Next Article in Journal
Foretinib Is Effective against Triple-Negative Breast Cancer Cells MDA-MB-231 In Vitro and In Vivo by Down-Regulating p-MET/HGF Signaling
Next Article in Special Issue
Genome-Wide Identification and Characterization of R2R3-MYB Provide Insight into Anthocyanin Biosynthesis Regulation Mechanism of Ananas comosus var. bracteatus
Previous Article in Journal
Metformin and Its Immune-Mediated Effects in Various Diseases
Previous Article in Special Issue
CRISPR/Cas9 Mutant Rice Ospmei12 Involved in Growth, Cell Wall Development, and Response to Phytohormone and Heavy Metal Stress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identification of NPF Family Genes in Brassica rapa Reveal Their Potential Functions in Pollen Development and Response to Low Nitrate Stress

1
School of Agriculture, Yunnan University, Kunming 650091, China
2
Industrial Crop Research Institute, Yunnan Academy of Agricultural Sciences, Kunming 650205, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(1), 754; https://doi.org/10.3390/ijms24010754
Submission received: 4 November 2022 / Revised: 25 December 2022 / Accepted: 29 December 2022 / Published: 1 January 2023

Abstract

:
Nitrate Transporter 1/Peptide Transporter Family (NPF) genes encode membrane transporters involved in the transport of diverse substrates. However, little is known about the diversity and functions of NPFs in Brassica rapa. In this study, 85 NPFs were identified in B. rapa (BrNPFs) which comprised eight subfamilies. Gene structure and conserved motif analysis suggested that BrNFPs were conserved throughout the genus. Stress and hormone-responsive cis-acting elements and transcription factor binding sites were identified in BrNPF promoters. Syntenic analysis suggested that tandem duplication contributed to the expansion of BrNPFs in B. rapa. Transcriptomic profiling analysis indicated that BrNPF2.6, BrNPF2.15, BrNPF7.6, and BrNPF8.9 were expressed in fertile floral buds, suggesting important roles in pollen development. Thirty-nine BrNPFs were responsive to low nitrate availability in shoots or roots. BrNPF2.10, BrNPF2.19, BrNPF2.3, BrNPF5.12, BrNPF5.16, BrNPF5.8, and BrNPF6.3 were only up-regulated in roots under low nitrate conditions, indicating that they play positive roles in nitrate absorption. Furthermore, many genes were identified in contrasting genotypes that responded to vernalization and clubroot disease. Our results increase understanding of BrNPFs as candidate genes for genetic improvement studies of B. rapa to promote low nitrate availability tolerance and for generating sterile male lines based on gene editing methods.

1. Introduction

Nitrate Transporter 1/Peptide Transporter (NRT1/PTR) family proteins are transporters of the Major Facilitator Superfamily (MFS) and are referred to as NPF (NRT1/PTR family) proteins that are present in all major domains of life [1]. The first identified NPF gene in plants was AtNPF6.3/AtNRT1.1/CHL1 (CHLORINA 1), and the protein was functionally characterized as a nitrate transporter [2]. Subsequently, nitrate has been considered the primary substrate of NPFs. However, further identification of NPFs from various plants has demonstrated that NPFs can transport a wider range of substrates, including nitrate, chloride, oligopeptide, IAA (Auxin), JA (Jasmonate), GA (Gibberllin), abscisic acid (ABA), glucosinolates, potassium, and sugar [3,4,5,6,7,8,9]. Further, some NPFs can transport multiple substrates, such as AtNPF6.3/AtNRT1.1, that has reported dual-affinity nitrate transport activity (i.e., response to both low- and high-nitrate concentrations) and is also involved in auxin transport [3,8].
During plant development and responses to environmental stresses, nutrients and other substrates are transported according to altered metabolic pathways involved in the synthesis, storage, mobilization, and conversion, thereby requiring transporters with capacities to transport a wide diversity of chemical substrates [10,11]. NPFs are one of the largest transporter groups in plants and can transport a wide range of substrates. For example, AtNPF2.10/GTR1 (glucosinolate transporters 1) and AtNPF2.11/GTR2 are essential for the translocation of glucosinolate defense compounds to seeds during maturation [5]. Furthermore, AtNPF2.3 functions as a root stele transporter that contributes to nitrate translocation to shoots during salt stress [12]. In addition, AtNPF2.8/FST1 (Flavonol-Sophoroside-Transporter 1) is expressed in the tapetum and is required for the accumulation of flavonol glycosides on pollen surfaces [13]. ZmNPF7.9 in maize promotes lipid and amino acid homeostasis activity during seed development, while ZmSUGCAR1 (Zea mays Sucrose and Glucose Carrier 1) is a paralog of AtNPF7.3 and is expressed in the basal endosperm transfer layer (BETL) of maize kernels, where it acts as a sugar transporter that imports glucose and sucrose into the endosperm [9,14]. Whenaken together, these observations suggest that NPFs play key roles in plant development and are responsive to environmental stresses, owing to their broad substrate specificity.
Given the important roles of NPFs in plants and the increasing availability of plant genomes, the systematic identification, and analysis of NPF genes have been conducted in many plants, leading to the identification of 53 NPFs in Arabidopsis [15], 80 in rice [16], 73 in apple [17], 178 in sugarcane [18], 169–199 in Brassica napus [19,20,21], 331 in hexaploid wheat (Triticum aestivum L.) [22], 57 in Spinach (Spinacia oleracea L.) [23], and 109 in Tea Plants (Camellia sinensis) [24]. Phylogenetic analysis has consistently divided NPFs into eight subfamilies, termed NPF1 to NPF8 [1]. Using Arabidopsis as a model, AtNPF1 and AtNPF2 from the NPF1 subfamily (comprising three members) are involved in the redistribution of nitrate into developing leaves [25]. The NPF2 family comprises 14 members that can transport numerous substrates, including nitrate, glucosinolates, and phytochromes [26]. Only one NPF gene, AtNPF3.1, is in the Arabidopsis NPF3 subfamily and functions in nitrite and GA transport [27,28]. Seven members comprise the NPF4 subfamily, and some can transport ABA and GA in addition to other substrates [4,28]. The NFP5 subfamily comprises 16 members that participate in the transport of nitrate, dipeptides, GA, JA, and ABA [29,30]. Four members comprise the NPF6 subfamily, including the first characterized NPF gene, AtNPF6.3, which primarily transports nitrate [31,32]. Three members comprise the NPF7 subfamily, all of which can transport nitrate [33,34,35], although cadmium and sodium could also be substrates of AtNPF7.2 and AtNPF7.3 [33,34]. Lastly, the NPF8 subfamily comprises five members, with some of them able to transport JA and dipeptides [36,37]. Notably, half of the AtNPFs with characterized functions are capable of transporting nitrate [35,38].
Brassica rapa is one of the most important crops in China, Korea, and other Asian countries, including several oil seed crops such as sarson and turnip rape, in addition to numerous vegetable crops such as turnip, Pak-choi, and Chinese cabbage [39]. The NPFs of B. rapa were previously isolated to investigate the functions of BnaNPFs in B. napus [19,20,21], although detailed characterizations of BrNPFs remain limited. In this study, BrNPFs were systematically identified and characterized from three Brassica genomes, with analysis of their gene composition, chromosomal locations, phylogenetic relationships, and the presence of cis-elements in their promoters. In silico and Semi-quantitative RT-PCR analyses indicated that BrNPF2.6, BrNPF2.15, BrNPF7.6, and BrNPF8.9 could be related to pollen development, as reflected by the co-expression of genes related to these functions. The responses of BrNPFs due to vernalization and clubroot were also investigated, while the expression of BrNPFs under low nitrate stress was evaluated using transcriptomic data. The results from this study provide a framework for better understanding the functions of BrNPFs during pollen development and B. rapa responses to vernalization, clubroot disease, and low nitrate stress.

2. Results

2.1. Bioinformatics Analysis of NPFs

2.1.1. Identification of NPF Proteins in Three Prototypical Diploid Species of Brassica

A total of 85, 110, and 97 NFP proteins were identified in the genomes of B. rapa, B. oleracea, and B. nigra, respectively (Table S1). The numbers of Br/Bol/BniNPF proteins were 1.6, 2.0, and 1.8 times that in Arabidopsis, respectively, owing to the gene expansion of Brassica during their evolution. The gene IDs, genome locations, coding sequence lengths, protein lengths, and other characteristics of the identified Br/Bol/BniNPF genes are listed in Table S1. The subcellular localization of these 292 proteins were predicted using the WoLF PSORT program, revealing that most of the NPF proteins were located on the plasma membrane (269), chloroplast (11), or in vacuoles (8) (Table S1). The lengths of BrNPFs ranged from 98 to 1070 amino acids, with an average length of 534 amino acids, and MWs ranging from 8.77 to 383.66 kDa, with an average weight of 62.08 kDa. Likewise, a wide range of PIs was also observed (4.73–9.74) which may be due to the considerable range in protein lengths (Table S1).

