (b,c) Boxplot of the branch number for two haplotypes in 2017 and 2018. Statistical significance (P = 6.0057 × 10−14 and P = 7.4744 × 10−10) of the difference between two haplotypes was determined by a two-sided Wilcoxon test. The center bold line represents the median; box edges indicate the upper and lower quantiles.
Figure 4. Branch number analysis of the different haplotypes in GmSPL9d. (a) Phylogenetic analysis of seven different haplotypes of GmSPL9d. (b,c) Boxplot of the branch number for two haplotypes in 2017 and 2018. Statistical significance (P = 6.0057 × 10−14 and P = 7.4744 × 10−10) of the difference between two haplotypes was determined by a two-sided Wilcoxon test. The center bold line represents the median; box edges indicate the upper and lower quantiles.
Ijms 25 05991 g004
Figure 5. Promoter activity analysis of pGmSPL9d H-I and pGmSPL9d H-II in GmSPL9d. (a) Correlation analysis between the branch number and the GmSPL9d expression level in the AM (V4 stage) of 82 soybean accessions. (b,c) Promoter activity analysis of GmSPL9d using sequences 2000 bp upstream from the translation initiation site (n  =  3 biologically independent replicates, * indicates a significant difference at 0.05 level by Student’s t-test). (d) Different tissues (SAM and AM) of GmSPL9d transgenic arabidopsis staining results in pGmSPL9d H-I and pGmSPL9d H-II. (e,f) Comparison of different tissue (SAM and AM) GUS activities between pGmSPL9d H-I and pGmSPL9d H-II. Different letters indicate a significant difference in the GUS activity between pGmSPL9d H-I and pGmSPL9d H-II at the 0.05 level by a one-way ANOVA test (n = 3).
Figure 5. Promoter activity analysis of pGmSPL9d H-I and pGmSPL9d H-II in GmSPL9d. (a) Correlation analysis between the branch number and the GmSPL9d expression level in the AM (V4 stage) of 82 soybean accessions. (b,c) Promoter activity analysis of GmSPL9d using sequences 2000 bp upstream from the translation initiation site (n  =  3 biologically independent replicates, * indicates a significant difference at 0.05 level by Student’s t-test). (d) Different tissues (SAM and AM) of GmSPL9d transgenic arabidopsis staining results in pGmSPL9d H-I and pGmSPL9d H-II. (e,f) Comparison of different tissue (SAM and AM) GUS activities between pGmSPL9d H-I and pGmSPL9d H-II. Different letters indicate a significant difference in the GUS activity between pGmSPL9d H-I and pGmSPL9d H-II at the 0.05 level by a one-way ANOVA test (n = 3).
Ijms 25 05991 g005
Figure 6. Effect of different GmSPL9d promoters on the branching of Arabidopsis. (a) Schematic diagram of different constructs using pCAMBIA1305 as the backbone vector, including 35S: GmSPL9d, PGmSPL9d H-I: GmSPL9d, and PGmSPL9d H-II: GmSPL9d. (b) qRT-PCR analysis of GmSPL9d/AtSPL9 expression in the transgenic Arabidopsis lines (20 days of axillary meristem). (c) Transgenic phenotype in Col-0. Scale bar, 5 cm. (d) Quantitative analysis of the branch number in wild-type plants (Col-0) and transgenic plants (n = 10). The data were collected at 45 days after emergence from the soil. ACTIN7 was used as an internal control for gene expression. All data are represented as mean ± SE. Different letters represent a significantly difference (one-way ANOVA test; p < 0.05).
Figure 6. Effect of different GmSPL9d promoters on the branching of Arabidopsis. (a) Schematic diagram of different constructs using pCAMBIA1305 as the backbone vector, including 35S: GmSPL9d, PGmSPL9d H-I: GmSPL9d, and PGmSPL9d H-II: GmSPL9d. (b) qRT-PCR analysis of GmSPL9d/AtSPL9 expression in the transgenic Arabidopsis lines (20 days of axillary meristem). (c) Transgenic phenotype in Col-0. Scale bar, 5 cm. (d) Quantitative analysis of the branch number in wild-type plants (Col-0) and transgenic plants (n = 10). The data were collected at 45 days after emergence from the soil. ACTIN7 was used as an internal control for gene expression. All data are represented as mean ± SE. Different letters represent a significantly difference (one-way ANOVA test; p < 0.05).
Ijms 25 05991 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, D.; Zheng, H.; Li, J.; Wan, M.; Shu, K.; Wang, W.; Hu, X.; Hu, Y.; Qiu, L.; Wang, X. Natural Variation in the Promoter of GmSPL9d Affects Branch Number in Soybean. Int. J. Mol. Sci. 2024, 25, 5991. https://doi.org/10.3390/ijms25115991

AMA Style

Zhao D, Zheng H, Li J, Wan M, Shu K, Wang W, Hu X, Hu Y, Qiu L, Wang X. Natural Variation in the Promoter of GmSPL9d Affects Branch Number in Soybean. International Journal of Molecular Sciences. 2024; 25(11):5991. https://doi.org/10.3390/ijms25115991

Chicago/Turabian Style

Zhao, Duo, Haowei Zheng, Jiajia Li, Mingyue Wan, Kuo Shu, Wenhui Wang, Xiaoyu Hu, Yu Hu, Lijuan Qiu, and Xiaobo Wang. 2024. "Natural Variation in the Promoter of GmSPL9d Affects Branch Number in Soybean" International Journal of Molecular Sciences 25, no. 11: 5991. https://doi.org/10.3390/ijms25115991

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