Next Article in Journal
miRNA Expression Profiles in Ovarian Endometriosis and Two Types of Ovarian Cancer—Endometriosis-Associated Ovarian Cancer and High-Grade Ovarian Cancer
Previous Article in Journal
Inflammatory and Prothrombotic Biomarkers Contribute to the Persistence of Sequelae in Recovered COVID-19 Patients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Maize Dek407 Encodes the Nitrate Transporter 1.5 and Is Required for Kernel Development

1
National Engineering Laboratory for Crop Molecular Breeding, Institute of Crop Sciences, Chinese Academy of Agricultural Sciences, Beijing 100081, China
2
National Key Laboratory of Wheat and Maize Crop Science, College of Agronomy, Henan Agricultural University, Zhengzhou 450002, China
3
The Shennong Laboratory, Zhengzhou 450002, China
4
Chongqing Academy of Agricultural Sciences, Chongqing 401329, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(24), 17471; https://doi.org/10.3390/ijms242417471
Submission received: 19 October 2023 / Revised: 30 November 2023 / Accepted: 11 December 2023 / Published: 14 December 2023
(This article belongs to the Section Molecular Plant Sciences)

Abstract

:
The kernel serves as the storage organ and harvestable component of maize, and it plays a crucial role in determining crop yield and quality. Understanding the molecular and genetic mechanisms of kernel development is of considerable importance for maize production. In this study, we obtained a mutant, which we designated defective kernel 407 (dek407), through ethyl methanesulfonate mutagenesis. The dek407 mutant exhibited reduced kernel size and kernel weight, as well as delayed grain filling compared with those of the wild type. Positional cloning and an allelism test revealed that Dek407 encodes a nitrate transporter 1/peptide transporter family (NPF) protein and is the allele of miniature 2 (mn2) that was responsible for a poorly filled defective kernel phenotype. A transcriptome analysis of the developing kernels showed that the mutation of Dek407 altered the expression of phytohormone-related genes, especially those genes associated with indole-3-acetic acid synthesis and signaling. Phytohormone measurements and analysis indicated that the endogenous indole-3-acetic acid content was significantly reduced by 66% in the dek407 kernels, which may be the primary cause of the defective phenotype. We further demonstrated that natural variation in Dek407 is associated with kernel weight and kernel size. Therefore, Dek407 is a potential target gene for improvement of maize yield.

1. Introduction

Maize is among the most important food crops worldwide, as well as an important feed source and bioenergy crop, thereby playing a pivotal role in addressing global food insecurity and facilitating the production of clean energy [1]. Progress on these issues can be effectively achieved through the improvement of maize yield. Maize yield is a complex trait that encompasses multiple secondary traits, but it is primarily determined by the number of ears per unit area, the number of kernels per ear, and the weight of the kernels. Maize kernel development directly affects the kernel weight. Thus, the identification and cloning of key regulatory genes involved in maize kernel development is instrumental in the genetic enhancement of maize yield.
To date, innumerable maize mutants that are defective in kernel development have been identified. These mutants can be categorized into a variety of kernel mutant types based on their mutant phenotypes. For example, the defective kernel (dek) mutants are characterized by impaired development of both the embryo and endosperm [2]. The empty pericarp (emp) mutants exhibit either an empty pericarp or a papery kernel [3]. The embryo-specific (emb) mutants display morphogenic effects that are specific to the embryo [4]. Numerous genes involved in maize kernel development have been identified through the characterization of these mutants. For example, the genes Dek605, Ppr2263, Smk1, PCW1, Emp8, and Emp10 encode PLS- or P-type pentatricopeptide repeat (PPR) proteins that regulate maize embryo and endosperm development by participating in mitochondrial RNA intron splicing or site-specific editing [5,6,7,8,9,10]. Smk2 encodes a glutaminase that regulates maize seed development by influencing the biosynthesis of vitamin B6 [11]. Mn1 and Meg1 encode a cell wall sucrose invertase and cysteine-rich polypeptide, respectively, and their mutation affects the development of cells in the basal endosperm transfer layer (BETL) [12,13]. Pro1 encodes a carboxylate synthase that influences kernel development by regulating proline synthesis and GCN2 (general control nonderepressible 2) kinase activity [14]. ZmUrb2, ZmShrek1, Reas1, and RCL1 encode diverse types of ribosome biosynthesis factors that participate in ribosome biogenesis during kernel development [15,16,17,18].
Nitrogen is a vital element for plant growth, metabolism, and heredity, and its nutritional status directly impacts crop yield. Nitrate is an abundant source of nitrogen in the soil and the primary form of nitrogen absorbed by plants [19,20]. The nitrate transporters (NRTs) play extensive, crucial roles in the uptake of nitrate from the soil and its subsequent translocation to various plant organs, thereby facilitating plant growth and development. In the past 30 years, four families of transporters involved in nitrate transport have been identified in plants, comprising the nitrate transporter 1 (NRT1)/peptide transporter (PTR) family (NPF), NRT2 family, chloride channel (CLC) family, and slowly activating anion channel (SLAC) family [21,22]. Their functions have been extensively described elsewhere [22,23].
The NPF gene family of Arabidopsis thaliana and rice (Oryza sativa) comprises 53 and 93 gene members, respectively, whereas in maize, the NPF gene family consists of 78 genes [23,24,25]. In higher plants, all members of the NPF protein family have 12 conserved transmembrane domains (TMs), with a large hydrophilic loop located between TM6 and TM7 [26]. The NPF family proteins, like AtNPF2.3, AtNPF6.3, AtNPF7.2, OsNPF2.2, and OsNPF2.4, were initially identified as nitrate or peptide transporters [27,28,29,30,31]. However, in recent years, more and more NPF proteins have been shown to transport a variety of substrates, ions including Cl and K+, amino acids, sucrose, glucose, and hormones including indole-3-butyric acid (IBA), indole-3-acetic acid (IAA), abscisic acid (ABA), jasmonates (JAs), and gibberellins (GAs) [32,33,34,35]. In Arabidopsis, AtNPF2.4 and AtNPF2.5 have been reported to exhibit chloride transport capacity [36,37]. Recent research revealed that the AtNPF7.3 transporter exhibits not only nitrate and K+ transport activities, but also the capability to transport IBA, which serves as a precursor in the synthesis of IAA [34,38]. The AtNPF7.3/NRT1.5 transporter may regulate root geotropism by mediating IBA uptake in root cells and maintaining an optimal auxin gradient within the root [34]. The ability of AtNPF3.1 and AtNPF4.6 to transport ABA was confirmed using a functional expression system in Xenopus laevis [39,40]. The AtNPF3 transporter is capable of efficiently transporting GA in vitro and in vivo, thereby affecting GA accumulation in the root endoderm [39]. Recent study in maize found that a NRT1/PTR-type transporter ZmSUGCAR1, the homologous protein of AtNPF7.3, also exhibits the additional capacity to transport sucrose, glucose, and K+, but not IAA [35]. The diversity of transport substrates of NPF transporters indicates that NPF transporters may perform extensive functions in plant growth and development [32].
To date, the mechanism by which NPF proteins regulate maize kernel development remains largely unknown. In the present study, we obtained the defective kernel 407 (dek407) mutant through ethyl methanesulfonate (EMS) mutagenesis. Using bulked segregant RNA-sequencing (BSR-seq) analysis, positional cloning, and an allelism test, we identified Dek407 as a variant of Mn2/ZmNPF7.9/ZmSUGCAR1 that was responsible for the poorly filled defective kernel phenotype and proposed to be a sucrose transporter [35,41,42]. Dek407 encodes an NPF protein with 12 conserved TMs. Our further analysis in the dek407 mutant showed that the G/A mutation in the third TM altered the expression of phytohormone-related genes and the phytohormone balance, especially the expression of IAA-related genes and the content of IAA. Finally, we showed that natural variation in Dek407 is associated with kernel weight and kernel size. This study provides a foundation for further elucidating the function of NPF protein involved in maize kernel development and hormone transport.

