Next Article in Journal
Artesunate Dry Emulsion Formulation Combined with Antibiotics for Treatment of Helicobacter pylori Infections: In Vitro/In Vivo Evaluation
Previous Article in Journal
Manipulating Microbiota in Inflammatory Bowel Disease Treatment: Clinical and Natural Product Interventions Explored
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Identification and Expression Analysis of Auxin Response Factor Gene Family in Linum usitatissimum

Institute of Crop, Gansu Academy of Agricultural Sciences, Lanzhou 730070, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(13), 11006; https://doi.org/10.3390/ijms241311006
Submission received: 24 May 2023 / Revised: 26 June 2023 / Accepted: 27 June 2023 / Published: 2 July 2023
(This article belongs to the Section Molecular Biology)

Abstract

:
Auxin response factors (ARFs) are critical components of the auxin signaling pathway, and are involved in diverse plant biological processes. However, ARF genes have not been investigated in flax (Linum usitatissimum L.), an important oilseed and fiber crop. In this study, we comprehensively analyzed the ARF gene family and identified 33 LuARF genes unevenly distributed on the 13 chromosomes of Longya-10, an oil-use flax variety. Detailed analysis revealed wide variation among the ARF family members and predicted nuclear localization for all proteins. Nineteen LuARFs contained a complete ARF structure, including DBD, MR, and CTD, whereas the other fourteen lacked the CTD. Phylogenetic analysis grouped the LuARFs into four (I–V) clades. Combined with sequence analysis, the LuARFs from the same clade showed structural conservation, implying functional redundancy. Duplication analysis identified twenty-seven whole-genome-duplicated LuARF genes and four tandem-duplicated LuARF genes. These duplicated gene pairs’ Ka/Ks ratios suggested a strong purifying selection pressure on the LuARF genes. Collinearity analysis revealed that about half of the LuARF genes had homologs in other species, indicating a relatively conserved nature of the ARFs. The promoter analysis identified numerous hormone- and stress-related elements, and the qRT-PCR experiment revealed that all LuARF genes were responsive to phytohormone (IAA, GA3, and NAA) and stress (PEG, NaCl, cold, and heat) treatments. Finally, expression profiling of LuARF genes in different tissues by qRT-PCR indicated their specific functions in stem or capsule growth. Thus, our findings suggest the potential functions of LuARFs in flax growth and response to an exogenous stimulus, providing a basis for further functional studies on these genes.

1. Introduction

Auxins are a class of phytohormones found extensively in plants. They play vital roles in regulating various plant processes, including seed germination, root architecture and development, shoot elongation, inflorescence development, xylem and phloem differentiation, fruit development, leaf expansion and senescence, shoot apical dominance, and plant stress response [1,2,3,4]. Research has demonstrated that the changes in auxin levels regulate the transient expression of auxin response genes, triggering the reprogramming of downstream genes and consequently influencing plant growth. The major auxin response genes include auxin/indole-3-acetic acid (AUX/IAA) family, auxin response factor (ARF) family, gretchenhagen 3 (GH3) family, and small auxin upregulated RNAs (SAURs) [5]. Among these, ARFs, the well-known transcription factors, are essential components of the auxin-mediated signaling pathways. They recognize and specifically bind with the auxin-response element (TGTCNN; AuxRE) in the promoter region of the target genes and regulate their transcription [6,7,8].
A typical ARF protein contains three conserved domains: an N-terminal DNA-binding B3 domain (DBD), a variable middle region (MR), and a C-terminal domain (CTD), which may form a type I/II Phox and Bem1p (PB1) protein–protein interaction domain [9,10,11]. The DBD of the ARF protein recognizes the AuxREs, while the MR acts as an activation domain (AD) or repression domain (RD). The amino acid composition of the MR determines whether ARFs activate or inhibit the transcription of downstream genes. Typically, glutamine (Q), serine (S), and leucine (L) are more in transcriptional activators, while serine (S), proline (P), glycine (G), and leucine (L) are abundant in transcriptional repressors [6,9]. However, CTD is present in only a few ARF proteins; it plays a vital role in plant-specific responses to auxin by mediating the interactions between Aux/IAA and ARF in a dose-dependent manner [12,13]. In plants, at low auxin levels or in the absence of auxin, the Aux/IAA proteins form multimers with ARFs; this complex binds to the co-repressor TOPLESS (TPL) and its family proteins (TPRs), inhibits the ARFs, and prevents the expression of downstream auxin-responsive genes [14]. However, when the auxin concentration is above a certain level, Aux/IAA is ubiquitinated by the ubiquitin ligase SCFTIR1/AFB subunit (transport inhibitor response 1, TIR1), and subsequently, degraded through the 26S proteasome pathway [15,16], releasing the ARFs to regulate the transcription of downstream genes.
Research has elucidated the functions of ARF genes in several plant species. In Arabidopsis thaliana, AtARF1 and AtARF2 regulate leaf senescence and floral abscission [17]; AtARF3 regulates shoot apical meristem maintenance [18]; AtARF5 regulates the patterning of vascular bundles, the development of hypocotyls, and the development of female and male gametophytes [19,20]; ARF6 and ARF8 regulate flower maturation [21]; and AtARF7 and AtARF19 regulate lateral root formation [22]. In rice, OsARF1 and OsARF12 play a vital role in regulating vegetative and reproductive development and phosphate homeostasis, respectively [23,24]. GmARF8a and GmARF8b of soybean influence nodulation and lateral root development [25], GhARF2 and GhARF18 of cotton regulate fiber cell initiation [26], and ARF7 of poplar regulates cambial activity [27]. Overexpressing of Magnolia sieboldii ARF5 in Arabidopsis has effects on vegetative and reproductive growth [28]. Silencing of LsARF8a in lettuce by the virus-induced gene silencing (VIGS) method indicated that LsARF8a conduced to the thermally induced bolting [29]. In addition, the ARF gene family of various plant species has been associated with responses to phytohormones and abiotic stresses [30,31,32]. For example, salicylic acid and heavy metals significantly upregulated BdARF4 and BdARF8 in Brachypodium distachyon [30]. The tomato knockout of SlARF4 exhibited improved water deficit resistance [33]. In Populus trichocarpa, 15 ARF genes have been associated with abiotic stress and ABA response [34].
ARF genes have been extensively identified and characterized in many plants, including Oryza sativa [35], Gossypium hirsutum [26], Setaria italica [36], Glycine max [37], and Magnolia sieboldii [28]; however, no study has been carried out in flax. As one of the first domesticated crops, flax (Linum usitatissimum L.) is extensively cultivated in various countries for its oil and fiber [38]. It is a diploid and self-pollinated crop (2n = 30) with a small genome (306 Mb) and a short life cycle [39,40]. With the increasing demand for seed and fiber, the production and quality of flax need to be improved. Recently, the flax genome has been released, providing a basis for developing high-yielding and high-quality varieties through molecular breeding [40]. Considering the vital role of ARF gene members in regulating growth and development in various species, it is worth investigating the gene family in flax (LuARF). Therefore, the present study conducted a genome-wide analysis to identify the flax ARF gene family members. We analyzed their chromosomal distribution, gene structure, conserved domain, evolutionary relationship, and cis-acting elements. We further assessed their expression patterns in various tissues and under phytohormone and abiotic stress treatments to determine the potential functions. Our results will provide comprehensive information on flax ARF genes and lay a foundation for further functional studies on the LuARF gene family.

2. Results

2.1. Identification of ARF Genes in Flax

The known ARF protein sequences from nine species were used as queries for the BLASTP search to identify the flax sequences (Table S1) [41]. Further, based on the presence and integrity of the conserved Auxin_resp and B3 domains in the Pfam and CDD databases, we identified 33 (named LuARF1LuARF33), 35, and 34 ARF members in Longya-10, Heiya-14, and pale flax, respectively (Table 1 and Table S2). Further analysis showed that the length of the ARF proteins varied from 470 (LuARF28) to 1123 (LuARF21) amino acids (aa), 401 to 1127 aa, 478 to 1058 aa in Longya-10, Heiya-14, and pale flax, respectively, while the predicted molecular weight (MW) ranged from 52.81 to 125.34 kDa, 44.04 to 125.79 kDa, and 52.85 to 117.28 kDa, and the isoelectric point (pI) varied from 5.45 to 8.59, 5.45 to 9.11, and 5.57 to 8.21, respectively. The subcellular localization analysis predicted that the ARF proteins in these three genomes were localized in the nucleus. All information, including gene name, locus ID, chromosome location, protein length, MW, pI, number of exons, and subcellular localization, of LuARF genes and the ARFs of Heiya-14 and pale flax are listed in Table 1 and Table S3, respectively, and the ARF protein sequences are shown in Table S4.

