Next Article in Journal
Novel Benzoxazoles Containing 4-Amino-Butanamide Moiety Inhibited LPS-Induced Inflammation by Modulating IL-6 or IL-1β mRNA Expression
Previous Article in Journal
Age-Related DNA Methylation in Normal Kidney Tissue Identifies Epigenetic Cancer Risk Susceptibility Loci in the ANKRD34B and ZIC1 Genes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Identification of Auxin Response Factors in Peanut (Arachis hypogaea L.) and Functional Analysis in Root Morphology

1
Key Laboratory of Plant Development and Environmental Adaptation Biology, Ministry of Education, School of Life Sciences, Shandong University, Qingdao 266237, China
2
State Key Laboratory of Crop Biology, Shandong Key Laboratory of Crop Biology, College of Agronomy, Shandong Agricultural University, Tai’an 271018, China
3
State Key Laboratory of Crop Biology, Shandong Key Laboratory of Crop Biology, College of Life Science, Shandong Agricultural University, Tai’an 271018, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(10), 5309; https://doi.org/10.3390/ijms23105309
Submission received: 14 April 2022 / Revised: 4 May 2022 / Accepted: 5 May 2022 / Published: 10 May 2022
(This article belongs to the Section Molecular Genetics and Genomics)

Abstract

:
Auxin response factors (ARFs) play important roles in plant growth and development; however, research in peanut (Arachis hypogaea L.) is still lacking. Here, 63, 30, and 30 AhARF genes were identified from an allotetraploid peanut cultivar and two diploid ancestors (A. duranensis and A. ipaensis). Phylogenetic tree and gene structure analysis showed that most AhARFs were highly similar to those in the ancestors. By scanning the whole-genome for ARF-recognized cis-elements, we obtained a potential target gene pool of AhARFs, and the further cluster analysis and comparative analysis showed that numerous members were closely related to root development. Furthermore, we comprehensively analyzed the relationship between the root morphology and the expression levels of AhARFs in 11 peanut varieties. The results showed that the expression levels of AhARF14/26/45 were positively correlated with root length, root surface area, and root tip number, suggesting an important regulatory role of these genes in root architecture and potential application values in peanut breeding.

1. Introduction

Auxin is a key hormone that plays an important role throughout the life cycle of plants from embryogenesis to fruit maturity [1,2,3]. Auxin response factors (ARFs) play a core role in auxin-mediated transcriptional regulation and auxin signaling transduction 2a ARF proteins contain three conserved domains. Briefly, the N-terminal DNA-binding domain (DBD) is responsible for recognizing the auxin-recognized cis-element (AuxRE) within target gene promoters [4,5]. The C-terminal Phox and Bem1 (PB1) domain mediates ARF–ARF or ARF–Aux/IAA (Auxin/INDOLE ACETIC ACID) dimerization. The middle domain determines whether ARF is a transcription activator or a transcription suppressor due to the variety of amino acids it contains [6,7]. At low-auxin levels, Aux/IAAs inhibit ARF transcriptional activity through recruit TOPLESS (TPL) and histone deacetylase (HDAC) [8]. With the increase of auxin and the degradation of Aux/IAA through the SCFTIR1-dependent pathway, the released ARF transcription factors regulate the expression of downstream targets and cause auxin-related plant growth and development [9,10].
There are multiple ARF members in different plants. For example, 23 in Arabidopsis, 25 in rice, and 31 members in maize [11,12,13]. Importantly, the role of some ARF-mediated auxin signaling in root growth and development has been well studied in both the model plant Arabidopsis and some crops. For example, ARF10/16 has been shown to regulate root cap cell formation and root stem cell identity in Arabidopsis [14]. ARF7/19-mediated auxin signaling promotes lateral root development by inducing the expression of LATERAL ORGAN BOUNDARIES DOMAIN (LBD16/18/29) and/or PR-1 homolog (PRH1) in Arabidopsis [15,16,17]. In soybean, GmARF8a/b negatively regulates lateral root development and nodule formation [18]. SlARF2 promotes lateral root formation in tomato [19]. Although ARFs appear to be conserved in seed plants, the question remains whether the functions of these proteins are also consistent across plant species. Therefore, it is necessary to analyze the expression pattern, DNA-binding specificity, and the relationship with important traits, such as root morphology, before replications in breeding [20,21].
Peanut (Arachis hypogaea L.) is one of the most important economic oil crops, grown in many Asian and African countries with an annual production of about 46 million tons (http://www.fao.org/faostat; accessed on 2 May 2021). Cultivar peanut (A. hypogaea, allotetraploid, AABB, 2n = 4x = 40) is assumed from hybridization of two wild diploid ancestors, Arachis duranensis (AA genome, 2n = 2x = 20) and Arachis ipaensis (BB genome, 2n =2x =20) [22]. Peanut cultivars exhibit extensive phenotypic and genetic variation, especially the root traits [23,24]. However, due to limited research methods, the regulation genes of important root traits are rarely identified in peanut. Their application in breeding is seriously lagging compared with other crops. Despite the importance of ARF proteins in root development, the role of peanut ARFs has not been characterized. Here, we identified 63 AhARF, 30 AdARF, and 30 AiARF genes from an allotetraploid peanut cultivar, A. duranensis, and A. ipaensis, respectively. Comprehensive analysis revealed that numerous predicted downstream targets of AhARFs are enriched in root development, especially AhARF14/26/45. Importantly, the expression of these genes is highly correlated with root architecture, implying the involvement of these members in root growth and development.

2. Results

2.1. Genome-Wide Identification of Peanut ARF Genes

Based on a HMMSCAN and BLASTP search, 30, 30, and 63 ARF genes were identified in A. duranensis (AdARFs), A. ipaensis (AiARFs), and A. hypogaea (allotetraploid peanut cultivar; AhARFs), respectively. These AhARF genes were numbered according to their locations in different chromosomes, and the AdARFs and AiARFs were named according to their gene IDs (Figure 1). Detailed information for all these genes, including gene ID, chromosome location, molecular weight (MW), and the isoelectric point (pI) of proteins are listed in Table S1.
There were 62 AhARF genes mapped to 18 chromosomes (Figure 1), and AhARF63 is located on scaffold 59. Similar to those of A. hypogaea, there was no ARF gene mapped to chromosome Aradu.A01 or Araip.B01. The number of AhARFs on each chromosome was approximately the same as that of the corresponding chromosome of A. duranensis and A. ipaensis. The absence of ARFs located on chromosomes Arahy.01 and Arahy.11 was also consistent with that of A. duranensis and A. ipaensis.

2.2. Phylogenetic Analysis of Peanut ARFs

An unrooted maximum-likelihood phylogenetic tree was constructed using ARFs from an allotetraploid peanut and its diploid ancestors. In most cases, a pair of ARFs from each ancestor and a pair of AhARFs clustered to a clade, and, usually, the two ancestor ARFs, showed a closer evolutionary relationship (green and blue dots in Figure 2). In addition, the gene structure of most AhARFs and their orthologs has a high similarity (Figure 2), mainly reflected in the number and position of introns. Interestingly, according to the annotation information of A. hypogaea cv. Tifrunner genome, alternative splicing (AS) events were identified in 15 peanut ARFs (Figure S1). A few AS events occurred in the untranslated regions (UTRs), which did not affect the protein structure. In fact, most AS events altered the coding of proteins and even led to a premature termination of translation (Figure S1). For example, the seventh and eighth exon of the AS1 of AhARF1 (arahy.Tifrunner.gnm1.ann1.TZ3SVZ.1) were not identified in other AS events, while the ninth exon corresponded to two exons in other AS events. There were five and four possible AS events identified in AhARF27 and AhARF60, respectively, and four of them (two for each) coding proteins with large C-terminal deletions. Considering that the C-terminal domain of ARFs is responsible for dimerization with other ARFs or Aux/IAA, these truncated ARFs may play different roles in planta. For example, AtARF3, an atypical ARF lacking the PB1 domain, can directly induce downstream gene expression in the presence of auxin, a more concise auxin transcriptional regulation mode.
To clarify which developmental or hormonal signals these ARFs are involved in, we then examined which cis-acting elements were present in their promoters. According to the prediction of a database search of plant promoter results, obtained from the PlantCARE website, auxin response elements were detected in the promoters of 42 peanut ARF genes, including 23 AhARFs, 9 AdARFs, and 10 AiARFs (Figure 2, black stars). In addition, other cis-elements were also found to participate in plant hormone responses and stress responses (Figure 2, inner cycle). In addition, species and the positions of cis-elements in genes belonging to the same clade are similar. AdARFs and AiARFs are similar to AhARFs, implying that this family is evolutionarily conservative.
The function of AhARFs can also be predicted based on the well-characterized function of their homologous proteins in other species, such as Arabidopsis thaliana, Oryza sativa, Solanum lycopersicum, and Glycine max [4,6,14,15,16,17,18,19,20,21,22,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66]. The phylogenetic analysis and function prediction results showed that most AhARFs were clustered with plant ARFs (Figure S2, Table S2). These results revealed that the functions of ARFs in different species might be conserved, and the well-functioning-identified ARFs in other species have an important reference value in peanut.