2. Results

2.1. Phenotypic and Genetic Analysis of the dek407 Mutant

To elucidate the molecular regulatory mechanisms of maize kernel development, we screened an EMS-mutagenized library in the maize B73 genetic background for mutants with kernel development defects [43]. One mutant, which we designated defective kernel 407 (dek407), showed severely defective kernel development. The mutant seeds were easily identified as having a reduced kernel size and delayed grain filling in the heterozygous ears at 12 days after pollination (DAP) (Figure 1A). The dek407 mature kernels exhibited a small, wrinkled kernel phenotype together with an empty pericarp at the top (Figure 1B,C and Figure S1D). Compared with wild-type (WT) kernels, the kernel length, width, thickness, and hundred-kernel weight of the dek407 mutant were significantly reduced by 6%, 21%, 36%, and 56%, respectively (Figure 1C and Figure S1A–C). When observed on a light box, the dek407 kernels exhibited greater light transmittance than the WT kernels (Figure 1F). Longitudinal sections of dek407 mature kernels revealed a much less relative floury endosperm area (Figure S1D,E). Paraffin-embedded sections of the developing kernels showed that the embryo in the mutant was reduced in size compared to the WT (Figure 1G), along with a small embryo in the mature dek407 kernels (Figure S1D). The dek407 seedlings showed a decrease in plant height, root length, and root number at 10 days after sowing (DAS) (Figure 1D). In addition, the starch content in the dek407 mutant kernels was 24% lower than that in the WT kernels (Figure S1E).
To further analyze the phenotype and segregation of dek407 in different genetic backgrounds, the dek407 homozygote was crossed with the Zheng58 and C01 inbred lines to generate two F2 populations. The mutant kernels in the two populations, dek407-1 (in the Zheng58 background) and dek407-2 (in the C01 background), also exhibited a small, wrinkled kernel phenotype, together with an empty pericarp at the top (Figure 2A–C,E–G). The kernel width, thickness, and hundred-kernel weight of dek407-1 were reduced by 20%, 33%, and 49%, respectively, compared with those of the WT, whereas the kernel length showed no significant difference (Figure S2A–E). In dek407-2, the kernel length, width, thickness, and hundred-kernel weight were reduced by 7%, 11%, 25%, and 45%, respectively, compared with those of the WT (Figure S2F–J). At 10 DAS, the seedling development of dek407-1 and dek407-2 was delayed compared with the WT seedlings (Figure 2D,H). The segregation ratio of WT kernels to mutant kernels in each of the two F2 populations was consistent with a 3:1 Mendelian segregation ratio (Table S1), suggesting that dek407 is a recessive mutation under monogenic control. All of the aforementioned results indicated that kernel development and seed size were affected in the dek407 mutant.

2.2. Positional Cloning and Identification of the Dek407 Gene

To preliminarily determine the chromosomal location of Dek407, we conducted bulked segregant RNA-sequencing (BSR-seq) using kernels from the homozygous dek407 mutant and WT kernels from segregated F2 ears at 15 DAP. The endosperm RNA isolated from 30 WT and 30 mutant seeds was mixed equally to generate two pools of extreme bulked segregants for BSR-seq analysis. We sequenced the two extreme bulked pools on an Illumina platform using 150 bp paired-end reads to obtained the polymorphic SNP markers. The genotype frequencies (SNP-index) of each polymorphic SNP marker were calculated after removing the non-polymorphic and low-quality marker. A 32.52 Mb region (22.12 Mb to 54.64 Mb) on chromosome 7 may be the causal locus for dek407 based on the occurrence of significant linkage disequilibrium (Figure 3A).
To further confirm and narrow the candidate interval, two insertion/deletion (InDel) markers (S12191 and S12341) and two simple sequence repeat (SSR) markers (umc1401 and umc2327) were developed to identify the genotype of the dek407-2 F2 population (145 mutant kernels). The Dek407 gene was mapped to a 5.84 Mb region containing 95 annotated genes based on the B73-REFERENCE-GRAMENE-4.0 genome assembly (https://www.maizegdb.org/genome/assembly/Zm-B73-REFERENCE-GRAMENE-4.0) (accessed on 1 January 2022) (Figure 3B). Among the 95 annotated genes, we observed that Zm00001d019294 encodes the previously reported nitrate transporter 1/peptide transporter family protein, which is responsible for the small kernel phenotype of mn2 [41]. We sequenced the full-length nucleotide sequence of Zm00001d019294 in the homozygous WT and mutant, which revealed the presence of a non-synonymous mutation (G/A) in the third exon (+727 bp from ATG) in dek407, leading to an alteration of the corresponding coding amino acid from cysteine (Cys) to tyrosine (Tyr) (Figure 3C). We also analyzed the specificity of this mutation site in other maize inbred lines and determined that only dek407 contained this mutation site (Figure 3D). These results suggested that the G/A mutation in the third exon of Zm00001d019294 may be responsible for the mutant phenotype.
To validate Zm00001d019294 as the allele of dek407, we crossed the heterozygous plant mn2-m1 (+/−) with dek407 (+/−) for an allelism test. The mn2-m1 mutant had a G/− deletion in the fifth exon of Zm00001d019294, resulting in premature termination of protein translation (Figure 3C) [41]. The allelism test in two separate ears revealed a 3:1 segregation ratio of WT to mutant kernel phenotypes, with 269 (or 146) WT and 89 (or 41) mutant kernels (Figure 4A,C and Table S2). Mutant kernels from the two ears were randomly selected for genotyping to verify the allelism test results (Figure 4B,D). These results confirmed that Zm00001d019294 was indeed the Dek407 gene.

2.3. Dek407 Encodes a Highly Conserved Nitrate Transporter Protein and Is Predominantly Expressed in Developing Seeds

Zm00001d019294 was previously annotated as ZmNRT1.5/ZmNPF7.9 of the NPF transporter family [41,42]. A sequence analysis and protein domain prediction of Zm00001d019294 based on TOPCONS (https://topcons.net) (accessed on 1 March 2022) and SMART (http://smart.embl-heidelberg.de) (accessed on 1 March 2022) revealed that it contains a 1923 bp open reading frame (ORF) that encodes a typical transmembrane protein containing 12 transmembrane domains (TMs) with a major facilitator superfamily (MFS) domain in the C terminal (Figure 5A and Figure S3). A phylogenetic tree was constructed based on the full-length DEK407 amino acid sequence and its homologs in 10 representative plant species. The homologs were resolved into monocotyledonous and dicotyledonous evolutionary groups (Figure 5B). DEK407 was most closely related to Sobic002G091800 from Sorghum bicolor, and it also exhibited a high amino acid sequence similarity with OsNPF7.9 and OsPTR2 of rice and with AtNPF7.3 and AtNPF7.2 from Arabidopsis. In maize, Zm00001d017666 and Zm00001d051637 shared a high sequence similarity with DEK407, but their most closely related homologs in rice were OsNPF7.9 and OsPTR2, respectively (Figure 5B). Detailed DEK407 amino acid sequence alignment with Zm00001d017666, Zm00001d051637, OsNPF7.9, OsPTR2, AtNPF7.3, AtNPF7.2, and Sobic002G091800 indicated that all of these homologs contained 12 conserved TMs, and the mutation site Cys/Tyr (C/Y) in dek407 was located in the TM3 region (Figure S3).
The N-terminus of the DEK407 protein contained an NRT/PTR domain typical of the NPF family. Thus, we further analyzed the evolutionary relationships of 30 and 50 well-known NPF family proteins in rice and Arabidopsis, as well as their homologs in maize. The phylogenetic tree revealed that these NPF family proteins in maize were grouped into seven NPF subfamilies based on their homology with proteins in rice and Arabidopsis (Figure S4). DEK407 was classified into the NPF7 subfamily and belonged to a relatively distinct subbranch compared with other NPF7 members. Zm00001d017666 and Zm00001d051637 were classified to the same subbranch with OsNPF7.9 and OsPTR2, respectively (Figure S4).
As Zm00001d017666 and Zm00001d051637 shared a high sequence similarity with DEK407, we analyzed their spatiotemporal expression patterns based on the datasets accessible from the qTeller platform (https://qteller.maizegdb.org) (accessed on 1 March 2022) [44]. The results revealed that Dek407 was predominantly expressed in developing seeds (Figure 5C). Zm00001d017666 exhibited high expression levels in the internode, leaf, and taproot, but it displayed minimal expression in the seed (Figure S5A). Zm00001d051637 was expressed specifically in the embryo (Figure S5B). Thus, these three homologs in maize potentially perform similar functions in different tissues owing to their tissue-specific expression.