2.2. Sequence Analysis of Flax ARFs

Further, to assess the genetic relationships among the LuARFs, a phylogenetic tree was constructed using the LuARF protein sequences (Figure 1a). The analysis showed that the LuARFs could be divided into four groups (groups A to D) and most LuARF members existed in pairs. Further analysis of the gene structure by comparing the cDNA sequences with the genomic DNA sequences using Gene Structure Display Server (GSDS) revealed that the length of LuARF genes varied significantly, with LuARF9 being the longest (26.0 kb) and LuARF13 being the shortest (2.4 kb) (Figure 1b and Table 1) [42]. The ARF genes in Heiya-14 and pale flax were 1.4 to 14.6 kb and 2.2 to 14.6 kb long, respectively (Table S3). These observations indicated significant differences among the members of the same cultivar. The number of exons was also remarkably different (2 to 14 or 15) among the ARF genes in the three genomes. Twenty-two LuARF genes (66.7%) had ≥10 exons, while seven had 2–5 exons, and the other four had 8–9 exons. The number of exons in the ARF genes of Heiya-14 and pale flax was similar to that of Longya-10.
Further analysis based on Pfam and CDD databases indicated that 19 out of 33 LuARF proteins had the typical ARF structure composed of a conserved DBD domain (B3 and AUX_RESP), a variable MR domain, and a CTD (Aux/IAA), while the remaining 14 lacked the CTD (Table S2). Analysis of the conserved motifs in LuARF proteins using the Multiple Expectation maximization for Motif Elicitation (MEME) program identified 92 motifs, named motif 1 to motif 92 (Figure 1c and Figure S1). Motifs 1, 2, and 6, which form the B3 domain, existed in all LuARFs. Motifs 5, 7, 8, 12, and 14 corresponded to the ARF domain; of them, motif 8 existed in all LuARF proteins, while the rest only existed in some LuARF proteins. Meanwhile, motifs 4, 9, and 16 constituted the CTD. Motifs 3, 6, and 11 existed in more than 90.9% of the gene members, indicating their significantly conserved nature among the LuARF proteins. Seventy-two motifs were found in less than half of LuARF proteins, and twenty-seven motifs existed only in 6% LuARFs. These differences in the distribution of motifs indicated functional diversity among the LuARF members. Further, based on the MR amino acid composition and the presence or absence of the CTD, LuARF proteins were divided into three groups: proteins with a DBD, an activator MR enriched with Q, S, and L, and a CTD (LuARF6, 9, 10, 11, 15, 17, 20, 21, 22, 23, 26, 32, and 33); proteins with a DBD, a repressor MR enriched with S, L, P, and G, and a CTD (LuARF1, 7, 14, 16, 24, and 31); and proteins with a DBD and a repressor MR, but lacking a CTD (LuARF2, 3, 4, 5, 8, 12, 13, 18, 19, 25, 27, 28, 29, and 30) (Table S5 and Figure S2).

2.3. Phylogenetic Analysis of ARF Proteins

Then, to assess the evolutionary significance of ARFs in flax domestication, a phylogenetic tree was constructed using the ARF sequences from Longya-10, Heiya-14, and pale flax (Figure S3). The analysis showed that most LuARFs had orthologs in both Heiya-14 and pale flax. Another phylogenetic tree was generated using 291 ARF protein sequences, including 7 from dicotyledons (L. usitatissimum, G. hirsutum, A. thaliana, G. max, Vitis vinifera, Eucalyptus grandis, and Medicago truncatula) and 3 from monocotyledons (O. sativa, Sorghum bicolor, and B. distachyon) to assess the evolution among the ARFs of diverse species (Figure 2). In this tree, the ARFs were clustered into four clades: clade I, clade II, clade III (clades IIIa, IIIb, and IIIc), and clade IV (clades IVa, IVb, and IVc) (Table S5). Except for clades IIIb, IVa, and IVb, all other clades contained ARFs from dicotyledons and monocotyledons. Most LuARF proteins clustered closely with the ARFs from G. hirsutum and A. thaliana. All the 13 LuARF proteins containing the DBD-MR activator-CTD structure, predicted to be transcriptional activators, clustered in clade III, while all the potential transcriptional repressors with the DBD-MR repressor-CTD structure belonged to clade I. In addition, the 14 LuARF proteins without a CTD were grouped as clades I, II, and IV.

2.4. Chromosomal Distribution and Synteny Analysis

We further investigated the chromosomal locations of LuARF genes (Figure 3). The analysis indicated that 30 LuARF genes were distributed unevenly on all chromosomes except for chromosomes 1 and 7. However, we could not map LuARF31, 32, and 33 onto the chromosomes. Chromosome 14, with six members, had the largest number of LuARFs. Chromosome 15 contained four LuARF genes, chromosomes 2 and 13 contained three genes each, chromosomes 4, 5, 8, 9, and 10 possessed two genes each, and chromosomes 3, 6, 11, and 12 each contained only one gene.
We then analyzed the duplication events to elucidate the mechanism underlying LuARF gene family expansion during evolution. The results showed that two gene pairs (LuARF10/LuARF11 and LuARF22/LuARF23) underwent tandem duplication (TD) and another eighteen gene pairs (LuARF1/14, LuARF2/12, LuARF3/27, LuARF4/5, LuARF4/13, LuARF5/25, LuARF5/8, LuARF8/25, LuARF6/26, LuARF7/16, LuARF9/17, LuARF10/22, LuARF15/20, LuARF15/32, LuARF20/32, LuARF32/33, LuARF19/LuARF29 and LuARF24/31) underwent segmental/whole-genome duplication (SD/WGD) (Figure 3 and Table 2). The genes that experienced SD had one to three syntenic gene pairs in the flax genome (Figure S4). Further, we calculated the nonsynonymous (Ka) and synonymous (Ks) values and used the Ka/Ks ratios to reveal the selective forces on the duplicated LuARF gene pairs. In general, Ka/Ks > 1 indicates a positive selection, Ka/Ks = 1 indicates a neutral evolution, and Ka/Ks < 1 shows a strong purifying selection [43]. We found that most Ks values of these collinear genes varied from 0.05 to 0.23, implying that the duplication events happened close to the recent WGD of flax (Ks = 0.13) [40]. The divergence time of these gene pairs was estimated as 4–18 Mya. However, the Ks values of LuARF15/32 and LuARF20/32 gene pairs were about 0.7, indicating that the duplication happened close to the early WGD event (Ks = 0.77), and the divergence time was estimated as 57–60 Mya. In addition, all duplicated gene pairs’ Ka/Ks ratios were less than 0.4, which suggested a strong purifying selection pressure on the LuARF genes.
Further, to explore the evolution of ARF genes, the collinearity relationships between flax and nine species (six dicots and three monocots) described above were analyzed (Figure 4 and Table S6). The analysis identified 10, 29, 47, 30, 15, 27, 5, 4, and 6 homologous pairs of ARFs between flax and Arabidopsis, soybean, cotton, barrel medic, Egrandis, grape, rice, sorghum, and B. distachyon, respectively. About 50% of LuARF genes had homologous genes in these nine plants, with collinear gene pairs ranging from one to five. In addition, the collinear genes between flax and monocots were far less than that between flax and dicots. A few LuARFs had more than three syntenic gene pairs between flax and another two plants (cotton, barrel medic). For example, LuARF13 and 24 had up to five syntenic gene pairs in cotton, and LuARF4, 13, and 25 had five pairs in barrel medic, indicating that these genes remained highly conserved during evolution. Also, LuARF4 and 13 had homologous genes in all the selected species, suggesting that these gene pairs might share a common ancestor and existed before the monocotyledon-dicotyledon split. LuARF25 had orthologs in all species except for Arabidopsis, while LuARF24 had orthologs in seven plants. Except for LuARF4, 13, 24, and 25, all other genes lacked orthologs in monocots but had homologs in dicots, implying that most genes might have appeared after the divergence of dicotyledons and monocotyledons.

2.5. Phytohormone- and Abiotic Stress-Related Cis-Acting Elements on LuARF Promoter

The cis-acting elements in the promoter regions are essential in regulating the transcription of genes. To determine the potential function of LuARF genes, the promoter sequences of all LuARFs were analyzed using the PlantCARE database [44]. The analysis revealed numerous light, phytohormone, and abiotic stress response elements on LuARF promoters and a few basic elements (TATA-box and CAAT-box) (Figure 5 and Table S7). Further analysis of the hormone- and stress-responsive elements identified five kinds of phytohormone-related elements, including those related to gibberellin (GARE-motif, TATC-box, and P-box), methyl jasmonate (MeJA, CGTCA-motif), salicylic acid (TCA-element and SARE), auxin (TGA-element, AuxRR-core, and TGA-box), and abscisic acid (ABRE). Among them, the gibberellin, MeJA, and abscisic acid-responsive elements were abundant in LuARF promoters (44, 63, and 68, respectively). The number of hormone response elements in each LuARF ranged from 4 to 12, implying an uneven distribution pattern. LuARF1, 2, 6, 8, 10, 13, and 32 had all five phytohormone-related elements, while LuARF20 and 27 contained only two types. LuARF13 and 22, LuARF1, and LuARF11 contained up to 4, 6, and 7 elements associated with gibberellin, MeJA, and abscisic acid, respectively. Additionally, three regulatory elements related to environmental stress, the MBS (drought inducibility), LTR (low temperature responsive), and TC-rich repeats (defense- and stress-responsive), were identified in the LuARF promoter; 17, 21, and 14 LuARF genes contained MBS, LTR, and TC-rich repeats, respectively. Among the 21 genes, LuARF11 had the most LTR elements.