2.3. Gene Duplication on the Expansion of ARF Genes

The important regulatory roles of gene duplication in evolution and the expansion of gene families and crop domestication have been well documented. To clarify the evolution process of the AhARF family after the formation of the allotetraploid peanut, we analyzed the gene duplication events of AhARFs. As shown in Figure 3A and Table S3, there were 22 orthologous, 14 paralogous, and 5 tandem gene pairs detected in A. hypogaea.
The substitution rates of nonsynonymous (Ka) versus synonymous (Ks) were subsequently calculated to explore the selection pressure of AhARFs, and the results showed that the Ka/Ks values of all orthologue pairs were less than 1.0, and 88% (22/25) were less than 0.5 (Table 1). This suggests that the orthologous AhARFs undergo intensive purifying selection pressure and remain conserved in both structure and function. The divergence time of these five gene pairs (AhARF40AhARF61, AhARF28AhARF40, AhARF37AhARF56, AhARF36AhARF57, AhARF39AhARF58) were estimated to be around 85–107 million years ago (Mya), while the other pairs were predicted to have diverged nearly 1.32–7.93 Mya.

2.4. Expression Profiles of AhARFs

To explore the possible biological functions of AhARFs, their expression patterns were analyzed based on RNA-seq data obtained from PeanutBase (https://www.peanutbase.org/ accessed on 21 April 2021). The AhARFs exhibited diverse expression patterns in different developmental stages and plant tissues; however, most of the gene pairs showed a similar expression pattern and were also clustered together in the heatmap (Figure 4). For example, AhARF16 and AhARF48 are highly similar in sequence and gene structure (Figure 1), and similarly, their expression patterns are almost identical in 22 plant tissues and at different developmental stages. However, there are exceptions, such as AhARF1 and AhARF31, which are highly homologous but are expressed differently in peg and fruit development.

2.5. Characterization of Downstream Target Genes of AhARFs

To predict the possible roles of these ARFs, we examined the distribution and function of target genes that may be regulated by these transcription factors. DR5 and IR8, two composite motifs, have been shown to be closely involved in the auxin-mediated transcriptional regulation. Therefore, we searched DR5, IR8, and single AuxRE (TGTCGG) motif in the 2000 bp upstream of each gene throughout the peanut genome (Figure 5A). There were 24,380 possible ARF-binding elements detected in 18,190 genes; among them, single AuxRE (TGTCGG) motif accounts for 65.44% (Table S4). There were also 8426 DR5 or IR8 in 7350 gene promoters, and most of these genes contain only one DR5 or IR8 (Table S4). The highest number of ARF-binding elements (DR5 and IR8 only) was found in the promoter of a predicted glucose-induced degradation protein (arahy.7YHY0R).
Subsequent gene ontology (GO) analysis showed that the predicted target genes mainly participate in plant hormone responses, biotic/abiotic stress responses, and tissue/organ development (Figure 5C). Besides being involved in auxin polar transport and auxin response, these targets also relate to seven other well-characterized plant hormones, including abscisic acid (ABA), salicylic acid (SA), brassinosteroids (BRs), jasmonic acid (JA), ethylene (ETH), gibberellins (GAs), and cytokinins (CKs). Notably, the function of these genes is significantly enriched in plant root development, especially root hair elongation and root meristem growth (Figure 5C).

2.6. The Possible Regulatory Roles of AhARFs in Peanut Root Morphology

To explore the relationship between the expression of AhARFs and the root morphology of peanut, we selected 11 peanut varieties (Table S5) and analyzed their root architecture and the expression levels of ARFs in their roots. Scanning the root systems of different varieties revealed variable root architectures among them (Figure 6A). For example, JuHua7 (JH7), EHua3 (EH3), and ShanHua11 (SH11) display a long primary root and less lateral roots, while FengHua 2 (FH2) and Jinkins Jumbo (JK) exhibit a long primary root with more lateral roots. After quantifying the total root length, root surface area, root mean diameter, and root tip number, we found that these indexes also differed among varieties (Figure 6B).
According to the expression patterns of AhARFs (Figure 4), seven members (AhARF8/14/26/38/39/45/56) were selected and their expression levels were further examined in the roots of these 11 varieties, including JK, Meiyinxuan 41165 (41165), Shixuan 64 (SX64), Juhua 27 (JH27), FH2, EH3, SH11, and Zhonghua 12 (ZH12). As shown in Figure 6C, these AhARFs exhibited different expression levels in different varieties, suggesting a possible link between the gene expression levels and root architecture. Among them, the expression of AhARF59 showed the greatest difference among different varieties. Subsequently, the Pearson’s correlation analysis was performed to explore the relationship between root morphology and the expression levels of the AhARFs (Figure 6D). The results showed that the expression levels of AhARF26 and AhARF45 are positively correlated with total root length; AhARF14/26/45 are positively correlated with root surface area; and AhARF26 is positively correlated with the root tip formation, indicating the important role of these AhARFs in peanut root development.
Subcellular localization of AhARF14 and AhARF26 proteins were further analyzed in Nicotiana benthamiana (N. benthamiana) epidermal cells. AhARF26 was specially localized in the nucleus, while AhARF14 mainly localized in the nucleus and weak signal of AhARF14-GFP, which was also detected in the cytoplasm (Figure 7). The subcellular localization of AhARF14 was further analyzed using the online tool DeepLoc-1.0, and it showed a likelihood of 0.587 to localize in cytoplasm (Figure S4).

3. Discussion

A previous study reported that there were 114, 28, and 28 ARFs in Shitouqi (allotetraploid peanut cultivar), A. duranensis, and A. ipaensis, respectively [67]. Another independent study identified 61 AhARFs from A. hypogaea, and then revealed the regulatory roles of AhARF6 in pod development [68]. Here, we supplemented and improved various data on this family, identified 63 members of A. hypogaea, and picked up the two missing members of AhARF1 and AhARF63. Multiple sequence alignment showed that the three AhARF1 AS formats and AhARF63 belong to the typical ARFs. In addition, we have provided some valuable information on the upstream regulation of cis-elements and downstream targets.
More importantly, we analyzed the potential relationship between the expression levels of AhARFs and the root morphology among varieties and found that expression levels of AhARF14/26/45 are positively correlated with peanut total root length, root surface area, or root tip number. Peanut is an important protein source and oil crop that can be grown in poor quality soil [69,70], while peanut yield is greatly restricted by adverse environmental conditions, such as drought, salt, and nutrient deficiencies [71,72,73]. Roots are the main organs of plants to absorb water and nutrients from soil. Therefore, root morphology and their spatial configuration will substantially determine the ability of a plant to secure edaphic resources [74]. Understanding the molecular mechanisms that regulate crop root morphology is conducive to improving root traits and increasing crop yield. Root morphology is a complex trait determined by both environmental and endogenous factors, and among them, auxin plays an important role [75,76,77,78]. Therefore, genome-wide identification and function prediction are essential to unlock the mechanisms of AhARF-mediated auxin signaling in peanut root development.
Root morphology traits have long been considered a key target by breeders for crop improvement; however, there are many challenges in root system measurement [79]. Although several root trait QTLs have been observed via analysis of crop populations, important genes are rarely reported [80,81,82,83,84]. Most favorable root traits are selected directly by visual test in breeding practices. The selection of germplasms represents better root traits, and finding their commonalities in genotypes or expression patterns is a promising practice. Through RNA-seq analysis, we found that seven AhARFs were highly expressed in roots, implying the involvement of these members in root growth regulation. Importantly, we showed that the expression levels of AhARF14/26/45 are largely positively correlated with root architecture. These three peanut genes exhibit high similarity with their homologous AtARF16, which has been reported to control root cap formation [14]. AhARF8/38, which were not related to the peanut root traits tested in our work, seem to display different functions from their homologous AtARF7/19. These results suggest that, although ARFs are conserved in different plant species, their functions are diversified. The identification of germplasms with effective genotypes and linked molecular markers that regulate root development is beneficial to molecular marker-assisted breeding. The finding that AhARF14/26/45 modulate peanut root architecture links their expressions to beneficial root traits, and it can be used directly in the selection of parental germplasms and their hybrid offspring, especially for drought-resistance and nutrition-efficient breeding. Interestingly, a weak AhARF14-GFP signal was detected in cytoplasm, which is inconsistent with previous studies on the subcellular localization of plant ARF proteins [68,85]. This may be caused by differences between species and may indicate new regulatory mechanisms for ARFs in peanut.