2.4. Comparative Transcriptome Analysis of Developing Kernels of the dek407 Mutant and WT

To investigate the impact of dek407 on kernel development at the whole-genome transcriptional level, we conducted a transcriptome analysis using total RNA isolated from WT and dek407 mutant kernels at 15 DAP, with three biological replicates. Total transcripts of 25,769 and 27,078 genes were detected in the WT and dek407 mutant, respectively, with fragments per kilobase of transcript per million mapped reads values higher than 0.1. Based on the criteria of a fold change (FC) > 2, fragments per kilobase million (FPKM) > 1, and false discovery rate (FDR) < 0.05, 3265 genes were identified as differentially expressed genes (DEGs). Among these genes, 2335 genes (72% of the 3265 DEGs) exhibited significantly increased transcript abundance in the dek407 mutant, and 930 genes (28% of the 3265 DEGs) showed significantly decreased transcript abundance (Figure 6A).
To analyze the functional annotations and classifications of the DEGs, we performed a gene ontology (GO) enrichment analysis using all the down- and up-regulated DEGs. The DEGs were categorized into the three main GO categories, namely biological processes (BPs), molecular functions (MFs), and cellular components (CCs) (Figure 6B). The most highly enriched GO terms classified in the BP category were involved in transmembrane transport (GO:0055085, p = 1.20 × 10−13), ion transport (GO:0006811, p = 2.70 × 10−7), nitrogen compound transport (GO:0071705, p = 5.00 × 10−6), oxidation–reduction process (GO:0055114, p = 7.00 × 10−8), and response to hormone (GO:0009725, p = 4.70 × 10−6). The GO terms enriched in the MF category mainly included transmembrane transporter activity (GO:0022857, p = 3.00 × 10−13), ion binding (GO:0043167, p = 1.10 × 10−7), oxidoreductase activity (GO:0016491, p = 1.60 × 10−8), and hydrolase activity, acting on glycosyl bonds (GO:0016798, p = 9.20 × 10−10). The enriched CC category mainly included the plasma membrane (GO:0005886, p = 3.30 × 10−11), membrane part (GO:0044425, p = 5.30 × 10−5), cell periphery (GO:0071944, p = 1.20 × 10−14), and extracellular region (GO:0005576, p = 1.20 × 10−8) (Figure 6B).
We also performed a Kyoto Encyclopedia of Genes and Genomes (KEGG) enrichment analysis to further explore the biological functions of the DEGs. Pathways with p < 0.05 were identified as significantly enriched pathways. The KEGG pathway analysis indicated that phenylpropanoid biosynthesis (zma00940, p = 1.62 × 10−9), plant hormone signal transduction (zma04075, p = 4.70 × 10−5), and starch and sucrose metabolism (zma00500, p = 7.80 × 10−5) were significantly enriched in the dek407 mutant (Figure 6C). These above findings revealed that the expression of genes involved in transmembrane transport, the oxidation–reduction process, ion binding, plant hormone signal transduction, and starch and sucrose metabolism were significantly altered in the dek407 mutant. Interestingly, the significantly enriched pathway involved in plant hormone signal transduction was not reported in zmsugcar1-1 [35].

2.5. Mutation of Dek407 Alters the Expression of Genes Associated with Phytohormones

Phytohormones, such as auxin, ethylene, ABA, and brassinosteroid (BR), play crucial roles in the development and maturation of plant seeds [45,46,47]. In the present study, the DEGs associated with hormone responses were identified as a significantly enriched GO subcategory (Figure 6B). In the KEGG enrichment analysis, the plant hormone signal transduction pathway was also identified as a significantly enriched pathway (Figure 6C). Therefore, we analyzed the expression of phytohormone-related DEGs in the dek407 mutant. Of the 34 DEGs associated with IAA synthesis and signaling in the dek407 mutant, seven DEGs encoding auxin response factor (ARF) family proteins and seven DEGs encoding auxin-responsive proteins were significantly up-regulated in the dek407 mutant. Additionally, four DEGs encoding polar auxin transport (PIN) proteins were significantly up-regulated in the dek407 mutant (except PIN12), and twelve DEGs encoding small auxin-up RNA (SAUR) family proteins were significantly down-regulated in the dek407 mutant (except SAUR56) (Figure 7A and Table S3). All seven DEGs associated with ethylene synthesis and signaling, as well as four DEGs associated with BR synthesis and signaling, were significantly up-regulated in the dek407 mutant (Figure 7B,C and Table S3). Three DEGs were involved in ABA signal transduction, comprising one up-regulated and two down-regulated genes (Figure 7D and Table S3). Altered expression of these genes may disturb hormone synthesis and signal transduction in the dek407 mutant, leading to defects in kernel development.

2.6. The Phytohormone Balance Was Altered in the dek407 Mutant

Auxin homeostasis has been reported to play indispensable roles in cereal endosperm development [48]. As the expression of genes associated with phytohormones was altered in the dek407 mutant, we measured the contents of endogenous phytohormones in WT and dek407 kernels using high-performance liquid chromatography–mass spectrometry (HPLC-MS). The results showed that the endogenous IAA content was significantly reduced by 66% in the dek407 mutant (Figure 8A). The endogenous ABA and salicylic acid (SA) contents in the dek407 mutant exhibited 1.32- and 2.14-fold increases, respectively, compared with those of the WT (Figure 8B,C). The contents of two important cytokinins, trans-zeatin (TZ) and trans-zeatin-riboside (TZR), in the dek407 mutant exhibited 1.43- and 4.13-fold increases, respectively, compared with the WT (Figure 8D,E). The content of aminocyclopropane-1-carboxylic acid (ACC), the precursor of ethylene, exhibited no significant disparity between the WT and dek407 mutant (Figure 8F). These results revealed that the phytohormone homeostasis was altered in the dek407 mutant.

2.7. Natural Variations in Dek407 Are Associated with Kernel Weight and Kernel Size

To analyze the potential association between natural variations in the Dek407 genomic region and traits associated with kernel weight and size, we performed a candidate gene association analysis using data from a panel of 503 maize genotypes. Six SNPs were observed to be significantly associated with kernel weight. Among these six SNPs, SNP25089804 and SNP25089818 were located in the 3′-untranslated region (3′-UTR), SNP25089963 (Ala/Thr), SNP25090099 (synonymous variant), and SNP25090645 (synonymous variant) were located in the fifth exon, and SNP25090842 was located in the fourth intron of Dek407 (Figure 9A). The available 372 maize inbred lines were classified into six haplotypes (Hap) based on the six SNPs (Figure 9B). Hap2hkw, Hap3hkw, Hap4hkw, and Hap5hkw exhibited no significant difference in their hundred-kernel weight (hkw), but all showed a significant reduction in their hundred-kernel weight compared with Hap1hkw (Figure 9C). The Hap6hkw exhibited no significant difference compared with any of the other haplotypes (Figure 9C). When Hap2hkw, Hap3hkw, Hap4hkw, and Hap5hkw were grouped together, the merged group still exhibited a significantly reduced hundred-kernel weight compared with that of Hap1hkw (Figure 9C). Therefore, Hap1hkw of Dek407 may be an excellent haplotype for kernel weight.
Three SNPs located in exon 4, SNP25090953 (Lys/Asn), SNP25090976 (Thr/Ala), and SNP25091118 (synonymous variant), were significantly associated with kernel thickness (kt) (Figure S6A). Based on these three SNPs, the available 448 maize inbreed lines were categorized into two haplotypes, Hap1kt and Hap2kt (Figure S6B). Hap2kt was considered to be an excellent haplotype for kernel thickness owing to its significant increase in kernel thickness compared with that of Hap1kt (Figure S6C). Eleven SNPs located in intron 2 (SNP25091853 and SNP25091855), exon 4 (SNP25090961 and SNP25091025), intron 4 (SNP25090842), and exon 5 (SNP25090099, SNP25090606, SNP25090645, SNP25090686, SNP25090738, and SNP25090788) were significantly associated with kernel width (kw) (Figure S7A). Among these SNPs, only SNP25090961 and SNP25090788 were missense variants, resulting in Gly/Ser and Phe/Leu mutations, respectively. Based on these 11 SNPs, the available 334 maize inbred lines were classified into seven haplotypes (Figure S7B). Hap1kw was considered to be an excellent haplotype owing to its significant increase in kernel width compared with that of Hap2kw and Hap3kw (Figure S7C).
Taken together, these results further suggested that natural variations in Dek407 are associated with kernel weight and kernel size.

3. Discussion

3.1. The Function of DEK407 Differs from Its Homologs in Rice and Arabidopsis

The primary mode of nitrogen absorption in plants is through the uptake of nitrates, which are subsequently transported via nitrate transporters, including NRT1/PTR family (NPF), NRT2 family, NRT3 family, CLC family, and SLAC family [21,22]. In Arabidopsis, there are 53 NRT1 members, and they all belong to the NPF family [24]. These NPFs are thought to be present in all organisms, and the functions of some of them have been well described. It has been reported that the number of NPF members in rice is as many as 93 [24]. In maize, a total of 78 NPF members belonging to eight subgroups have been identified [25].
According to the present results, there are four homologs of DEK407 in Arabidopsis and rice, namely AtNPF7.3, AtNPF7.2, OsPTR2, and OsNPF7.9 (Figure 5B). With regard to OsPTR2 and OsNPF7.9, the most closely related homologs in maize were identified as Zm00001d017666 and Zm00001d051637, excluding DEK407 (Figure S5). Based on the expression data accessible in the gramene database (https://ensembl.gramene.org) (accessed on 1 May 2022), OsPTR2 and Zm00001d017666 are highly expressed in the embryo, OsNPF7.9 and Zm00001d051637 exhibit high expression levels in the leaf, and only Dek407 is highly expressed exclusively in developing seeds. These results indicate that DEK407 exhibits functional differentiation from its two homologs in rice, although these proteins share a high amino acid sequence similarity. AtNPF7.3 was identified as the most closely related homolog of DEK407 in Arabidopsis. AtNPF7.3 is a low-affinity, pH-dependent, bidirectional nitrate transporter and is expressed in root pericycle cells close to the xylem. AtNPF7.3 participates in root xylem loading of nitrate by influencing the amount of nitrate transported from the root to the shoot, with no reported function in seed development [32]. This finding suggests that the function of DEK407 also differs from that of its homolog in Arabidopsis. The function of DEK407 could be special for seed development.