2.6. Tissue-Specific Expression of LuARF Genes

To assess the potential functions of LuARF genes, we first analyzed their expression patterns in different tissues (stem and capsule) and cultivars (Longya-10 and Heiya-14) using the previously published transcriptome data (Figure 6a) [40]. The analysis detected the transcripts of 26 LuARF genes in both stems and capsules of Longya-10 and Heiya-14, but not that of LuARF9, 11, 23, and 28. We found that six LuARF genes (LuARF2, 4, 6, 18, 25, and 26) were highly expressed in capsules, while nine (LuARF3, 8, 13, 15, 21, 22, 27, 32, and 33) had higher expression levels in stems. Further, we investigated these genes’ expression patterns in flax stem and seeds by qRT-PCR. As shown in Figure 6b, only LuARF4/13 exhibited differences in the expression patterns between the two cultivars, Longya-14 and Zhangya-2, with different oil content. Subsequent analysis of the correlation between oil content and the expression levels of LuARF4/13 in seeds at different developmental stages indicated that the expression level in 10 d seeds was positively related to the oil content in Longya-14, while the expression level in 25 d and 40 d seeds was positively associated with the oil content in Zhangya-2 (Table S8).
To further confirm the importance of the nine genes explicitly expressed in the stem, their expression levels were tested in both whole stems and stem peels at nine positions along the Longya-10 stem (Figure 6c). The nine LuARF genes exhibited significantly similar expression patterns. Above the snap point (SA to C), these genes exhibited higher expression in the whole stem than in the stem peel, while the opposite was found at point A. Below the snap point (D to H), LuARF8 and LuARF32/33 showed lower expression at position G than C in the stem peel, while the other genes showed obviously higher expression levels at all positions than that at C. The expression patterns of these genes in the whole stem were opposite to that in the stem peel, and the levels diminished from C to D. In general, the nine LuARFs exhibited significantly higher expression in the stem peel than in the whole stem at points D to H, with the expression patterns of LuARF8 and 32/33 at point G and LuARF21 at point E as the exceptions. LuARF8 expression at point G was lower in the stem peel than in the whole stem, and LuARF32/33 expression at point E was lower in the stem peel than in the whole stem.

2.7. Expression Profiling of LuARFs under Various Stresses

Finally, to assess the potential functions of LuARF genes in response to phytohormones and abiotic stresses, we investigated their expression levels in flax seedlings under IAA, NAA, and GA3 treatments and salinity, drought, cold, and heat stresses. The analysis showed that all LuARFs were significantly induced 4 h after exposure to cold and heat stress, and the response of all genes, except LuARF6/26, to cold stress was more pronounced than to heat stress (Figure 7a). Under salinity stress (NaCl), all genes were significantly induced at least at one time point after exposure. Most genes exhibited significantly higher expression than the control at 48–72 h after treatment, while LuARF8 showed higher expression at 9 and 24 h (Figure 7b). After PEG treatment, all LuARF genes exhibited an expression peak at 9 h, except for LuARF7/16, LuARF18, LuARF19/29, and LuARF28 (Figure 7c). Notably, the expression of LuARF18 and LuARF28 did not exceed that of the control at any point after PEG treatment. In addition, LuARF genes exhibited differential expression patterns under GA3 and IAA treatments (Figure 7d,e). Most genes were significantly upregulated at 3 or 9 h after GA3 and IAA treatments. LuARF7/16 expression after GA3 treatment was higher than the control at all time points except for 12 h, while LuARF10, LuARF11/23, and LuARF32/33 expression levels after IAA exposure were significantly higher than the control at up to six time points. Most LuARFs were highly induced at 12 or 24 h after NAA treatment; LuARF6/26, LuARF7/16, LuARF10, LuARF21, and LuARF24/31 were significantly induced at 12, 48, or 72 h, and another three time points (3, 6, 9, or 24 h) after exposure to NAA (Figure 7f). At 48 h recovery, different LuARFs exhibited different expression patterns under phytohormone and stress treatments; a few LuARFs showed higher expression than the control, while some members showed lower expression than the control.

3. Discussion

ARFs are pivotal in controlling various biological processes associated with plant growth and development [2]. Although researchers have identified and reported the functions of ARFs across multiple plant species, this present work reports the genes in the oilseed and fiber crop, flax. This study comprehensively analyzed the flax ARF gene family to elucidate their potential roles in regulating flax growth and response to environmental stimuli.

3.1. ARF Genes in Flax

This current research identified 33, 35, and 34 ARF members in the genomes of three representative flax cultivars, the conventional oil cultivar Longya-10, the conventional fiber cultivar Heiya-14, and their ancestor pale flax. The three genomes had almost the same number of ARF genes, indicating that the ARF gene family remained relatively conserved during flax domestication. However, the number of ARF genes varied considerably among plant species, such as 33 in flax, 25 in rice [35], 19 in grape [45], 56 in cotton [26], and 40 in M. truncatula [46], probably due to the deletion and expansion of genes during the evolutionary process. Further analysis showed extensive variations in gene structure, protein length, MW, and pI of ARFs in the three genomes, implying functional diversity. In addition, all ARF proteins identified in this study were predicted to be located in the nucleus. These observations of LuARFs are consistent with the reports on ARFs of other species [30,32,36,37].
Domain characteristics are crucial for predicting the functions of transcription factors [7]. The present study found that all LuARF proteins had a typical B3 domain (motifs 1, 2, and 6) and an ARF domain (motifs 5, 7, 8, 12, and 14); however, only 19 of the 33 LuARFs had the CTD (motifs 4, 9, and 16), which is not necessary for binding promoter regions, but is necessary for interaction with Aux/IAA proteins to mediate transcriptional levels of downstream genes [7]. The percentage of LuARF without CTD (42.4%) was higher than that in A. thaliana (17.39%) [47], O. sativa (24%) [35], and M. truncatula (37.5%) [46]. Generally, the ARFs without CTD are regulated via interaction with other transcription factors but not Aux/IAA and might be insensitive to auxin and function in an auxin-independent way, such as the Arabidopsis AtARF3 [7,9,48]. Therefore, it can be speculated that the 14 LuARFs without the CTD might regulate biological processes in an auxin-independent manner. In addition, we found that 13 out of 19 LuARF members with complete ARF structure contained an MR domain enriched with Q, S, and L, implying their role as transcriptional activators [6,49,50]. The other six LuARFs with CTD and all the LuARFs without CTD might function as transcriptional repressors as their MR domains were enriched with S, L, P, and G, similar to those in other plants [36,46]. These observations showed that the ratio of transcriptional activators and repressors among the LuARF proteins were 0.65, demonstrating that LuARF members evolved distinct functions, as in other plants [34].
Furthermore, the three phylogenetic trees generated in this study demonstrated similar evolutionary relationships for the LuARFs, indicating the accuracy of the phylogeny analysis. All clades (I, II, IIIa, IIIc, and IVc) contained ARFs from both dicots and monocots, suggesting relatively conserved ARF functions among the species. However, clade IIIb and IVb were dicot-specific subgroups and clade IVa only had one monocot ARF member, suggesting specific roles for these ARFs. Our study also found that the LuARFs within the same clade had similar exon–intron structures and domain composition, similar to those of other species [32,36,46], indicating functional redundancy. We then found 4 LuARF genes (2 gene pairs) experienced TD events, while 27 (81.8%, 18 gene pairs) experienced SD events. Studies have demonstrated the significance of SD/WGD and TD in gene expansion and functional differentiation of multiple gene families [51,52]. Generally, SD results in functional redundancy, while TD results in novel functions, which help the species adapt to the fast-changing environment [53]. Subsequent Ks analysis implied that the most recent WGD event was the leading cause of the expansion of the LuARF gene family [40]. In S. italica, the WGD event caused the expansion of the ARF gene family [36]. Thus, our observations suggest that the continual duplication of LuARF genes might have contributed to function specificity or redundancy. Additionally, the Ka/Ks ratios of all the duplicated gene pairs were less than 0.4, indicating a purifying selection force on LuARF genes (Table 2). Research has proven that ARF members demonstrate relatively conserved homologous relationships among species [54]. In the present study, comparing the LuARFs with the genes of six dicots and three monocots revealed that about half of LuARFs had homologous genes in the analyzed species. In addition, monocots had fewer collinear gene pairs than the dicots, implying that ARF genes have experienced a wide range of evolutionary and replication events after the split of monocot and dicot plants, as in Prunus avium [32], probably leading to the unique features of monocot and dicot plants.