2.1.2. Phylogenetic, Conserved Motif, and Gene Structure Analysis of BrNFPs

In order to explore the evolutionary characteristics and classification of BrNPF proteins, phylogenetic analysis with neighbor-joining methods was conducted with 292 Br/Bol/BniNPF protein sequences and 53 AtNPFs. All NFP proteins were classified into eight subfamilies, based on the scheme presented in [1], and belonged to the families NPF1-8 (Figure 1). The NPF5 subfamily comprised the most NFP genes (109 members), followed by the subfamilies NPF2 (85 members), NPF4 (39 members), NPF8 (35 members), NPF6 (27 members), NPF7 (23 members), NPF1 (17 members), and NPF3 (10 members). An ortholog in the Brassica genomes was not found in the subfamily NFP2 for AtNPF2.1, AtNPF2.2, and AtNPF2.5, indicating gene loss during evolution (Figure 1). Six homologs of AtNPF5.2 were identified in B. rapa (BrNPF5.4, BrNPF5.5, BrNPF5.6, BrNPF5.16, BrNPF5.17, and BrNPF5.18), six in B. oleracea (BolNPF5.6, BolNPF5.7, BolNPF5.8, BolNPF5.39, BolNPF5.40, and BolNPF5.41) and four in B. nigra (BniNPF5.11, BniNPF5.12, BniNPF5.13, and BniNPF5.14) (Figure 1), indicating the expansion of AtNPF5.2 in three prototypical Brassica diploid species over evolutionary time.
The conserved motifs of the BrNPFs were analyzed with the MEME program, yielding the identification of 10 conserved motifs among 85 BrNPFs from B. rapa (Figure 2 and Table S2). The 10 motifs of typical BrNPF proteins followed the order of Motif9-Motif4-Motif2-Motif10-Motif6-Motif5-Motif3-Motif7-Motif1-Motif8. Apart from motifs 7, 8, and 9, the other motifs contained the core PTR2 domain conserved sequence (Table S3). No significantly conserved motifs were identified for the 7, 8, and 9 domains based on BLAST searches with NCBI and Pfam.
In order to investigate the structural diversity of BrNPFs, exon-intron organizations were analyzed. The number of BrNPF introns ranged from 0 to 14. The most common exon-intron organization comprised four exons separated by three introns, which were present in 41 BrNPFs (Figure 2 and Table S2). Except for BrNPF2.8, BrNPF2.18, and BrNPF7.4, most BrNPFs contained more than one intron, indicating the possible existence of alternative splicing during expression. Eight types of gene structures were identified in the NPF5 subfamily, implying they exhibited diverse functions. BrNPF members from the same subgroups exhibited similar gene structures for the other subfamilies, indicating the potential for conserved functions (Figure 2).

2.1.3. BrNPF Chromosomal Location and Gene Duplication Analysis

Eighty-five BrNPF genes were present on the ten chromosomes of Brassica rapa and were non-uniformly distributed (Figure A1). Chromosome A09 harbored the largest number of BrNPFs (18 members), followed by chromosomes A07, A02, and A06, which contained 14, 13, and 11 BrNPFs, respectively. Chromosome A04 carried the smallest number of BrNPFs (two). BLAST and MCScanX analysis indicated that BrNPF gene duplication events were present in the B. rapa genome. Briefly, 22 tandem duplicated genes (25.9%) were identified that belonged to ten clusters (Figure 3 and Figure A1). Among the tandem duplicated genes, two, two, and three clusters were located on chromosomes A02, A07, and A09, respectively. The other three clusters were located on chromosomes A03, A05, and A06, respectively (Figure 3 and Figure A1). These results suggest that tandem duplication is related to NPF expansion in Brassica genomes.
To evaluate the collinear relationships of all NFP genes in Arabidopsis, B. rapa, B. nigra, and B. oleracea, collinear gene pairs were identified using the MCScanX software package. A total of 26, 18, and 13 gene pairs were identified between B. rapa and B. oleracea, B. rapa and B. nigra, and B. rapa and Arabidopsis, respectively (Figure 3 and Table S3). All syntenic NPF genes in B. rapa were located on chromosomes A01, A02, A07, and A09 (Figure 3). Six segmental duplication events were also identified within the B. oleracea genome (Figure 3), indicating that the greater numbers of NPFs in the B. oleracea genome could be due to segmental gene duplication events in its genome.

2.1.4. Cis-Elements in Promoters of BrNPFs

In order to identify the cis-regulatory elements in BrNPF promoters, cis-elements were analyzed using the PlantCARE platform (https://bioinformatics.psb.ugent.be/webtools/plantcare/html/ accessed on 16 August 2022). A total of 904 phytohormone-responsive elements, 1605 environmental-responsive elements, 205 plant growth, and development-related elements, and 979 transcriptional factor binding sites were predicted within the promoters of the 85 BrNPFs (Table S4). Among these, light-responsive, MYB transcription-binding, MYC transcription-binding, and ABA-responsive cis elements were the four most prevalent (Table S4). Thus, most BrNPFs respond to diverse environmental stresses and are regulated by various transcriptional factors (TFs).

2.2. Tissue Expression of BrNPFs Reveals Their Potential Functions during Pollen Development

In order to investigate the expression of BrNPFs, their expression patterns were compared using RNA-sequencing data from 59 different organ or tissues samples, including callus, roots, stems, stem leaves, flowers, siliques, head leaves (24 samples), developmental stages of floral buds (10 samples), pistils (four samples), unfertilized ovules, embryos (seven samples), and seed coats (seven samples) (Table S5). BrNPFs were differentially expressed among groups of the 59 tissue samples (Figure 4). The BrNPF1.1, BrNPF1.2, BrNPF1.3, and BrNPF6.5 genes were generally predominantly expressed in all tissues except embryos and seed coats. BrNPF2.5 and BrNPF2.21 were mostly expressed in stem leaves and opened flowers. BrNPF7.3 and BrNPF7.5 were predominantly expressed in roots, while BrNFP7.1 was predominantly expressed in callus tissues. BrNPF2.5 and BrNPF2.21 were mostly expressed in all developmental floral buds and late development seed coats. Further, BrNPF2.6, BrNPF2.15, BrNPF7.6, and BrNPF8.9 were predominantly expressed in fertile buds from the uninucleate to binucleate pollen stages. The primary difference between fertile and sterile floral buds is the presence of pollen grains [40]. Accordingly, BrNPF2.6, BrNPF2.15, BrNPF7.6, and BrNPF8.9 were predominantly expressed per normal pollen development in B. rapa. Semi-quantitative RT-PCR was conducted to confirm the expression patterns of these four genes based on the different developmental stages of floral buds from male genetic sterility (GMS) lines (Figure 5A,B). The expression patterns of these four genes were similar between RNA-Seq and RT-PCR datasets. Briefly, BrNPF2.6, BrNFP7.6, and BrNPF8.9 were specifically expressed in the floral buds from F2 (floral buds containing tread stage pollen) to F3 (floral buds after the tetrad stage, but before containing mature pollen) stages (Figure 5B). In addition, BrNPF2.15 was only specifically expressed in F3 floral buds (Figure 5B).
In order to identify the functions of BrNPF2.6, BrNPF2.15, BrNPF7.6, and BrNPF8.9 during pollen development, co-expression analysis was conducted with the RNA-seq data from five different types of sterile male lines [41,42,43,44]. Using a Pearson correlation coefficient (PCC) value criterion between -0.6 and 0.6, a total of 1627, 1796, 1829, and 2309 genes were co-expressed with BrNPF2.6, BrNPF2.15, BrNPF7.6, and BrNPF8.9, respectively (Table S6). GO enrichment analysis was subsequently conducted to assess biological processes related to these co-expressed genes (Figure 5C). Pollen development, gametophyte development, and sexual reproduction were represented by the genes co-expressed with BrNPF2.6 and BrNPF8.9 (Figure 5C). In addition, pollen tube development, pollination, and pollen tube growth processes were represented by genes co-expressed with BrNPF7.6 and BrNPF8.9 (Figure 5C). Co-expression analysis and reference to Arabidopsis revealed that several pollen coats or tapetum development-related genes were co-expressed with BrNPF2.6 and BrNPF8.9, including BrSHT (homolog of Arabidopsis Spermidine Hydroxycinnamoyl Transferase), BrA7 (homolog of Arabidopsis Thaliana Anther 7), and BrPTEN1 (homolog of Arabidopsis Phosphatase And Tensin Homolog Deleted On Chromosome Ten 1) (Figure 5D) [45,46,47]. In addition, BrPRK2 (homolog of Arabidopsis Pollen Receptor such as Kinase 2), BrARO1 (homolog of Arabidopsis Armadillo Repeat Only 1), BrRABA4D (homolog of Arabidopsis Rab Gtpase Homolog A4d, Atraba4D), and other important genes for pollen tube growth were represented in the genes co-expressed with BrNPF7.6 and BrNPF8.9 (Figure 5E) [48,49,50].

2.3. Expression of BrNPFs during B. rapa Growth during Vernalization and P. brassicae

NPFs have been previously suggested to be involved in plant responses during vernalization [21]. In order to identify BrNPFs that might be responsive to vernalization in B. rapa, RNA-Seq data for BrNPFs during vernalization were re-calculated, as previously described [51]. Briefly, BrNPF2.5, BrNPF2.21, BrNPF3.3, BrNPF6.1, BrNPF6.3, BrNPF7.1, and BrNPF7.2 were up-regulated during vernalization in both genotypes, while BrNPF7.3 and BrNPF7.5 were only up-regulated in the late bolting type (Table S7 and Figure 6A). BrNPF2.5 and BrNFP2.21 are homologs of AtNPF2.13 that are responsible for source-to-sink remobilization of nitrate [15], indicating that they might play important roles during nitrate remobilization during vernalization. BrNPF3.3 is a homolog of AtNPF3.1 that enables the transport of GA [28], indicating that BrNPF3.3 might play important role in GA transport during the developmental phase transition.
The development of clubroot disease can be influenced by nitrogen fertilization [52], and NPFs are one of the main transporters of nitrogen, indicating that NPFs might be involved in a response mechanism to P. brassicae infection. RNA-seq data suggested that many NPF members were expressed after inoculation with P. brassicae (Table S7 and Figure 6B). Briefly, BrNPF2.23 was only responsive to P. brassicae in the susceptible line, while BrNPF5.3 was induced in the susceptible line after infection by P. brassicae (Figure 6B). BrNPF2.21 and BrNPF7.4 expressions were both up-regulated in the resistant line compared to the susceptible line (Figure 6B). BrNPF2.24 exhibited down regulation in both the susceptible and resistant lines. Thus, NPF genes may participate in clubroot disease responses via the transport of their specific substrates.