4. Materials and Methods

4.1. Plant Materials and Growth Conditions

The A. hypogea L. cultivars and germplasms used in this study were selected from a natural peanut population with extensive phenotypic and genetic variation preserved by our group [23]. Mature seeds were soaked in 0.1% H2O2 for 6 h, and then washed 3–5 times with sterile ddH2O. Sterilized seeds were sown on water-wet degreasing cotton, extended in seedling cultivation disks to germinate at 26 °C in darkness for 3 days. Seedlings were then exposed to long-day conditions (16 h light and 8 h dark cycle, 11.4 K LuX light intensity) for another 2 or 3 days. Two-functional-leaf seedlings were transplanted to the hydroponic box and cultured with 1/5 Hoagland’s nutrient solution until further use [86].

4.2. Genome-Wide Identification of AhARFs, AdARFs, and AiARFs

Whole-genome sequences of allotetraploid peanut A. hypogaea cv. Tifrunner (version 1) [87] and the diploid ancestors (A. duranensis and A. ipaensis) [79] were downloaded from PeanutBase (https://www.peanutbase.org/peanut_genome; accessed on 21 April 2021 ). The protein sequences of AtARFs were downloaded from the TAIR database (http://www.arabidopsis.org/, accessed on 21 April 2021), while those of rice, soybean, Medicago truncatula, and tomato were downloaded from Ensembl Plants (http://plants.ensembl.org/index.html; accessed on 21 April 2021). The hidden Markov model (HMM) profile of the B3-type DNA-binding domain (DBD, PF02362), auxin-response domain (PF06507), and Aux-IAA domain (PF02309) were retrieved from the Pfam database (http://pfam.xfam.org/; accessed on 15 March 2021) and used for peptide-searching through BLASTP analysis. The predicted peanut ARFs were confirmed using the SMART program (http://smart.embl-heidelberg.de/; accessed on 21 April 2021) and the Conserved Domain Database (CDD; https://www.ncbi.nlm.nih.gov/cdd; accessed on 11 April 2021). The protein sequences of peanut ARFs were submitted to the ProParam tool in ExPASy (https://web.expasy.org/protparam/; accessed on 30 April 2021) to calculate the molecular weight (MW) and theoretical isoelectric point (pI).

4.3. Chromosome Distribution, Gene Structure, Phylogenetic Analysis, and cis-Elements Analysis

Information on peanut ARFs loci on chromosome were derived from annotation gff3 files downloaded from PeanutBase (https://www.peanutbase.org/; accessed on 18 April 2021). The Gene Structure Display Server (https://gsds.cbi.pku.edu.cn/; accessed on 18 April 2021) was employed for analyzing AhARF structure according to the gff3 data. MEGA7 [88] was used to perform sequence alignment and maximum-likelihood phylogenetic tree construction with the bootstrap method (number of bootstrap replications = 1000). To analyze the regulatory region of the AhARFs, the 2000 bp genomic sequences located upstream of the start codon ATG were analyzed in the PlantCARE website (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/; accessed on 18 April 2021). Visualization and further editing of the phylogenetic tree were performed on the website tool iTOL (http://itol.embl.de/; accessed on 25 April 2021).

4.4. Syntenic and Genome Duplication Analysis of Peanut ARFs

The Multiple Collinearity Scan toolkit (MCScanX) was used to identify gene duplication events according to a previous study [89]. The amino acid sequences of duplicated ARF pairs were aligned first, and then used to guide the alignment of cDNA sequences with in-house Perl-scripts. The nonsynonymous (Ka) and synonymous (Ks) substitution ratios were calculated by the Ka/Ks calculator using the YN model. The divergence times (Mya) of A. hypogaea and the diploid ancestors (A. duranensis and A. ipaensis) were calculated with the formula T = Ks/2r. The r (neutral substitution rate) was taken to be 8.12 ×10–9 according to a previous study [79].

4.5. Genome-Wide Search for Downstream Targets of AhARFs

Upstream regions (2000 bp upstream of the start codon) of genes in the peanut genome were used for screening. Searching of the AuxRE (DR5 and IR8) was performed on MEME Suite (Motif-based sequence analysis tools, http://meme-suite.org/; accessed on 13 May 2021) using Motif Cluster Alignment and Search Tool (MCAST) [90,91]. The match p-value was set at <10–4. The peptide sequences were downloaded from PeanutBase. The GO annotation of genes carrying AuxRE repeats was performed using OmicsBox (Version 1.4.11).

4.6. Expression Analysis of AhARFs

The expression atlas of 22 A. hypogaea tissues was downloaded from PeanutBase (https://www.peanutbase.org/gene_expression/atlas; accessed on 21 April 2021) [79]. In these RNA-Seq data, the normalized reads were mapped to an in silico amphidiploid genome assembled from the genome of the diploid ancestors A. duranensis and A. ipaensis [92]. BLAST was performed to identify the homologous genes of AhARFs in A. duranensis and A. ipaensis. Only the AdARFs or AiARFs showed the highest similarity in amino acid sequence with AhARF was defined as homologous gene of AhARF. The IDs of the homologous gene were used to extract the fragments per kilobase of transcript per million mapped reads (FPKM) values from the tissue expression atlas.
Quick RNA isolation kit (Waryong, Beijing, China) was used to isolate the total RNA following the manufacturer’s instructions. Samples were quantified by NanoDrop 2000 microvolume spectrophotometry (Thermo, Wilmington, DE, USA), and 1 μg RNA was reverse-transcribed by the PrimeScript RT reagent Kit with DNA Eraser (Takara Bio, Dalian, China) following the manufacturer protocol. Total RNA and cDNA were stored at −80 °C and −20 °C, respectively. qRT-PCR was performed on an ABI StepOne Real-Time PCR Systerms (Thermo, Wilmington, DE, USA) using SYBR Premix Ex Taq (Takara, Dalian, China). Details of primers are provided in Table S6. AhUKN1 was used as the internal reference [93]. Three technical replicates were included in each biological replicate, and three biological replicates were performed.

4.7. Root Morphology Analysis

For root morphology analysis, roots of 21-day-old seedlings were harvest and washed with deionized water. The WinRHIZO LA-S image analysis system (WinRHizo LA-S, Regent Instr. Inc., Quebec, Canada) was used for scanning and quantization of total root length (TRL), root surface area (RSA), root mean diameter per plant (ARD), and root tip number (RTN). Three independent biological replications were performed for each germplasm, and three technical repeats were included in each biological replication. For each technical repeat, roots from 6 plants were scanned and analyzed.

4.8. Statistical Analysis

GraphPad Prism version 8.0.2 (GraphPad, San Diego, CA, USA) was used to prepare the figures. Data analysis was performed by SPSS 21.0 (SPSS Inc., Chicago, IL, USA) using one-way ANOVA. Differences were considered significant at a probability level of p < 0.05.

4.9. Subcellular Localization Analysis

The full-length CDS of AhARF14 and AhARF26 were amplified from JK using cDNA as a template. Primers used are listed in Table S6. The purified PCR products were cut by BamH I and Sal I (TaKaRa, Dalian, China) and cloned into pBI122 vector. The sequenced pBI122-35S::AhARFs-GFP vectors were transformed into Agrobacterium tumefaciens strain GV3101 and coinfected into N. benthamiana leaf epidermal cells. Plants were grown at 26 °C for 48–60 h, and GFP signals in the leaves were detected at 488 nm using an LSM 51 confocal laser scanning microscope (Carl Zeiss, Jena, Germany). The pBI122-35S::GFP vector was used as a control. Three independent biological repeats were performed. Subcellular localization of AhARF14 and AhARF26 was predicted using online tool DeepLoc-1.0 (https://services.healthtech.dtu.dk/service.php?DeepLoc-1.0; accessed on 21 January 2022).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms23105309/s1.