3.2. LuARFs’ Possible Role in Regulating the Development of Organs and Responses to Exogenous Stimuli

Spatial expression profiling of genes helps elucidate their functions. In this study, we generated a heatmap based on the transcriptome data of capsules and stems and analyzed the expression of a few selected LuARF genes (specifically expressed) in seeds and stems at different development stages. LuARF4 showed higher expression in capsules and exhibited different expression patterns in cultivars with different oil content (Figure 6b). Moreover, its expression level was positively related to oil content (Table S8). Thus, LuARF4 was speculated to regulate genes related to fatty acid synthesis. In flax, the specification and elongation of fiber occur above the snap point, while the thickening of the fiber cell wall occurs below [55]. Studies have indicated that genes with increased expression levels below the snap point in stem peels but decreased levels in the whole stem and whose expression levels in stem peels at least were as high as those in whole stems at equivalent positions might be involved in the strengthening of fiber cell wall when the fiber cells stop elongating [56,57]. In Arabidopsis, CESA4 and CESA8 with high expression in phloem peel have been associated with cellulose microfibril synthesis in the secondary cell wall [58]. In flax, LuPME1 showed increased expression below the snap point in stem peels, with an expression 50 times higher than in the whole stem at the same position [56]. These observations suggested its role in the thickening of the cell wall when the cells stop elongating. In this study, below the snap point, nine LuARF genes showed higher expression in stem peel than in the whole stem at the same position. Thus, based on the earlier reports and the current findings, LuARF3, 8, 13, 15, 21, 22, 27, 32, and 33 could be associated with the thickening of the flax fiber cell wall.
The ARF proteins are known to play key roles in the auxin signaling pathway, and their activities are regulated by auxin concentrations [59]. Additionally, studies have identified that ARF genes, with cis-acting elements related to hormone or abiotic stresses, regulate plant responses to auxin or abiotic stress [30,34,46]. The present study identified 6 to 16 cis-acting elements associated with the response to hormones and environmental stimuli in the LuARF genes, indicating their vital role in hormone and stress responses. Therefore, we further analyzed the effect of environmental stresses and exogenous hormones on LuARF expression to elucidate their roles in stress and phytohormone responses. Our experiments demonstrated that all LuARF genes were significantly upregulated after exposure to IAA, NAA, and GA3, probably because exogenous hormones affect in vivo auxin homeostasis and, in turn, the expression of these genes [59]. These observations are consistent with the findings on the response of ARF genes in other species to IAA, NAA, and GA3 treatments [30,36,60]. We also found that, under the same treatment, different LuARF members showed different expression patterns, indicating differences in their functions in the auxin signal transduction pathway. The expression patterns of LuARF genes were also different under the three hormone treatments, which implied distinct regulatory modes of LuARF genes when subjected to different treatments. In addition, all LuARF genes were significantly upregulated under cold and heat stress, consistent with the observations in B. distachyon [60]. In addition, all LuARF members were responsive to NaCl and PEG treatment, consistent with the ARF genes of tomato under salt and drought stress [61]. For example, IbARF5 of sweet potato significantly improved salt and drought tolerance of Arabidopsis, and this response was associated with ABA biosynthesis [62]. Previous studies have also indicated that the ARF genes might resist heat stress [63] and respond to salt and drought stresses [36] through ABA-mediated signaling pathways. Therefore, it could be speculated that LuARF genes regulate osmotic stress response ABA-dependently.

4. Materials and Methods

4.1. Plant Growth and Abiotic Stress Treatments

The flax genotypes Longya-10, Longya-14, Longya-15, and Zhangya-2 were chosen for this study. Longya-14, Zhangya-2, and Longya-10 were cultivated in the experimental field of Gansu Academy of Agricultural Sciences in Lanzhou, Gansu Province (103°40′ E, 34° N). The seeds collected from Longya-14 and Zhangya-2 plants at 5, 10, 15, 20, 25, 30, 40, and 50 days post anthesis and the stem tissues collected from Longya-10 plants (52~54 cm) at about 5 weeks post-emergence were used to determine the expression patterns of LuARF genes. Here, according to the sampling method described by Pinzon-Latorre and Deyholos [56], the whole stem (1-cm-long sections) or the stem peel from Longya-10 plants was collected from nine positions based on the fiber development stage as follows: 0 to 1 cm (shoot apex, SA), 1 to 2 cm (A), 2 to 3 cm (B), 3 to 4 cm (C), 10 to 11 cm (D), 20 to 21 cm (E), 30 to 31 cm (F), 40 to 41 cm (G), and 50 to 51 cm (H).
Then, to investigate the expression patterns of LuARF genes under various phytohormone and abiotic stress treatments, seeds of Longya-15 were allowed to germinate in Petri dishes at 25 °C under 16 h:8 h light/dark conditions in a climate chamber. After 10 days of growth, uniformly grown seedlings were selected, transferred to 1/2 MS liquid medium, and cultivated for 3 days under the same conditions. Then, the medium was replaced with fresh medium containing 150 mM NaCl and 20% PEG to induce salinity and drought treatment, respectively, and 15 µM indole-3-acetic acid (IAA) or 5 µM naphthalene acetic acid (NAA) or 5 µM gibberellic acid (GA3) for the phytohormone treatment. The seedlings were subjected to 42 °C and 4 °C to induce heat and cold stress, respectively. The plants used for treatments were all from a single sowing. For each treatment, three biological replicates were set up. The leaves were collected from seedlings under heat and cold stress after 3 h of exposure and from drought, salinity, and phytohormone treatment at 3, 6, 9, 12, 24, 48, and 72 h under stress and 48 h after subsequent growth in 1/2 MS liquid medium without supplement. The leaves from the control seedlings were collected at 0 h and all other time points. For each treatment, samples were collected from three plants per replicate at each time point. All samples were frozen in liquid nitrogen immediately after collection and stored at −80 °C for RNA extraction.

4.2. Identification of ARF Genes

The genome sequences of L. usitatissimum (Longya-10 and Heiya-14) and Linum bienne (pale flax) were obtained from NCBI (https://www.ncbi.nlm.nih.gov/, accessed on 13 December 2022), and their annotation files were downloaded from Figshare (https://figshare.com/, accessed on 13 December 2022). The ARF protein sequences of G. hirsutum, A. thaliana, G. max, V. vinifera, E. grandis, M.truncatula, O. sativa, S. bicolor, and B. distachyon were obtained from Phytozome v13 (https://phytozome-next.jgi.doe.gov/, accessed on 13 December 2022) and used as queries to identify the ARF genes in Longya-10, Heiya-14, and pale flax using BLASTP analysis (e-value of 1 × 10−5) [41]. All candidate ARF proteins were further examined for the presence and integrity of the conserved domains [Pfam 06507: ARF (AUX_RESP); Pfam 02362: B3 DNA binding domain (B3)] using the Pfam (http://pfam-legacy.xfam.org/search#tabview=tab1, accessed on 15 December 2022) and the NCBI Conserved Domain Databases (CDD, https://www.ncbi.nlm.nih.gov/cdd/, accessed on 15 December 2022). Finally, the non-redundant LuARF genes were obtained using one gene model per locus. The MW and pI of all LuARF proteins were computed using the ExPASy ProtoParam (http://web.wxoasy.org/protparam/, accessed on 18 December 2022) tool [64] and the subcellular localization was predicted using Plant-mPloc (http://www.csbio.sjtu.edu.cn/ bioinf/plant-multi/, accessed on 18 December 2022) [65]. MEME program was used to identify and analyze the conserved motifs of the LuARF proteins using default parameters (http://alternate.meme-suite.org/tools/meme, accessed on 24 December 2022) [66].

4.3. Chromosomal Mapping and Collinearity Analysis

The chromosomal location of LuARF genes was obtained from the annotation files downloaded from Figshare, and the chromosome distribution map was drawn using Map Gene 2 Chromosome (MG2C, http://mg2c.iask.in/mg2c_v2.0/, accessed on 4 January 2023). The MCScanX package was adopted to analyze the LuARF gene duplication events using default parameters [67], and the TBtools software (version 1.9876.0.0) was used to assess and visualize the collinearity among the ARFs [68]. The Ka and Ks substitution rates per site of the duplicated gene pairs were calculated using DnaSP software (version 6.0) [69], and the Ka/Ks ratio was used to assess the mode and strength of the selective pressure [43]. For each gene pair, the divergence time (million years ago, Mya) was also calculated using the following formula [70]:
Mya = Ks/(2 × 6.1 × 10−9) × 10−6.