2.4. Expression of BrNPF Responses to Low Nitrate Conditions

Most NPF genes in Arabidopsis are related to nitrate transport, indicating that nitrate is the primary substrate of NPF proteins [53]. To assess the potential functions of BrNPFs in nitrate uptake and use, B. rapa seedlings (accession Chiifu-401-42) were hydroponically cultured in Hoagland’s nutrient solution [21] and treated with low nitrate conditions. After treatment, plant heights, leaf areas, fresh weights, and nitrogen contents were significantly lower, while root lengths increased compared to normal growth conditions (Figure 7A–F). The root and shoot components of seedlings were then separately sampled for RNA-seq analysis. Considering the criteria of TPM ≥ 1 and fold change values > 2.0, a total of 39 BrNPFs were identified as responsive to low nitrate conditions either in the shoots or roots (Table S8 and Figure 7G). Among these, ten BrNPFs (BrNPF2.12, BrNPF2.14, BrNPF2.21, BrNPF2.22, BrNPF2.25, BrNPF2.7, BrNPF3.2, BrNPF5.17, BrNPF7.1, and BrNPF7.2) were up-regulated in both shoots and roots under low nitrate conditions, while BrNPF1.3 and BrNPF6.5 were down-regulated in both shoots and roots. Twelve BrNFPs were specifically responsive to low nitrate conditions in shoots, of which eleven genes (BrNPF2.13, BrNPF2.23, BrNPF2.24, BrNPF2.5, BrNPF3.3, BrNPF4.8, BrNPF5.19, BrNPF5.22, BrNPF5.3, BrNPF7.5, and BrNPF8.2) were up-regulated, indicating their potential positive function in nitrate homeostasis (Table S8 and Figure 7G). Seven genes (BrNPF2.10, BrNPF2.19, BrNPF2.3, BrNPF5.12, BrNPF5.16, BrNPF5.8, and BrNPF6.3) were only up-regulated in roots indicating that they may play positive roles in nitrate absorption. In addition, five BrNFPs (BrNPF1.4, BrNPF2.9, BrNPF3.1, BrNPF5.1, and BrNPF8.4) were down-regulated in roots, suggesting they may play negative roles in nitrate absorption (Table S8 and Figure 7G). Three BrNPFs (BrNPF6.6, BrNPF6.7, and BrNPF7.3) were up-regulated in shoots, but down-regulated in roots, indicating that they may exhibit contrasting roles in roots and shoots during low nitrate conditions.

3. Discussion

3.1. Identification and Analysis of BrNPFs

NFPs are one of the largest groups of transporter family genes in plants. Genome-wide identification of NFP family genes has been conducted in many plants based on sequence and motif conservation, including in Arabidopsis [15], rice [54], apple [17], sugarcane [18], Brassica napus [19,20,21], hexaploid wheat (Triticum aestivum L.) [22], spinach (Spinacia oleracea L.) [23] and tea plants (Camellia sinensis) [24]. In this study, 85 NFPs were identified in B. rapa, in addition to 110 members in B. oleracea and 97 in B. nigra (Table S1). Ninety-five NPFs were previously identified in B. rapa (“Chiifu-401”, version 1.5) [1]. The differences in identification could be explained by the use of a higher quality genome version (“Chiifu-401”, version 3.0) [55], leading to greater accuracy in BrNPF identification. Similar results were also observed for B. napus, in which 169, 193, and 199 NPFs were identified based on different versions of B. napus genomic data [19,20,21]. Brassica species have undergone an extra genomic duplication event compared to Arabidopsis [56]. Thus, one Arabidopsis gene should theoretically have one to three orthologs in Brassica genomes. However, the expansion of NPFs in B. rapa, B. oleracea, and B. nigra represent 1.6-, 2.0, and 1.8-fold increases relative to the Arabidopsis genome, respectively. Thus, duplicated genes may have been lost during Brassica evolution. Consistently, AtNPF2.1, AtNPF2.2, and AtNPF2.5 did not have any orthologs in the three species of Brassica analyzed here, while six, six, and four orthologs of AtNPF5.2 were identified in the B. rapa, B. oleracea, and B. nigra genomes, respectively (Figure 1 and Table S1).
Segmental duplication and tandem duplication are two major mechanisms of gene family duplication in plants [57]. In this study, 22 tandem duplicated genes (25.9%) were identified in the B. rapa genome, while no segmental duplication events were identified (Figure S1). Thus, tandem duplication may be a primary driving force in the expansion of BrNPFs. In contrast, six segmental duplication events were observed within the genome of B. oleracea, which may explain the greater number of NPFs in the B. oleracea genome compared to the B. rapa and B. nigra genomes (Figure 3).

3.2. BrNPF Functions

NPF genes represent one of the largest transporter gene families in plants and participate in the transport of diverse substrates across membranes over short or long distances [11,19]. Cis-elements play key roles in controlling gene expression during plant development and are responsive to stress [58]. Numerous cis-elements have been previously observed in the promoter regions of NPFs from B. napus, tea plant, apple, and spinach plants, in addition to others [14,19,20,23,52]. In the present study, many phytohormone-responsive, environmental-responsive, plant growth and development-related, and transcriptional factor binding cis-elements were identified in the promoters of the BrNPFs (Table S4). The existence of abundant and diverse cis-elements in NPF gene promoters could be related to their multiple functions during plant development and responses to environmental stresses.
Gene expression patterns can provide clues for predicting gene functions. Consequently, expression profiles for BrNPFs from 59 diverse tissues were analyzed, in addition to expressional profiles responsive to vernalization and P. brassicae infection between contrasting genotypes and under low nitrate availability stress (Figure 4, Figure 6 and Figure 7). Some BrNPFs exhibited tissue-specific expression, while others exhibited differential expression due to vernalization, P. brassicae infection, and low nitrate availability. For example, the homolog of AtNPF3.1, BrNPF3.3, exhibited induction by vernalization (Figure 6A). Nitrate was previously reported to delay flowering time via the GA signaling pathway [59]. Further, AtNPF3.1 in Arabidopsis has been reported to be involved in GA transport [60], with GA increasing during B. rapa vernalization [61]. The up-regulation of BrNPF3.3 due to vernalization suggests that it might contribute to GA presence during vernalization. Further, nitrogen fertilization has been shown to affect the susceptibility of B. napus [52]. Homologs of AtNPF2.13 and BrNPF2.21 exhibited up-regulation in the resistant genotype (Figure 6B). AtNPF2.13 is involved in the remobilization of nitrate from sources to sinks [62], indicating that BrNPF2.21 might respond to clubroot disease through the redistribution of nitrate. These results collectively provide new insights for the future functional prediction and characterizations of BrNPFs.

3.3. BrNPFs and Pollen Development

AtNPF2.8 was previously suggested by co-expression analysis to be involved in the transport of flavonol-3-O-sophoroside from tapetum cells to pollen walls in Arabidopsis [13]. Two orthologs of AtNPF2.8 were identified in B. rapa, including BrNFP2.6 and BrNPF2.15 (Figure 1 and Table S1). Both orthologs exhibited expression only in fertile floral buds of MS lines (Figure 5B). Further, BrNPF2.6 was highly expressed in floral buds containing pollen grains from tetrads prior to the mature stage (from the F2 to F3 stages). In addition, BrNPF2.15 only exhibited expression in F3-stage floral buds that contain pollen grains after the tetrad to before the mature stages (Figure 5B). Additionally, gene co-expression profiles and associated GO enrichment biological processes differed between BrNFP2.6 and BrNPF2.15. For example, pollen development was represented by genes co-expressed with BrNFP2.6, while pollen tube development was suggested by genes co-expressed with BrNPF2.15 (Figure 5C,D). Thus, the function of AtNPF2.8 might have expanded or diversified over Brassica evolution. Flavonol diglucosides are essential for maintaining pollen fertility and increasing pollen tolerance to environmental stresses [63]. Thus, expanded AtNPF2.8 genes suggest that the regulatory network related to flavonol diglucoside metabolism during pollen development might be more complex in B. rapa compared to Arabidopsis.
BrNFP7.6 and BrNFP8.9 were also specifically expressed in fertile floral buds (Figure 4 and Figure 5B) and were the only identified orthologs of AtNPF7.1 and AtNPF8.2, respectively (Figure 1). AtNPF8.2 is also referred to as AtPRT5 and facilitates peptide transport into germinating pollen and possibly into maturing pollen, ovules, and seeds [36]. Pollen development and pollen tube development were suggested by genes co-expressed with BrNFP8.9 (Figure 5C–E). Based on the phylogenetic and Semi-quantitative RT-PCR analysis results, the functions of AtNPF8.2 and BrNFP8.9 may be highly conserved during pollen development and pollen tube growth. The processes of pollination, pollen tube growth, lipid oxidation, transport, and response to nutrient levels were also represented by genes co-expressed with BrNPF7.6, indicating that they might also play important roles in pollen development. Taken together, the fertile floral buds exhibited genes that were specifically expressed in these tissues, and these may provide new breeding targets for creating sterile male lines in B. rapa.