3.2. Reduction in Auxin in the dek407 Mutant May Be a New Cause for the Defective Kernel Phenotype

Auxin homeostasis plays a crucial role in the endosperm development of cereal crops. The processes of IAA biosynthesis, conjugation, oxidation, and transport can affect auxin homeostasis [49]. In rice, the expression levels of the IAA biosynthesis-related genes OsYUC1, OsYUC9, and OsYUC11 gradually increase during development of the endosperm, which is coupled with an increase in the IAA content in the grain [50]. Mutations in both OsYUC9 and OsYUC11 can result in defective grain filling [51]. IAA is the most abundant endogenous hormone in maize kernels, and it plays an important role in the entire process of kernel development. Mutation of ZmYUC1 leads to a decrease in the IAA content, abnormal differentiation of the BETL, defective endosperm filling, and a reduction in kernel size [52]. ZmEHD1 encodes a C-terminal Eps15 homology domain (EHD) protein, and it regulates auxin homeostasis by mediating clathrin-mediated endocytosis through its interaction with the ZmAP2s subunit [53]. In kernels of the ehd1 mutant, the content of IAA is significantly reduced compared with that of WTs, ultimately resulting in defective kernel development and vegetative growth [53].
In the transcriptome analysis of the dek407 mutant, DEGs associated with hormone responses were identified among the significantly enriched GO categories (Figure 6B). In addition, the plant hormone signal transduction pathway was also identified as a significantly enriched KEGG pathway (Figure 6C). A total of 34 DEGs were involved in IAA synthesis and signaling in the dek407 mutant (Figure 7A). However, the endogenous IAA content in the dek407 mutant was only 34% of that of the WT (Figure 8A). Thus, we predicted that the significant reduction in IAA synthesis in the dek407 mutant may be a new contributor to the defective kernel phenotype.

3.3. The Amino Acid Residues in Different TMs May Confer Their Specificity for Various Substrate Transportation

Compared with animals, higher plants possess a greater number of NPF genes. These NPF proteins are capable of transporting a variety of substrates, ions including Cl and K+, amino acids, sucrose, glucose, and hormones including IBA, IAA, ABA, JA, and GA [32,33,34,35]. However, the specific amino acid residues and domains of these NPFs that determine the specific binding of these different substrates are still unknown.
In Arabidopsis, AtNPF7.3 is a low-affinity, pH-dependent, bidirectional nitrate transporter [38]. A recent study reported that the AtNPF7.3 protein could also function as a transporter of IBA, a precursor of the endogenous auxin [34]. When expressed in yeast, AtNPF7.3 mediated cellular uptake of IBA. The loss-of-function mutant of AtNPF7.3 showed defective root gravitropism with reduced IBA contents and auxin responses. Nevertheless, the phenotype was restored through exogenous application of IAA but not through exogenous IBA treatment [34]. As the homologous protein of AtNPF7.3 in maize, ZmSUGCAR1 has been proposed to have the capacity to transport sucrose and glucose but not IAA. The premature stop mutation in its TM9 affects sugar accumulation in the endosperm of zmsugcar1-1 [35]. Our results showed that the G/A mutation in the third transmembrane domain (TM3) of ZmSUGCAR1 altered the expression of phytohormone-related genes and the phytohormone balance, especially the expression of IAA-related genes and the content of IAA. Then, it can be inferred that the TM3 of ZmSUGCAR1 likely plays a crucial role in facilitating IBA or IAA transport and maintaining intracellular hormone homeostasis in developing kernels.
Multidrug resistance protein 1 (MRP1) belongs to the ABC transporter superfamily. Mutagenesis studies of MRP1 have confirmed the importance of residues in different TMs in determining substrate specificity and overall activity [54]. As one of the typical members of the RND (resistance–nodulation–division) superfamily, the Escherichia coli OqxAB multidrug efflux pump confers resistance to antimicrobial agents, such as olaquindox and fluoroquinolone. A molecular docking analysis demonstrated that residues from different domains of OqxAB exhibit diverse functions in the transportation of various substrates [55]. In the zmsugcar1-1 mutant, the premature stop mutation in the TM9 of ZmSUGCAR1 resulted in the loss of the last three TMs but had no effect on the other TMs [35]. In the dek407 mutant, the G/A mutation in TM3 of ZmSUGCAR1 leads to an alteration of the corresponding coding amino acid from Cys to Tyr, subsequently altering the expression of phytohormone-related genes and the phytohormone balance, especially the expression of IAA-related genes and the content of IAA. This new finding, which differs from zmsugcar1-1, suggests that different amino acid residues or different TMs of ZmSUGCAR1 may be responsible for the specific recognition and interaction with different substrates. Whether ZmSUGCAR1 could transport IAA or its precursor IBA and cooperate with other proteins in vivo through specific amino acid residues remains unknown and is an attractive possibility. This intriguing hypothesis requires further investigation in future studies.

3.4. Dek407 Is a Potential Target for the Improvement of Maize Yield

The domestication of maize can be traced to approximately 9000 years ago in Southern Mexico. Through a long process of selection and domestication, the progenitor teosinte has been transformed into today’s high-yielding, high-quality, and stress-resistant maize varieties [56]. During this process, variations in the gene loci that contribute to yield and stress resistance were selected by breeders, and the frequency of these elite allelic variants were continuously enriched. A recent study has shown that natural variation in ZmUrb2 is associated with maize kernel size, and that inbred lines with haplotype 3 have significantly longer kernels than those with other haplotypes [15]. These identified loci or genes can serve as direct targets of genetic selection for maize trait improvement. In the present study, we determined that maize inbred lines with Hap1hkw of Dek407 had a significantly higher hundred-kernel weight than those lines with Hap2hkw, Hap3hkw, Hap4hkw, and Hap5hkw. Thus, we infer that an elite haplotype of Dek407 may be helpful to increase maize yield. Therefore, Dek407 is a potential target for the improvement of maize yield.

4. Materials and Methods

4.1. Plant Materials

The maize (Zea mays L.) mutant dek407 was isolated by screening the EMS mutagenesis library of the B73 inbred line [43]. The dek407 mutant was crossed with the Zheng58 and C01 genetic backgrounds to generate F2 populations for genetic analysis and gene mapping. All the plants were grown in the experimental field of Henan Agricultural University under natural conditions (Zhengzhou, China).

4.2. Light Microscopy of Cytological Sections

Wild-type and dek407 mutant kernels from the same heterozygous ear at 12 DAP were collected for light microscopy visualization of cytological sections. To prepare the paraffin sections of the kernels, 12 DAP seeds were fixed overnight at 4 °C in a formalin-acetic acid–alcohol (FAA) solution. The fixed materials were then dehydrated in an ethanol gradient series of 70%, 75%, 80%, 85%, 90%, 95%, and 100% ethanol for 2 h at each step. Afterwards, the samples were treated with xylene, embedded in paraffin wax via infiltration, and cut into 10–12 µm thick sections using a microtome (Leica RM2235, Heidelberg, Germany). The sections were stained with toluidine blue (Sinopharm Chemical Reagent Co., Ltd. Shanghai, China) and examined under a Leica M165FC stereomicroscope (Leica M165FC, Heidelberg, Germany).

4.3. Measurement of Hundred-Kernel Weight and Starch Content

The self-pollinated heterozygous ears were harvested from the field to isolate wild-type and mutant (dek407, dek407-1, and dek407-2) kernels and dried to a constant weight. Three ears were utilized as individual biological replicates. The weight of one hundred WT and mutant kernels from each ear was measured using a laboratory scale. For starch content analysis, the WT and dek407 seeds harvested at 45 DAP were dried at 80 °C to a constant weight and subsequently pulverized into a fine powder using a pestle and mortar. For each sample, 50 mg of flour was used to measure the starch content using a starch content assay kit (Beijing Solarbio Science & Technology Co., Ltd. BC0700, Beijing, China). All the measurements were conducted on three biological replicates.

4.4. Bulked Segregant RNA-Seq Analysis

The dek407 mutant was crossed with C01, and self-pollinated F2 populations were used for BSR-seq. WT and mutant kernels were collected from the same ear at 15 DAP. The pericarps were removed, and 30 kernels of either the WT or mutant phenotype were pooled for the BSR analysis. The total RNA was extracted from each pool using TransZol Plant (TransGen, ET121-01, Beijing, China) and subjected to Illumina sequencing in Hiseq2000 (Berry Genomics, Fuzhou, China). A BSR-seq analysis was performed following the method of Liu et al. [57]. In brief, the sequences were mapped against the B73 genome (AGPv4.32). Single-nucleotide polymorphism (SNP) callings were processed using SAMtools (https://www.htslib.org/ accessed on 1 May 2022). High-quality SNPs were selected for the SNP index analysis. Δ (SNP-index) was calculated by subtracting the SNP index of the mutant pool from the WT pool. A graph of the average values of Δ (SNP-index) was plotted using a 1 Mb window size and a 1 kb window step size.