4.4. Gene Structure and Promoter Analysis

The exon–intron organization of the LuARF gene family members was determined based on the genome and coding sequences and visualized using the GSDS (http://gsds.cbi.pku.edu.cn/, accessed on 4 January 2023) [42]. The 2000 bp region upstream of each LuARF start codon was extracted from the Longya-10 genome, and the cis-acting elements were determined using the PlantCARE database (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/, accessed on 4 January 2023) [44].

4.5. Phylogenetic Analysis

The full-length amino acid sequences of the ARFs of L. usitatissimum, V. vinifera, G. max, G. hirsutum, A. thaliana, E. grandis, M. truncatula, S. bicolor, B. distachyon, and O. sativa were used to construct a phylogenetic tree. First, a multiple sequence alignment was performed with ClustalX in MEGA (6.0) using default parameters. Then, a phylogenetic tree was built using the alignment files based on the neighbor-joining (NJ) method with 1000 bootstrap replicates [71].

4.6. Gene Expression Analysis

Previously published transcriptome data on the capsule and stem of Longya-10 and Heiya-14 (PRJNA505721) were initially used to determine the expression patterns of LuARFs in different tissues and genetic backgrounds [40], and the TBtools software (version 1.9876.0.0) was used to draw a heat map based on these expression patterns. Then, the LuARF genes with high expression levels in the capsule or stem were further analyzed using qRT-PCR. The expression profiles of LuARFs under various treatments were also determined using qRT-PCR. Total RNA was extracted by using EZgeneTM Plant Easy Spin RNA Miniprep Kit following the manufacturer’ s instructions (BIOMIGA, San Diego, CA, USA), and cDNA synthesis was performed by using PrimeScriptTM RT reagent Kit with gDNA Eraser (Takara, San Jose, CA, USA). The specific primer pairs used for qRT-PCR were designed by Primer Premier 5.0 (PREMIER Biosoft International, Palo Alto, CA, USA) and listed in Table S9. qRT-PCR was performed by using TB GreenTM Premix Ex TaqTM II (Takara, San Jose, CA, USA) and conducted on Eco Real-Time PCR System (Illumine). The reaction condition was as follows: 2 min at 50 °C, 10 min at 95 °C, with 40 cycles of 15 s at 95 °C, 15 s at 60 °C, and 15 s at 72 °C, followed by 10 s at 95 °C, 1 min at 65 °C, and 1 s at 97 °C. Each sample had three replicates. GAPDH was used as an internal reference gene [72], and the relative expression levels of the LuARF genes were calculated following the 2−ΔΔCt method [73]. The coding sequences of the LuARF1/14, LuARF2/12, LuARF3/27, LuARF4/13, LuARF5/25, LuARF6/26, LuARF7/16, LuARF9/17, LuARF11/23, LuARF15/20, LuARF19/29/30, LuARF24/31, and LuARF32/33 gene pairs were extremely similar, and therefore, it was impossible to design primers specific for each gene of the pair. Thus, only one primer pair was designed to amplify the genes of each pair.

5. Conclusions

In conclusion, the present study identified 33 LuARF genes in Longya-10 (oil-use flax variety), and their gene structure, protein features, phylogeny, promoter elements, and expression patterns were analyzed. The results showed that LuARFs can function as either transcriptional activators or repressors, and the evolution and expansion of LuARFs were mainly driven by WGD events. Moreover, a positive association between LuARF gene expression and oil content was observed. All these results revealed the important roles of LuARFs in evolution, organ development, and responses to stress and phytohormone.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241311006/s1.