3.4. Responses of BrNPFs during Low Nitrate Stress

Nitrate was previously identified as the main substrate of NPF genes [22,53]. Here 45.9% of BrNPFs (39 of 85) exhibited differential expression due to low nitrate availability stress (Figure 7G). Among these, ten were induced in both shoots and roots, eleven were specifically induced in shoots, and seven were only up-regulated in roots, indicating that they might play positive roles in nitrate absorption, uptake, homeostasis, and redistribution under low nitrate availability conditions (Table S8 and Figure 7G). AtNPF2.3 in Arabidopsis functions as a root stele transporter and contributes to nitrate translocation to shoots during salt stress [12]. In this study, a homolog of AtNPF2.3, BrNPF2.3, exhibited specific induction in roots, indicating that it might serve a similar function (Figure 1 and Figure 7G). AtNPF6.2 plays key role in regulating leaf nitrate homeostasis [32]. The expression of its homolog, BrNPF6.5, was repressed in both shoots and roots, indicating that low nitrate availability may lead to decreased NPF transporter activity (Figure 1 and Figure 7G). AtNPF2.13 is reportedly involved in the transport of nitrate and GA [15]. A homolog of AtNPF2.13, BrNPF2.21, was up-regulated in both shoots and roots, indicating that BrNPF2.21 might function in response to low nitrate availability by coupling to hormone signaling. AtNPF6.3 can repress lateral root growth under low nitrate availability by promoting basipetal auxin transport out of roots [3]. In this study, a homolog of AtNPF6.3, BrNPF6.6, was down-regulated in roots but up-regulated in shoots (Figure 1 and Figure 7G). Taken together, the expression of BrNPFs under low nitrate availability conditions suggests that there is a crosstalk between low nitrate stress responses and phytohormone signaling pathways, consistent with results from previous studies [59].

4. Materials and Methods

4.1. Plant Growth and Low Nitrate Treatments

Uniform B. rapa seeds (accession Chiifu-401-42) were germinated in Petri dishes at 23 ± 1 °C in the dark for two days, followed by hydroponic cultivation of germinated seeds in Hoagland’s nutrient solution for one week [64]. To establish low nitrate treatments, KNO3 and Ca(NO3)2 in Hoagland’s solutions were replaced with KCl and CaCl2, respectively. The final concentration of NO3 in the treatments was 0.1 mM. During cultivation, growth conditions within growth chambers were set as previously described [65]. After treatments, the shoots and roots in the low nitrogen and control treatments were individually harvested and immediately frozen in liquid nitrogen, followed by storage at −80 °C.
In order to collect materials from male sterile (MS) lines, seeds of MS lines from our previous study were germinated in Petri dishes at 23 ± 1 °C in the dark [40]. Vernalization was then induced with germinated seeds at 4 °C in the dark for 30 days. After vernalization, seeds were sown into pots (15 × 15 × 18 cm) containing potting soil and transferred to a greenhouse, followed by growth at 23 ± 1 °C with a light intensity of 6000–7000 Lux under a long day photoperiod (light/dark, 16 h/8 h). After flowering, floral buds were collected from MS line plants using three biological replicates and with previously reported criteria [40]. Root and shoot tissues were collected from three-week-old seedlings without vernalization. Stem and leaf tissues were sampled from plants one week after bolting. The siliques were collected two weeks after pollination. After sampling, tissues were immediately frozen in liquid nitrogen and stored at −80 °C until further analysis.

4.2. Identification of BrNPFs in Three Prototypical Diploid Species of Brassica

To identify NPF proteins from three prototypical diploid species of Brassica, all putative protein sequences encoded by the B. rapa (“Chiifu-401”, version 3.0) [55], B. oleracea (“JZS,” version 2.0) [66] and B. nigra (“Ni100”, version 2.0) [67] genomes were downloaded from the Brassicaceae Database (BRAD; www.brassicadb.cn accessed on 20 April 2022). The previously identified 53 AtNPFs were used as queries to search against Brassica protein sequences (using an E-value < 10−5 and identity > 20%). Search hits without a proton-dependent transport 2 (PTR2) domain (PF00854) were excluded based on HMM analysis with an E-value cut-off of 10−5. In order to identify all potential NPFs among the three Brassica, the search baits were used as BLAST queries for searching against the Phytozome 13 and NCBI databases with an E-value < 10−5 and identity > 20%. No additional predicted NPFs were identified at this stage. All NPFs were identified according to previously reported rules [1].

4.3. Phylogenetic and Bioinformatic Analysis of BrNPFs

Phylogenetic analysis was conducted with the NPF protein sequences from Arabidopsis, B. oleracea, B. nigra, and B. rapa after alignment with the MUSCLE program, implemented in MEGA6 with default parameters [68]. An unrooted phylogenetic reconstruction was then constructed using MEGA6 with neighbor-joining methods and analysis parameters, including pairwise taxa deletion, 1000 bootstrap replicates, and the use of the Jones Taylor Thornton (JTT) amino acid substitution model [69]. The chromosomal positions of each NPF from the three genomes were identified among those from the BRAD (www.brassicadb.cn accessed on 16 August 2022) and visualized with a custom Python script. The isoelectric point (PI) and molecular weights (MWs) of the NPFs were analyzed using the ProtParam tool (Expasy, the Swiss Bioinformatics Resource Portal, https://web.expasy.org/protparam/ accessed on 16 August 2022) [70]. Subcellular localization predictions of NPFs were conducted using the WoLF PSORT software package (https://wolfpsort.hgc.jp/ accessed on 16 August 2022) with default settings. Conserved motifs in the BrNPFs were identified using the MEME software program (Suite 5.1.1, http://meme-suite.org/ accessed on 16 August 2022) [71]. BrNPFs gene structures were drawn using the Gene Structure Display Server (GSDS; version 2.0, http://gsds.cbi.pku.edu.cn/ accessed on 16 August 2022) [72].

4.4. RNA Extraction, Leaf Area, and Nitrate Content

Total RNA was isolated from 100 mg of homogenized leaves using the RNAiso Plus Reagent (Takara Biomedical Technology Co., Ltd., Beijing, China) according to the manufacturer’s instructions. The outermost leaves of Chiifu seedlings were dissected for leaf area determination using a Yaxin-1241 leaf meter (Beijing Yaxinliyi, Beijing, China), following the manufacturer’s instructions. The nitrogen concentrations of oven-dried shoots were measured using the Kjeldahl method with a JK9830 Kjeldahl Auto Analyzer (ELITE-Lab Instrument Co., Ltd., Jinan, China) [73] and are expressed as concentrations of per hundred dry matter (g/100 g).

4.5. RNA-Sequencing and Assembly

RNA samples from low nitrate treatments were sent to Gene Denovo Biotechnology Co., Ltd. (Guangzhou, China) for RNA-Seq analysis. The RNA libraries were constructed and sequenced on the Illumina platform. Sequencing and analyses were conducted following standard protocols at Gene Denovo Biotechnology Co., Ltd. (Guangzhou, China). Filtered clean reads were then mapped to the reference genome (“Chiifu-401”, version 3.0) using the HISAT2 software program [74], and transcripts per million (TPM) values were calculated using DESeq [75]. Raw sequencing data were deposited in the China National Center for Bioinformation (CNCB) under the project ID PRJCA012597.

4.6. Expression of BrNPFs within RNA-Seq Data

To analyze the expression of BrNPFs among various tissues, publicly available RNA-seq data from 59 different organs or tissues of B. rapa were retrieved from the NCBI database via the Bioproject accessions PRJNA185152, PRJNA778186, PRJNA641876, and PRJNA473318 [76,77,78,79]. Data were included from tissues comprising callus, stems, stem leaves, opened flowers, siliques, different developmental stages of heading leaves, floral buds, pistils, embryos, and seed coats. Gene expression levels were re-calculated using the transcripts per million (TPM) metric. Heatmaps for expression profiles of BrNPFs were generated using the TBtools software program (version 1.0987663) [80].
The expression of BrNPFs after vernalization was recalculated for a previous study (BioProject PRJNA615255) using the TPM metric [51]. To achieve vernalization, two inbred B. rapa accessions, including a late bolting type (JWW) and an early bolting type (XBJ), were investigated. Prior to vernalization, both inbred lines were grown at 25 ± 2 °C for 32 days under natural light conditions. Both inbred lines were then transferred to a growth chamber at 4 °C with 150 µmol m−2s−1 light intensity under long daylight conditions (16/8 h, day/night) for vernalization, followed by a collection of the third fully expanded leaves from the center for subsequent analyses. JWW leaves were collected after 0, 25, 30, 35, 45, and 50 days following treatment. XBJ leaves were collected 0, 10, 15, 25, 40, and 50 days after treatment.
The TPM values of BrNPF genes were recalculated after infection with Plasmodiophora brassicae, as described in previous studies (Bio-Project PRJNA743585) [81]. In order to initiate P. brassicae infection, 20-day-old healthy plants of resistant (BrT24) and susceptible (Y510-9) B. rapa genotypes were inoculated with 20 mL of a P. brassicae (race 4) solution. For the control group, 20 mL of sterile water was used for inoculation. The root samples for each genotype were then collected at 0, 3, 9, and 20 d after inoculation, based on the four-time points of disease development [81].

4.7. Semi-Quantitative RT-PCR

First-strand cDNA was synthesized using the PrimeScript™ RT reagent Kit (Takara Biomedical Technology Co., Ltd., Beijing, China) using 1 μg of total RNA. Synthesized cDNA was then diluted to 10 ng/μL for PCR amplification. Semi-quantitative RT-PCR assay reactions (20 µL) contained: 2 µL (20 ng) template cDNA, 1.0 µL (10 pmol) of forward primer, 1.0 µL (10 pmol) of the reverse primer, 10 µL 2 × Tag PCR StarMix (GenStar Biosolutions Co., Ltd., Beijing, China) and 6 µL distilled water. The thermocycling conditions were: denaturation at 94 °C for 5 min, followed by 94 °C for 30 s with 28 cycles, then 55 °C for 30 s, and 72 °C for 60 s. PCR primer sequences used for semi-quantitative RT-PCR are shown in Table S9. Semi-quantitative RT-PCR products were separated on 1.5% agarose gels and stained with ethidium bromide to evaluate amplification success.