4.5. Map-Based Cloning of Dek407

The F2 mapping population was derived from a cross between dek407 and the maize inbred line C01. Map-based cloning was performed using 288 mutant individuals from the F2 mapping population. For fine mapping, InDel and SSR molecular markers (Table S4) were developed to narrow the Dek407 locus to a 5.9 Mb region. To identify the mutation site, the candidate genes were amplified from the mutant and WT kernels using Phanta Max Super-Fidelidy DNA polymerase (Vazyme, P505, Nanjing, China) and were subsequently sequenced.

4.6. Sequence BLAST and Phylogenetic Tree Construction

The DEK407 protein was used as the query for BLAST searches against the Zea mays, Sorghum bicolor, Setaria italica, Oryza sativa, Arabidopsis thaliana, Glycine max, Hordeum vulgare, Prunus persica, Populus tremula, Poncirus trifoliata, and Paspalum vaginatum genomes on phytozome (https://phytozome.jgi.doe.gov) (accessed on 1 May 2022). A multiple sequence alignment was performed using the protein sequences encoded by 30 and 50 well-known NRT1/PTR family (NPF) genes in rice and Arabidopsis thaliana and their homologous in maize. These protein sequences were aligned using the MUSCLE algorithm implemented in MEGA 11 (https://www.megasoftware.net/ accessed on 1 May 2022). The phylogenetic tree was subsequently constructed based on the alignment results using the neighbor-joining (NJ) method implemented in MEGA 11. Bootstrapping was performed with 1000 replications.

4.7. RNA-Sequencing Analysis

The total RNA of three biological replicates was extracted from 15 DAP WT and dek407 kernels with the pericarp removed using TransZol Plant (TransGen, ET121-01, Beijing, China). Clean reads were obtained using the Illumina HiSeq X Ten platform (Berry Genomics, Fuzhou, China) and mapped to the B73 reference genome (RefGen_V4). The gene expression value was normalized as fragments per kilobase of transcript per million mapped reads (FPKM). The DEGs were produced using the threshold of a p-value < 0.05 in the DESeq software package (http://bioconductor.org/ accessed on 1 May 2022). GO enrichment of DEGs was performed using the DESeq R package based on the hypergeometric distribution.

4.8. Measurement of Endogenous Phytohormones

Approximately 2 g of WT and dek407 kernels was ground into fine powder in the presence of liquid nitrogen for the measurement of endogenous phytohormones. According to the methods of Xin et al., the powder was extracted using 1.0 mL of 80% methanol (v/v) at 4 °C for 12 h [58]. The supernatant was obtained after centrifugation at 10,000× g for 20 min at 4 °C. The solvent was evaporated under a gentle nitrogen stream at 35 °C and then re-dissolved in 40% methanol with a volume of 100 μL. The filtrate was filtered through a 0.22 µm filter membrane after shock mixing. Subsequently, the phytohormones were quantified using ultra-high-performance liquid chromatography–tandem mass spectrometry (UPLC-MS/MS) using a Xevo TQ-XS system (Waters, Milford, MA, USA). The flow rate of the mobile phase, composed of solvent A (0.1% formic acid, water) and solvent B (methanol), was set to 0.3 mL/min. The linear-gradient system was set as follows: 0–2 min, 2% B; 2–10 min, up to 80% B; 10–12 min, 80% B; 12–13 min, down to 2% B; 13–15 min, 2% B. The autosampler temperature was set to 4 °C, and the sample injection volume was 10 μL. Three biological replicates were prepared for each measurement.

4.9. Candidate Gene-Based Association Analysis

A total of 503 maize inbred lines were utilized for the candidate gene-based association analysis, with corresponding genotyping data of Dek407 and phenotypic data acquired from MaizeGo (http://www.maizego.org/Resources.html) (accessed on 1 May 2022). The SNP data were filtered based on a minor allele frequency (MAF) of >0.05 and a missing rate of <20%. The candidate gene-based association analysis was conducted using TASSEL 5.0. A linkage disequilibrium (LD) plot was generated using HAPLOVIEW software (https://sourceforge.net/projects/haploview/) (accessed on 1 May 2022). Differences in the phenotypic values and haplotypes were analyzed using a one-way ANOVA or Student’s t-tests.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms242417471/s1.

Author Contributions

Conceptualization, H.W., J.T. and W.-X.L.; methodology, H.W. and J.T.; software, Q.D. and P.Y.; validation, X.Y., J.X., X.M., X.L. and H.L.; formal analysis, G.L.; investigation, Z.F.; resources, H.W.; data curation, H.W. and J.T.; writing—original draft preparation, H.W. and X.Y.; writing—review and editing, J.T. and W.-X.L.; visualization, X.Y.; supervision, J.T.; project administration, H.W. and J.T.; funding acquisition, H.W., G.L. and Z.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (31901562 to H.W. and 32101783 to G.L.), the Key Technologies R&D Program of Henan Province (212102110043 to H.W.), the Key Scientific Research Project of Higher Education of Henan Province (22A210007 to H.W.), the Youth Talent Lifting Project of Henan Province (2022HYTP032 to H.W.), the China Postdoctoral Science Foundation (2018M640211 to H.W.), and the Open Project of Co-construction State Key Laboratory of Wheat and Maize Crop Science (SKL2021KF08 to Z.F.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article and the Supplementary Materials.