Author Contributions

Conceptualization, J.Z. and L.W.; methodology, Y.Q.; software, W.L. (Wenjuan Li) and Z.D.; validation, L.Z., W.L. (Wen Li), C.Y. and C.X.; formal analysis, Y.Q., Z.D. and W.Z.; data curation, Y.Q., L.W. and Y.X.; writing—original draft preparation, Y.Q.; writing—review and editing, J.Z.; funding acquisition, J.Z., Y.Q. and W.L. (Wenjuan Li). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Industrial Technology System of Characteristics Oil of China, MOF and MARA (CARS-14-1-05), the Modern Biology Breeding Project of Gansu Academy of Agricultural Sciences (2022GAAS04), the Science and technology project of Gansu Province (21JR7RA722), Major Project of Gansu Province (21ZD4NA022-02), and the National Natural Science Foundation of China (31760426).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are reported in the article and the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hagen, G.; Guilfoyle, T. Auxin-responsive gene expression: Genes, promotersand regulatory factors. Plant Mol. Biol. 2002, 49, 373–385. [Google Scholar] [CrossRef]
  2. Li, S.B.; Xie, Z.Z.; Hu, C.G.; Zhang, J.Z. A review of auxin response factors (ARFs) in plants. Front. Plant Sci. 2016, 7, 47. [Google Scholar] [CrossRef] [Green Version]
  3. Roosjen, M.; Paque, S.; Weijers, D. Auxin Response Factors: Output control in auxin biology. J. Exp. Bot. 2018, 69, 179–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Casanova-Sáez, R.; Voß, U. Auxin Metabolism Controls Developmental Decisions in Land Plants. Trends Plant Sci. 2019, 24, 741–754. [Google Scholar] [CrossRef] [PubMed]
  5. Hagen, G. Auxin Signal Transduction. Essays Biochem. 2015, 58, 1–12. [Google Scholar]
  6. Ulmasov, T.; Hagen, G.; Guilfoyle, T.J. Activation and repression of transcription by auxin-response factors. Proc. Natl. Acad. Sci. USA 1999, 96, 5844–5849. [Google Scholar] [CrossRef] [PubMed]
  7. Guilfoyle, T.J.; Hagen, G. Auxin response factor. Curr. Opin. Plant Biol. 2007, 10, 453–460. [Google Scholar] [CrossRef]
  8. 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]
  9. Tiwari, S.B.; Hagen, G.; Guilfoyle, T. The roles of auxin response factor domains in auxin-responsive transcription. Plant Cell 2003, 15, 533–543. [Google Scholar] [CrossRef] [Green Version]
  10. Boer, D.R.; Freire-Rios, A.; van den Berg, W.A.; Saaki, T.; Manfield, I.W.; Kepinski, S.; López-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]
  11. Guilfoyle, T.J. The PB1 domain in auxin response factor and aux/IAA proteins: A versatile protein interaction module in the auxin response. Plant Cell 2015, 27, 33–43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Ulmasov, T.; Hagen, G.; Guilfoyle, T.J. ARF1, a transcription factor that binds to auxin response elements. Science 1997, 276, 1865–1868. [Google Scholar] [CrossRef]
  13. Weijers, D.; Wagner, D. Transcriptional Responses to the Auxin Hormone. Annu. Rev. Plant Biol. 2016, 67, 539–574. [Google Scholar] [CrossRef] [PubMed]
  14. Leyser, O. Auxin signaling. Plant Physiol. 2018, 176, 465–479. [Google Scholar] [CrossRef] [Green Version]
  15. Kepinski, S.; Leyser, O. The Arabidopsis F-box protein TIR1 is an auxin receptor. Nature 2005, 435, 446–451. [Google Scholar] [CrossRef] [PubMed]
  16. Tan, X.; Calderon-Villalobos, L.I.A.; Sharon, M.; Zheng, C.; Robinson, C.V.; Estelle, M.; Zheng, N. Mechanism of auxin perception by the TIR1 ubiquitin ligase. Nature 2007, 446, 640–645. [Google Scholar] [CrossRef] [PubMed]
  17. Ellis, C.M.; Nagpal, P.; Young, J.C.; Hagen, G.; Guifoyle, 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]
  18. Zhang, K.; Zhang, H.; Pan, Y.; Niu, Y.; Guo, L.; Ma, Y.; Tian, S.; Wei, J.; Wang, C.; Yang, X.; et al. Cell- and noncell-autonomous AUXIN RESPONSE FACTOR3 controls meristem proliferation and phyllotactic patterns. Plant Physiol. 2022, 190, 2335–2349. [Google Scholar] [CrossRef]
  19. 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]
  20. Liu, Z.N.; Miao, L.M.; Huo, R.X.; Song, X.Y.; Johnson, C.; Kong, L.J.; Sundaresan, V.; Yu, X. ARF2ARF4 and ARF5 are essential for female and male gametophyte development in Arabidopsis. Plant Cell Physiol. 2018, 59, 179–189. [Google Scholar] [CrossRef] [Green Version]
  21. 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, 127, 3877–3888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Okushima, Y.; Fukaki, H.; Onoda, M.; Theologis, A.; Tasaka, M. ARF7 and ARF19 regulate lateral root formation via direct activation of LBD/ASL genes in Arabidopsis. Plant Cell 2007, 19, 118–130. [Google Scholar] [CrossRef] [Green Version]
  23. 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, anauxin response factor, in transgenic rice. Curr. Issues Mol. Biol. 2009, 11, i29–i34. [Google Scholar]
  24. Wang, S.; Zhang, S.; Sun, C.; Xu, Y.; Chen, Y.; Yu, C.; Qian, Q.; Jiang, D.; Qi, Y. (OsARF12), a novel regulator for phosphate homeostasis inrice (Oryza sativa). New Phytol. 2014, 201, 91–103. [Google Scholar] [CrossRef] [PubMed]
  25. 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 (Glycine max L.) nodulation and lateral root development. Plant Physiol. 2015, 168, 984–999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Xiao, G.; He, P.; Zhao, P.; Liu, H.; Zhang, L.; Pang, C.; Yu, J. Genome-wide identification of the GhARF gene family reveals that GhARF2 and GhARF18 are involved in cotton fibre cell initiation. J. Exp. Bot. 2018, 69, 4323–4337. [Google Scholar] [CrossRef]
  27. Hu, J.; Su, H.; Cao, H.; Wei, H.; Fu, X.; Jiang, X.; Song, Q.; He, X.; Xu, C.; Luo, K. AUXIN RESPONSE FACTOR7 integrates gibberellin and auxin signaling via interactions between DELLA and AUX/IAA proteins to regulate cambial activity in poplar. Plant Cell 2022, 34, 2688–2707. [Google Scholar] [CrossRef]
  28. Mei, M.; Ai, W.; Liu, L.; Xu, X.; Lu, X. Genome-wide identification of the auxin response factor (ARF) gene family in Magnolia sieboldii and functional analysis of MsARF5. Front. Plant Sci. 2022, 13, 958816. [Google Scholar] [CrossRef] [PubMed]
  29. Hu, M.; Qi, Z.; Ren, Z.; Tong, J.; Wang, B.; Wu, Z.; Hao, J.; Liu, N. Genome-Wide Analysis of Auxin Response Factors in Lettuce (Lactuca sativa L.) Reveals the Positive Roles of LsARF8a in Thermally Induced Bolting. Int. J. Mol. Sci. 2022, 23, 13509. [Google Scholar] [CrossRef] [PubMed]
  30. Liu, N.; Dong, L.; Deng, X.; Liu, D.; Liu, Y.; Li, M.; Hu, Y.; Yan, Y. Genome-wide identification, molecular evolution, and expression analysis of auxin response factor (ARF) gene family in Brachypodium distachyon L. BMC Plant Biol. 2018, 18, 336. [Google Scholar] [CrossRef] [Green Version]
  31. Die, J.V.; Gil, J.; Millan, T. Genome-wide identification of the auxin response factor gene family in Cicer arietinum. BMC Genom. 2018, 19, 301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Hou, Q.; Qiu, Z.; Wen, Z.; Zhang, H.; Li, Z.; Hong, Y.; Qiao, G.; Wen, X. Genome-Wide Identification of ARF Gene Family Suggests a Functional Expression Pattern during Fruitlet Abscission in Prunus avium L. Int. J. Mol. Sci. 2021, 22, 11968. [Google Scholar] [CrossRef] [PubMed]
  33. Chen, M.; Zhu, X.; Liu, X.; Wu, C.; Yu, C.; Hu, G.; Chen, L.; Chen, R.; Bouzayen, M.; Zouine, M.; et al. Knockout of Auxin Response Factor SlARF4 Improves Tomato Resistance to Water Deficit. Int. J. Mol. Sci. 2021, 22, 3347. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, Y.; Wang, R.; Yu, J.; Huang, S.; Zhang, Y.; Wei, H.; Wei, Z. Genome-Wide Identification and Characterization of Auxin Response Factor (ARF) Gene Family Involved in Wood Formation and Response to Exogenous Hormone Treatment in Populus trichocarpa. Int. J. Mol. Sci. 2023, 24, 740. [Google Scholar] [CrossRef]
  35. 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]
  36. Chen, Y.; Liu, B.; Zhao, Y.; Yu, W.; Si, W. Whole-Genome Duplication and Purifying Selection Contributes to the Functional Redundancy of Auxin Response Factor (ARF) Genes in Foxtail Millet (Setaria italica L.). Int. J. Genom. 2021, 2021, 2590665. [Google Scholar] [CrossRef]
  37. Ali, S.; Wang, W.; Zhang, Z.; Xie, L.; Boer, D.R.; Khan, N. Genome-Wide Identification, Expression and Interaction Analysis of ARF and AUX/IAA Gene Family in Soybean. Front. Biosci. 2022, 7, 251. [Google Scholar] [CrossRef] [PubMed]
  38. van Zeist, W.; Bakker-Heeres, J.A.H. Evidence for linseed cultivation before 6000 BC. J. Archaeol. Sci. 1975, 2, 215–219. [Google Scholar] [CrossRef]
  39. Zohary, D. Monophyletic vs. polyphyletic origin of the crops on which agriculture was founded in the near east. Genet. Resour. Crop Evol. 1999, 46, 133–142. [Google Scholar] [CrossRef]
  40. Zhang, J.; Qi, Y.; Wang, L.; Wang, L.; Yan, X.; Dang, Z.; Li, W.; Zhao, W.; Pei, X.; Li, X. Genomic Comparison and Population Diversity Analysis Provide Insights into the Domestication and Improvement of Flax. iScience 2020, 23, 100967. [Google Scholar] [CrossRef]