5. Conclusions

Here, a total of 85, 110, and 97 NFP proteins were identified in the genomes of B. rapa, B. oleracea, and B. nigra, respectively. The gene structures, chromosomal locations, conserved motifs, cis-elements, evolutionary relationships, gene duplications, and expression patterns of the BrNFPs were systematically analyzed. These results provide new targets for future studies to elucidate the molecular mechanisms underlying BrNPF functions in pollen development, nitrate utilization, responses to vernalization, and P. brassicae infection response in B. rapa, especially for BrNPF2.6, BrNPF2.15, BrNPF7.6, and BrNPF8.9 showing potential for generating sterile male lines based on gene editing methods in B. rapa and, possibly, other crops.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24010754/s1.

Author Contributions

Conceptualization: X.Y.; data curation: X.Y., J.G. and X.D.; formal analysis: X.Y. and W.H.; funding acquisition: X.D. and G.L.; project administration: X.D.; supervision: X.D.; validation: J.Q., W.H., Y.L., Y.Z., J.W. and X.C.; original draft writing: X.Y.; review and editing of manuscript: J.G. and X.D. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financially supported by the Key Program for Applied Basic Research of the Yunnan Province of China (202102AE090002) and the Applied Basic Research Project of Yunnan (grant nos. 202001BB050017 and 202101AT070016).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The datasets generated or analyzed in the present study are available in the supplementary data or within the CNCB (China National Center for Bioinformation) under project ID PRJCA012597.

Acknowledgments

We are grateful to anonymous reviewers for their constructive comments on the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Figure A1. Chromosomal locations of BrNPFs identified in this study. The chromosomal number is indicated above each chromosome. Black ovals on each chromosome indicate centromeric regions. Tandem duplicated genes are indicated by purple rectangles.
Figure A1. Chromosomal locations of BrNPFs identified in this study. The chromosomal number is indicated above each chromosome. Black ovals on each chromosome indicate centromeric regions. Tandem duplicated genes are indicated by purple rectangles.
Ijms 24 00754 g0a1