Acknowledgments

We would like to acknowledge Haiying Guan of the Shandong Academy of Agricultural Sciences for providing us with the mn2 allele mutant. We extend our appreciation to the anonymous reviewers for their valuable suggestions to help improve this article. We are also grateful to Chunyi Zhang from the Institute of Biotechnology of the Chinese Academy of Agricultural Sciences and Xiaoduo Lu from Qilu Normal University for providing us with mutant materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shiferaw, B.; Prasanna, B.M.; Hellin, J.; Bänziger, M. Crops that feed the world 6. Past successes and future challenges to the role played by maize in global food security. Food Sec. 2011, 3, 307–327. [Google Scholar] [CrossRef]
  2. Neuffer, M.G.; Sheridan, W.F. Defective Kernel Mutants of Maize. I. Genetic and Lethality Studies. Genetics 1980, 95, 929–944. [Google Scholar] [CrossRef]
  3. Scanlon, M.J.; Stinard, P.S.; James, M.G.; Myers, A.M.; Robertson, D.S. Genetic analysis of 63 mutations affecting maize kernel development isolated from Mutator stocks. Genetics 1994, 136, 281–294. [Google Scholar] [CrossRef]
  4. Clark, J.K.; Sheridan, W.F. Isolation and Characterization of 51 embryo-specific Mutations of Maize. Plant Cell 1991, 3, 935–951. [Google Scholar] [CrossRef]
  5. Fan, K.; Peng, Y.; Ren, Z.; Li, D.; Zhen, S.; Hey, S.; Cui, Y.; Fu, J.; Gu, R.; Wang, J.; et al. Maize Defective Kernel605 Encodes a Canonical DYW-Type PPR Protein that Edits a Conserved Site of nad1 and Is Essential for Seed Nutritional Quality. Plant Cell Physiol. 2020, 61, 1954–1966. [Google Scholar] [CrossRef]
  6. Sosso, D.; Mbelo, S.; Vernoud, V.; Gendrot, G.; Dedieu, A.; Chambrier, P.; Dauzat, M.; Heurtevin, L.; Guyon, V.; Takenaka, M.; et al. PPR2263, a DYW-Subgroup Pentatricopeptide repeat protein, is required for mitochondrial nad5 and cob transcript editing, mitochondrion biogenesis, and maize growth. Plant Cell 2012, 24, 676–691. [Google Scholar] [CrossRef]
  7. Li, X.-J.; Zhang, Y.-F.; Hou, M.; Sun, F.; Shen, Y.; Xiu, Z.-H.; Wang, X.; Chen, Z.-L.; Sun, S.S.M.; Small, I.; et al. Small kernel 1 encodes a pentatricopeptide repeat protein required for mitochondrial nad7 transcript editing and seed development in maize (Zea mays) and rice (Oryza sativa). Plant J. 2014, 79, 797–809. [Google Scholar] [CrossRef]
  8. Wang, Y.; Li, H.; Huang, Z.Q.; Ma, B.; Yang, Y.Z.; Xiu, Z.H.; Wang, L.; Tan, B.C. Maize PPR-E proteins mediate RNA C-to-U editing in mitochondria by recruiting the trans deaminase PCW1. Plant Cell 2023, 35, 529–551. [Google Scholar] [CrossRef]
  9. Sun, F.; Zhang, X.; Shen, Y.; Wang, H.; Liu, R.; Wang, X.; Gao, D.; Yang, Y.Z.; Liu, Y.; Tan, B.C. The pentatricopeptide repeat protein EMPTY PERICARP8 is required for the splicing of three mitochondrial introns and seed development in maize. Plant J. 2018, 95, 919–932. [Google Scholar] [CrossRef]
  10. Cai, M.; Li, S.; Sun, F.; Sun, Q.; Zhao, H.; Ren, X.; Zhao, Y.; Tan, B.C.; Zhang, Z.; Qiu, F. Emp10 encodes a mitochondrial PPR protein that affects the cis-splicing of nad2 intron 1 and seed development in maize. Plant J. 2017, 91, 132–144. [Google Scholar] [CrossRef]
  11. Yang, Y.Z.; Ding, S.; Wang, Y.; Li, C.L.; Shen, Y.; Meeley, R.; McCarty, D.R.; Tan, B.C. Small kernel2 Encodes a Glutaminase in Vitamin B6 Biosynthesis Essential for Maize Seed Development. Plant Physiol. 2017, 174, 1127–1138. [Google Scholar] [CrossRef]
  12. Vilhar, B.; Kladnik, A.; Blejec, A.; Chourey, P.S.; Dermastia, M. Cytometrical evidence that the loss of seed weight in the miniature1 seed mutant of maize is associated with reduced mitotic activity in the developing endosperm. Plant Physiol. 2002, 129, 23–30. [Google Scholar] [CrossRef]
  13. Gutiérrez-Marcos, J.F.; Costa, L.M.; Biderre-Petit, C.; Khbaya, B.; O’Sullivan, D.M.; Wormald, M.; Perez, P.; Dickinson, H.G. maternally expressed gene1 Is a novel maize endosperm transfer cell-specific gene with a maternal parent-of-origin pattern of expression. Plant Cell 2004, 16, 1288–1301. [Google Scholar] [CrossRef]
  14. Wang, G.; Zhang, J.; Wang, G.; Fan, X.; Sun, X.; Qin, H.; Xu, N.; Zhong, M.; Qiao, Z.; Tang, Y.; et al. Proline responding1 Plays a Critical Role in Regulating General Protein Synthesis and the Cell Cycle in Maize. Plant Cell 2014, 26, 2582–2600. [Google Scholar] [CrossRef]
  15. Wang, H.; Wang, K.; Du, Q.; Wang, Y.; Fu, Z.; Guo, Z.; Kang, D.; Li, W.X.; Tang, J. Maize Urb2 protein is required for kernel development and vegetative growth by affecting pre-ribosomal RNA processing. New Phytol. 2018, 218, 1233–1246. [Google Scholar] [CrossRef]
  16. Liu, H.; Xiu, Z.; Yang, H.; Ma, Z.; Yang, D.; Wang, H.; Tan, B.C. Maize Shrek1 encodes a WD40 protein that regulates pre-rRNA processing in ribosome biogenesis. Plant Cell 2022, 34, 4028–4044. [Google Scholar] [CrossRef]
  17. Qi, W.; Zhu, J.; Wu, Q.; Wang, Q.; Li, X.; Yao, D.; Jin, Y.; Wang, G.; Wang, G.; Song, R. Maize reas1 Mutant Stimulates Ribosome Use Efficiency and Triggers Distinct Transcriptional and Translational Responses. Plant Physiol. 2016, 170, 971–988. [Google Scholar] [CrossRef]
  18. Wang, T.; Chang, Y.; Zhao, K.; Dong, Q.; Yang, J. Maize RNA 3′-terminal phosphate cyclase-like protein promotes 18S pre-rRNA cleavage and is important for kernel development. Plant Cell 2022, 34, 1957–1979. [Google Scholar] [CrossRef]
  19. Massel, K.; Campbell, B.C.; Mace, E.S.; Tai, S.; Tao, Y.; Worland, B.G.; Jordan, D.R.; Botella, J.R.; Godwin, I.D. Whole Genome Sequencing Reveals Potential New Targets for Improving Nitrogen Uptake and Utilization in Sorghum bicolor. Front. Plant Sci. 2016, 7, 1544. [Google Scholar] [CrossRef]
  20. O’Brien, J.A.; Vega, A.; Bouguyon, E.; Krouk, G.; Gojon, A.; Coruzzi, G.; Gutiérrez, R.A. Nitrate Transport, Sensing, and Responses in Plants. Mol. Plant 2016, 9, 837–856. [Google Scholar] [CrossRef]
  21. Vidal, E.A.; Alvarez, J.M.; Araus, V.; Riveras, E.; Brooks, M.D.; Krouk, G.; Ruffel, S.; Lejay, L.; Crawford, N.M.; Coruzzi, G.M.; et al. Nitrate in 2020: Thirty Years from Transport to Signaling Networks. Plant Cell 2020, 32, 2094–2119. [Google Scholar] [CrossRef]
  22. Fan, X.; Naz, M.; Fan, X.; Xuan, W.; Miller, A.J.; Xu, G. Plant nitrate transporters: From gene function to application. J. Exp. Bot. 2017, 68, 2463–2475. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, Y.Y.; Cheng, Y.H.; Chen, K.E.; Tsay, Y.F. Nitrate Transport, Signaling, and Use Efficiency. Annu. Rev. Plant Biol. 2018, 69, 85–122. [Google Scholar] [CrossRef]
  24. Léran, 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]
  25. Jia, L.; Hu, D.; Wang, J.; Liang, Y.; Li, F.; Wang, Y.; Han, Y. Genome-Wide Identification and Functional Analysis of Nitrate Transporter Genes (NPF, NRT2 and NRT3) in Maize. Int. J. Mol. Sci. 2023, 24, 12941. [Google Scholar] [CrossRef]
  26. 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] [PubMed]
  27. 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]
  28. Ho, C.H.; Lin, S.H.; Hu, H.C.; Tsay, Y.F. CHL1 functions as a nitrate sensor in plants. Cell 2009, 138, 1184–1194. [Google Scholar] [CrossRef]
  29. 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]
  30. Li, Y.; Ouyang, J.; Wang, Y.Y.; Hu, R.; Xia, K.; Duan, J.; Wang, Y.; Tsay, Y.F.; Zhang, M. Disruption of the rice nitrate transporter OsNPF2.2 hinders root-to-shoot nitrate transport and vascular development. Sci. Rep. 2015, 5, 9635. [Google Scholar] [CrossRef]
  31. 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]
  32. Corratgé-Faillie, C.; Lacombe, B. Substrate (un)specificity of Arabidopsis NRT1/PTR FAMILY (NPF) proteins. J. Exp. Bot. 2017, 68, 3107–3113. [Google Scholar] [CrossRef]
  33. Chiba, Y.; Shimizu, T.; Miyakawa, S.; Kanno, Y.; Koshiba, T.; Kamiya, Y.; Seo, M. Identification of Arabidopsis thaliana NRT1/PTR FAMILY (NPF) proteins capable of transporting plant hormones. J. Plant Res. 2015, 128, 679–686. [Google Scholar] [CrossRef]
  34. Watanabe, S.; Takahashi, N.; Kanno, Y.; Suzuki, H.; Aoi, Y.; Takeda-Kamiya, N.; Toyooka, K.; Kasahara, H.; Hayashi, K.I.; Umeda, M.; et al. The Arabidopsis NRT1/PTR FAMILY protein NPF7.3/NRT1.5 is an indole-3-butyric acid transporter involved in root gravitropism. Proc. Natl. Acad. Sci. USA 2020, 117, 31500–31509. [Google Scholar] [CrossRef]
  35. 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]
  36. Li, B.; Byrt, C.; Qiu, J.; Baumann, U.; Hrmova, M.; Evrard, A.; Johnson, A.A.; Birnbaum, K.D.; Mayo, G.M.; Jha, D.; et al. Identification of a Stelar-Localized Transport Protein That Facilitates Root-to-Shoot Transfer of Chloride in Arabidopsis. Plant Physiol. 2016, 170, 1014–1029. [Google Scholar] [CrossRef]