Author Contributions

Conceptualization, F.L. and L.L.; Methodology, L.L. and Q.W.; Software, S.Z.; Validation, L.L. and Q.W.; Formal Analysis, L.L. and X.Z.; Investigation, L.L. and Q.W.; Resources, Y.W.; Data Curation, K.Z.; Writing–Original Draft Preparation, L.L. and Q.W.; Writing–Review and Editing, Z.Y., Z.D. and L.L.; Visualization, L.L. and Q.W.; Supervision, Y.W.; Project Administration, Z.D. and F.L.; Funding Acquisition, F.L. and K.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Peanut Seed Industry Project in Shandong Province, China (No. 2020LZGC001), an earmarked fund for the Agriculture Research System in Shandong Province, China (No. SDAIT-04-03), the Natural Science Foundation of China (NSFC, 31871561), the earmarked fund for China Agriculture Research System (CARS-13), and Postdoctoral Funding of Shandong University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study or in the collection, analyses, or interpretation of the data.

References

  1. Mironova, V.; Teale, W.; Shahriari, M.; Dawson, J.; Palme, K. The Systems Biology of Auxin in Developing Embryos. Trends Plant Sci. 2017, 22, 225–235. [Google Scholar] [CrossRef] [PubMed]
  2. Gallei, M.; Luschnig, C.; Friml, J. Auxin signalling in growth: Schrodinger’s cat out of the bag. Curr. Opin. Plant Biol. 2020, 53, 43–49. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, L.; Yu, Z.; Xu, Y.; Guo, C.; Zhang, L.; Wu, C.; Yang, G.; Huang, J.; Yan, K.; Shu, H.; et al. Function identification of MdTIR1 in apple root growth benefited from the predicted MdPPI network. J. Integr. Plant Biol. 2020, 63, 723–739. [Google Scholar] [CrossRef]
  4. Chandler, J.W. Auxin response factors. Plant Cell Environ. 2016, 39, 1014–1028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Boer, D.R.; Freire-Rios, A.; van den Berg, W.A.; Saaki, T.; Manfield, I.W.; Kepinski, S.; Lopez-Vidrieo, I.; Franco-Zorrilla, J.M.; de Vries, S.C.; Solano, R.; et al. Structural basis for DNA binding specificity by the auxin-dependent ARF transcription factors. Cell 2014, 156, 577–589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Finet, C.; Berne-Dedieu, A.; Scutt, C.P.; Marletaz, F. Evolution of the ARF gene family in land plants: Old domains, new tricks. Mol. Biol. Evol. 2013, 30, 45–56. [Google Scholar] [CrossRef] [Green Version]
  7. Roosjen, M.; Paque, S.; Weijers, D. Auxin Response Factors: Output control in auxin biology. J. Exp. Bot. 2018, 69, 179–188. [Google Scholar] [CrossRef]
  8. Weijers, D.; Wagner, D. Transcriptional Responses to the Auxin Hormone. Annu. Rev. Plant Biol. 2016, 67, 539–574. [Google Scholar] [CrossRef]
  9. Swarup, R.; Bhosale, R. Developmental Roles of AUX1/LAX Auxin Influx Carriers in Plants. Front. Plant Sci. 2019, 10, 1306. [Google Scholar] [CrossRef] [Green Version]
  10. Hayashi, K.I. Chemical Biology in Auxin Research. Cold Spring Harb. Perspect. Biol. 2021, 13, a040105. [Google Scholar] [CrossRef]
  11. Xing, H.; Pudake, R.N.; Guo, G.; Xing, G.; Hu, Z.; Zhang, Y.; Sun, Q.; Ni, Z. Genome-wide identification and expression profiling of auxin response factor (ARF) gene family in maize. BMC Genom. 2011, 12, 178. [Google Scholar] [CrossRef] [Green Version]
  12. Okushima, Y.; Overvoorde, P.J.; Arima, K.; Alonso, J.M.; Chan, A.; Chang, C.; Ecker, J.R.; Hughes, B.; Lui, A.; Nguyen, D.; et al. Functional genomic analysis of the AUXIN RESPONSE FACTOR gene family members in Arabidopsis thaliana: Unique and overlapping functions of ARF7 and ARF19. Plant Cell 2005, 17, 444–463. [Google Scholar] [CrossRef] [Green Version]
  13. Wang, D.; Pei, K.; Fu, Y.; Sun, Z.; Li, S.; Liu, H.; Tang, K.; Han, B.; Tao, Y. Genome-wide analysis of the auxin response factors (ARF) gene family in rice (Oryza sativa). Gene 2007, 394, 13–24. [Google Scholar] [CrossRef]
  14. Wang, J.W.; Wang, L.J.; Mao, Y.B.; Cai, W.J.; Xue, H.W.; Chen, X.Y. Control of root cap formation by MicroRNA-targeted auxin response factors in Arabidopsis. Plant Cell 2005, 17, 2204–2216. [Google Scholar] [CrossRef] [Green Version]
  15. Wilmoth, J.C.; Wang, S.; Tiwari, S.B.; Joshi, A.D.; Hagen, G.; Guilfoyle, T.J.; Alonso, J.M.; Ecker, J.R.; Reed, J.W. NPH4/ARF7 and ARF19 promote leaf expansion and auxin-induced lateral root formation. Plant J. Cell Mol. Biol. 2005, 43, 118–130. [Google Scholar] [CrossRef]
  16. Zhang, F.; Tao, W.; Sun, R.; Wang, J.; Li, C.; Kong, X.; Tian, H.; Ding, Z. PRH1 mediates ARF7-LBD dependent auxin signaling to regulate lateral root development in Arabidopsis thaliana. PLoS Genet. 2020, 16, e1008044. [Google Scholar] [CrossRef]
  17. Lee, H.W.; Cho, C.; Pandey, S.K.; Park, Y.; Kim, M.J.; Kim, J. LBD16 and LBD18 acting downstream of ARF7 and ARF19 are involved in adventitious root formation in Arabidopsis. BMC Plant Biol. 2019, 19, 46. [Google Scholar] [CrossRef] [Green Version]
  18. Wang, Y.; Li, K.; Chen, L.; Zou, Y.; Liu, H.; Tian, Y.; Li, D.; Wang, R.; Zhao, F.; Ferguson, B.J.; et al. MicroRNA167-Directed Regulation of the Auxin Response Factors GmARF8a and GmARF8b Is Required for Soybean Nodulation and Lateral Root Development. Plant Physiol. 2015, 168, 984–999. [Google Scholar] [CrossRef] [Green Version]
  19. Ren, Z.; Liu, R.; Gu, W.; Dong, X. The Solanum lycopersicum auxin response factor SlARF2 participates in regulating lateral root formation and flower organ senescence. Plant Sci. Int. J. Exp. Plant Biol. 2017, 256, 103–111. [Google Scholar] [CrossRef]
  20. Stigliani, A.; Martin-Arevalillo, R.; Lucas, J.; Bessy, A.; Vinos-Poyo, T.; Mironova, V.; Vernoux, T.; Dumas, R.; Parcy, F. Capturing Auxin Response Factors Syntax Using DNA Binding Models. Mol. Plant 2019, 12, 822–832. [Google Scholar] [CrossRef]
  21. Leyser, O. Auxin Signaling. Plant Physiol. 2018, 176, 465–479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Moretzsohn, M.C.; Gouvea, E.G.; Inglis, P.W.; Leal-Bertioli, S.C.; Valls, J.F.; Bertioli, D.J. A study of the relationships of cultivated peanut (Arachis hypogaea) and its most closely related wild species using intron sequences and microsatellite markers. Ann. Bot. 2013, 111, 113–126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zhang, X.; Zhu, S.; Zhang, K.; Wan, Y.; Liu, F.; Sun, Q.; Li, Y. Establishment and evaluation of a peanut association panel and analysis of key nutritional traits. J. Integr. Plant Biol. 2018, 60, 195–215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Zhang, K.; Liu, Y.; Luo, L.; Zhang, X.; Li, G.; Wan, Y.; Liu, F. Root traits of peanut cultivars with different drought resistant under drought stress at flowering and pegging phase. Acta Agric. Scand. Sect. B—Soil Plant Sci. 2021, 71, 363–376. [Google Scholar] [CrossRef]