  41. Camacho, C.; Coulouris, G.; Avagyan, V.; Ma, N.; Papadopoulos, J.; Bealer, K.; Madden, T.L. BLAST+: Architecture and applications. BMC Bioinform. 2009, 10, 421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Hu, B.; Jin, J.; Guo, A.Y.; Zhang, H.; Luo, J.; Gao, G. GSDS 2.0: Anupgraded gene feature visualization server. Bioinformatics 2015, 31, 1296–1297. [Google Scholar] [CrossRef] [Green Version]
  43. Nekrutenko, A.; Makova, K.D.; Li, W.H. The KA/KS ratio test for assessing the protein-coding potential of genomic regions: An empirical and simulation study. Genome Res. 2002, 12, 198–202. [Google Scholar] [CrossRef] [Green Version]
  44. Lescot, M.; Dehais, P.; Thijs, G.; Marchal, K.; Moreau, Y.; Van de Peer, Y.; Rouzé, P.; Rombauts, S. PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res. 2002, 30, 325–327. [Google Scholar] [CrossRef]
  45. Wan, S.; Li, W.; Zhu, Y.; Liu, Z.; Huang, W.; Zhan, J. Genome-wide identification, characterization and expression analysis of the auxin response factor gene family in Vitis vinifera. Plant Cell Rep. 2014, 33, 1365–1375. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, R.; Guo, Z.; Lu, S. Genome-Wide Identification and Expression Analysis of the Aux/IAA and Auxin Response Factor Gene Family in Medicago truncatula. Int. J. Mol. Sci. 2021, 22, 10494. [Google Scholar] [CrossRef] [PubMed]
  47. 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]
  48. Kubeš, M.; Napier, R. Non-canonical auxin signalling: Fast and curious. J. Exp. Bot. 2019, 70, 2609. [Google Scholar] [CrossRef]
  49. Shen, C.; Wang, S.; Bai, Y.; Wu, Y.; Zhang, S.; Chen, M.; Guilfoyle, T.J.; Wu, P.; Qi, Y. Functional analysis of the structural domain of ARF proteins in rice (Oryza sativa L.). J. Exp. Bot. 2010, 61, 3971–3981. [Google Scholar] [CrossRef] [Green Version]
  50. 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]
  51. Zhu, Y.; Wu, N.; Song, W.; Yin, G.; Qin, Y.; Yan, Y. Soybean (Glycine max) expansin gene superfamily origins: Segmental and tandem duplication events followed by divergent selection among subfamilies. BMC Plant Biol. 2014, 43, 14–93. [Google Scholar] [CrossRef] [Green Version]
  52. Ren, R.; Wang, H.; Guo, C.; Zhang, N.; Zeng, L.; Chen, Y.; Ma, H.; Qi, J. Widespread Whole Genome Duplications Contribute to Genome Complexity And Species Diversity in Angiosperms. Mol. Plant 2018, 11, 414–428. [Google Scholar] [CrossRef] [Green Version]
  53. Qiao, X.; Li, Q.; Yin, H.; Qi, K.; Li, L.; Wang, R.; Zhang, S.; Paterson, A.H. Gene duplication and evolution in recurring polyploidization-diploidization cycles in plants. Genome Biol. 2019, 20, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Matthes, M.S.; Best, N.B.; Robil, J.M.; Malcomber, S.; Gallavotti, A.; McSteen, P. Auxin EvoDevo: Conservation and Diversification of Genes Regulating Auxin Biosynthesis, Transport, and Signaling. Mol. Plant 2019, 12, 298–320. [Google Scholar] [CrossRef] [Green Version]
  55. Gorshkova, T.A.; Sal’nikova, V.V.; Chemikosova, S.B.; Ageeva, M.V.; Pavlencheva, N.V.; van Dam, J.E.G. The snap point: A transition point in Linum usitatissimum bast fiber development. Ind. Crops Prod. 2003, 18, 213–221. [Google Scholar] [CrossRef]
  56. Pinzon-Latorre, D.; Deyholos, M.K. Pectinmethylesterases (PME) and Pectinmethylesterase Inhibitors (PMEI) Enriched during Phloem Fiber Development in Flax (Linum usitatissimum). PLoS ONE 2014, 9, e105386. [Google Scholar] [CrossRef] [PubMed]
  57. Guo, Y.; Qiu, C.; Long, S.; Chen, P.; Hao, D.; Preisner, M.; Wang, H.; Wang, Y. Digital gene expression profiling of flax (Linum usitatissimum L.) stem peel identifies genes enriched in fiber-bearing phloem tissue. Gene 2017, 626, 32–40. [Google Scholar] [CrossRef] [PubMed]
  58. Nairn, C.J.; Haselkorn, T. Three loblolly pine CesA genes expressed in developing xylem are orthologous to secondary cell wall CesA genes of angiosperms. New Phytol. 2005, 166, 907–915. [Google Scholar] [CrossRef]
  59. Shibasaki, K.; Uemura, M.; Tsurumi, S.; Rahman, A. Auxin response in Arabidopsis under cold stress: Underlying molecular mechanisms. Plant Cell 2009, 21, 3823–3838. [Google Scholar] [CrossRef] [Green Version]
  60. Zhou, X.; Wu, X.; Li, T.; Jia, M.; Liu, X.; Zou, Y.; Liu, Z.; Wen, F. Identification, characterization, and expression analysis of auxin response factor (ARF) gene family in Brachypodium distachyon. Funct. Integr. Genom. 2018, 18, 709–724. [Google Scholar] [CrossRef]
  61. Bouzroud, S.; Gouiaa, S.; Hu, N.; Bernadac, A.; Mila, I.; Bendaou, N.; Smouni, A.; Bouzayen, M.; Zouine, M. Auxin response factors (ARFs) are potential mediators of auxin action in tomato response to biotic and abiotic stress (Solanum lycopersicum). PLoS ONE 2018, 13, e0193517. [Google Scholar] [CrossRef] [Green Version]
  62. Kang, C.; He, S.; Zhai, H.; Li, R.; Zhao, N.; Liu, Q. A sweet potato auxin response factor gene (IbARF5) is involved in carotenoid biosynthesis and salt and drought tolerance in transgenic Arabidopsis. Front. Plant Sci. 2018, 9, 1307. [Google Scholar] [CrossRef]
  63. Talanova, V.V.; Titov, A.F. Endogenous abscisic acid content in cucumber leaves under the influence of unfavourable temperatures and salinity. J. Exp. Bot. 1994, 45, 1031–1033. [Google Scholar] [CrossRef]
  64. Wilkins, M.R.; Gasteiger, E.; Bairoch, A.; Sanchez, J.C.; Williams, K.L.; Appel, R.D.; Hochstrasser, D.F. Protein identification and analysis tools on the ExPASy server. Methods Mol. Biol. 1999, 112, 531–552. [Google Scholar] [PubMed]
  65. Chou, K.C.; Shen, H.B. Plant-mPLoc: A top-down strategy to augment the power for predicting plant protein subcellular localization. PLoS ONE 2010, 5, e11335. [Google Scholar] [CrossRef] [Green Version]
  66. Bailey, T.L.; Johnson, J.; Grant, C.E.; Noble, W.S. The MEME suite. Nucleic Acids Res. 2015, 43, W39–W49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Wang, Y.; Li, J.; Paterson, A.H. MCScanX-transposed: Detecting transposed gene duplications based on multiple colinearity scans. Bioinformatics 2013, 29, 1458–1460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Chen, C.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.; Xia, R. TBtools: An integrative toolkit developed for interactive analyses of big biological data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef]
  69. Rozas, J.; Ferrer-Mata, A.; Sánchez-DelBarrio, J.C.; Guirao-Rico, S.; Librado, P.; Ramos-Onsins, S.E.; Sánchez-Gracia, A. DnaSP 6: DNA Sequence Polymorphism Analysis of Large Data Sets. Mol. Biol. Evol. 2017, 34, 3299–3302. [Google Scholar] [CrossRef]
  70. Lynch, M.; Conery, J.S. The evolutionary fate and consequences of duplicate genes. Science 2000, 290, 1151–1155. [Google Scholar] [CrossRef] [Green Version]
  71. Tamura, K.; Stecher, G.; Peterson, D.; Filipski, A.; Kumar, S. MEGA6: Molecular evolutionary genetics analysis version 6.0. Mol. Biol. Evol. 2013, 30, 2725–2729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Huis, R.; Hawkins, S.; Neutelings, G. Selection of reference genes for Quantitative gene expression normalization in flax (Linum usitatissimum L.). BMC Plant Biol. 2010, 10, 71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Antonov, J.; Goldstein, D.R.; Oberli, A.; Baltzer, A.; Pirotta, M.; Fleischmann, A.; Altermatt, H.J.; Jaggi, R. Reliable gene expression measurements from degraded RNA by quantitative real-time PCR depend on short amplicons and a proper normalization. Lab. Investig. 2005, 85, 1040–1050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Evolutionary relationship, gene structure, and conserved motifs of LuARFs. (a) Phylogenetic tree of LuARF proteins conducted using the NJ method through MEGA 6.0. (b) Exon–intron organization of LuARF genes. The exons, introns, and upstream/downstream are represented by yellow filled boxes, black lines, and blue filled boxes, respectively. (c) Motif patterns of LuARF proteins detected using MEME. Different colors indicate different motifs.
Figure 1. Evolutionary relationship, gene structure, and conserved motifs of LuARFs. (a) Phylogenetic tree of LuARF proteins conducted using the NJ method through MEGA 6.0. (b) Exon–intron organization of LuARF genes. The exons, introns, and upstream/downstream are represented by yellow filled boxes, black lines, and blue filled boxes, respectively. (c) Motif patterns of LuARF proteins detected using MEME. Different colors indicate different motifs.
Ijms 24 11006 g001
Figure 2. Phylogenetic tree of ARF proteins from flax, Arabidopsis, soybean, cotton, barrel medic, Egrandis, grape, rice, sorghum, and B. distachyon. All ARFs are divided into four main clades, and clades III and IV were further divided into three subgroups, each indicated by a different color. ARF proteins obtained from different species were marked by circles filled with different colors.
Figure 2. Phylogenetic tree of ARF proteins from flax, Arabidopsis, soybean, cotton, barrel medic, Egrandis, grape, rice, sorghum, and B. distachyon. All ARFs are divided into four main clades, and clades III and IV were further divided into three subgroups, each indicated by a different color. ARF proteins obtained from different species were marked by circles filled with different colors.