  37. Li, B.; Qiu, J.; Jayakannan, M.; Xu, B.; Li, Y.; Mayo, G.M.; Tester, M.; Gilliham, M.; Roy, S.J. AtNPF2.5 Modulates Chloride (Cl) Efflux from Roots of Arabidopsis thaliana. Front. Plant Sci. 2017, 7, 2013. [Google Scholar] [CrossRef]
  38. 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]
  39. Tal, I.; Zhang, Y.; Jørgensen, 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]
  40. Zhang, L.; Yu, Z.; Xu, Y.; Yu, M.; Ren, Y.; Zhang, S.; Yang, G.; Huang, J.; Yan, K.; Zheng, C.; et al. Regulation of the stability and ABA import activity of NRT1.2/NPF4.6 by CEPR2-mediated phosphorylation in Arabidopsis. Mol. Plant 2021, 14, 633–646. [Google Scholar] [CrossRef]
  41. Guan, H.Y.; Dong, Y.B.; Shou-Ping, L.U.; Liu, T.S.; Wang, L.M. Characterization and map-based cloning of miniature2-m1, a gene controlling kernel size in maize. J. Integr. Agric. 2020, 19, 1961–1973. [Google Scholar] [CrossRef]
  42. 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]
  43. Lu, X.; Liu, J.; Ren, W.; Yang, Q.; Chai, Z.; Chen, R.; Wang, L.; Zhao, J.; Lang, Z.; Wang, H.; et al. Gene-Indexed Mutations in Maize. Mol. Plant 2018, 11, 496–504. [Google Scholar] [CrossRef] [PubMed]
  44. Stelpflug, S.C.; Sekhon, R.S.; Vaillancourt, B.; Hirsch, C.N.; Buell, C.R.; de Leon, N.; Kaeppler, S.M. An Expanded Maize Gene Expression Atlas based on RNA Sequencing and its Use to Explore Root Development. Plant Genome. 2016, 9, 1–16. [Google Scholar] [CrossRef] [PubMed]
  45. Finkelstein, R.R. Plant Hormones; Springer: Berlin/Heidelberg, Germany, 2010; pp. 549–550. [Google Scholar]
  46. Ali, F.; Qanmber, G.; Li, F.; Wang, Z. Updated role of ABA in seed maturation, dormancy, and germination. J. Adv. Res. 2021, 35, 199–214. [Google Scholar] [CrossRef] [PubMed]
  47. Locascio, A.; Roig-Villanova, I.; Bernardi, J.; Varotto, S. Current perspectives on the hormonal control of seed development in Arabidopsis and maize: A focus on auxin. Front. Plant Sci. 2014, 5, 412. [Google Scholar] [CrossRef]
  48. Liu, J.; Wu, M.W.; Liu, C.M. Cereal Endosperms: Development and Storage Product Accumulation. Annu. Rev. Plant Biol. 2022, 73, 255–291. [Google Scholar] [CrossRef]
  49. Basunia, M.A.; Nonhebel, H.M. Hormonal regulation of cereal endosperm development with a focus on rice (Oryza sativa). Funct. Plant Biol. 2019, 46, 493–506. [Google Scholar] [CrossRef]
  50. Abu-Zaitoon, Y.M.; Bennett, K.; Normanly, J.; Nonhebel, H.M. A large increase in IAA during development of rice grains correlates with the expression of tryptophan aminotransferase OsTAR1 and a grain-specific YUCCA. Plant Physiol. 2012, 146, 487–499. [Google Scholar] [CrossRef]
  51. Xu, X.; E, Z.; Zhang, D.; Yun, Q.; Zhou, Y.; Niu, B.; Chen, C. OsYUC11-mediated auxin biosynthesis is essential for endosperm development of rice. Plant Physiol. 2021, 185, 934–950. [Google Scholar] [CrossRef]
  52. Bernardi, J.; Lanubile, A.; Li, Q.B.; Kumar, D.; Kladnik, A.; Cook, S.D.; Ross, J.J.; Marocco, A.; Chourey, P.S. Impaired auxin biosynthesis in the defective endosperm18 mutant is due to mutational loss of expression in the ZmYuc1 gene encoding endosperm-specific YUCCA1 protein in maize. Plant Physiol. 2012, 160, 1318–1328. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, Y.; Liu, W.; Wang, H.; Du, Q.; Fu, Z.; Li, W.X.; Tang, J. ZmEHD1 Is Required for Kernel Development and Vegetative Growth through Regulating Auxin Homeostasis. Plant Physiol. 2020, 182, 1467–1480. [Google Scholar] [CrossRef] [PubMed]
  54. Deeley, R.G.; Cole, S.P. Substrate recognition and transport by multidrug resistance protein 1 (ABCC1). FEBS Lett. 2006, 580, 1103–1111. [Google Scholar] [CrossRef] [PubMed]
  55. Xu, S.; Chen, G.; Liu, Z.; Xu, D.; Wu, Z.; Li, Z.; Hong, M. Site-Directed Mutagenesis Reveals Crucial Residues in Escherichia coli Resistance-Nodulation-Division Efflux Pump OqxB. Microb. Drug Resist. 2020, 26, 550–560. [Google Scholar] [CrossRef]
  56. Doebley, J.; Gaut, B.S.; Smith, D.R. The biology and evolution of maize. Annu. Rev. Genet. 2004, 38, 207–223. [Google Scholar] [CrossRef]
  57. Liu, X.; Bi, B.; Xu, X.; Li, B.; Tian, S.; Wang, J.; Zhang, H.; Wang, G.; Han, Y.; McElroy, J.S. Rapid identification of a candidate nicosulfuron sensitivity gene (Nss) in maize (Zea mays L.) via combining bulked segregant analysis and RNA-seq. Theor. Appl. Genet. 2019, 132, 1351–1361. [Google Scholar] [CrossRef]
  58. Xin, P.; Guo, Q.; Li, B.; Cheng, S.; Yan, J.; Chu, J. A Tailored High-Efficiency Sample Pretreatment Method for Simultaneous Quantification of 10 Classes of Known Endogenous Phytohormones. Plant Commun. 2020, 1, 100047. [Google Scholar] [CrossRef]
Figure 1. Phenotypic features of the maize dek407 mutant. (A) Self-pollinated dek407 heterozygote ears at 12 days after pollination (DAP). The red arrows indicate the dek407 mutant kernels. Bar = 2 cm. (B) Mature dek407 heterozygote ears. The red arrows indicate the dek407 mutant kernels. Bar = 2 cm. (C) The kernel length, kernel width, and kernel thickness of WT and dek407 mature kernels, which were randomly selected from the heterozygote ears. Bar = 1 cm. (D) Phenotypes of WT and dek407 seedlings at 10 days after sowing (DAS). Bar = 5 cm. (E) Hundred-kernel weight of the WT and dek407 mutant. Error bars indicate SE. ***, p < 0.001 (t-test), significant difference from the WT. (F) WT and dek407 kernels viewed on a light box. Bar = 1 cm. (G) Micrograph of longitudinal sections of WT and dek407 kernels stained with toluidine blue at 12 DAP. Bar = 1 mm.
Figure 1. Phenotypic features of the maize dek407 mutant. (A) Self-pollinated dek407 heterozygote ears at 12 days after pollination (DAP). The red arrows indicate the dek407 mutant kernels. Bar = 2 cm. (B) Mature dek407 heterozygote ears. The red arrows indicate the dek407 mutant kernels. Bar = 2 cm. (C) The kernel length, kernel width, and kernel thickness of WT and dek407 mature kernels, which were randomly selected from the heterozygote ears. Bar = 1 cm. (D) Phenotypes of WT and dek407 seedlings at 10 days after sowing (DAS). Bar = 5 cm. (E) Hundred-kernel weight of the WT and dek407 mutant. Error bars indicate SE. ***, p < 0.001 (t-test), significant difference from the WT. (F) WT and dek407 kernels viewed on a light box. Bar = 1 cm. (G) Micrograph of longitudinal sections of WT and dek407 kernels stained with toluidine blue at 12 DAP. Bar = 1 mm.
Ijms 24 17471 g001
Figure 2. Phenotypic features of maize dek407 mutants in Zheng58 (dek407-1) and C01 (dek407-2) backgrounds. (A) Mature F2 ear of dek407 × Zheng58. The red arrows indicate the mutant kernels (dek407-1). Bar = 1 cm. (B) Longitudinal and transection sections of WT and dek407-1 kernels. The part circled by the red dotted line represents the embryo. Bar = 2 mm. (C) WT and dek407-1 kernels viewed on a light box. Bar = 1 cm. (D) Phenotypes of WT and dek407-1 seedlings at 10 DAS. Bar = 5 cm. (E) Mature F2 ear of dek407 × C01. The red arrows indicate the mutant kernels (dek407-2). Bar = 1 cm. (F) Longitudinal and transection sections of WT and dek407-2 kernels. The part circled by the red dotted line represents the embryo. Bar = 2 mm. (G) WT and dek407-2 kernels viewed on a light box. Bar = 1 cm. (H) Phenotypes of WT and dek407-2 seedlings at 10 DAS. Bar = 5 cm.
Figure 2. Phenotypic features of maize dek407 mutants in Zheng58 (dek407-1) and C01 (dek407-2) backgrounds. (A) Mature F2 ear of dek407 × Zheng58. The red arrows indicate the mutant kernels (dek407-1). Bar = 1 cm. (B) Longitudinal and transection sections of WT and dek407-1 kernels. The part circled by the red dotted line represents the embryo. Bar = 2 mm. (C) WT and dek407-1 kernels viewed on a light box. Bar = 1 cm. (D) Phenotypes of WT and dek407-1 seedlings at 10 DAS. Bar = 5 cm. (E) Mature F2 ear of dek407 × C01. The red arrows indicate the mutant kernels (dek407-2). Bar = 1 cm. (F) Longitudinal and transection sections of WT and dek407-2 kernels. The part circled by the red dotted line represents the embryo. Bar = 2 mm. (G) WT and dek407-2 kernels viewed on a light box. Bar = 1 cm. (H) Phenotypes of WT and dek407-2 seedlings at 10 DAS. Bar = 5 cm.
Ijms 24 17471 g002