  25. Kong, X.; Zhang, C.; Zheng, H.; Sun, M.; Zhang, F.; Zhang, M.; Cui, F.; Lv, D.; Liu, L.; Guo, S.; et al. Antagonistic Interaction between Auxin and SA Signaling Pathways Regulates Bacterial Infection through Lateral Root in Arabidopsis. Cell Rep. 2020, 32, 108060. [Google Scholar] [CrossRef] [PubMed]
  26. Attia, K.A.; Abdelkhalik, A.F.; Ammar, M.H.; Wei, C.; Yang, J.; Lightfoot, D.A.; El-Sayed, W.M.; El-Shemy, H.A. Antisense phenotypes reveal a functional expression of OsARF1, an auxin response factor, in transgenic rice. Curr. Issues Mol. Biol. 2009, 11 (Suppl. 1), i29–i34. [Google Scholar] [PubMed]
  27. Tabata, R.; Ikezaki, M.; Fujibe, T.; Aida, M.; Tian, C.E.; Ueno, Y.; Yamamoto, K.T.; Machida, Y.; Nakamura, K.; Ishiguro, S. Arabidopsis auxin response factor6 and 8 regulate jasmonic acid biosynthesis and floral organ development via repression of class 1 KNOX genes. Plant Cell Physiol. 2010, 51, 164–175. [Google Scholar] [CrossRef] [Green Version]
  28. Hardtke, C.S.; Berleth, T. The Arabidopsis gene MONOPTEROS encodes a transcription factor mediating embryo axis formation and vascular development. EMBO J. 1998, 17, 1405–1411. [Google Scholar] [CrossRef]
  29. Huang, K.L.; Ma, G.J.; Zhang, M.L.; Xiong, H.; Wu, H.; Zhao, C.Z.; Liu, C.S.; Jia, H.X.; Chen, L.; Kjorven, J.O.; et al. The ARF7 and ARF19 Transcription Factors Positively Regulate PHOSPHATE STARVATION RESPONSE1 in Arabidopsis Roots. Plant Physiol. 2018, 178, 413–427. [Google Scholar] [CrossRef] [Green Version]
  30. Wang, S.; Zhang, S.; Sun, C.; Xu, Y.; Chen, Y.; Yu, C.; Qian, Q.; Jiang, D.A.; Qi, Y. Auxin response factor (OsARF12), a novel regulator for phosphate homeostasis in rice (Oryza sativa). New Phytol. 2014, 201, 91–103. [Google Scholar] [CrossRef]
  31. Schruff, M.C.; Spielman, M.; Tiwari, S.; Adams, S.; Fenby, N.; Scott, R.J. The AUXIN RESPONSE FACTOR 2 gene of Arabidopsis links auxin signalling, cell division, and the size of seeds and other organs. Development 2006, 133, 251–261. [Google Scholar] [CrossRef] [Green Version]
  32. Breitel, D.A.; Chappell-Maor, L.; Meir, S.; Panizel, I.; Puig, C.P.; Hao, Y.; Yifhar, T.; Yasuor, H.; Zouine, M.; Bouzayen, M.; et al. AUXIN RESPONSE FACTOR 2 Intersects Hormonal Signals in the Regulation of Tomato Fruit Ripening. PLoS Genet. 2016, 12, e1005903. [Google Scholar] [CrossRef] [Green Version]
  33. Liu, X.; Dinh, T.T.; Li, D.; Shi, B.; Li, Y.; Cao, X.; Guo, L.; Pan, Y.; Jiao, Y.; Chen, X. AUXIN RESPONSE FACTOR 3 integrates the functions of AGAMOUS and APETALA2 in floral meristem determinacy. Plant J. Cell Mol. Biol. 2014, 80, 629–641. [Google Scholar] [CrossRef] [Green Version]
  34. Yuan, Y.; Xu, X.; Gong, Z.; Tang, Y.; Wu, M.; Yan, F.; Zhang, X.; Zhang, Q.; Yang, F.; Hu, X.; et al. Auxin response factor 6A regulates photosynthesis, sugar accumulation, and fruit development in tomato. Hortic. Res. 2019, 6, 85. [Google Scholar] [CrossRef] [Green Version]
  35. Zhang, S.; Wang, S.; Xu, Y.; Yu, C.; Shen, C.; Qian, Q.; Geisler, M.; Jiang de, A.; Qi, Y. The auxin response factor, OsARF19, controls rice leaf angles through positively regulating OsGH3-5 and OsBRI1. Plant Cell Environ. 2015, 38, 638–654. [Google Scholar] [CrossRef] [Green Version]
  36. Ellis, C.M.; Nagpal, P.; Young, J.C.; Hagen, G.; Guilfoyle, T.J.; Reed, J.W. AUXIN RESPONSE FACTOR1 and AUXIN RESPONSE FACTOR2 regulate senescence and floral organ abscission in Arabidopsis thaliana. Development 2005, 132, 4563–4574. [Google Scholar] [CrossRef] [Green Version]
  37. Zhang, K.; Wang, R.; Zi, H.; Li, Y.; Cao, X.; Li, D.; Guo, L.; Tong, J.; Pan, Y.; Jiao, Y.; et al. AUXIN RESPONSE FACTOR3 Regulates Floral Meristem Determinacy by Repressing Cytokinin Biosynthesis and Signaling. Plant Cell 2018, 30, 324–346. [Google Scholar] [CrossRef] [Green Version]
  38. Goetz, M.; Vivian-Smith, A.; Johnson, S.D.; Koltunow, A.M. AUXIN RESPONSE FACTOR8 is a negative regulator of fruit initiation in Arabidopsis. Plant Cell 2006, 18, 1873–1886. [Google Scholar] [CrossRef] [Green Version]
  39. Xu, X.F.; Wang, B.; Feng, Y.F.; Xue, J.S.; Qian, X.X.; Liu, S.Q.; Zhou, J.; Yu, Y.H.; Yang, N.Y.; Xu, P.; et al. AUXIN RESPONSE FACTOR17 Directly Regulates MYB108 for Anther Dehiscence. Plant Physiol. 2019, 181, 645–655. [Google Scholar] [CrossRef] [Green Version]
  40. Nagpal, P.; Ellis, C.M.; Weber, H.; Ploense, S.E.; Barkawi, L.S.; Guilfoyle, T.J.; Hagen, G.; Alonso, J.M.; Cohen, J.D.; Farmer, E.E.; et al. Auxin response factors ARF6 and ARF8 promote jasmonic acid production and flower maturation. Development 2005, 132, 4107–4118. [Google Scholar] [CrossRef] [Green Version]
  41. Chung, Y.; Zhu, Y.; Wu, M.F.; Simonini, S.; Kuhn, A.; Armenta-Medina, A.; Jin, R.; Ostergaard, L.; Gillmor, C.S.; Wagner, D. Auxin Response Factors promote organogenesis by chromatin-mediated repression of the pluripotency gene SHOOTMERISTEMLESS. Nat. Commun. 2019, 10, 886. [Google Scholar] [CrossRef] [Green Version]
  42. Zhang, X.; Yan, F.; Tang, Y.; Yuan, Y.; Deng, W.; Li, Z. Auxin Response Gene SlARF3 Plays Multiple Roles in Tomato Development and is Involved in the Formation of Epidermal Cells and Trichomes. Plant Cell Physiol. 2015, 56, 2110–2124. [Google Scholar] [CrossRef] [Green Version]
  43. Pashkovskiy, P.P.; Kartashov, A.V.; Zlobin, I.E.; Pogosyan, S.I.; Kuznetsov, V.V. Blue light alters miR167 expression and microRNA-targeted auxin response factor genes in Arabidopsis thaliana plants. Plant Physiol. Biochem. PPB 2016, 104, 146–154. [Google Scholar] [CrossRef]
  44. Zhang, H.; Li, L.; He, Y.; Qin, Q.; Chen, C.; Wei, Z.; Tan, X.; Xie, K.; Zhang, R.; Hong, G.; et al. Distinct modes of manipulation of rice auxin response factor OsARF17 by different plant RNA virumses for infection. Proc. Natl. Acad. Sci. USA 2020, 117, 9112–9121. [Google Scholar] [CrossRef]
  45. Bouzroud, S.; Gasparini, K.; Hu, G.; Barbosa, M.A.M.; Rosa, B.L.; Fahr, M.; Bendaou, N.; Bouzayen, M.; Zsogon, A.; Smouni, A.; et al. Down Regulation and Loss of Auxin Response Factor 4 Function Using CRISPR/Cas9 Alters Plant Growth, Stomatal Function and Improves Tomato Tolerance to Salinity and Osmotic Stress. Genes 2020, 11, 272. [Google Scholar] [CrossRef] [Green Version]
  46. Liu, N.; Wu, S.; Van Houten, J.; Wang, Y.; Ding, B.; Fei, Z.; Clarke, T.H.; Reed, J.W.; van der Knaap, E. Down-regulation of AUXIN RESPONSE FACTORS 6 and 8 by microRNA 167 leads to floral development defects and female sterility in tomato. J. Exp. Bot. 2014, 65, 2507–2520. [Google Scholar] [CrossRef] [Green Version]