Ijms 24 11006 g002
Figure 3. Chromosomal distribution of LuARF genes. Red rectangles and gray lines show the gene pairs underwent tandem and segmental duplication, respectively.
Figure 3. Chromosomal distribution of LuARF genes. Red rectangles and gray lines show the gene pairs underwent tandem and segmental duplication, respectively.
Ijms 24 11006 g003
Figure 4. Collinearity analyses of ARF genes between flax and other nine species (Arabidopsis, soybean, cotton, barrel medic, Egrandis, grape, rice, sorghum, and B. distachyon). Gray lines indicate the syntenic blocks between flax and other plant genomes, whereas the red lines represent the collinear ARF gene pairs.
Figure 4. Collinearity analyses of ARF genes between flax and other nine species (Arabidopsis, soybean, cotton, barrel medic, Egrandis, grape, rice, sorghum, and B. distachyon). Gray lines indicate the syntenic blocks between flax and other plant genomes, whereas the red lines represent the collinear ARF gene pairs.
Ijms 24 11006 g004
Figure 5. Promoter analysis of LuARF genes. The number of different cis elements are presented in the form of column charts, which are indicated by different colors.
Figure 5. Promoter analysis of LuARF genes. The number of different cis elements are presented in the form of column charts, which are indicated by different colors.
Ijms 24 11006 g005
Figure 6. Expression patterns of LuARF genes in different tissues of flax. (a) Expression analysis of LuARFs in different tissues and varieties using transcriptome data. (b) Relative expression levels of LuARF4/13 in seeds of two flax varieties at different developmental stages. (c) Relative expression levels of LuARF genes in whole stem and stem peel. SA, A, to H represent the whole stem or the stem peel collected from 0 to 1 cm, 1 to 2 cm, 2 to 3 cm, 3 to 4 cm, 10 to 11 cm, 20 to 21 cm, 30 to 31 cm, 40 to 41 cm, and 50 to 51 cm of Longya-10 plants based on the fiber development stage, respectively. All the results were based on three biological replicates. Error bars indicate ±SD for three biological replicates.
Figure 6. Expression patterns of LuARF genes in different tissues of flax. (a) Expression analysis of LuARFs in different tissues and varieties using transcriptome data. (b) Relative expression levels of LuARF4/13 in seeds of two flax varieties at different developmental stages. (c) Relative expression levels of LuARF genes in whole stem and stem peel. SA, A, to H represent the whole stem or the stem peel collected from 0 to 1 cm, 1 to 2 cm, 2 to 3 cm, 3 to 4 cm, 10 to 11 cm, 20 to 21 cm, 30 to 31 cm, 40 to 41 cm, and 50 to 51 cm of Longya-10 plants based on the fiber development stage, respectively. All the results were based on three biological replicates. Error bars indicate ±SD for three biological replicates.
Ijms 24 11006 g006
Figure 7. Expression profiles of LuARF genes under (a) cold and heat, (b) NaCl, (c) PEG, (d) GA3, (e) IAA, and (f) NAA treatments. CK indicates samples collected from the control seedlings. All the results were based on three biological replicates. Error bars indicate ±SD for three biological replicates. * and ** represents the significant differences at the 0.05 and 0.01 level, respectively.
Figure 7. Expression profiles of LuARF genes under (a) cold and heat, (b) NaCl, (c) PEG, (d) GA3, (e) IAA, and (f) NAA treatments. CK indicates samples collected from the control seedlings. All the results were based on three biological replicates. Error bars indicate ±SD for three biological replicates. * and ** represents the significant differences at the 0.05 and 0.01 level, respectively.
Ijms 24 11006 g007aIjms 24 11006 g007bIjms 24 11006 g007cIjms 24 11006 g007d
Table 1. Basic characteristics of the LuARF genes in flax (Longya-10).
Table 1. Basic characteristics of the LuARF genes in flax (Longya-10).
Gene NameLocus IDChromosome LocationLength (aa)MW (KDa)pIExonsSubcellular Localization
LuARF1L.us.o.m.scaffold38.111Chr2: 4056063-4061396(−)65473.746.4313Nucleus
LuARF2L.us.o.m.scaffold177.42Chr2:12271477-12280264(+)60265.956.052Nucleus
LuARF3L.us.o.m.scaffold33.48Chr2:14271938-14276080(+)68875.216.989Nucleus
LuARF4L.us.o.m.scaffold20.385Chr3:12023683-12026552(+)65972.317.854Nucleus
LuARF5L.us.o.m.scaffold183.107Chr4:2988619-2992012(+)68275.766.984Nucleus
LuARF6L.us.o.m.scaffold336.4Chr4:11172052-11176570(−)937102.775.6014Nucleus
LuARF7L.us.o.m.scaffold2.53Chr5:13559053-13563212(−)68375.985.8114Nucleus
LuARF8L.us.o.m.scaffold255.60Chr5:17150878-17155311(−)64972.378.095Nucleus
LuARF9L.us.o.m.scaffold53.226Chr6:9052662-9078679(+)1010112.576.6710Nucleus
LuARF10L.us.o.m.scaffold82.237Chr8:1405577-1411201(+)84793.846.0415Nucleus
LuARF11L.us.o.m.scaffold82.236Chr8:1413136-1418471(+)948105.436.3914Nucleus
LuARF12L.us.o.m.scaffold33.141Chr9:8160867-8164989(−)59565.116.762Nucleus
LuARF13L.us.o.m.scaffold131.61Chr9:13620512-13622865(−)65772.436.814Nucleus
LuARF14L.us.o.m.scaffold9.422Chr10:2282901-2288310(+)69878.456.7813Nucleus
LuARF15L.us.o.m.scaffold199.64Chr10:13744807-13750216(+)82091.326.1614Nucleus
LuARF16L.us.o.m.scaffold5.303Chr11:9965217-9970709(−)68476.115.7214Nucleus
LuARF17L.us.o.m.scaffold3.144Chr12:12422846-12437639(+)1032114.796.2811Nucleus
LuARF18L.us.o.m.scaffold75.41Chr13:524613-527571(−)47852.828.4912Nucleus
LuARF19L.us.o.m.scaffold71.2Chr13:1846189-1849447(−)78487.266.338Nucleus
LuARF20L.us.o.m.scaffold83.44Chr13:14011405-14017053(−)82091.396.1214Nucleus
LuARF21L.us.o.m.scaffold48.39Chr14:1479346-1484254(−)1123125.346.3111Nucleus
LuARF22L.us.o.m.scaffold48.78Chr14:1639106-1644136(+)87697.135.8714Nucleus
LuARF23L.us.o.m.scaffold48.79Chr14:1646009-1651254(+)924102.676.2314Nucleus
LuARF24L.us.o.m.scaffold56.318Chr14:4404281-4408082(−)84994.306.3113Nucleus
LuARF25L.us.o.m.scaffold168.120Chr14:12632364-12636289(+)68275.576.554Nucleus
LuARF26L.us.o.m.scaffold122.81Chr14:16412628-16419332(+)936102.875.4514Nucleus
LuARF27L.us.o.m.scaffold107.81Chr15:7797623-7802834−)67774.046.8810Nucleus
LuARF28L.us.o.m.scaffold0.555Chr15:11542664-11545214(−)47052.818.5912Nucleus
LuARF29L.us.o.m.scaffold239.1Chr15: 13529432-13532699(−)79288.226.458Nucleus
LuARF30L.us.o.m.scaffold156.62Chr15: 13548791-13552058(−)79288.226.388Nucleus
LuARF31L.us.o.m.scaffold81.17 85094.286.4513Nucleus
LuARF32L.us.o.m.scaffold165.136 79187.685.7314Nucleus
LuARF33L.us.o.m.scaffold434.2 79287.795.8614Nucleus
Note: LuARF31, LuARF32, and LuARF33 were not mapped on any chromosome.
Table 2. Duplicate events, selection pressure, and divergence time of LuARF genes.
Table 2. Duplicate events, selection pressure, and divergence time of LuARF genes.
Gene PairsDuplication EventKaKsKa/KsSelection TypeDivergence Time (Mya)
LuARF10/LuARF11TD0.04950.33110.1495Purifying selection27.1393
LuARF22/LuARF23TD0.05660.30280.1869Purifying selection24.8197
LuARF1/LuARF14SD0.02240.10090.2220Purifying selection8.2705
LuARF2/LuARF12SD0.03750.10970.3418Purifying selection8.9918
LuARF3/LuARF27SD0.03190.17310.1843Purifying selection14.1885
LuARF4/LuARF5SD0.31732.75050.1154Purifying selection225.4508
LuARF4/LuARF13SD0.04040.11770.3432Purifying selection9.6475
LuARF5/LuARF25SD0.01960.10730.1827Purifying selection8.7951
LuARF5/LuARF8SD0.27941.93600.1443Purifying selection158.6885
LuARF8/LuARF25SD0.27892.44930.1139Purifying selection200.7623
LuARF6/LuARF26SD0.00900.07350.1224Purifying selection6.0246
LuARF7/LuARF16SD0.01590.05560.2860Purifying selection4.5574
LuARF9/LuARF17SD0.01580.10570.1495Purifying selection8.6639
LuARF10/LuARF22SD0.01920.16120.1191Purifying selection13.2131
LuARF15/LuARF20SD0.03720.11410.3260Purifying selection9.3525
LuARF15/LuARF32SD0.10290.69600.1478Purifying selection57.0492
LuARF20/LuARF32SD0.10480.72680.1442Purifying selection59.5738
LuARF32/LuARF33SD0.00620.22260.0279Purifying selection18.2459
LuARF19/LuARF29SD0.02500.16600.1506Purifying selection13.6066
LuARF24/LuARF31SD0.00670.06810.0984Purifying selection5.5820
Note: TD indicates tandem duplication; SD indicates segmental duplication.
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

Qi, Y.; Wang, L.; Li, W.; Dang, Z.; Xie, Y.; Zhao, W.; Zhao, L.; Li, W.; Yang, C.; Xu, C.; et al. Genome-Wide Identification and Expression Analysis of Auxin Response Factor Gene Family in Linum usitatissimum. Int. J. Mol. Sci. 2023, 24, 11006. https://doi.org/10.3390/ijms241311006

AMA Style

Qi Y, Wang L, Li W, Dang Z, Xie Y, Zhao W, Zhao L, Li W, Yang C, Xu C, et al. Genome-Wide Identification and Expression Analysis of Auxin Response Factor Gene Family in Linum usitatissimum. International Journal of Molecular Sciences. 2023; 24(13):11006. https://doi.org/10.3390/ijms241311006

Chicago/Turabian Style

Qi, Yanni, Limin Wang, Wenjuan Li, Zhao Dang, Yaping Xie, Wei Zhao, Lirong Zhao, Wen Li, Chenxi Yang, Chenmeng Xu, and et al. 2023. "Genome-Wide Identification and Expression Analysis of Auxin Response Factor Gene Family in Linum usitatissimum" International Journal of Molecular Sciences 24, no. 13: 11006. https://doi.org/10.3390/ijms241311006

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