Figure 3. Cloning and identification of Dek407. (A) Frequency variance analysis result of bulk segregant RNA-seq (BSR-seq). The 10 chromosomes are represented with different colors, and the candidate region is indicated by a red arrow. The red line indicates the threshold. (B) The dek407 locus was mapped to a 5.84 Mb region on chromosome 7 between the SSR marker umc2337 and the InDel marker S12341. (C) Schematic diagram of the dek407 and mn2-m1 coding regions. Black boxes indicate exons. (D) Analysis of dek407 mutant site in other maize inbred lines. The red rectangle indicates the mutant site in other maize inbred lines. Red star indicates mutant site in dek407.
Figure 3. Cloning and identification of Dek407. (A) Frequency variance analysis result of bulk segregant RNA-seq (BSR-seq). The 10 chromosomes are represented with different colors, and the candidate region is indicated by a red arrow. The red line indicates the threshold. (B) The dek407 locus was mapped to a 5.84 Mb region on chromosome 7 between the SSR marker umc2337 and the InDel marker S12341. (C) Schematic diagram of the dek407 and mn2-m1 coding regions. Black boxes indicate exons. (D) Analysis of dek407 mutant site in other maize inbred lines. The red rectangle indicates the mutant site in other maize inbred lines. Red star indicates mutant site in dek407.
Ijms 24 17471 g003
Figure 4. Allelism test of dek407 with mn2-m1. (A) Allelism test ear of heterozygous dek407 × mn2-m1. The red arrows indicate the mutant kernels. Bar = 1 cm. (B) Sequence analysis of the mutant kernels from the allelism test ear. The locations of the dek407 and mn2-m1 mutant sites are shown in the red box. C/T indicates the mutant site in dek407 and G/− indicates the mutant site in mn2-m1. (C) Allelism test ear of heterozygous mn2-m1 × dek407. The red arrows indicate the mutant kernels. Bar = 1 cm. (D) Sequence analysis of the mutant kernels from the allelism test ear. The locations of the dek407 and mn2-m1 mutant sites are shown in the red box.
Figure 4. Allelism test of dek407 with mn2-m1. (A) Allelism test ear of heterozygous dek407 × mn2-m1. The red arrows indicate the mutant kernels. Bar = 1 cm. (B) Sequence analysis of the mutant kernels from the allelism test ear. The locations of the dek407 and mn2-m1 mutant sites are shown in the red box. C/T indicates the mutant site in dek407 and G/− indicates the mutant site in mn2-m1. (C) Allelism test ear of heterozygous mn2-m1 × dek407. The red arrows indicate the mutant kernels. Bar = 1 cm. (D) Sequence analysis of the mutant kernels from the allelism test ear. The locations of the dek407 and mn2-m1 mutant sites are shown in the red box.
Ijms 24 17471 g004
Figure 5. Phylogenetic relationship and expression pattern of DEK407. (A) The predicted major facilitator superfamily (MFS) domain of the DEK407 protein. (B) Phylogenetic tree of DEK407 and its homologs in different species. The bold represents the DEK407 protein in maize. (C) Expression levels of Dek407 in different maize organs from qTeller platform [44].
Figure 5. Phylogenetic relationship and expression pattern of DEK407. (A) The predicted major facilitator superfamily (MFS) domain of the DEK407 protein. (B) Phylogenetic tree of DEK407 and its homologs in different species. The bold represents the DEK407 protein in maize. (C) Expression levels of Dek407 in different maize organs from qTeller platform [44].
Ijms 24 17471 g005
Figure 6. GO and KEGG pathway enrichment analysis of DEGs between WT and dek407. (A) Volcano plot used to visualize RNA-seq data. Each point corresponds to a DEG. Red and blue dots represent up-regulated and down-regulated genes in dek407, respectively. Black dots represent non DEGs. (B) Enriched GO terms of the DEGs. (C) Enriched KEGG pathways of the DEGs. The size of the circles denotes the number of genes, and the color of the circles denotes the range of the p-value.
Figure 6. GO and KEGG pathway enrichment analysis of DEGs between WT and dek407. (A) Volcano plot used to visualize RNA-seq data. Each point corresponds to a DEG. Red and blue dots represent up-regulated and down-regulated genes in dek407, respectively. Black dots represent non DEGs. (B) Enriched GO terms of the DEGs. (C) Enriched KEGG pathways of the DEGs. The size of the circles denotes the number of genes, and the color of the circles denotes the range of the p-value.
Ijms 24 17471 g006
Figure 7. Heat map of plant hormone-related DEGs in WT and dek407. (A) IAA synthesis and signaling related DEGs. (B) Ethylene synthesis and signaling related DEGs. (C) BR synthesis and signaling related DEGs. (D) ABA signaling related DEGs. The red and blue box at the top of the heat map represent the three repetitions of the WT and dek407, respectively. The colors denote the range of the Z-Score.
Figure 7. Heat map of plant hormone-related DEGs in WT and dek407. (A) IAA synthesis and signaling related DEGs. (B) Ethylene synthesis and signaling related DEGs. (C) BR synthesis and signaling related DEGs. (D) ABA signaling related DEGs. The red and blue box at the top of the heat map represent the three repetitions of the WT and dek407, respectively. The colors denote the range of the Z-Score.
Ijms 24 17471 g007
Figure 8. Measurement of endogenous hormones in WT and dek407 kernels. (A) IAA (indole-3-acetic acid) content in WT and dek407 kernels. (B) ABA (abscisic acid) content in WT and dek407 kernels. (C) SA (salicylic acid) content in WT and dek407 kernels. (D) TZ (trans-zeatin) content in WT and dek407 kernels. (E) TZR (trans-zeatin riboside) content in WT and dek407 kernels. (F) ACC (aminocyclopropane-1-carboxylic acid) content in WT and dek407 kernels. Data are means ± SD (*** p < 0.001; * p < 0.05; n.s., not significant; Student’s t-test).
Figure 8. Measurement of endogenous hormones in WT and dek407 kernels. (A) IAA (indole-3-acetic acid) content in WT and dek407 kernels. (B) ABA (abscisic acid) content in WT and dek407 kernels. (C) SA (salicylic acid) content in WT and dek407 kernels. (D) TZ (trans-zeatin) content in WT and dek407 kernels. (E) TZR (trans-zeatin riboside) content in WT and dek407 kernels. (F) ACC (aminocyclopropane-1-carboxylic acid) content in WT and dek407 kernels. Data are means ± SD (*** p < 0.001; * p < 0.05; n.s., not significant; Student’s t-test).
Ijms 24 17471 g008
Figure 9. Natural variations in Dek407 are significantly associated with maize hundred-kernel weight. (A) Dek407-based genetic variations associated with maize hundred-kernel weight and pairwise linkage disequilibrium (LD) analysis of the variations. Blue dots represent SNPs that are not significantly associated with hundred-kernel weight.(B) Haplotypes of Dek407 hundred-kernel weight among these natural variations. Small letters hkw represents hundred-kernel weight. (C) Statistical analysis of hundred-kernel weight in the six haplotypes. Small letters hkw represents hundred-kernel weight. Means with the same small letter are not significantly different at p < 0.05 according to the LSD test.
Figure 9. Natural variations in Dek407 are significantly associated with maize hundred-kernel weight. (A) Dek407-based genetic variations associated with maize hundred-kernel weight and pairwise linkage disequilibrium (LD) analysis of the variations. Blue dots represent SNPs that are not significantly associated with hundred-kernel weight.(B) Haplotypes of Dek407 hundred-kernel weight among these natural variations. Small letters hkw represents hundred-kernel weight. (C) Statistical analysis of hundred-kernel weight in the six haplotypes. Small letters hkw represents hundred-kernel weight. Means with the same small letter are not significantly different at p < 0.05 according to the LSD test.
Ijms 24 17471 g009
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

Wang, H.; Yan, X.; Du, Q.; Yan, P.; Xi, J.; Meng, X.; Li, X.; Liu, H.; Liu, G.; Fu, Z.; et al. Maize Dek407 Encodes the Nitrate Transporter 1.5 and Is Required for Kernel Development. Int. J. Mol. Sci. 2023, 24, 17471. https://doi.org/10.3390/ijms242417471

AMA Style

Wang H, Yan X, Du Q, Yan P, Xi J, Meng X, Li X, Liu H, Liu G, Fu Z, et al. Maize Dek407 Encodes the Nitrate Transporter 1.5 and Is Required for Kernel Development. International Journal of Molecular Sciences. 2023; 24(24):17471. https://doi.org/10.3390/ijms242417471

Chicago/Turabian Style

Wang, Hongqiu, Xiaolan Yan, Qingguo Du, Pengshuai Yan, Jinjin Xi, Xiaoruo Meng, Xuguang Li, Huijian Liu, Guoqin Liu, Zhongjun Fu, and et al. 2023. "Maize Dek407 Encodes the Nitrate Transporter 1.5 and Is Required for Kernel Development" International Journal of Molecular Sciences 24, no. 24: 17471. https://doi.org/10.3390/ijms242417471

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