  47. Kelley, D.R.; Arreola, A.; Gallagher, T.L.; Gasser, C.S. ETTIN (ARF3) physically interacts with KANADI proteins to form a functional complex essential for integument development and polarity determination in Arabidopsis. Development 2012, 139, 1105–1109. [Google Scholar] [CrossRef] [Green Version]
  48. Vert, G.; Walcher, C.L.; Chory, J.; Nemhauser, J.L. Integration of auxin and brassinosteroid pathways by Auxin Response Factor 2. Proc. Natl. Acad. Sci. USA 2008, 105, 9829–9834. [Google Scholar] [CrossRef] [Green Version]
  49. Su, Y.H.; Liu, Y.B.; Zhou, C.; Li, X.M.; Zhang, X.S. The microRNA167 controls somatic embryogenesis in Arabidopsis through regulating its target genes ARF6 and ARF8. Plant Cell, Tissue and Organ Culture 2016, 124, 405–417. [Google Scholar] [CrossRef]
  50. Luo, L.; Zeng, J.; Wu, H.; Tian, Z.; Zhao, Z. A Molecular Framework for Auxin-Controlled Homeostasis of Shoot Stem Cells in Arabidopsis. Mol. Plant 2018, 11, 899–913. [Google Scholar] [CrossRef] [Green Version]
  51. Ghelli, R.; Brunetti, P.; Napoli, N.; De Paolis, A.; Cecchetti, V.; Tsuge, T.; Serino, G.; Matsui, M.; Mele, G.; Rinaldi, G.; et al. A Newly Identified Flower-Specific Splice Variant of AUXIN RESPONSE FACTOR8 Regulates Stamen Elongation and Endothecium Lignification in Arabidopsis. Plant Cell 2018, 30, 620–637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Shen, C.; Yue, R.; Sun, T.; Zhang, L.; Yang, Y.; Wang, H. OsARF16, a transcription factor regulating auxin redistribution, is required for iron deficiency response in rice (Oryza sativa L.). Plant Sci. Int. J. Exp. Plant Biol. 2015, 231, 148–158. [Google Scholar] [CrossRef] [PubMed]
  53. Li, Y.; Li, J.; Chen, Z.; Wei, Y.; Qi, Y.; Wu, C. OsmiR167a-targeted auxin response factors modulate tiller angle via fine-tuning auxin distribution in rice. Plant Biotechnol. J. 2020, 18, 2015–2026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Kalve, S.; Sizani, B.L.; Markakis, M.N.; Helsmoortel, C.; Vandeweyer, G.; Laukens, K.; Sommen, M.; Naulaerts, S.; Vissenberg, K.; Prinsen, E.; et al. Osmotic stress inhibits leaf growth of Arabidopsis thaliana by enhancing ARF-mediated auxin responses. New Phytol. 2020, 226, 1766–1780. [Google Scholar] [CrossRef]
  55. Finet, C.; Fourquin, C.; Vinauger, M.; Berne-Dedieu, A.; Chambrier, P.; Paindavoine, S.; Scutt, C.P. Parallel structural evolution of auxin response factors in the angiosperms. Plant J. Cell Mol. Biol. 2010, 63, 952–959. [Google Scholar] [CrossRef]
  56. Cheng, Z.J.; Wang, L.; Sun, W.; Zhang, Y.; Zhou, C.; Su, Y.H.; Li, W.; Sun, T.T.; Zhao, X.Y.; Li, X.G.; et al. Pattern of auxin and cytokinin responses for shoot meristem induction results from the regulation of cytokinin biosynthesis by AUXIN RESPONSE FACTOR3. Plant Physiol. 2013, 161, 240–251. [Google Scholar] [CrossRef] [Green Version]
  57. Orosa-Puente, B.; Leftley, N.; von Wangenheim, D.; Banda, J.; Srivastava, A.K.; Hill, K.; Truskina, J.; Bhosale, R.; Morris, E.; Srivastava, M.; et al. Root branching toward water involves posttranslational modification of transcription factor ARF7. Science 2018, 362, 1407–1410. [Google Scholar] [CrossRef] [Green Version]
  58. Nziengui, H.; Lasok, H.; Kochersperger, P.; Ruperti, B.; Rebeille, F.; Palme, K.; Ditengou, F.A. Root Gravitropism Is Regulated by a Crosstalk between para-Aminobenzoic Acid, Ethylene, and Auxin. Plant Physiol. 2018, 178, 1370–1389. [Google Scholar] [CrossRef] [Green Version]
  59. Sagar, M.; Chervin, C.; Mila, I.; Hao, Y.; Roustan, J.P.; Benichou, M.; Gibon, Y.; Biais, B.; Maury, P.; Latche, A.; et al. SlARF4, an auxin response factor involved in the control of sugar metabolism during tomato fruit development. Plant Physiol. 2013, 161, 1362–1374. [Google Scholar] [CrossRef] [Green Version]
  60. Yuan, Y.; Mei, L.; Wu, M.; Wei, W.; Shan, W.; Gong, Z.; Zhang, Q.; Yang, F.; Yan, F.; Zhang, Q.; et al. SlARF10, an auxin response factor, is involved in chlorophyll and sugar accumulation during tomato fruit development. J. Exp. Bot. 2018, 69, 5507–5518. [Google Scholar] [CrossRef] [Green Version]
  61. De Jong, M.; Wolters-Arts, M.; Schimmel, B.C.; Stultiens, C.L.; de Groot, P.F.; Powers, S.J.; Tikunov, Y.M.; Bovy, A.G.; Mariani, C.; Vriezen, W.H.; et al. Solanum lycopersicum AUXIN RESPONSE FACTOR 9 regulates cell division activity during early tomato fruit development. J. Exp. Bot. 2015, 66, 3405–3416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Reed, J.W.; Wu, M.F.; Reeves, P.H.; Hodgens, C.; Yadav, V.; Hayes, S.; Pierik, R. Three Auxin Response Factors Promote Hypocotyl Elongation. Plant Physiol. 2018, 178, 864–875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Hunter, C.; Willmann, M.R.; Wu, G.; Yoshikawa, M.; de la Luz Gutierrez-Nava, M.; Poethig, S.R. Trans-acting siRNA-mediated repression of ETTIN and ARF4 regulates heteroblasty in Arabidopsis. Development 2006, 133, 2973–2981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Roodbarkelari, F.; Du, F.; Truernit, E.; Laux, T. ZLL/AGO10 maintains shoot meristem stem cells during Arabidopsis embryogenesis by down-regulating ARF2-mediated auxin response. BMC Biol. 2015, 13, 74. [Google Scholar] [CrossRef] [Green Version]
  65. Wang, X.; Yu, R.; Wang, J.; Lin, Z.; Han, X.; Deng, Z.; Fan, L.; He, H.; Deng, X.W.; Chen, H. The Asymmetric Expression of SAUR Genes Mediated by ARF7/19 Promotes the Gravitropism and Phototropism of Plant Hypocotyls. Cell Rep. 2020, 31, 107529. [Google Scholar] [CrossRef]
  66. Peng, J.; Berbel, A.; Madueno, F.; Chen, R. AUXIN RESPONSE FACTOR3 Regulates Compound Leaf Patterning by Directly Repressing PALMATE-LIKE PENTAFOLIATA1 Expression in Medicago truncatula. Front. Plant Sci. 2017, 8, 1630. [Google Scholar] [CrossRef] [Green Version]
  67. Zhuang, W.; Chen, H.; Yang, M.; Wang, J.; Pandey, M.K.; Zhang, C.; Chang, W.C.; Zhang, L.; Zhang, X.; Tang, R.; et al. The genome of cultivated peanut provides insight into legume karyotypes, polyploid evolution and crop domestication. Nat. Genet. 2019, 51, 865–876. [Google Scholar] [CrossRef]
  68. Li, P.; Ma, Q.; Qu, C.; Zhu, S.; Zhao, K.; Ma, X.; Li, Z.; Zhang, X.; Gong, F.; Yin, D. Genome-wide identification and expression analysis of auxin response factors in peanut (Arachis hypogaea L.). PeerJ 2021, 9, e12319. [Google Scholar] [CrossRef]
  69. Xu, Y.; Zhang, G.; Ding, H.; Ci, D.; Dai, L.; Zhang, Z. Influence of salt stress on the rhizosphere soil bacterial community structure and growth performance of groundnut (Arachis hypogaea L.). Int. Microbiol. 2020, 23, 453–465. [Google Scholar] [CrossRef]