References

  1. Leran, S.; Varala, K.; Boyer, J.C.; Chiurazzi, M.; Crawford, N.; Daniel-Vedele, F.; David, L.; Dickstein, R.; Fernandez, E.; Forde, B.; et al. A unified nomenclature of NITRATE TRANSPORTER 1/PEPTIDE TRANSPORTER family members in plants. Trends Plant Sci. 2014, 19, 5–9. [Google Scholar] [CrossRef] [PubMed]
  2. Tsay, Y.F.; Schroeder, J.I.; Feldmann, K.A.; Crawford, N.M. The herbicide sensitivity gene CHL1 of Arabidopsis encodes a nitrate-inducible nitrate transporter. Cell 1993, 72, 705–713. [Google Scholar] [CrossRef]
  3. Krouk, G.; Lacombe, B.; Bielach, A.; Perrine-Walker, F.; Malinska, K.; Mounier, E.; Hoyerova, K.; Tillard, P.; Leon, S.; Ljung, K.; et al. Nitrate-regulated auxin transport by NRT1.1 defines a mechanism for nutrient sensing in plants. Dev. Cell 2010, 18, 927–937. [Google Scholar] [CrossRef] [PubMed]
  4. Kanno, Y.; Hanada, A.; Chiba, Y.; Ichikawa, T.; Nakazawa, M.; Matsui, M.; Koshiba, T.; Kamiya, Y.; Seo, M. Identification of an abscisic acid transporter by functional screening using the receptor complex as a sensor. Proc. Natl. Acad. Sci. USA 2012, 109, 9653–9658. [Google Scholar] [CrossRef] [Green Version]
  5. Nour-Eldin, H.H.; Andersen, T.G.; Burow, M.; Madsen, S.R.; Jorgensen, M.E.; Olsen, C.E.; Dreyer, I.; Hedrich, R.; Geiger, D.; Halkier, B.A. NRT/PTR transporters are essential for translocation of glucosinolate defence compounds to seeds. Nature 2012, 488, 531–534. [Google Scholar] [CrossRef] [PubMed]
  6. Boursiac, Y.; Leran, S.; Corratge-Faillie, C.; Gojon, A.; Krouk, G.; Lacombe, B. ABA transport and transporters. Trends Plant Sci. 2013, 18, 325–333. [Google Scholar] [CrossRef]
  7. Li, H.; Yu, M.; Du, X.Q.; Wang, Z.F.; Wu, W.H.; Quintero, F.J.; Jin, X.H.; Li, H.D.; Wang, Y. NRT1.5/NPF7.3 Functions as a Proton-Coupled H(+)/K(+) Antiporter for K(+) Loading into the Xylem in Arabidopsis. Plant Cell 2017, 29, 2016–2026. [Google Scholar] [CrossRef] [Green Version]
  8. Wang, W.; Hu, B.; Li, A.; Chu, C. NRT1.1s in plants: Functions beyond nitrate transport. J. Exp. Bot. 2020, 71, 4373–4379. [Google Scholar] [CrossRef]
  9. Yang, B.; Wang, J.; Yu, M.; Zhang, M.; Zhong, Y.; Wang, T.; Liu, P.; Song, W.; Zhao, H.; Fastner, A.; et al. The sugar transporter ZmSUGCAR1 of the Nitrate Transporter 1/Peptide Transporter family is critical for maize grain filling. Plant Cell 2022, 34, 4232–4254. [Google Scholar] [CrossRef]
  10. Gani, U.; Vishwakarma, R.A.; Misra, P. Membrane transporters: The key drivers of transport of secondary metabolites in plants. Plant Cell Rep. 2021, 40, 1–18. [Google Scholar] [CrossRef]
  11. Kanstrup, C.; Nour-Eldin, H.H. The emerging role of the nitrate and peptide transporter family: NPF in plant specialized metabolism. Curr. Opin. Plant Biol. 2022, 68, 102243. [Google Scholar] [CrossRef] [PubMed]
  12. Taochy, C.; Gaillard, I.; Ipotesi, E.; Oomen, R.; Leonhardt, N.; Zimmermann, S.; Peltier, J.B.; Szponarski, W.; Simonneau, T.; Sentenac, H.; et al. The Arabidopsis root stele transporter NPF2.3 contributes to nitrate translocation to shoots under salt stress. Plant J. 2015, 83, 466–479. [Google Scholar] [CrossRef] [PubMed]
  13. Grunewald, S.; Marillonnet, S.; Hause, G.; Haferkamp, I.; Neuhaus, H.E.; Vess, A.; Hollemann, T.; Vogt, T. The Tapetal Major Facilitator NPF2.8 Is Required for Accumulation of Flavonol Glycosides on the Pollen Surface in Arabidopsis thaliana. Plant Cell 2020, 32, 1727–1748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Wei, Y.M.; Ren, Z.J.; Wang, B.H.; Zhang, L.; Zhao, Y.J.; Wu, J.W.; Li, L.G.; Zhang, X.S.; Zhao, X.Y. A nitrate transporter encoded by ZmNPF7.9 is essential for maize seed development. Plant Sci. 2021, 308, 110901. [Google Scholar] [CrossRef] [PubMed]
  15. Fan, S.C.; Lin, C.S.; Hsu, P.K.; Lin, S.H.; Tsay, Y.F. The Arabidopsis nitrate transporter NRT1.7, expressed in phloem, is responsible for source-to-sink remobilization of nitrate. Plant Cell 2009, 21, 2750–2761. [Google Scholar] [CrossRef] [Green Version]
  16. Tsay, Y.-F.; Chiu, C.-C.; Tsai, C.-B.; Ho, C.-H.; Hsu, P.-K. Nitrate transporters and peptide transporters. FEBS Lett. 2007, 581, 2290–2300. [Google Scholar] [CrossRef] [Green Version]
  17. Wang, Q.; Liu, C.; Dong, Q.; Huang, D.; Li, C.; Li, P.; Ma, F. Genome-Wide Identification and Analysis of Apple NITRATE TRANSPORTER 1/PEPTIDE TRANSPORTER Family (NPF) Genes Reveals MdNPF6.5 Confers High Capacity for Nitrogen Uptake under Low-Nitrogen Conditions. Int. J. Mol. Sci. 2018, 19, 2761. [Google Scholar] [CrossRef] [Green Version]
  18. Wang, J.; Li, Y.; Zhu, F.; Ming, R.; Chen, L.-Q. Genome-Wide Analysis of Nitrate Transporter (NRT/NPF) Family in Sugarcane Saccharum spontaneum L. Trop. Plant Biol. 2019, 12, 133–149. [Google Scholar] [CrossRef]
  19. Wen, J.; Li, P.F.; Ran, F.; Guo, P.C.; Zhu, J.T.; Yang, J.; Zhang, L.L.; Chen, P.; Li, J.N.; Du, H. Genome-wide characterization, expression analyses, and functional prediction of the NPF family in Brassica napus. BMC Genom. 2020, 21, 871. [Google Scholar] [CrossRef]
  20. Zhang, H.; Li, S.; Shi, M.; Wang, S.; Shi, L.; Xu, F.; Ding, G. Genome-Wide Systematic Characterization of the NPF Family Genes and Their Transcriptional Responses to Multiple Nutrient Stresses in Allotetraploid Rapeseed. Int. J. Mol. Sci. 2020, 21, 5947. [Google Scholar] [CrossRef]
  21. Chao, H.; He, J.; Cai, Q.; Zhao, W.; Fu, H.; Hua, Y.; Li, M.; Huang, J. The Expression Characteristics of NPF Genes and Their Response to Vernalization and Nitrogen Deficiency in Rapeseed. Int. J. Mol. Sci. 2021, 22, 4944. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, H.; Wan, Y.; Buchner, P.; King, R.; Ma, H.; Hawkesford, M.J. Phylogeny and gene expression of the complete NITRATE TRANSPORTER 1/PEPTIDE TRANSPORTER FAMILY in Triticum aestivum. J. Exp. Bot. 2020, 71, 4531–4546. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, X.; Cai, X.; Xu, C.; Wang, Q. Identification and characterization of the NPF, NRT2 and NRT3 in spinach. Plant Physiol. Biochem. 2021, 158, 297–307. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, Y.; Wei, K.; Ruan, L.; Bai, P.; Wu, L.; Wang, L.; Cheng, H. Systematic Investigation and Expression Profiles of the Nitrate Transporter 1/Peptide Transporter Family (NPF) in Tea Plant (Camellia sinensis). Int. J. Mol. Sci. 2022, 23, 6663. [Google Scholar] [CrossRef]
  25. Hsu, P.K.; Tsay, Y.F. Two phloem nitrate transporters, NRT1.11 and NRT1.12, are important for redistributing xylem-borne nitrate to enhance plant growth. Plant Physiol. 2013, 163, 844–856. [Google Scholar] [CrossRef] [Green Version]
  26. Jorgensen, M.E.; Xu, D.; Crocoll, C.; Ernst, H.A.; Ramirez, D.; Motawia, M.S.; Olsen, C.E.; Mirza, O.; Nour-Eldin, H.H.; Halkier, B.A. Origin and evolution of transporter substrate specificity within the NPF family. Elife 2017, 6, e19466. [Google Scholar] [CrossRef] [PubMed]
  27. Sugiura, M.; Georgescu, M.N.; Takahashi, M. A nitrite transporter associated with nitrite uptake by higher plant chloroplasts. Plant Cell Physiol. 2007, 48, 1022–1035. [Google Scholar] [CrossRef] [Green Version]
  28. Tal, I.; Zhang, Y.; Jorgensen, M.E.; Pisanty, O.; Barbosa, I.C.; Zourelidou, M.; Regnault, T.; Crocoll, C.; Olsen, C.E.; Weinstain, R.; et al. The Arabidopsis NPF3 protein is a GA transporter. Nat. Commun. 2016, 7, 11486. [Google Scholar] [CrossRef] [Green Version]
  29. Karim, S.; Holmstrom, K.O.; Mandal, A.; Dahl, P.; Hohmann, S.; Brader, G.; Palva, E.T.; Pirhonen, M. AtPTR3, a wound-induced peptide transporter needed for defence against virulent bacterial pathogens in Arabidopsis. Planta 2007, 225, 1431–1445. [Google Scholar] [CrossRef]
  30. Leran, S.; Edel, K.H.; Pervent, M.; Hashimoto, K.; Corratge-Faillie, C.; Offenborn, J.N.; Tillard, P.; Gojon, A.; Kudla, J.; Lacombe, B. Nitrate sensing and uptake in Arabidopsis are enhanced by ABI2, a phosphatase inactivated by the stress hormone abscisic acid. Sci. Signal. 2015, 8, ra43. [Google Scholar] [CrossRef]
  31. Okamoto, M.; Vidmar, J.J.; Glass, A.D. Regulation of NRT1 and NRT2 gene families of Arabidopsis thaliana: Responses to nitrate provision. Plant Cell Physiol. 2003, 44, 304–317. [Google Scholar] [CrossRef]
  32. Chiu, C.C.; Lin, C.S.; Hsia, A.P.; Su, R.C.; Lin, H.L.; Tsay, Y.F. Mutation of a nitrate transporter, AtNRT1:4, results in a reduced petiole nitrate content and altered leaf development. Plant Cell Physiol. 2004, 45, 1139–1148. [Google Scholar] [CrossRef]
  33. Li, J.Y.; Fu, Y.L.; Pike, S.M.; Bao, J.; Tian, W.; Zhang, Y.; Chen, C.Z.; Zhang, Y.; Li, H.M.; Huang, J.; et al. The Arabidopsis nitrate transporter NRT1.8 functions in nitrate removal from the xylem sap and mediates cadmium tolerance. Plant Cell 2010, 22, 1633–1646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Chen, C.Z.; Lv, X.F.; Li, J.Y.; Yi, H.Y.; Gong, J.M. Arabidopsis NRT1.5 is another essential component in the regulation of nitrate reallocation and stress tolerance. Plant Physiol. 2012, 159, 1582–1590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Babst, B.A.; Gao, F.; Acosta-Gamboa, L.M.; Karve, A.; Schueller, M.J.; Lorence, A. Three NPF genes in Arabidopsis are necessary for normal nitrogen cycling under low nitrogen stress. Plant Physiol. Biochem. 2019, 143, 1–10. [Google Scholar] [CrossRef]
  36. Komarova, N.Y.; Thor, K.; Gubler, A.; Meier, S.; Dietrich, D.; Weichert, A.; Suter Grotemeyer, M.; Tegeder, M.; Rentsch, D. AtPTR1 and AtPTR5 transport dipeptides in planta. Plant Physiol. 2008, 148, 856–869. [Google Scholar] [CrossRef] [PubMed]
  37. Choi, M.G.; Kim, E.J.; Song, J.Y.; Choi, S.B.; Cho, S.W.; Park, C.S.; Kang, C.S.; Park, Y.I. Peptide transporter2 (PTR2) enhances water uptake during early seed germination in Arabidopsis thaliana. Plant Mol. Biol. 2020, 102, 615–624. [Google Scholar] [CrossRef] [Green Version]
  38. He, F.; Karve, A.A.; Maslov, S.; Babst, B.A. Large-Scale Public Transcriptomic Data Mining Reveals a Tight Connection between the Transport of Nitrogen and Other Transport Processes in Arabidopsis. Front. Plant Sci. 2016, 7, 1207. [Google Scholar] [CrossRef] [Green Version]
  39. Gómez-Campo, C.; Prakash, S. 2 Origin and domestication. In Developments in Plant Genetics and Breeding; Gómez-Campo, C., Ed.; Elsevier: Amsterdam, The Netherlands, 1999; Volume 4, pp. 33–58. [Google Scholar]
  40. Dong, X.; Feng, H.; Xu, M.; Lee, J.; Kim, Y.K.; Lim, Y.P.; Piao, Z.; Park, Y.; Ma, H.; Hur, Y. Comprehensive Analysis of Genic Male Sterility-Related Genes in Brassica rapa Using a Newly Developed Br300K Oligomeric Chip. PLoS ONE 2013, 8, e72178. [Google Scholar] [CrossRef] [Green Version]
  41. Liu, C.; Liu, Z.; Li, C.; Zhang, Y.; Feng, H. Comparative transcriptome analysis of fertile and sterile buds from a genetically male sterile line of Chinese cabbage. Vitr. Cell. Dev. Biol. Plant 2016, 52, 130–139. [Google Scholar] [CrossRef]
  42. Zhou, X.; Liu, Z.; Ji, R.; Feng, H. Comparative transcript profiling of fertile and sterile flower buds from multiple-allele-inherited male sterility in Chinese cabbage (Brassica campestris L. ssp. pekinensis). Mol. Genet. Genom. 2017, 292, 967–990. [Google Scholar] [CrossRef] [PubMed]
  43. Shen, X.; Xu, L.; Liu, Y.; Dong, H.; Zhou, D.; Zhang, Y.; Lin, S.; Cao, J.; Huang, L. Comparative transcriptome analysis and ChIP-sequencing reveals stage-specific gene expression and regulation profiles associated with pollen wall formation in Brassica rapa. BMC Genom. 2019, 20, 264. [Google Scholar] [CrossRef]
  44. Huang, S.; Peng, S.; Liu, Z.; Li, C.; Tan, C.; Yao, R.; Li, D.; Li, X.; Hou, L.; Feng, H. Investigation of the genes associated with a male sterility mutant (msm) in Chinese cabbage (Brassica campestris ssp. pekinensis) using RNA-Seq. Mol. Genet. Genom. 2020, 295, 233–249. [Google Scholar] [CrossRef] [PubMed]
  45. Gupta, R.; Ting, J.T.; Sokolov, L.N.; Johnson, S.A.; Luan, S. A tumor suppressor homolog, AtPTEN1, is essential for pollen development in Arabidopsis. Plant Cell 2002, 14, 2495–2507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Huang, M.D.; Chen, T.L.; Huang, A.H. Abundant type III lipid transfer proteins in Arabidopsis tapetum are secreted to the locule and become a constituent of the pollen exine. Plant Physiol. 2013, 163, 1218–1229. [Google Scholar] [CrossRef] [Green Version]
  47. Elejalde-Palmett, C.; de Bernonville, T.D.; Glevarec, G.; Pichon, O.; Papon, N.; Courdavault, V.; St-Pierre, B.; Giglioli-Guivarc’h, N.; Lanoue, A.; Besseau, S. Characterization of a spermidine hydroxycinnamoyltransferase in Malus domestica highlights the evolutionary conservation of trihydroxycinnamoyl spermidines in pollen coat of core Eudicotyledons. J. Exp. Bot. 2015, 66, 7271–7285. [Google Scholar] [CrossRef] [Green Version]
  48. Gebert, M.; Dresselhaus, T.; Sprunck, S. F-actin organization and pollen tube tip growth in Arabidopsis are dependent on the gametophyte-specific Armadillo repeat protein ARO1. Plant Cell 2008, 20, 2798–2814. [Google Scholar] [CrossRef] [Green Version]
  49. Szumlanski, A.L.; Nielsen, E. The Rab GTPase RabA4d regulates pollen tube tip growth in Arabidopsis thaliana. Plant Cell 2009, 21, 526–544. [Google Scholar] [CrossRef] [Green Version]
  50. Zhao, X.Y.; Wang, Q.; Li, S.; Ge, F.R.; Zhou, L.Z.; McCormick, S.; Zhang, Y. The juxtamembrane and carboxy-terminal domains of Arabidopsis PRK2 are critical for ROP-induced growth in pollen tubes. J. Exp. Bot. 2013, 64, 5599–5610. [Google Scholar] [CrossRef] [Green Version]
  51. Dai, Y.; Zhang, S.; Sun, X.; Li, G.; Yuan, L.; Li, F.; Zhang, H.; Zhang, S.; Chen, G.; Wang, C.; et al. Comparative Transcriptome Analysis of Gene Expression and Regulatory Characteristics Associated with Different Vernalization Periods in Brassica rapa. Genes 2020, 11, 392. [Google Scholar] [CrossRef]
  52. Aigu, Y.; Daval, S.; Gazengel, K.; Marnet, N.; Lariagon, C.; Laperche, A.; Legeai, F.; Manzanares-Dauleux, M.J.; Gravot, A. Multi-Omic Investigation of Low-Nitrogen Conditional Resistance to Clubroot Reveals Brassica napus Genes Involved in Nitrate Assimilation. Front. Plant Sci. 2022, 13, 790563. [Google Scholar] [CrossRef]
  53. Corratge-Faillie, C.; Lacombe, B. Substrate (un)specificity of Arabidopsis NRT1/PTR FAMILY (NPF) proteins. J. Exp. Bot. 2017, 68, 3107–3113. [Google Scholar] [CrossRef]
  54. Xia, X.; Fan, X.; Wei, J.; Feng, H.; Qu, H.; Xie, D.; Miller, A.J.; Xu, G. Rice nitrate transporter OsNPF2.4 functions in low-affinity acquisition and long-distance transport. J. Exp. Bot. 2015, 66, 317–331. [Google Scholar] [CrossRef] [PubMed]