  70. Xu, Y.; Zhang, D.; Dai, L.; Ding, H.; Ci, D.; Qin, F.; Zhang, G.; Zhang, Z. Influence of Salt Stress on Growth of Spermosphere Bacterial Communities in Different Peanut (Arachis hypogaea L.) Cultivars. Int. J. Mol. Sci. 2020, 21, 2131. [Google Scholar] [CrossRef] [Green Version]
  71. Zhao, S.-c.; Lü, J.-l.; Xu, X.-p.; Lin, X.-m.; Luiz, M.R.; Qiu, S.-j.; Ciampitti, I.; He, P. Peanut yield, nutrient uptake and nutrient requirements in different regions of China. J. Integr. Agric. 2021, 20, 2502–2511. [Google Scholar] [CrossRef]
  72. Dai, L.; Zhang, G.; Yu, Z.; Ding, H.; Xu, Y.; Zhang, Z. Effect of Drought Stress and Developmental Stages on Microbial Community Structure and Diversity in Peanut Rhizosphere Soil. Int. J. Mol. Sci. 2019, 20, 2265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ci, D.; Tang, Z.; Ding, H.; Cui, L.; Zhang, G.; Li, S.; Dai, L.; Qin, F.; Zhang, Z.; Yang, J.; et al. The synergy effect of arbuscular mycorrhizal fungi symbiosis and exogenous calcium on bacterial community composition and growth performance of peanut (Arachis hypogaea L.) in saline alkali soil. J. Microbiol. 2021, 59, 51–63. [Google Scholar] [CrossRef] [PubMed]
  74. De Dorlodot, S.; Forster, B.; Pages, L.; Price, A.; Tuberosa, R.; Draye, X. Root system architecture: Opportunities and constraints for genetic improvement of crops. Trends Plant Sci. 2007, 12, 474–481. [Google Scholar] [CrossRef]
  75. Ding, T.; Zhang, F.; Wang, J.; Wang, F.; Liu, J.; Xie, C.; Hu, Y.; Shani, E.; Kong, X.; Ding, Z.; et al. Cell-type action specificity of auxin on Arabidopsis root growth. Plant J. 2021, 106, 928–941. [Google Scholar] [CrossRef]
  76. Lv, B.; Wei, K.; Hu, K.; Tian, T.; Zhang, F.; Yu, Z.; Zhang, D.; Su, Y.; Sang, Y.; Zhang, X.; et al. MPK14-mediated auxin signaling controls lateral root development via ERF13-regulated very-long-chain fatty acid biosynthesis. Mol. Plant 2021, 14, 285–297. [Google Scholar] [CrossRef]
  77. Jing, H.; Strader, L.C. Interplay of Auxin and Cytokinin in Lateral Root Development. Int. J. Mol. Sci. 2019, 20, 486. [Google Scholar] [CrossRef] [Green Version]
  78. Yu, Z.; Zhang, F.; Friml, J.; Ding, Z. Auxin signaling: Research advances over the past 30 years. J. Integr. Plant Biol. 2022, 64, 371–392. [Google Scholar] [CrossRef]
  79. Bertioli, D.J.; Cannon, S.B.; Froenicke, L.; Huang, G.; Farmer, A.D.; Cannon, E.K.; Liu, X.; Gao, D.; Clevenger, J.; Dash, S.; et al. The genome sequences of Arachis duranensis and Arachis ipaensis, the diploid ancestors of cultivated peanut. Nat. Genet. 2016, 48, 438–446. [Google Scholar] [CrossRef]
  80. Fletcher, R.S.; Mullen, J.L.; Heiliger, A.; McKay, J.K. QTL analysis of root morphology, flowering time, and yield reveals trade-offs in response to drought in Brassica napus. J. Exp. Bot. 2015, 66, 245–256. [Google Scholar] [CrossRef] [Green Version]
  81. Soriano, J.M.; Alvaro, F. Discovering consensus genomic regions in wheat for root-related traits by QTL meta-analysis. Sci. Rep. 2019, 9, 10537. [Google Scholar] [CrossRef] [Green Version]
  82. Chen, H.; Kumawat, G.; Yan, Y.; Fan, B.; Xu, D. Mapping and validation of a major QTL for primary root length of soybean seedlings grown in hydroponic conditions. BMC Genom. 2021, 22, 132. [Google Scholar] [CrossRef]
  83. Fan, X.; Zhang, W.; Zhang, N.; Chen, M.; Zheng, S.; Zhao, C.; Han, J.; Liu, J.; Zhang, X.; Song, L.; et al. Identification of QTL regions for seedling root traits and their effect on nitrogen use efficiency in wheat (Triticum aestivum L.). TAG Theor. Appl. Genet. Theor. Angew. Genet. 2018, 131, 2677–2698. [Google Scholar] [CrossRef]
  84. Wu, Y.; Zhang, S.; Zhang, H.; Li, F.; Li, G.; Fan, C.; Sun, R.; Zhang, S. QTL Mapping and Candidate Gene Identification of Swollen Root Formation in Turnip. Int. J. Mol. Sci. 2021, 22, 653. [Google Scholar] [CrossRef]
  85. Diao, D.; Hu, X.; Guan, D.; Wang, W.; Yang, H.; Liu, Y. Genome-wide identification of the ARF (auxin response factor) gene family in peach and their expression analysis. Mol. Biol. Rep. 2020, 47, 4331–4344. [Google Scholar] [CrossRef]
  86. Pan, C.L.; Yao, S.C.; Xiong, W.J.; Luo, S.Z.; Wang, Y.L.; Wang, A.Q.; Xiao, D.; Zhan, J.; He, L.F. Nitric Oxide Inhibits Al-Induced Programmed Cell Death in Root Tips of Peanut (Arachis hypogaea L.) by Affecting Physiological Properties of Antioxidants Systems and Cell Wall. Front. Physiol. 2017, 8, 1037. [Google Scholar] [CrossRef] [Green Version]
  87. Bertioli, D.J.; Jenkins, J.; Clevenger, J.; Dudchenko, O.; Gao, D.; Seijo, G.; Leal-Bertioli, S.C.M.; Ren, L.; Farmer, A.D.; Pandey, M.K.; et al. The genome sequence of segmental allotetraploid peanut Arachis hypogaea. Nat. Genet. 2019, 51, 877–884. [Google Scholar] [CrossRef] [Green Version]
  88. 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] [Green Version]
  89. Wang, Y.; Tang, H.; Debarry, J.D.; Tan, X.; Li, J.; Wang, X.; Lee, T.H.; Jin, H.; Marler, B.; Guo, H.; et al. MCScanX: A toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Res. 2012, 40, e49. [Google Scholar] [CrossRef] [Green Version]
  90. Bailey, T.L.; Noble, W.S. Searching for statistically significant regulatory modules. Bioinformatics 2003, 19, ii16–ii25. [Google Scholar] [CrossRef] [Green Version]
  91. Freire-Rios, A.; Tanaka, K.; Crespo, I.; van der Wijk, E.; Sizentsova, Y.; Levitsky, V.; Lindhoud, S.; Fontana, M.; Hohlbein, J.; Boer, D.R.; et al. Architecture of DNA elements mediating ARF transcription factor binding and auxin-responsive gene expression in Arabidopsis. Proc. Natl. Acad. Sci. USA 2020, 117, 24557–24566. [Google Scholar] [CrossRef] [PubMed]
  92. Clevenger, J.; Chu, Y.; Scheffler, B.; Ozias-Akins, P. A Developmental Transcriptome Map for Allotetraploid Arachis hypogaea. Front. Plant Sci. 2016, 7, 1446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Lv, Y.; Zhang, X.; Luo, L.; Yang, H.; Li, P.; Zhang, K.; Liu, F.; Wan, Y. Characterization of glycerol-3-phosphate acyltransferase 9 (AhGPAT9) genes, their allelic polymorphism and association with oil content in peanut (Arachis hypogaea L.). Sci. Rep. 2020, 10, 14648. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chromosome distribution of ARFs in cultivar (A) and wild (B) peanuts. Chromosomes belong to the A sub-genome of A. hypogaea (Arahy.02–10) and their homologues from A. duranensis are colored in yellow, while the chromosomes that belong to the B sub-genome of A. hypogaea (Arahy.12–20) and their homologues from A. ipaensis are colored in green. Chromosomes that do not contain ARFs and ARFs located on the scaffold are not shown.