  55. Zhang, L.; Cai, X.; Wu, J.; Liu, M.; Grob, S.; Cheng, F.; Liang, J.; Cai, C.; Liu, Z.; Liu, B.; et al. Improved Brassica rapa reference genome by single-molecule sequencing and chromosome conformation capture technologies. Hortic. Res. 2018, 5, 50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Cai, X.; Chang, L.; Zhang, T.; Chen, H.; Zhang, L.; Lin, R.; Liang, J.; Wu, J.; Freeling, M.; Wang, X. Impacts of allopolyploidization and structural variation on intraspecific diversification in Brassica rapa. Genome Biol. 2021, 22, 166. [Google Scholar] [CrossRef]
  57. Xu, G.; Guo, C.; Shan, H.; Kong, H. Divergence of duplicate genes in exon–intron structure. Proc. Natl. Acad. Sci. USA 2012, 109, 1187–1192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Biłas, R.; Szafran, K.; Hnatuszko-Konka, K.; Kononowicz, A.K. Cis-regulatory elements used to control gene expression in plants. Plant Cell Tissue Organ Cult. 2016, 127, 269–287. [Google Scholar] [CrossRef] [Green Version]
  59. Vega, A.; O’Brien, J.A.; Gutierrez, R.A. Nitrate and hormonal signaling crosstalk for plant growth and development. Curr. Opin. Plant Biol. 2019, 52, 155–163. [Google Scholar] [CrossRef]
  60. David, L.C.; Berquin, P.; Kanno, Y.; Seo, M.; Daniel-Vedele, F.; Ferrario-Mery, S. N availability modulates the role of NPF3.1, a gibberellin transporter, in GA-mediated phenotypes in Arabidopsis. Planta 2016, 244, 1315–1328. [Google Scholar] [CrossRef]
  61. Shang, M.; Wang, X.; Zhang, J.; Qi, X.; Ping, A.; Hou, L.; Xing, G.; Li, G.; Li, M. Genetic Regulation of GA Metabolism during Vernalization, Floral Bud Initiation and Development in Pak Choi (Brassica rapa ssp. chinensis Makino). Front. Plant Sci. 2017, 8, 1533. [Google Scholar] [CrossRef]
  62. Chen, K.E.; Chen, H.Y.; Tseng, C.S.; Tsay, Y.F. Improving nitrogen use efficiency by manipulating nitrate remobilization in plants. Nat. Plants 2020, 6, 1126–1135. [Google Scholar] [CrossRef] [PubMed]
  63. Paupiere, M.J.; Muller, F.; Li, H.; Rieu, I.; Tikunov, Y.M.; Visser, R.G.F.; Bovy, A.G. Untargeted metabolomic analysis of tomato pollen development and heat stress response. Plant Reprod. 2017, 30, 81–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Hoagland, D.R.; Arnon, D.I. The water culture method for growing plants without soil. Calif. Agric. Exp. Stn. Circ. 1950, 347, 32. [Google Scholar]
  65. Cai, Y.; Qi, J.; Li, C.; Miao, K.; Jiang, B.; Yang, X.; Han, W.; Wang, Y.; Gao, J.; Dong, X. Genome-Wide Analysis of Purple Acid Phosphatase Genes in Brassica rapa and Their Association with Pollen Development and Phosphorus Deprivation Stress. Horticulturae 2021, 7, 363. [Google Scholar] [CrossRef]
  66. Cai, X.; Wu, J.; Liang, J.; Lin, R.; Zhang, K.; Cheng, F.; Wang, X. Improved Brassica oleracea JZS assembly reveals significant changing of LTR-RT dynamics in different morphotypes. Theor Appl Genet 2020, 133, 3187–3199. [Google Scholar] [CrossRef] [PubMed]
  67. Perumal, S.; Koh, C.S.; Jin, L.; Buchwaldt, M.; Higgins, E.E.; Zheng, C.; Sankoff, D.; Robinson, S.J.; Kagale, S.; Navabi, Z.K.; et al. A high-contiguity Brassica nigra genome localizes active centromeres and defines the ancestral Brassica genome. Nat Plants 2020, 6, 929–941. [Google Scholar] [CrossRef]
  68. Edgar, R.C. MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004, 32, 1792–1797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. 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] [Green Version]
  70. Wilkins, M.R.; Gasteiger, E.; Bairoch, A.; Sanchez, J.C.; Williams, K.L.; Appel, R.D.; Hochstrasser, D.F. Protein identification and analysis tools in the ExPASy server. Methods Mol. Biol. 1999, 112, 531–552. [Google Scholar] [CrossRef]
  71. Bailey, T.L.; Boden, M.; Buske, F.A.; Frith, M.; Grant, C.E.; Clementi, L.; Ren, J.; Li, W.W.; Noble, W.S. MEME SUITE: Tools for motif discovery and searching. Nucleic Acids Res. 2009, 37, W202–W208. [Google Scholar] [CrossRef]
  72. Hu, B.; Jin, J.; Guo, A.Y.; Zhang, H.; Luo, J.; Gao, G. GSDS 2.0: An upgraded gene feature visualization server. Bioinformatics 2015, 31, 1296–1297. [Google Scholar] [CrossRef]
  73. Eastin, E.F. Total nitrogen determination for plant material containing nitrate. Anal. Biochem. 1978, 85, 591–594. [Google Scholar] [CrossRef]
  74. Pertea, M.; Kim, D.; Pertea, G.M.; Leek, J.T.; Salzberg, S.L. Transcript-level expression analysis of RNA-seq experiments with HISAT, StringTie and Ballgown. Nat. Protoc. 2016, 11, 1650–1667. [Google Scholar] [CrossRef]
  75. Wang, L.; Feng, Z.; Wang, X.; Wang, X.; Zhang, X. DEGseq: An R package for identifying differentially expressed genes from RNA-seq data. Bioinformatics 2010, 26, 136–138. [Google Scholar] [CrossRef] [Green Version]
  76. Tong, C.; Wang, X.; Yu, J.; Wu, J.; Li, W.; Huang, J.; Dong, C.; Hua, W.; Liu, S. Comprehensive analysis of RNA-seq data reveals the complexity of the transcriptome in Brassica rapa. BMC Genom. 2013, 14, 689. [Google Scholar] [CrossRef] [Green Version]
  77. Guo, X.; Liang, J.; Lin, R.; Zhang, L.; Wu, J.; Wang, X. Series-Spatial Transcriptome Profiling of Leafy Head Reveals the Key Transition Leaves for Head Formation in Chinese Cabbage. Front. Plant Sci. 2021, 12, 787826. [Google Scholar] [CrossRef]
  78. Huang, L.; Dong, H.; Zhou, D.; Li, M.; Liu, Y.; Zhang, F.; Feng, Y.; Yu, D.; Lin, S.; Cao, J. Systematic identification of long non-coding RNAs during pollen development and fertilization in Brassica rapa. Plant J. 2018, 96, 203–222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Gao, P.; Quilichini, T.D.; Yang, H.; Li, Q.; Nilsen, K.T.; Qin, L.; Babic, V.; Liu, L.; Cram, D.; Pasha, A.; et al. Evolutionary divergence in embryo and seed coat development of U’s Triangle Brassica species illustrated by a spatiotemporal transcriptome atlas. New Phytol. 2022, 233, 30–51. [Google Scholar] [CrossRef]
  80. 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] [PubMed]
  81. Wei, X.; Zhang, Y.; Zhao, Y.; Xie, Z.; Hossain, M.R.; Yang, S.; Shi, G.; Lv, Y.; Wang, Z.; Tian, B.; et al. Root Transcriptome and Metabolome Profiling Reveal Key Phytohormone-Related Genes and Pathways Involved Clubroot Resistance in Brassica rapa L. Front. Plant Sci. 2021, 12, 759623. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Phylogenetic reconstruction of 345 NPF proteins identified in the genomes of Brassica rapa (n = 85), B. oleracea (110), B. nigra (97), and Arabidopsis (53). The neighbor-joining phylogenetic tree was generated in MEGA6 with full-length NPF protein sequences, and branch support was evaluated with 1000 bootstrap replicates.
Figure 1. Phylogenetic reconstruction of 345 NPF proteins identified in the genomes of Brassica rapa (n = 85), B. oleracea (110), B. nigra (97), and Arabidopsis (53). The neighbor-joining phylogenetic tree was generated in MEGA6 with full-length NPF protein sequences, and branch support was evaluated with 1000 bootstrap replicates.
Ijms 24 00754 g001
Figure 2. Phylogenetic tree (A), gene structures (B), and MEME motifs (C) for 85 NPFs identified in B. rapa. (B) Yellow boxes represent exons, and black lines indicate introns. (C) The colored boxes indicate motifs, as shown by the legend on the right.
Figure 2. Phylogenetic tree (A), gene structures (B), and MEME motifs (C) for 85 NPFs identified in B. rapa. (B) Yellow boxes represent exons, and black lines indicate introns. (C) The colored boxes indicate motifs, as shown by the legend on the right.
Ijms 24 00754 g002
Figure 3. Syntenic relationships among NPF genes of Brassica rapa, B. nigra, B. oleracea, and Arabidopsis. The chromosomes of B. rapa (chromosomes A01 to A10), B. nigra (B1 to B8), B. oleracea (C01 to C09), and Arabidopsis (Chr1 to Chr5) are shown in red, orange, green, and brown, respectively. Collinear gene pairs were mapped onto chromosomes and are connected to each other. The collinear gene pairs from B. rapa and B. oleracea, B. rapa and B. nigra, B. rapa and Arabidopsis, B. nigra and B. oleracea, B. nigra and Arabidopsis, B. oleracea and Arabidopsis are connected by light blue, green, pink, red, purple, and orange lines, respectively. Segmental duplication genes in B. oleracea are connected by dark blue lines. The locations of tandem duplicated genes are indicated with black lines. Only the names of collinear gene pairs are shown.
Figure 3. Syntenic relationships among NPF genes of Brassica rapa, B. nigra, B. oleracea, and Arabidopsis. The chromosomes of B. rapa (chromosomes A01 to A10), B. nigra (B1 to B8), B. oleracea (C01 to C09), and Arabidopsis (Chr1 to Chr5) are shown in red, orange, green, and brown, respectively. Collinear gene pairs were mapped onto chromosomes and are connected to each other. The collinear gene pairs from B. rapa and B. oleracea, B. rapa and B. nigra, B. rapa and Arabidopsis, B. nigra and B. oleracea, B. nigra and Arabidopsis, B. oleracea and Arabidopsis are connected by light blue, green, pink, red, purple, and orange lines, respectively. Segmental duplication genes in B. oleracea are connected by dark blue lines. The locations of tandem duplicated genes are indicated with black lines. Only the names of collinear gene pairs are shown.
Ijms 24 00754 g003
Figure 4. Expression of BrNPFs in different tissues and organs of Brassica rapa. Expression data were subjected to log2(TPM+1) normalization. Comparisons of stem leaf, stem, and root tissues were conducted by comparison against seven-week-old Chiifu (B. rapa cv. Chiifu) plants. Flower tissue was also generated from blooming plants without floral shoots, while silique tissues were obtained 15 days after pollination. Further, 24 samples from the heading leaves of eleven-week-old plants were collected. All leaves from the heading leaf samples were divided into eleven whorls extending from the inside to the outside. Leaves with a length < 2 cm and shoot apical meristem (SAM) are identified as SAMs. Leaves from whorls one, two, three, five, seven, and nine were identified as L1, L2, L3, L5, L7, and L9, respectively. L2 samples were divided into leaf petiole (L2R2) and leaf blade (L2R1) samples, while L3, L5, L7, and L9 were divided into regions including the top region (R1), outer margin region (R2), the middle region of the blade (R3), top region of the petiole (R4), and middle region of the petiole (R5). FS1-5 and SS1-5 indicate the floral buds from the ‘Bcajh97-01A/B’ GMS A/B line of B. rapa, representing the pollen mother cells, tetrad, uninucleate pollen, binucleate pollen, and mature pollen stages, respectively. 1 HAP, 3 HAP, and 10 HAP indicate pistils at 1, 3, and 10 h after pollination in the fertile line, respectively.
Figure 4. Expression of BrNPFs in different tissues and organs of Brassica rapa. Expression data were subjected to log2(TPM+1) normalization. Comparisons of stem leaf, stem, and root tissues were conducted by comparison against seven-week-old Chiifu (B. rapa cv. Chiifu) plants. Flower tissue was also generated from blooming plants without floral shoots, while silique tissues were obtained 15 days after pollination. Further, 24 samples from the heading leaves of eleven-week-old plants were collected. All leaves from the heading leaf samples were divided into eleven whorls extending from the inside to the outside. Leaves with a length < 2 cm and shoot apical meristem (SAM) are identified as SAMs. Leaves from whorls one, two, three, five, seven, and nine were identified as L1, L2, L3, L5, L7, and L9, respectively. L2 samples were divided into leaf petiole (L2R2) and leaf blade (L2R1) samples, while L3, L5, L7, and L9 were divided into regions including the top region (R1), outer margin region (R2), the middle region of the blade (R3), top region of the petiole (R4), and middle region of the petiole (R5). FS1-5 and SS1-5 indicate the floral buds from the ‘Bcajh97-01A/B’ GMS A/B line of B. rapa, representing the pollen mother cells, tetrad, uninucleate pollen, binucleate pollen, and mature pollen stages, respectively. 1 HAP, 3 HAP, and 10 HAP indicate pistils at 1, 3, and 10 h after pollination in the fertile line, respectively.
Ijms 24 00754 g004
Figure 5. Confirmation and analysis of BrNPFs related to pollen development. (A), The phenotype of male Brassica rapa sterile lines. (B), Semi-quantitative RT-PCR analysis of BrNPFs related to pollen development within various tissues. (C), GO enrichment analysis of co-expression of BrNPFs. (D), Network of pollen development-related genes based on co-expression analysis of BrNPF8.9 and BrNPF2.6. (E), Network of pollen tube development-related genes based on co-expression analysis of BrNPF8.9 and BrNPF7.6.
Figure 5. Confirmation and analysis of BrNPFs related to pollen development. (A), The phenotype of male Brassica rapa sterile lines. (B), Semi-quantitative RT-PCR analysis of BrNPFs related to pollen development within various tissues. (C), GO enrichment analysis of co-expression of BrNPFs. (D), Network of pollen development-related genes based on co-expression analysis of BrNPF8.9 and BrNPF2.6. (E), Network of pollen tube development-related genes based on co-expression analysis of BrNPF8.9 and BrNPF7.6.
Ijms 24 00754 g005
Figure 6. Expression patterns of BrNPFs responsive to vernalization (A) and P. brassicae infection stress (B). DAT, day after treatment. DAI, the day after inoculation.
Figure 6. Expression patterns of BrNPFs responsive to vernalization (A) and P. brassicae infection stress (B). DAT, day after treatment. DAI, the day after inoculation.
Ijms 24 00754 g006
Figure 7. Phenotypic characterization and transcriptome profiling of BrNPFs under low nitrate availability stress. (A), Brassica rapa phenotype after low nitrate availability treatment. (BF), Comparisons of plant height, leaf area, root length, fresh weight, and total nitrate concentrations between normal treatment and low nitrate-treated B. rapa seedlings. (G), Heatmap visualization of BrNFPs differential expression due to low nitrate stress.
Figure 7. Phenotypic characterization and transcriptome profiling of BrNPFs under low nitrate availability stress. (A), Brassica rapa phenotype after low nitrate availability treatment. (BF), Comparisons of plant height, leaf area, root length, fresh weight, and total nitrate concentrations between normal treatment and low nitrate-treated B. rapa seedlings. (G), Heatmap visualization of BrNFPs differential expression due to low nitrate stress.
Ijms 24 00754 g007
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

Yang, X.; Han, W.; Qi, J.; Li, Y.; Chen, X.; Zhang, Y.; Wu, J.; Li, G.; Gao, J.; Dong, X. Identification of NPF Family Genes in Brassica rapa Reveal Their Potential Functions in Pollen Development and Response to Low Nitrate Stress. Int. J. Mol. Sci. 2023, 24, 754. https://doi.org/10.3390/ijms24010754

AMA Style

Yang X, Han W, Qi J, Li Y, Chen X, Zhang Y, Wu J, Li G, Gao J, Dong X. Identification of NPF Family Genes in Brassica rapa Reveal Their Potential Functions in Pollen Development and Response to Low Nitrate Stress. International Journal of Molecular Sciences. 2023; 24(1):754. https://doi.org/10.3390/ijms24010754

Chicago/Turabian Style

Yang, Xiaoshuang, Wenyu Han, Jiao Qi, Yueying Li, Xingbo Chen, Yiwen Zhang, Jingyu Wu, Genze Li, Jing Gao, and Xiangshu Dong. 2023. "Identification of NPF Family Genes in Brassica rapa Reveal Their Potential Functions in Pollen Development and Response to Low Nitrate Stress" International Journal of Molecular Sciences 24, no. 1: 754. https://doi.org/10.3390/ijms24010754

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