Figure 1. Chromosome distribution of ARFs in cultivar (A) and wild (B) peanuts. Chromosomes belong to the A sub-genome of A. hypogaea (Arahy.02–10) and their homologues from A. duranensis are colored in yellow, while the chromosomes that belong to the B sub-genome of A. hypogaea (Arahy.12–20) and their homologues from A. ipaensis are colored in green. Chromosomes that do not contain ARFs and ARFs located on the scaffold are not shown.
Ijms 23 05309 g001
Figure 2. Phylogenetic tree, cis-elements, and gene structure analysis of ARFs in wild and cultivar peanuts. Gene structures are shown in the outer cycle, and cis-elements are shown in the cycle inside the gene names. The red, green, and blue dots represent AhARFs, AdARFs, and AiARFs, respectively.
Figure 2. Phylogenetic tree, cis-elements, and gene structure analysis of ARFs in wild and cultivar peanuts. Gene structures are shown in the outer cycle, and cis-elements are shown in the cycle inside the gene names. The red, green, and blue dots represent AhARFs, AdARFs, and AiARFs, respectively.
Ijms 23 05309 g002
Figure 3. Duplication derived ARFs in wild and cultivar peanuts. (A) The orthologous genes (black lines), paralogous genes (red), and tandem repeat (blue) in cultivar peanut. (B) The Ks value distribution of the duplicated orthologous genes between the wild and cultivar peanut.
Figure 3. Duplication derived ARFs in wild and cultivar peanuts. (A) The orthologous genes (black lines), paralogous genes (red), and tandem repeat (blue) in cultivar peanut. (B) The Ks value distribution of the duplicated orthologous genes between the wild and cultivar peanut.
Ijms 23 05309 g003
Figure 4. Heatmap illustration of tissue expression patterns of AhARFs based on FPKM values. The log2 transformation of FPKM values and visualization were performed by TBtools. Scalebar on top right indicates the levels of gene expression.
Figure 4. Heatmap illustration of tissue expression patterns of AhARFs based on FPKM values. The log2 transformation of FPKM values and visualization were performed by TBtools. Scalebar on top right indicates the levels of gene expression.
Ijms 23 05309 g004
Figure 5. Genome-wide prediction of ARF-binding elements. (A) Definition of DR5 and IR8 elements. N indicates A, C, G, or T. (B) Number of DR5, IR8, and AuxRE in peanut genomes. (C) GO analysis of genes carrying DR5, IR8, or AuxRE in their promoters, and only the GO terms of those containing at least 40 genes are shown here.
Figure 5. Genome-wide prediction of ARF-binding elements. (A) Definition of DR5 and IR8 elements. N indicates A, C, G, or T. (B) Number of DR5, IR8, and AuxRE in peanut genomes. (C) GO analysis of genes carrying DR5, IR8, or AuxRE in their promoters, and only the GO terms of those containing at least 40 genes are shown here.
Ijms 23 05309 g005
Figure 6. Root morphology and their correlation with AhARF expression levels in peanut germplasms. (A) Root morphology of peanut germplasms. JK, Jinkins Jumbo; 41165, Meiyinxuan 41165; SX64, Shixuan 64; JH27, Juhua 27; FH2, Fenghua 2; EH3, Ehua 3; SH11, Shanhua 11; ZH12, Zhonghua 12. (B) Total root length, root surface area, root mean diameter, and root tip number of peanut germplasms. Lower case letters (a, b, c, d) indicate statistically significant differences between cultivars. (C) Relative expression levels of AhARFs in peanut roots. (D) Pearson’s correlation analysis between root morphology and expression of AhARFs. R2, Pearson’s correlation coefficient.
Figure 6. Root morphology and their correlation with AhARF expression levels in peanut germplasms. (A) Root morphology of peanut germplasms. JK, Jinkins Jumbo; 41165, Meiyinxuan 41165; SX64, Shixuan 64; JH27, Juhua 27; FH2, Fenghua 2; EH3, Ehua 3; SH11, Shanhua 11; ZH12, Zhonghua 12. (B) Total root length, root surface area, root mean diameter, and root tip number of peanut germplasms. Lower case letters (a, b, c, d) indicate statistically significant differences between cultivars. (C) Relative expression levels of AhARFs in peanut roots. (D) Pearson’s correlation analysis between root morphology and expression of AhARFs. R2, Pearson’s correlation coefficient.
Ijms 23 05309 g006
Figure 7. Subcellular localization analysis of AhARFs in N. benthamiana leaves.
Figure 7. Subcellular localization analysis of AhARFs in N. benthamiana leaves.
Ijms 23 05309 g007
Table 1. The nonsynonymous (Ka) and synonymous (Ks) substitution rates of orthologue AhARF pairs in A. hypogaea.
Table 1. The nonsynonymous (Ka) and synonymous (Ks) substitution rates of orthologue AhARF pairs in A. hypogaea.
Gene 1Gene 2KaKsKa/KsDivergence Time (Mya)
AhARF3AhARF340.03770.05170.72886.37
AhARF1AhARF310.01160.01660.69892.04
AhARF18AhARF460.03490.05710.61027.03
AhARF17AhARF470.00830.01690.49112.08
AhARF26AhARF590.01410.03040.46413.74
AhARF6AhARF370.02950.06440.45857.93
AhARF28AhARF610.00450.01070.42341.32
AhARF12AhARF430.01320.03200.41103.94
AhARF16AhARF480.01650.04160.39735.12
AhARF15AhARF490.00970.03230.30153.98
AhARF25AhARF580.01380.04610.29845.68
AhARF40AhARF610.17410.69580.250285.69
AhARF28AhARF400.18440.74190.248691.36
AhARF37AhARF560.16380.69230.236785.25
AhARF19AhARF500.00490.02060.23622.54
AhARF36AhARF570.16180.75330.214892.77
AhARF14AhARF450.00770.04420.17415.44
AhARF5AhARF360.00250.01500.16451.85
AhARF39AhARF580.13350.86810.1538106.91
AhARF11AhARF420.00230.01880.12342.32
AhARF2AhARF330.00250.03110.08053.83
AhARF22AhARF560.00220.02730.07933.36
AhARF8AhARF380.00340.05250.06476.47
AhARF9AhARF400.00130.02040.06172.51
AhARF20AhARF510.00050.03290.01544.06
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Luo, L.; Wan, Q.; Yu, Z.; Zhang, K.; Zhang, X.; Zhu, S.; Wan, Y.; Ding, Z.; Liu, F. Genome-Wide Identification of Auxin Response Factors in Peanut (Arachis hypogaea L.) and Functional Analysis in Root Morphology. Int. J. Mol. Sci. 2022, 23, 5309. https://doi.org/10.3390/ijms23105309

AMA Style

Luo L, Wan Q, Yu Z, Zhang K, Zhang X, Zhu S, Wan Y, Ding Z, Liu F. Genome-Wide Identification of Auxin Response Factors in Peanut (Arachis hypogaea L.) and Functional Analysis in Root Morphology. International Journal of Molecular Sciences. 2022; 23(10):5309. https://doi.org/10.3390/ijms23105309

Chicago/Turabian Style

Luo, Lu, Qian Wan, Zipeng Yu, Kun Zhang, Xiurong Zhang, Suqing Zhu, Yongshan Wan, Zhaojun Ding, and Fengzhen Liu. 2022. "Genome-Wide Identification of Auxin Response Factors in Peanut (Arachis hypogaea L.) and Functional Analysis in Root Morphology" International Journal of Molecular Sciences 23, no. 10: 5309. https://doi.org/10.3390/ijms23105309

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